This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

Expanding bipartite Bell inequalities for maximum multi-partite randomness

Lewis Wooltorton lewis.wooltorton@york.ac.uk Department of Mathematics, University of York, Heslington, York, YO10 5DD, United Kingdom Quantum Engineering Centre for Doctoral Training, H. H. Wills Physics Laboratory and Department of Electrical & Electronic Engineering, University of Bristol, Bristol BS8 1FD, United Kingdom    Peter Brown peter.brown@telecom-paris.fr Télécom Paris, LTCI, Institut Polytechnique de Paris, 19 Place Marguerite Perey, 91120 Palaiseau, France    Roger Colbeck roger.colbeck@york.ac.uk Department of Mathematics, University of York, Heslington, York, YO10 5DD, United Kingdom
(12th12^{\text{th}} December 2024)
Abstract

Nonlocal tests on multi-partite quantum correlations form the basis of protocols that certify randomness in a device-independent (DI) way. Such correlations admit a rich structure, making the task of choosing an appropriate test difficult. For example, extremal Bell inequalities are tight witnesses of nonlocality, but achieving their maximum violation places constraints on the underlying quantum system, which can reduce the rate of randomness generation. As a result there is often a trade-off between maximum randomness and the amount of violation of a given Bell inequality. Here, we explore this trade-off for more than two parties. More precisely, we study the maximum amount of randomness that can be certified by correlations exhibiting a violation of the Mermin-Ardehali-Belinskii-Klyshko (MABK) inequality. We find that maximum quantum violation and maximum randomness are incompatible for any even number of parties, with incompatibility diminishing as the number of parties grows, and conjecture the precise trade-off. We also show that maximum MABK violation is not necessary for maximum randomness for odd numbers of parties. To obtain our results, we derive new families of Bell inequalities certifying maximum randomness from a technique for randomness certification, which we call “expanding Bell inequalities”. Our technique allows a bipartite Bell expression to be used as a seed, and transformed into a multi-partite Bell inequality tailored for randomness certification, showing how intuition learned in the bipartite case can find use in more complex scenarios.

\twocolumngrid

1 Introduction

Distant measurements on a shared quantum system can display correlations inaccessible to any locally realistic model [1, 2]. Such correlations are termed nonlocal and provide a device-independent (DI) witness of useful quantum characteristics, opening up a new paradigm of information processing [3]. Tasks such as randomness expansion [4, 5, 6, 7, 8], amplification [9] and key distribution [10, 11, 12, 13, 14, 15] can all be achieved in the DI regime, where security is proven based on the observed correlations and without trusting the inner workings of the devices. Moreover, nonlocal correlations can be used to prove the presence of a particular state or sets of measurements, known as self-testing [16, 17, 18, 19, 20, 21].

Certification of DI randomness typically uses a Bell inequality, and different Bell inequalities give different amounts of certified randomness. It is natural to consider extremal Bell inequalities, which form the minimal set separating local and non-local correlations. The set of correlations compatible with high violation is tightly constrained and can negatively affect the certification of randomness. For example, in the simplest bipartite scenario, correlations with higher violation of the only extremal Bell inequality, the Clauser-Horne-Shimony-Holt (CHSH) inequality, cannot certify maximal DI randomness, while those with a lower violation can [22, 23]. Further complications arise when scenarios with more inputs, outputs or parties are considered because the number of different classes of extremal Bell inequalities grows and it becomes more difficult to navigate the trade-offs between extremal Bell violation and maximum certifiable randomness. Alongside foundational interest, this question is also motivated practically, in finding more robust device-independent randomness expansion (DIRE) protocols [23].

In this work we study the family of multi-partite Bell inequalities due to Mermin, Ardehali, Belinskii, and Klyshko (MABK) [24, 25, 26] and their application to DIRE. These all have two inputs and two outputs per party. The one and two sided randomness given a tripartite MABK violation was bounded in [27, 28], and compared to other tripartite inequalities in [29]. Considering global randomness, reference [30] gave a bipartite construction for certification of MM bits of randomness using MM copies of almost unentangled states.

Reference [31] showed how, for an odd number of parties NN, maximum MABK violation can certify NN bits of DI randomness, whilst reference [32] provided an alternative family of Bell inequalities certifying NN bits in the even case. This leaves the open the following questions: is MABK suitable for maximum randomness generation when NN is even, and, if not, what is the compatibility trade-off between MABK violation and maximum randomness for arbitrary NN? Are there correlations with lower MABK that achieve maximal randomness when NN is odd?

To analyse multi-partite nonlocality, we require tools to construct multi-partite Bell tests. To do so, one approach is to generalize intuition from the well understood CHSH scenario; this has been done to extend the number of measurements per party [33, 34], or the number of parties [35], where it was shown how to construct multi-partite Bell inequalities from a bipartite seed. Self-testing results using projections onto bipartite subsystems [36, 37, 38] suggest stronger capabilities of the technique presented in reference [35], prompting the question: can it be leveraged for multi-partite randomness certification?

Our work seeks to answer these questions. We first show maximum MABK violation certifies N+1/2log2(1+2)/2N0.4N+1/2-\log_{2}(1+\sqrt{2})/\sqrt{2}\approx N-0.4 bits of global randomness when NN is even, contrasting the NN bits certified when NN is odd [31]. We then investigate the limits on certifiable randomness from quantum behaviours that violate an MABK inequality. To do so, we construct new families of NN-partite Bell inequalities whose maximum violation certifies NN bits for arbitrary NN. When NN is even, we show for any MABK violation mm up to a non-maximal value mm^{*}, there exists a quantum behaviour achieving mm which maximally violates an inequality from our family, and hence certifies maximum global randomness. For violations above mm^{*}, we find (using a second family of inequalities) that randomness monotonically decreases with MABK violation. Our results lower bound this trade-off, and we provide numerical evidence of tightness, indicating an extension of the known tight bound when N=2N=2 [23] to arbitrary NN. Additionally, we show that this trade-off diminishes (in the sense that the gap between mm^{*} and maximal MABK violation decreases) as the number of parties grows. Finally, for odd NN, we use our family of Bell inequalities to show that NN bits of randomness can be certified for a range of MABK violations.

Our constructions are formed from a type of concatenation of bipartite Bell inequalities, allowing us to lift our intuition derived from the bipartite case [23] to the multipartite setting. Specifically, we extend the technique presented in reference [35], originally introduced to witness genuinely multi-partite nonlocality. Our extension exploits self-testing properties of the seed to certify randomness. Specifically, our main technical contribution is a decoupling lemma (see Eq.˜13), which shows that under certain conditions the maximum violation of an expanded Bell expression implies a decoupling between the honest parties and Eve. Such decoupling occurs for extremal quantum correlations, and guarantees security [39]. We envisage that this technique will be a useful tool in multi-partite DI cryptography and of independent interest.

The paper is structured as followed. In Section˜2 we provide the necessary background and notation. In Section˜3 we detail our enhancement of the technique from [35], along with the decoupling lemma. In Section˜4 we describe our constructions, and show that they certify NN bits of randomness for a broad range of MABK violations, which we conjecture to be optimal; this result is summarized in Lemmas˜6 and 7. We additionally show how the conjectured highest value of MABK at which maximum randomness remains possible tends to the maximum quantum value as the number of parties increases (see Proposition˜6). In Section˜5 we consider MABK values beyond the conjectured threshold for maximum randomness, up to the maximum quantum value. Here we give a candidate form for the exact trade-off, which is summarized in Lemma˜10. We conclude with a discussion in Section˜6. All technical proofs can be found in the appendices.

2 Background

2.1 Multiparty DI-scenario

We consider an NN party, 2-input 2-output DI scenario, where NN isolated devices are given a random input, labeled xk{0,1}x_{k}\in\{0,1\}, and produce an output ak{0,1}a_{k}\in\{0,1\} stored in a classical register RkR_{k}, where k{1,,N}k\in\{1,...,N\} indexes the party. We use bold to denote tuples; for example, 𝒂=(a1,,aN)\bm{a}=(a_{1},...,a_{N}) denotes an NN bit string of device outputs. The behaviour of the devices is then described by the joint conditional probability distribution p(𝒂|𝒙)p(\bm{a}|\bm{x}), which must be no-signalling following the isolation of the devices.

A behaviour p(𝒂|𝒙)p(\bm{a}|\bm{x}) is quantum if there exists (following Naimark’s dilation theorem [40]) a pure state and sets of orthonormal projectors that can reproduce the distribution via the Born rule. Specifically, we consider an adversary Eve, who wishes to guess the device outputs, and let |Ψ𝑸~E|\Psi\rangle_{\tilde{\bm{Q}}E} denote the global state in the Hilbert space 𝑸~E\mathcal{H}_{\tilde{\bm{Q}}}\otimes\mathcal{H}_{E}, where E\mathcal{H}_{E} is held by Eve, 𝑸~=k=1NQ~k\mathcal{H}_{\tilde{\bm{Q}}}=\bigotimes_{k=1}^{N}\mathcal{H}_{\tilde{Q}_{k}} and Q~k\mathcal{H}_{\tilde{Q}_{k}} is held by device kk. Throughout the text, tildes will denote elements pertaining to physical Hilbert spaces Q~k\mathcal{H}_{\tilde{Q}_{k}} (whose dimension is unknown), whilst no tilde will describe qubit Hilbert spaces Qk\mathcal{H}_{Q_{k}}. We let {P~ak|xk(k)}ak\{\tilde{P}_{a_{k}|x_{k}}^{(k)}\}_{a_{k}} be binary-outcome projective measurements on Q~k\mathcal{H}_{\tilde{Q}_{k}}, and we denote the corresponding observables of each party by A~xk(k)=P~0|xk(k)P~1|xk(k)\tilde{A}_{x_{k}}^{(k)}=\tilde{P}_{0|x_{k}}^{(k)}-\tilde{P}_{1|x_{k}}^{(k)}; here bracketed superscripts will typically keep track of the party to which the object belongs (i.e., the Hilbert space on which it acts). Following measurements 𝒙\bm{x} by the NN honest parties, we obtain the classical quantum state ρ𝑹E|𝒙=𝒂|𝒂𝒂|𝑹ρE𝒂|𝒙\rho_{\bm{R}E|\bm{x}}=\sum_{\bm{a}}|\bm{a}\rangle\!\langle\bm{a}|_{\bm{R}}\otimes\rho_{E}^{\bm{a}|\bm{x}}, where ρE𝒂|𝒙=Tr𝑸~[(P~𝒂|𝒙𝕀E)|ΨΨ|𝑸~E]\rho_{E}^{\bm{a}|\bm{x}}=\mathrm{Tr}_{\tilde{\bm{Q}}}\big{[}(\tilde{P}_{\bm{a}|\bm{x}}\otimes\mathbb{I}_{E})|\Psi\rangle\!\langle\Psi|_{\tilde{\bm{Q}}E}\big{]} and P~𝒂|𝒙=k=1NP~ak|xk(k)\tilde{P}_{\bm{a}|\bm{x}}=\bigotimes_{k=1}^{N}\tilde{P}_{a_{k}|x_{k}}^{(k)}. The behaviour (as seen by the NN honest parties) is recovered by p(𝒂|𝒙)=Tr[ρE𝒂|𝒙]p(\bm{a}|\bm{x})=\mathrm{Tr}\big{[}\rho_{E}^{\bm{a}|\bm{x}}\big{]}.

2.2 Multiparty Nonlocality

Given an observed distribution p(𝒂|𝒙)p(\bm{a}|\bm{x}), it will be useful to quantify its distance from the local boundary. When N=2N=2, the CHSH inequality, as the only non-trivial facet, is a natural choice. The multiparty scenario is more complex however, and the number of classes of Bell inequality increases rapidly with NN [41]. Instead, a general way to quantify this distance for an arbitrary scenario can be used, which is related to how much the local polytope needs to be diluted to encompass a given non-local correlation. Computing this can be done using linear programming (see Section˜B.6).

In this work, we choose to study one such CHSH generalization and its relationship to DI randomness; the MABK family [24, 25, 26],

MN=21N2𝒙cos[π2(N12k=1Nxk)]A~𝒙,\langle M_{N}\rangle=2^{\frac{1-N}{2}}\sum_{\bm{x}}\cos\Big{[}\frac{\pi}{2}\Big{(}\frac{N-1}{2}-\sum_{k=1}^{N}x_{k}\Big{)}\Big{]}\langle\tilde{A}_{\bm{x}}\rangle, (1)

where A~𝒙=𝒂(1)k=1Nakp(𝒂|𝒙)=Ψ|A~x1(1)A~xN(N)𝕀E|Ψ\langle\tilde{A}_{\bm{x}}\rangle=\sum_{\bm{a}}(-1)^{\sum_{k=1}^{N}a_{k}}p(\bm{a}|\bm{x})=\langle\Psi|\tilde{A}_{x_{1}}^{(1)}\otimes\cdots\otimes\tilde{A}_{x_{N}}^{(N)}\otimes\mathbb{I}_{E}|\Psi\rangle when the behaviour is quantum. [In effect the prefactor (1)k=1Nak(-1)^{\sum_{k=1}^{N}a_{k}} corresponds to relabelling the outcomes 010\mapsto 1 and 111\mapsto-1 to match the usual formulation of observables with eigenvalues ±1\pm 1.] The local bound is given by MN1\langle M_{N}\rangle\leq 1, and the maximum quantum value is 2(N1)/22^{(N-1)/2}. Note that the MABK functional is characterized by the fact that the coefficients c𝒙=cos[π2(N12k=1Nxk)]c_{\bm{x}}=\cos\Big{[}\frac{\pi}{2}\Big{(}\frac{N-1}{2}-\sum_{k=1}^{N}x_{k}\Big{)}\Big{]} are equal for all strings 𝒙\bm{x} with the same Hamming weight (that is, the number of 11s in the string 𝒙\bm{x}). When NN is even, c𝒙{1/2,1/2}c_{\bm{x}}\in\{-1/\sqrt{2},1/\sqrt{2}\}, and when NN is odd c𝒙{1,0,+1}c_{\bm{x}}\in\{-1,0,+1\}. This is illustrated by writing out the first few cases. To ease notation, we let Ax(1)=Ax,Ay(2)=By,Az(3)=CzA_{x}^{(1)}=A_{x},\ A_{y}^{(2)}=B_{y},\ A_{z}^{(3)}=C_{z} and At(4)=DtA_{t}^{(4)}=D_{t}.

M2\displaystyle\langle M_{2}\rangle =12[A0B0+A0B1+A1B0A1B1],\displaystyle=\frac{1}{2}\Big{[}\langle A_{0}B_{0}\rangle+\langle A_{0}B_{1}\rangle+\langle A_{1}B_{0}\rangle-\langle A_{1}B_{1}\rangle\Big{]}, (2)
M3\displaystyle\langle M_{3}\rangle =12[A0B0C1+A0B1C0+A1B0C0A1B1C1],\displaystyle=\frac{1}{2}\Big{[}\langle A_{0}B_{0}C_{1}\rangle+\langle A_{0}B_{1}C_{0}\rangle+\langle A_{1}B_{0}C_{0}\rangle-\langle A_{1}B_{1}C_{1}\rangle\Big{]},
M4\displaystyle\langle M_{4}\rangle =14[A0B0C0D0+A0B0C0D1+A0B0C1D0+A0B1C0D0+A1B0C0D0+A0B0C1D1\displaystyle=\frac{1}{4}\Big{[}-\langle A_{0}B_{0}C_{0}D_{0}\rangle+\langle A_{0}B_{0}C_{0}D_{1}\rangle+\langle A_{0}B_{0}C_{1}D_{0}\rangle+\langle A_{0}B_{1}C_{0}D_{0}\rangle+\langle A_{1}B_{0}C_{0}D_{0}\rangle+\langle A_{0}B_{0}C_{1}D_{1}\rangle
+A1B1C0D0+A0B1C0D1+A1B0C1D0+A0B1C1D0+A1B0C0D1\displaystyle\hskip 25.6073pt+\langle A_{1}B_{1}C_{0}D_{0}\rangle+\langle A_{0}B_{1}C_{0}D_{1}\rangle+\langle A_{1}B_{0}C_{1}D_{0}\rangle+\langle A_{0}B_{1}C_{1}D_{0}\rangle+\langle A_{1}B_{0}C_{0}D_{1}\rangle
(A1B1C1D0+A1B1C0D1+A1B0C1D1+A0B1C1D1)A1B1C1D1].\displaystyle\hskip 25.6073pt-\big{(}\langle A_{1}B_{1}C_{1}D_{0}\rangle+\langle A_{1}B_{1}C_{0}D_{1}\rangle+\langle A_{1}B_{0}C_{1}D_{1}\rangle+\langle A_{0}B_{1}C_{1}D_{1}\rangle\big{)}-\langle A_{1}B_{1}C_{1}D_{1}\rangle\Big{]}.

Above, M2\langle M_{2}\rangle is the CHSH functional, whereas M3\langle M_{3}\rangle is the Mermin functional, associated with the GHZ paradox [42]. The property that all input strings of the same Hamming weight have the same coefficient makes the MABK expressions invariant under party relabelings. Additionally, note that for NN even, every correlator A𝒙\langle A_{\bm{x}}\rangle for 𝒙{0,1}n\bm{x}\in\{0,1\}^{n} is included in the expression, whereas for NN odd, exactly half of the correlators are included.

2.3 Self-testing

It is known that maximum quantum violation of the MABK family is uniquely achieved, up to local isometries, by the GHZ state and pairs of maximally anticommuting observables [43, 44]. This form of uniqueness arising from a Bell expression is known as self-testing. In this work, we take the choice |ψGHZ=(|0N+i|1N)/2|\psi_{\mathrm{GHZ}}\rangle=(|0\rangle^{\otimes N}+i|1\rangle^{\otimes N})/\sqrt{2}, A0(k)=cosθN+σX+sinθN+σYA_{0}^{(k)}=\cos\theta_{N}^{+}\,\sigma_{X}+\sin\theta_{N}^{+}\,\sigma_{Y}, and A1(k)=cosθNσX+sinθNσYA_{1}^{(k)}=\cos\theta_{N}^{-}\,\sigma_{X}+\sin\theta_{N}^{-}\,\sigma_{Y}, where θN±=(π/4)(1/N±1)\theta_{N}^{\pm}\leavevmode\nobreak\ =\leavevmode\nobreak\ (\pi/4)(1/N\pm 1).

Whilst self-testing statements can be formally defined between NN parties (see, e.g., [36, 20]), in our formulation it will only be necessary to use a bipartite definition.

Definition 1 (Bipartite self-test).

Let k,l{1,,N}k,l\in\{1,...,N\} index two distinct parties. Define the sets of target qubit projectors {Pak|xk(k)}ak\{P_{a_{k}|x_{k}}^{(k)}\}_{a_{k}}, {Pal|xl(l)}al\{P_{a_{l}|x_{l}}^{(l)}\}_{a_{l}}, and a target two qubit state |ΦQkQl|\Phi\rangle_{Q_{k}Q_{l}}. Let I(k,l)I^{(k,l)} be a Bell operator between parties kk and ll, with maximum quantum value ηQ\eta^{\mathrm{Q}}. The inequality I(k,l)ηQ\langle I^{(k,l)}\rangle\leq\eta^{\text{Q}} self-tests the target state and measurements if, for all physical quantum strategies (ρ~Q~kQ~l,P~ak|xk(k),P~al|xl(l))(\tilde{\rho}_{\tilde{Q}_{k}\tilde{Q}_{l}},\tilde{P}_{a_{k}|x_{k}}^{(k)},\tilde{P}_{a_{l}|x_{l}}^{(l)}) that satisfy I(k,l)=ηQ\langle I^{(k,l)}\rangle=\eta^{\mathrm{Q}}, there exists a local isometery V:Q~kQ~lEQkQlJunkV:\mathcal{H}_{\tilde{Q}_{k}}\otimes\mathcal{H}_{\tilde{Q}_{l}}\otimes\mathcal{H}_{E}\rightarrow\mathcal{H}_{Q_{k}}\otimes\mathcal{H}_{Q_{l}}\otimes\mathcal{H}_{\mathrm{Junk}} and ancillary state |ξJunk|\xi\rangle_{\mathrm{Junk}}, such that

V(P~ak|xk(k)P~al|xl(l)𝕀E)|ΨQ~kQ~lE=(Pak|xk(k)Pal|xl(l))|ΦQkQl|ξJunk,V\Big{(}\tilde{P}_{a_{k}|x_{k}}^{(k)}\otimes\tilde{P}_{a_{l}|x_{l}}^{(l)}\otimes\mathbb{I}_{E}\Big{)}|\Psi\rangle_{\tilde{Q}_{k}\tilde{Q}_{l}E}\\ =\Big{(}P_{a_{k}|x_{k}}^{(k)}\otimes P_{a_{l}|x_{l}}^{(l)}\Big{)}|\Phi\rangle_{Q_{k}Q_{l}}\otimes|\xi\rangle_{\mathrm{Junk}}\,, (3)

for all aka_{k}, ala_{l}, xkx_{k}, xlx_{l}, where |Ψ|\Psi\rangle is a purification of ρ~\tilde{\rho}.

We define the shifted Bell operator as I¯(k,l)=ηQ𝕀I(k,l)\bar{I}^{(k,l)}=\eta^{\mathrm{Q}}\mathbb{I}-I^{(k,l)}, and we say I¯(k,l)\bar{I}^{(k,l)} admits a sum-of-squares (SOS) decomposition if there exists a set of polynomials, {Mi(k,l)}i\{M^{(k,l)}_{i}\}_{i}, of the operators P~ak|xk(k),P~al|xl(l)\tilde{P}_{a_{k}|x_{k}}^{(k)},\tilde{P}_{a_{l}|x_{l}}^{(l)}, satisfying

I¯(k,l)=iMi(k,l)Mi(k,l).\bar{I}^{(k,l)}=\sum_{i}M^{(k,l)\dagger}_{i}M^{(k,l)}_{i}. (4)

SOS decompositions can be used to enforce algebraic constraints on any quantum state ρ~\tilde{\rho}, and measurements P~ak|xk(k),P~al|xl(l)\tilde{P}_{a_{k}|x_{k}}^{(k)},\tilde{P}_{a_{l}|x_{l}}^{(l)}, satisfying I¯(k,l)=0\langle\bar{I}^{(k,l)}\rangle=0. For example, when ρ~=|ψψ|\tilde{\rho}=|\psi\rangle\!\langle\psi| is pure, it must satisfy

Mi(k,l)|ψ=0,i,M_{i}^{(k,l)}|\psi\rangle=0,\ \forall i, (5)

(or Tr[(Mi(k,l))Mi(k,l)ρ~]=0\mathrm{Tr}[(M_{i}^{(k,l)})^{\dagger}M_{i}^{(k,l)}\tilde{\rho}]=0 more generally). Typically, we find that constraints of the form in Eq.˜5 are only satisfied by a unique state (up to local isometeries), and when that is the case we call Eq.˜5 the self-testing criteria.

2.4 DI randomness certification

We use the conditional von Neumann entropy, H(𝑹|𝑿=𝒙,E)ρ𝑹E|𝒙H(\bm{R}|\bm{X}=\bm{x}^{*},E)_{\rho_{\bm{R}E|\bm{x}^{*}}} to measure the DI randomness generation rate. This is the correct quantity for spot-checking DI random number generation, where 𝒙\bm{x}^{*} is a specified measurement choice used to generate randomness (see [45] for a discussion of this and other possibilities). The asymptotic rate of DI randomness generation is given by

r=inf|Ψ𝑸~E,{{P~ak|xk(k)}ak}kcompatiblewithfi(Pobs)H(𝑹|𝑿=𝒙,E)ρ𝑹E|𝒙r=\inf_{\begin{subarray}{c}|\Psi\rangle_{\tilde{\bm{Q}}E},\big{\{}\{\tilde{P}_{a_{k}|x_{k}}^{(k)}\}_{a_{k}}\big{\}}_{k}\\ \mathrm{compatible\ with}\ f_{i}(P_{\mathrm{obs}})\end{subarray}}H(\bm{R}|\bm{X}=\bm{x}^{*},E)_{\rho_{\bm{R}E|\bm{x}^{*}}} (6)

and we require lower bounds on this quantity over states and measurements compatible with the observed statistics, PobsP_{\mathrm{obs}}, or linear functions fi(Pobs)f_{i}(P_{\mathrm{obs}}) of them, such as the value of some Bell expression. We will typically only consider one functional, ff, which will be a Bell inequality with observed value f(Pobs)=ωf(P_{\mathrm{obs}})=\omega; then we denote the infimum in Eq.˜6 as the function Rf(ω)R_{f}(\omega). The asymptotic rate can be used as a basis for rates with finite statistics using tools such as the entropy accumulation theorem [46, 47, 48].

Because we study fundamental limits on certifiable DI randomness we work in the noiseless scenario, using self-testing statements to certify DI randomness. Here, f(Pobs)f(P_{\mathrm{obs}}) is a Bell expression, and we show all states and measurements that achieve its maximum violation correspond to a post-measurement state in tensor product with Eve, ρ𝑹E|𝒙=ρ𝑹|𝒙ρE\rho_{\bm{R}E|\bm{x}^{*}}=\rho_{\bm{R}|\bm{x}^{*}}\otimes\rho_{E}. This allows us to directly evaluate the conditional entropy H(𝑹|𝑿=𝒙,E)=H(𝑹|𝑿=𝒙)H(\bm{R}|\bm{X}=\bm{x}^{*},E)=H(\bm{R}|\bm{X}=\bm{x}^{*}), and the infimum becomes trivial since all compatible strategies give rise to the same classical distribution (see Appendix A.3 for details).

3 Expanding Bell inequalities

In this section we discuss and enhance the technique presented in [35], which allows us to derive the main results of this work. Ref. [35] introduced a method for building multi-partite Bell inequalities by expanding a bipartite inequality, called the “seed”, which can be used to witness genuinely multi-partite nonlocality. A new Bell expression is constructed by summing the seed over different subsets of parties, whilst the remaining parties perform some fixed measurement. The more multi-partite nonlocal the correlations, the more bipartite terms are violated. This resembles other recent results that enable multi-partite self-testing by projections onto bipartite subsystems [38, 36, 37].

In this work we extend this technique to make it suitable for DI cryptographic purposes. More precisely, we consider cases where the seed is a bipartite self-test, and use the maximum quantum violation of the expanded Bell expression, constructed in an equivalent manner to [35], to draw conclusions about the post-measurement state held between the honest parties and Eve. These conclusions allow us to derive rates for randomness certification.

3.1 Definition

We begin by introducing a generic formulation for expanding Bell expressions, shortly followed by an example.

Definition 2 (Expanded Bell expressions).

Let 𝑪\bm{C} be an N×NN\times N nonzero matrix with entries ck,lc_{k,l} satisfying ck,l{0,1}c_{k,l}\in\{0,1\} if k<lk<l, and ck,l=0c_{k,l}=0 otherwise. For each pair k,lk,l for which ck,l0c_{k,l}\neq 0, let {I𝝁(k,l)}𝝁\{I^{(k,l)}_{\bm{\mu}}\}_{\bm{\mu}} denote a set of bipartite Bell expressions between parties kk and ll, where 𝝁\bm{\mu} indexes over the N2N-2 measurement outcomes of all parties excluding k,lk,l, and each Bell inequality is equivalent up to output relabellings, with a local bound ηL\eta^{\mathrm{L}} strictly less than the maximum quantum value ηQ\eta^{\mathrm{Q}}. Then the expanded Bell operator II from the seed I𝝁(k,l)I^{(k,l)}_{\bm{\mu}} is defined as

I=k,lck,l(𝝁P~𝝁|𝟎(k,l)¯I𝝁(k,l)),I=\sum_{k,l}c_{k,l}\Bigg{(}\sum_{\bm{\mu}}\tilde{P}_{\bm{\mu}|\bm{0}}^{\overline{(k,l)}}I^{(k,l)}_{\bm{\mu}}\Bigg{)}, (7)

where P~𝝁|𝟎(k,l)¯=k=1,kk,lNP~μk|0(k)\tilde{P}_{\bm{\mu}|\bm{0}}^{\overline{(k,l)}}=\prod_{k^{\prime}=1,k^{\prime}\neq k,l}^{N}\tilde{P}_{\mu_{k^{\prime}}|0}^{(k^{\prime})} is the projector for all parties excluding kk and ll, corresponding to the joint setting 𝟎\bm{0} and joint outcome 𝝁\bm{\mu}111We have written the Bell operator in Eq. 7 in terms of projectors, which is convenient since we consider quantum strategies. We can however rewrite it in a theory independent way, by taking the expectation value and making the substitution p(𝒂|𝒙)=P~𝒂|𝒙p(\bm{a}|\bm{x})=\langle\tilde{P}_{\bm{a}|\bm{x}}\rangle. .

Consider the following tripartite example. To ease notation we label the observables of the three parties Ax=P0|xAP1|xAA_{x}=P^{A}_{0|x}-P^{A}_{1|x} and similarly for BB and CC. From reference [23], observing saturation of the following Bell expression certifies 2 bits of global randomness,

33IA,B33,-3\sqrt{3}\leq\langle I^{A,B}\rangle\leq 3\sqrt{3}, (8)

where IA,B=A0B0+2(A0B1+A1B0A1B1)I^{A,B}=A_{0}B_{0}+2\big{(}A_{0}B_{1}+A_{1}B_{0}-A_{1}B_{1}\big{)}. The bound IA,B=±33\langle I^{A,B}\rangle=\pm 3\sqrt{3} is uniquely achieved, up to local isometeries, by the bipartite strategy

ρQAQB=|Φ±Φ±|,|Φ±=|00±i|112,A0=B0=σX,A1=B1=σX+3σY2.\begin{gathered}\rho_{Q_{A}Q_{B}}=|\Phi_{\pm}\rangle\!\langle\Phi_{\pm}|,\ |\Phi_{\pm}\rangle=\frac{|00\rangle\pm i|11\rangle}{\sqrt{2}},\\ A_{0}=B_{0}=\sigma_{X},\ A_{1}=B_{1}=\frac{-\sigma_{X}+\sqrt{3}\sigma_{Y}}{2}.\end{gathered} (9)

Now consider a tripartite extension of the above strategy,

ρQAQBQC=|ψGHZψGHZ|,A0=B0=C0=σX,A1=B1=C1=σX+3σY2.\begin{gathered}\rho_{Q_{A}Q_{B}Q_{C}}=|\psi_{\mathrm{GHZ}}\rangle\!\langle\psi_{\mathrm{GHZ}}|,\\ A_{0}=B_{0}=C_{0}=\sigma_{X},\ A_{1}=B_{1}=C_{1}=\frac{-\sigma_{X}+\sqrt{3}\sigma_{Y}}{2}.\end{gathered} (10)

One can verify that when all three parties measure x=y=z=0x=y=z=0, their outcomes are uniformly distributed, resulting in 3 bits of raw randomness; how can we certify this device-independently? Notice |ψGHZ|\psi_{\mathrm{GHZ}}\rangle has the following property: when a single party, say AA, measures σX\sigma_{X}, the leftover state held by BCBC is |Φ±|\Phi_{\pm}\rangle where the sign depends on the parity of AA’s measurement outcome μ{0,1}\mu\in\{0,1\}. Parties BB and CC can now saturate one of the bounds IB,C=±33\langle I^{B,C}\rangle=\pm 3\sqrt{3} conditioned on AA’s measurement. By using IμA,B=(1)μIA,BI^{A,B}_{\mu}=(-1)^{\mu}I^{A,B} as the seed, and choosing the matrix 𝑪\bm{C} with party indices k,l{A,B,C}k,l\in\{A,B,C\},

𝑪=[001001000],\bm{C}=\begin{bmatrix}0&0&1\\ 0&0&1\\ 0&0&0\end{bmatrix}, (11)

we can construct the expanded Bell expression according to Footnote˜1:

I=Pμ=0|x=0AIB,CPμ=1|x=0AIB,Ck=A,l=C+Pμ=0|y=0BIA,CPμ=1|y=0BIA,Ck=B,l=C=A0IB,C+B0IA,C.I=\underbrace{P_{\mu=0|x=0}^{A}I^{B,C}-P_{\mu=1|x=0}^{A}I^{B,C}}_{k=A,l=C}\\ +\underbrace{P_{\mu=0|y=0}^{B}I^{A,C}-P_{\mu=1|y=0}^{B}I^{A,C}}_{k=B,l=C}\\ =A_{0}I^{B,C}+B_{0}I^{A,C}. (12)

Due to the properties of |ψGHZ|\psi_{\mathrm{GHZ}}\rangle discussed above, we find I=233\langle I\rangle=2\cdot 3\sqrt{3} is the maximum quantum value of I\langle I\rangle, and achieved by the strategy in Eq.˜10. Moreover, it can be shown that this is strictly greater than the maximum local value of I\langle I\rangle, implying I\langle I\rangle is a nontrivial Bell functional. In later sections, we will prove that I=233\langle I\rangle=2\cdot 3\sqrt{3} is a sufficient condition for certifying maximum randomness.

Returning to the general form of II from Eq.˜7, we call II an expanded Bell expression, since it is built by combining a bipartite seed, I𝝁(k,l)I_{\bm{\mu}}^{(k,l)}, conditioned on fixed measurement settings 𝟎\bm{0} and outcomes 𝝁\bm{\mu} for the remaining N2N-2 parties. We then sum over all possible outcomes of this N2N-2 party measurement, changing the Bell expression accordingly, and over different combinations of parties (k,l)(k,l) chosen according to 𝑪\bm{C}. There needs to be a gap between the local and quantum bounds of I\langle I\rangle for II to define a nontrivial Bell inequality; typically, we find the following upper bound is achievable, which is strictly greater than the local bound:

Lemma 1.

Let II be an expanded Bell expression according to Footnote˜1. The maximum quantum value of I\langle I\rangle is upper bounded by ηNQ:=k,lck,lηQ\eta^{\mathrm{Q}}_{N}:=\sum_{k,l}c_{k,l}\eta^{\mathrm{Q}}.

A proof of Lemma˜1 is given in Section˜A.1. If achievable, the only strategy that can give I=ηNQ\langle I\rangle=\eta^{\mathrm{Q}}_{N} is the one satisfying 𝝁P~𝝁|𝟎(k,l)¯I𝝁(k,l)=ηQ\big{\langle}\sum_{\bm{\mu}}\tilde{P}_{\bm{\mu}|\bm{0}}^{\overline{(k,l)}}I^{(k,l)}_{\bm{\mu}}\big{\rangle}=\eta^{\mathrm{Q}} for all pairing combinations k,lk,l with ck,l>0c_{k,l}>0, i.e., the reduced state held between parties kk and ll, following the projection P~𝝁|𝟎(k,l)¯\tilde{P}_{\bm{\mu}|\bm{0}}^{\overline{(k,l)}} on the global state, must achieve the maximum quantum value of I𝝁(k,l)=ηQ\langle I^{(k,l)}_{\bm{\mu}}\rangle=\eta^{\mathrm{Q}}. This explains why we must include a dependence on 𝝁\bm{\mu} for the Bell expression, since it should be tailored to the bipartite state following the outcome 𝝁\bm{\mu}. For simplicity we assume each bipartite state is equivalent up to local unitaries, meaning we effectively use a single seed I𝝁(k,l)I_{\bm{\mu}}^{(k,l)}. In principle however, one could use different seeds depending on the value of 𝝁\bm{\mu}, which tailors the construction to non-symmetric states [35]. For our work we only consider the symmetric case. Additionally, one could further generalize by choosing specific measurement choices 𝒚\bm{y} dependent on the choice of non-projecting parties k,lk,l, instead of fixing 𝒚=𝟎\bm{y}=\bm{0} as we have done here.

3.2 Expanding Bell expressions and entropy evaluation

Next we present the characteristic of expanded Bell expressions that allows us to certify DI randomness from witnessing its maximum quantum value, forming the main technical contribution of our work.

Lemma 2 (Decoupling lemma).

Let II be an expanded NN-party Bell expression defined in Footnote˜1 with binary inputs and outputs, and ck,N=1c_{k,N}=1 if k<Nk<N and zero otherwise. Suppose for every I𝛍(k,N)I_{\bm{\mu}}^{(k,N)}, there exists an SOS decomposition that self-tests the same pure bipartite entangled state |Φ|\Phi\rangle between parties kk and NN, along with some ideal measurements Pak|xk(k),PaN|xN(N)P^{(k)}_{a_{k}|x_{k}},P^{(N)}_{a_{N}|x_{N}}, according to Eq.˜3, satisfying Φ|Pak|0(k)PaN|0(N)|Φ>0\langle\Phi|P^{(k)}_{a_{k}|0}\otimes P^{(N)}_{a_{N}|0}|\Phi\rangle>0 for all ak,aNa_{k},a_{N}. Then for any strategy |Ψ𝐐~E,{{P~ak|xk(k)}xk}k|\Psi\rangle_{\tilde{\bm{Q}}E},\big{\{}\{\tilde{P}_{a_{k}|x_{k}}^{(k)}\}_{x_{k}}\big{\}}_{k} that achieves I=ηNQ\langle I\rangle=\eta^{\mathrm{Q}}_{N}, the post-measurement state ρ𝐑E|𝐱\rho_{\bm{R}E|\bm{x}}, for measurement settings 𝐱=𝟎\bm{x}=\bm{0}, admits the tensor product decomposition,

ρ𝑹E|𝟎=ρ𝑹|𝟎ρE.\rho_{\bm{R}E|\bm{0}}=\rho_{\bm{R}|\bm{0}}\otimes\rho_{E}. (13)

Having established that Eve is decoupled it is then straightforward to evaluate the rate in Eq.˜6 conditioned on observing the maximum quantum value of II.

Lemma 3.

Let I,ηNQI,\eta_{N}^{\mathrm{Q}} be defined as in Eq.˜13. Then

RI(ηNQ)=H({p(𝒂|𝟎)}),R_{I}(\eta_{N}^{\mathrm{Q}})=H(\{p(\bm{a}|\bm{0})\}), (14)

where H({pi})H(\{p_{i}\}) is the Shannon entropy of a distribution {pi}i\{p_{i}\}_{i}.

The proof can be found in Section˜A.3. Eq.˜13 allows one to relate observing the maximum quantum value of the expanded Bell expression II, to a condition on the post-measurement state held by the parties and Eve, namely that it must be in tensor product with the purifying system EE. This allows one to directly evaluate the conditional entropy according to Lemma˜3.

4 Certifying NN bits of DI randomness

We now consider constructions for generating NN bits of DI randomness. As previously mentioned, when NN is odd the MABK family can be used [31]. For an arbitrary number of parties, we apply the previously outlined techniques to derive a suitable Bell expression for this task. This will involve generalizing a one parameter family of quantum strategies, symmetric under permutation of the parties.

4.1 NN odd

Using symmetry arguments, the authors of Ref. [31] showed that maximum violation of the MABK inequality, for NN odd, implies maximum randomness.

Proposition 1 (Maximum randomness for NN odd [31]).

When NN is odd, maximum quantum violation of the MABK family of Bell inequalities, given by Eq.˜1, certifies NN bits of global DI randomness, i.e.

RMN(2(N1)/2)=N.R_{M_{N}}(2^{(N-1)/2})=N. (15)

From a self-testing point of view, maximum violation of the MABK family can only be achieved with a GHZ state and maximally anticommuting observables [43, 44]. For the form given in Eq.˜1, recall the following strategy achieves maximum violation:

ρ𝑸=|ψGHZψGHZ|A0(k)=cosθN+σX+sinθN+σY,A1(k)=cosθNσX+sinθNσY,k{1,,N},\begin{gathered}\rho_{\bm{Q}}=|\psi_{\mathrm{GHZ}}\rangle\!\langle\psi_{\mathrm{GHZ}}|\\ A_{0}^{(k)}=\cos\theta_{N}^{+}\,\sigma_{X}+\sin\theta_{N}^{+}\,\sigma_{Y},\\ A_{1}^{(k)}=\cos\theta_{N}^{-}\,\sigma_{X}+\sin\theta_{N}^{-}\,\sigma_{Y},\\ k\in\{1,...,N\},\end{gathered} (16)

with θN±=π/(4N)±π/4\theta_{N}^{\pm}=\pi/(4N)\pm\pi/4, and satisfies A0(k)A1(k)+A1(k)A0(k)=0A_{0}^{(k)}A_{1}^{(k)}+A_{1}^{(k)}A_{0}^{(k)}=0 for each kk. Following this self-testing property, the infimum in Eq.˜15 reduces to a single point up to symmetries, and the fact that the parties hold a pure state implies Eve must be uncorrelated, i.e., ρ𝑹E|𝒙=ρ𝑹|𝒙ρ𝑬\rho_{\bm{R}E|\bm{x}^{*}}=\rho_{\bm{R}|\bm{x}^{*}}\otimes\rho_{\bm{E}}. By direct calculation, one can verify for N=4n1N=4n-1, n=1,2,3,n=1,2,3,..., p(𝒂|𝟎)=1/2N,𝒂p(\bm{a}|\bm{0})=1/2^{N},\forall\bm{a}, hence H({p(𝒂|𝟎)})=NH(\{p(\bm{a}|\bm{0})\})=N. Similarly for N=4n+1N=4n+1, n=1,2,3,n=1,2,3,..., p(𝒂|𝟏)=1/2N,𝒂p(\bm{a}|\bm{1})=1/2^{N},\forall\bm{a}, implying H({p(𝒂|𝟏)})=NH(\{p(\bm{a}|\bm{1})\})=N.

As an example, taking N=3N=3 in the above construction we have M3=(A~1(0)A~0(1)A~0(2)+A~0(0)A~1(1)A~0(2)+A~0(0)A~0(1)A~1(2)A~1(0)A~1(1)A~1(2))/2\langle M_{3}\rangle=\big{(}\langle\tilde{A}_{1}^{(0)}\tilde{A}_{0}^{(1)}\tilde{A}_{0}^{(2)}\rangle+\langle\tilde{A}_{0}^{(0)}\tilde{A}_{1}^{(1)}\tilde{A}_{0}^{(2)}\rangle+\langle\tilde{A}_{0}^{(0)}\tilde{A}_{0}^{(1)}\tilde{A}_{1}^{(2)}\rangle-\langle\tilde{A}_{1}^{(0)}\tilde{A}_{1}^{(1)}\tilde{A}_{1}^{(2)}\rangle\big{)}/2, and the strategy in Eq.˜16 reaches the algebraic bound of 22. We therefore have that M3\langle M_{3}\rangle implies 33 bits of randomness when all three parties use measurement 0. On the other hand, the correlators in the inequality must take the values ±1\pm 1 to reach the algebraic bound, so no string of inputs which appears as a correlator in the inequality can generate more than 2 bits of global randomness (in fact they give exactly 2 bits when M3=2\langle M_{3}\rangle=2).

4.2 NN even

When NN is even, all combinations A~x1(1)A~x2(2)A~xN(N)\langle\tilde{A}_{x_{1}}^{(1)}\tilde{A}_{x_{2}}^{(2)}\cdots\tilde{A}_{x_{N}}^{(N)}\rangle appear in MN\langle M_{N}\rangle, and all must have non-zero weight to achieve the largest possible quantum value. This is incompatible with maximum randomness from one input combination. Hence, achieving the quantum bound does not certify maximum randomness; by direct calculation, maximum MABK violation certifies N+1/2log2(1+2)/2N0.4N+1/2-\log_{2}(1+\sqrt{2})/\sqrt{2}\approx N-0.4 bits of global DI randomness when all parties use measurement 0.

We now proceed to construct new Bell expressions which certify NN bits of randomness in the even case. Specifically, we use the techniques introduced in Section˜3 to generalize the N=2N=2 case, which was addressed in [23].

4.2.1 Bipartite self-tests

To begin with, we summarize the results obtained for the N=2N=2 case.

Lemma 4 (IθI_{\theta} family of self-tests).

Define the set 𝒢=(π/4,π/2)(π/2,3π/4)(5π/4,3π/2)(3π/2,7π/4)\mathcal{G}=(\pi/4,\pi/2)\cup(\pi/2,3\pi/4)\cup(5\pi/4,3\pi/2)\cup(3\pi/2,7\pi/4)222Note that 𝒢={θ[0,2π]cos(2θ)<0,cos(θ)0}\mathcal{G}=\left\{\theta\in[0,2\pi]\mid\cos(2\theta)<0,\ \cos(\theta)\neq 0\right\}.. Let θ𝒢\theta\in\mathcal{G}, and define the family of Bell expressions parameterized by θ\theta, for parties k,l{1,,N}k,l\in\{1,...,N\},

Iθ(k,l)=cosθcos2θA0(k)A0(l)cos2θ(A0(k)A1(l)+A1(k)A0(l))+cosθA1(k)A1(l).\langle I_{\theta}^{(k,l)}\rangle=\cos\theta\cos 2\theta\langle A_{0}^{(k)}A_{0}^{(l)}\rangle-\\ \cos 2\theta\big{(}\langle A_{0}^{(k)}A_{1}^{(l)}\rangle+\langle A_{1}^{(k)}A_{0}^{(l)}\rangle\big{)}+\cos\theta\langle A_{1}^{(k)}A_{1}^{(l)}\rangle. (17)

Then we have the following:

  1. (i)

    The local bounds are given by ±ηθL\pm\eta_{\theta}^{\mathrm{L}}, where

    ηθL=max(|cosθ(1cos2θ)|,|cosθ(1+cos2θ)|+|2cos2θ|).\eta_{\theta}^{\mathrm{L}}=\mathrm{max}\Big{(}|\cos\theta(1-\cos 2\theta)|,\\ |\cos\theta\big{(}1+\cos 2\theta\big{)}|+|2\cos 2\theta|\Big{)}. (18)
  2. (ii)

    The quantum bounds are given by ±ηθQ\pm\eta_{\theta}^{\mathrm{Q}}, where ηθQ=2sin3θ\eta_{\theta}^{\mathrm{Q}}=2\sin^{3}\theta.

  3. (iii)

    Up to local isometries, there exists a unique strategy that achieves Iθ=ηθQ\langle I_{\theta}\rangle=\eta_{\theta}^{\mathrm{Q}}:

    ρQkQl\displaystyle\rho_{Q_{k}Q_{l}} =|ψψ|,where|ψ=12(|00+i|11)\displaystyle=|\psi\rangle\!\langle\psi|,\ \mathrm{where}\ |\psi\rangle=\frac{1}{\sqrt{2}}\big{(}|0\rangle+i|1\rangle\big{)} (19)
    A0(k)\displaystyle A_{0}^{(k)} =A0(l)=σX\displaystyle=A_{0}^{(l)}=\sigma_{X}
    A1(k)\displaystyle A_{1}^{(k)} =A1(l)=cosθσX+sinθσY.\displaystyle=A_{1}^{(l)}=\cos\theta\,\sigma_{X}+\sin\theta\,\sigma_{Y}.

The above lemma is a rewriting of Proposition 1 in [23], and can be recovered as a special case of Lemma˜8 to follow. Importantly, the self-testing property implies that when their inputs are 0, both parties measure σX\sigma_{X} on |ψ|\psi\rangle, which results in 2 bits of randomness, H({p(ab|00)})=2H(\{p(ab|00)\})=2. We also remark that the state |ψ=(|00i|11)/2|\psi^{\prime}\rangle=(|00\rangle-i|11\rangle)/\sqrt{2} with the same measurements achieves Iθ=ηθQ\langle I_{\theta}\rangle=-\eta_{\theta}^{\mathrm{Q}}. This is a symmetry of the case above, since relabeling the measurement outcomes of party kk yields the new Bell expression Iθ=IθI_{\theta}^{\prime}=-I_{\theta}. Therefore Iθ=ηθQ\langle I_{\theta}\rangle=-\eta_{\theta}^{\mathrm{Q}} implies Iθ=ηθQ\langle I_{\theta}^{\prime}\rangle=\eta_{\theta}^{\mathrm{Q}}, which self-tests |ψ|\psi\rangle and the negated measurements, or equivalently |ψ|\psi^{\prime}\rangle and the original measurements.

4.2.2 Target strategy

For the NN partite case, we generalize the above bipartite strategy:

ρ𝑸=|ψGHZψGHZ|A0(k)=σX,A1(k)=cosθσX+sinθσY,k{1,,N},\begin{gathered}\rho_{\bm{Q}}=|\psi_{\mathrm{GHZ}}\rangle\!\langle\psi_{\mathrm{GHZ}}|\\ A_{0}^{(k)}=\sigma_{X},\\ A_{1}^{(k)}=\cos\theta\,\sigma_{X}+\sin\theta\,\sigma_{Y},\ k\in\{1,...,N\},\end{gathered} (20)

where θ𝒢\theta\in\mathcal{G} is a free parameter. As before, we find p(𝒂|𝟎)=1/2N,𝒂p(\bm{a}|\bm{0})=1/2^{N},\forall\bm{a}, hence this is a target strategy for certifying NN bits of global randomness. Our objective now will be to design a tailored Bell inequality that can certify NN bits by witnessing its maximum value, attained by the strategy in Eq.˜20.

4.2.3 Constructing an NN-partite Bell inequality

We use the techniques introduced in Section˜3 to construct the desired Bell expression by expanding the seed IθI_{\theta}. Studying the strategy in Eq.˜20, we notice that when N2N-2 parties perform their first measurement, σX\sigma_{X}, depending on the outcome parity the post-measurement state for the remaining parties is (|00±i|11)/2(|00\rangle\pm i|11\rangle)/\sqrt{2}. Now, performing the measurements in Eq.˜20 on that post-measurement state achieves Iθ=±ηθQ\langle I_{\theta}\rangle=\pm\eta_{\theta}^{\mathrm{Q}}, which, according to the self-testing properties of IθI_{\theta}, will certify 2 bits of randomness. Since the strategy is symmetric under party permutations, we can repeat this argument over different subsets of N2N-2 parties, and conclude that, when every bipartite self-test is satisfied, the output setting 𝑿=𝟎\bm{X}=\bm{0} must generate uniform DI randomness. The corresponding Bell inequality is constructed according to the following lemma.

Lemma 5 (Bell inequality for maximum randomness).

Let 𝛍\bm{\mu} be a tuple of N2N-2 measurement outcomes for all parties excluding k,lk,l, and n𝛍{0,1}n_{\bm{\mu}}\in\{0,1\} be the parity of 𝛍\bm{\mu}. Let θ𝒢\theta\in\mathcal{G} and {I𝛍(k,l)}𝛍\{I_{\bm{\mu}}^{(k,l)}\}_{\bm{\mu}} be a set of bipartite Bell expressions between parties k,lk,l, where

I𝝁(k,l)=(1)n𝝁Iθ(k,l).I_{\bm{\mu}}^{(k,l)}=(-1)^{n_{\bm{\mu}}}I_{\theta}^{(k,l)}. (21)

Then the expanded Bell expression given by

Iθ=k=1N1(𝝁P~𝝁|𝟎(k,N)¯I𝝁(k,N))I_{\theta}=\sum_{k=1}^{N-1}\Bigg{(}\sum_{\bm{\mu}}\tilde{P}^{\overline{(k,N)}}_{\bm{\mu}|\bm{0}}I_{\bm{\mu}}^{(k,N)}\Bigg{)} (22)

has quantum bounds ±ηN,θQ\pm\eta_{N,\theta}^{\mathrm{Q}} where ηN,θQ=2(N1)sin3θ\eta_{N,\theta}^{\mathrm{Q}}=2(N-1)\sin^{3}\theta. Moreover, Iθ=ηN,θQ\langle I_{\theta}\rangle=\eta_{N,\theta}^{\mathrm{Q}} is achieved up to relabelings by the strategy in Eq.˜20, and cannot be achieved classically.

The proof can be found in Section˜B.1.

We can now state our first main result:

Proposition 2 (Maximum randomness certification).

Achieving the maximum quantum value of the Bell inequality in Lemma˜5 certifies NN bits of global randomness, i.e.,

RIθ(ηN,θQ)=N.R_{I_{\theta}}(\eta_{N,\theta}^{\mathrm{Q}})=N. (23)

The proof is obtained directly by relating maximal Bell violation to decoupling Eve from the post-measurement state, proven in Eqs.˜13 and 3.

4.3 MABK value and maximum randomness

Above we gave a new one parameter construction for certifying NN bits of device-independent randomness in any NN-party 2-input 2-output scenario. Now we will study some properties of the correlations used to certify maximum randomness. In particular we’re interested in how maximal randomness and nonlocality tradeoff against each other. For instance, how nonlocal can correlations be whilst certifying maximal randomness?

4.3.1 Achievable MABK values with maximum randomness

When NN is odd, we have discussed that one can always certify NN bits of DI randomness from maximum MABK violation; for the even case, it is unclear how large the MABK value can be whilst certifying maximum randomness. Using the fact that, for the strategy in Eq.˜20, A𝒙=sinnθ\langle A_{\bm{x}}\rangle=\sin n\theta, where n=k=1Nxkn=\sum_{k=1}^{N}x_{k}, its MABK value is given by333This can be established using the identities cos(x)=(eix+eix)/2,sin(x)=(eixeix)/(2i)\cos(x)=(e^{ix}+e^{-ix})/2,\ \sin(x)=(e^{ix}-e^{-ix})/(2i) and (1+eix)N=2NeixN/2cosN(x/2)(1+e^{ix})^{N}=2^{N}e^{ixN/2}\cos^{N}(x/2).

MN(θ)=2N12(cosN(θ/2+π/4)sin(Nθ/2+π/4)+cosN(θ/2π/4)sin(Nθ/2π/4)).\langle M_{N}(\theta)\rangle=2^{\frac{N-1}{2}}\Big{(}\cos^{N}\big{(}\theta/2+\pi/4\big{)}\sin\big{(}N\theta/2+\pi/4\big{)}\\ +\cos^{N}\big{(}\theta/2-\pi/4\big{)}\sin\big{(}N\theta/2-\pi/4\big{)}\Big{)}. (24)

We begin with the following proposition:

Proposition 3.

For even NN, let

θN=2πtNN+1,\theta^{*}_{N}=\frac{2\pi t_{N}}{N+1}, (25)

where tNt_{N} is the (N/2)th(N/2)^{\mathrm{th}} element of the sequence 1,1,5,7,3,3,11,13,5,5,1,1,5,7,3,3,11,13,5,5,... given by

tN={N/4+1/2,ifN=8n+2,N/4,ifN=8n+4,3N/4+1/2,ifN=8n+6,3N/4+1,ifN=8n+8,n0.t_{N}=\begin{cases}N/4+1/2,\ \mathrm{if}\ N=8n+2,\\ N/4,\ \mathrm{if}\ N=8n+4,\\ 3N/4+1/2,\ \mathrm{if}\ N=8n+6,\\ 3N/4+1,\ \mathrm{if}\ N=8n+8,\ n\in\mathbb{N}_{0}.\end{cases} (26)

Then θN𝒢\theta^{*}_{N}\in\mathcal{G}.

Proof can be found in Section˜B.2. Using this result, we are able to prove the following. It will be convenient to define

mN=MN(θN).m_{N}^{*}=\langle M_{N}(\theta^{*}_{N})\rangle. (27)
Proposition 4.

Let NN be an even integer. For every MABK value ss in the range (1,mN](1,m_{N}^{*}], there exists a θs𝒢\theta_{s}\in\mathcal{G} that satisfies s=MN(θs)s=\langle M_{N}(\theta_{s})\rangle.

The proof of Proposition 4 can be found in Section˜B.3. In Fig.˜1, we plot the MABK value for a given NN as a function of θ𝒢\theta\in\mathcal{G}, showing the full range of violations (1,mN](1,m_{N}^{*}] is accessible. Specifically, the dashed lines in Fig.˜1 indicate an MABK value of 1, while the peak of each graph is mNm_{N}^{*}. As θ\theta varies, the graph shows a continuous curve between these values. We next state the following technical conjecture, which holds in the case N=2N=2 and we provide evidence for general NN in Appendix B.10.

Conjecture 1.

The maximum MABK value achievable by quantum strategies with uniformly random outputs on input 𝟎\bm{0}, i.e., p(𝐚|𝟎)=1/2N,𝐚p(\bm{a}|\bm{0})=1/2^{N},\ \forall\bm{a}, is mNm_{N}^{*}.

Fig.˜1 illustrates that mNm_{N}^{*} is the largest MABK value that can be achieved by the family of strategies in Eq.˜20, all of which generate maximum randomness. However, we have not ruled out the existence of a strategy outside this family, also generating maximum randomness, achieving an MABK value greater than mNm_{N}^{*}. ˜1 is that this is not the case, and mNm_{N}^{*} is indeed the true maximum. Based on this, we can now state the range of MABK violations for which maximum DI randomness can be certified when NN is even.

Lemma 6 (MABK violations achievable with maximum DI randomness, even case).

For even NN, we have:

  1. (i)

    The maximum amount of device-independent randomness that can be certified for the range of MABK values (1,mN](1,m_{N}^{*}] is NN bits.

Suppose ˜1 holds. Then we additionally have:

  1. (ii)

    mNm_{N}^{*} is the largest MABK value compatible with strategies generating NN bits of device-independent randomness.

Part (i)(i) can be established by the following reasoning: by varying θ𝒢\theta\in\mathcal{G}, the family of strategies in Eq.˜20 achieves every MABK value in the interval (1,mN](1,m_{N}^{*}] by Proposition˜4. Since θ𝒢\theta\in\mathcal{G} for all these values, the corresponding bipartite strategy is self-testable according to Lemma˜4; we can hence apply Lemma˜5 to expand the Bell expression to NN parties, and it follows from Lemma˜3 that this Bell expression certifies NN bits of DI randomness.

An implication of Proposition˜3 is that the MABK value mNm_{N}^{*} becomes an achievable lower bound on the maximum MABK value for quantum correlations certifying maximum randomness. This follows from the same reasoning outlined in the previous paragraph, since θN𝒢\theta_{N}^{*}\in\mathcal{G} implies maximum DI randomness can be certified from the correlations generated by the associated strategy in Eq.˜20. Part (ii) of Lemma 6 is that this is optimal (if ˜1 holds), in the sense that one cannot achieve a larger MABK value whilst simultaneously certifying maximum randomness. As evidence we derive a numerical technique that can generate upper bounds on this MABK value, and show these upper bounds match the lower bounds for some values NN. See Section˜B.4 for the details. We also remark that ˜1 is known to hold for the case N=2N=2 [23].

We now consider the case of odd MABK expressions.

Proposition 5.

Let NN be an odd integer. For every MABK value ss in the range (1,2(N1)/2)(1,2^{(N-1)/2}), there exists a θs𝒢\theta_{s}\in\mathcal{G} that satisfies s=MN(θs)s=\langle M_{N}(\theta_{s})\rangle.

Poof can be found in Section˜B.3. Similar to the even case, in Fig.˜2 we plot the MABK value as a function of θ\theta for different NN, and see that all quantum achievable values can be obtained by Eq.˜20 for some θ𝒢\theta\in\mathcal{G}. We then have the following consequence:

Lemma 7 (MABK violations achievable with maximum DI randomness, odd case).

For odd NN the maximum amount of device-independent randomness that can be certified for the range of MABK values (1,2(N1)/2](1,2^{(N-1)/2}] is NN bits.

The above lemma asserts that when NN is odd, for all quantum-achievable MABK values maximum randomness can be realized simultaneously. For the case of MN=2(N1)/2\langle M_{N}\rangle=2^{(N-1)/2}, i.e., the maximum quantum value, randomness can be certified according to Eq.˜15. For all other MABK values, Proposition˜5 implies maximum randomness can be certified by IθI_{\theta} for some θ𝒢\theta\in\mathcal{G}.

We remark on the nontrivial maximal MABK violation compatible with maximum randomness when NN is even, contrasting the odd case; as discussed in Section˜4.1, this can be seen as a consequence MABK expressions for even NN containing every NN party correlator. When maximum randomness is being certified, one of these correlators must be zero, which restricts the maximum MABK violation to be strictly less than the optimal quantum value.

Refer to caption
Refer to caption
Figure 1: Using the family of strategies in (20) for even NN we plot the MABK value renormalized by the maximum quantum value over all strategies (i.e., 2(N1)/22^{(N-1)/2}) in terms of θ[π/4,3π/4]\theta\in[\pi/4,3\pi/4], and θ[5π/4,7π/4]\theta\in[5\pi/4,7\pi/4] respectively. The dashed lines indicate where the strategy becomes local. All points in this interval, excluding the boundaries and center point (since π/2\pi/2 and 3π/23\pi/2 are not in 𝒢\mathcal{G}), correspond to strategies that certify NN bits of randomness device-independently using our expanded Bell expressions.
Refer to caption
Refer to caption
Figure 2: A similar plot to that of Fig. 1 for odd NN and second measurement angle of all parties θ[π/4,3π/4]\theta\in[\pi/4,3\pi/4], and θ[5π/4,7π/4]\theta\in[5\pi/4,7\pi/4] respectively. All points in this interval, excluding the boundaries, correspond to strategies that can certify NN bits of randomness device-independently, using our expanded Bell expressions, or using the MABK inequality for the center points. For N=3,7,11,15N=3,7,11,15, we use the MABK expression given by Eq.˜1, and for N=5,9,13,17N=5,9,13,17 we use the same expression after relabeling every parties inputs followed by their first measurement’s output.

4.3.2 Asymptotic behaviour

We now consider the behaviour of the conjectured maximal MABK value achievable with maximum randomness, mNm_{N}^{*}, for increasingly large even NN. We show that mNm_{N}^{*} converges to the largest possible quantum value in this limit. Note this is not based on a conjecture; mNm_{N}^{*} is an achievable lower bound, and in the following proposition we show this lower bound tends towards the global upper bound, namely the maximal quantum MABK value.

Proposition 6 (Maximum randomness in the asymptotic limit).

In the limit of large even NN, one can achieve arbitrarily close to the maximum quantum violation of the NN party MABK inequality, 2(N1)/22^{(N-1)/2}, whilst certifying maximum device-independent randomness.

This is proven in Section˜B.5.

4.3.3 Nonlocality and maximum randomness

Whilst we have studied the MABK values achieved by the strategies in Eq.˜20, we also consider how nonlocal they are, quantified by how much the local set needs to be “diluted” to contain them. We refer to this measure as the local dilution, which can be computed via a linear program for small NN, and we give details in Section˜B.6. Our findings are presented in Fig.˜3.

Refer to caption
Figure 3: Nonlocality, as measured using local dilution, of the strategies in Eq.˜20, for second measurement angle of all parties θ[0,π]\theta\in[0,\pi]. All values of θ(π/4,π/2)(π/2,3π/4)\theta\in(\pi/4,\pi/2)\cup(\pi/2,3\pi/4) correspond to strategies which can certify NN bits of maximum randomness using our technique for expanding Bell expressions. θ=π/2\theta=\pi/2 can also correspond to maximum randomness when NN is odd by testing the MABK inequality.

Interestingly, for N>2N>2 the correlations of the strategy in Eq.˜20 are bounded away from the local boundary for any θ𝒢\theta\in\mathcal{G}, even for small MABK violation; whilst the violation of a given MABK inequality can be made arbitrarily small, the correlations are still distant from the local boundary. In fact, by comparing Fig.˜3 with Fig.˜1 and Fig.˜2, one can see how correlations achieving at or below the maximum local MABK value, MN1\langle M_{N}\rangle\leq 1, can still be nonlocal and certify maximum randomness. To illustrate this point, consider the N=3N=3 (blue) curve in Fig.˜2. Provided θ(π/4,π/2)𝒢\theta\in(\pi/4,\pi/2)\subset\mathcal{G}, NN bits of DI randomness can be certified following previous arguments. However, the blue dashed line in Fig.˜2, indicating the local bound (M3=1\langle M_{3}\rangle=1), intersects the solid curve inside the region (π/4,π/2)(\pi/4,\pi/2), implying the corresponding strategy in Eq.˜20 achieves M31\langle M_{3}\rangle\leq 1 while simultaneously certifying maximum randomness. This is explained by the N=3N=3 (red) curve in Fig.˜3, where it can be seen that in the closed region [π/4,π/2][\pi/4,\pi/2] the correlations are nonlocal and hence must violate some other Bell inequality. This illustrates the complexity of the tradeoff between randomness versus nonlocality in the multi-partite scenario.

5 Trade-off between DI randomness and MABK violation

In this section, we present an achievable lower bound on the maximum device-independent randomness that can be generated by correlations achieving any given MABK violation. Moreover, we conjecture this lower bound to be tight based on numerical evidence. This is achieved by introducing a new family of two parameter quantum strategies along with their self-testing Bell expressions. Using this as the seed, we construct a multi-partite Bell expression which certifies their randomness.

5.1 Constructing the Bell expression

5.1.1 Bipartite self-tests

We begin by introducing a versatile family of bipartite self-tests which generalize the results in [23].

Lemma 8 (Jϕ,θJ_{\phi,\theta} family of self-tests).

Let (ϕ,θ)2(\phi,\theta)\in\mathbb{R}^{2} such that cos(2θ)cos(2ϕ)<0\cos(2\theta)\cos(2\phi)<0 and cos(θϕ)0\cos(\theta-\phi)\neq 0444The reason for this condition is explained in Section B.7.. Define the family of Bell expressions parameterized by ϕ\phi and θ\theta,

Jϕ,θ(k,l)=cos2θcos(θϕ)A0(k)A0(l)cos2θcos2ϕ(A0(k)A1(l)+A1(k)A0(l))+cos2ϕcos(θϕ)A1(k)A1(l).\langle J_{\phi,\theta}^{(k,l)}\rangle=\cos 2\theta\cos(\theta-\phi)\langle A_{0}^{(k)}A_{0}^{(l)}\rangle\\ -\cos 2\theta\cos 2\phi\big{(}\langle A_{0}^{(k)}A_{1}^{(l)}\rangle+\langle A_{1}^{(k)}A_{0}^{(l)}\rangle\big{)}\\ +\cos 2\phi\cos(\theta-\phi)\langle A_{1}^{(k)}A_{1}^{(l)}\rangle. (28)

Then we have the following:

  1. (i)

    The local bounds are given by ±ηϕ,θL\pm\eta_{\phi,\theta}^{\mathrm{L}}, where

    ηϕ,θL=max{|cos(θϕ)(cos2θcos2ϕ)|,|cos(θϕ)(cos2θ+cos2ϕ)|+|2cos2ϕcos2θ|}.\eta_{\phi,\theta}^{\mathrm{L}}=\max\big{\{}|\cos(\theta-\phi)\big{(}\cos 2\theta-\cos 2\phi\big{)}|,\\ |\cos(\theta-\phi)\big{(}\cos 2\theta+\cos 2\phi\big{)}|+|2\cos 2\phi\cos 2\theta|\big{\}}. (29)
  2. (ii)

    The quantum bounds are given by ±ηϕ,θQ\pm\eta_{\phi,\theta}^{\mathrm{Q}}, where ηϕ,θQ=2sin2(θ+ϕ)sin(θϕ)\eta_{\phi,\theta}^{\mathrm{Q}}=2\sin^{2}(\theta+\phi)\sin(\theta-\phi).

  3. (iii)

    |ηϕ,θQ|>ηϕ,θL|\eta^{\mathrm{Q}}_{\phi,\theta}|>\eta^{\mathrm{L}}_{\phi,\theta}.

  4. (iv)

    Up to local isometries, there exists a unique strategy that achieves Jϕ,θ(k,l)=ηϕ,θQ\langle J_{\phi,\theta}^{(k,l)}\rangle=\eta_{\phi,\theta}^{\mathrm{Q}}:

    ρQkQl\displaystyle\rho_{Q_{k}Q_{l}} =|ψψ|,where|ψ=12(|00+i|11)\displaystyle=|\psi\rangle\!\langle\psi|,\ \mathrm{where}\ |\psi\rangle=\frac{1}{\sqrt{2}}\big{(}|0\rangle+i|1\rangle\big{)} (30)
    A0(k)\displaystyle A_{0}^{(k)} =A0(l)=cos(ϕ)σXsin(ϕ)σY\displaystyle=A_{0}^{(l)}=\cos(\phi)\,\sigma_{X}-\sin(\phi)\,\sigma_{Y}
    A1(k)\displaystyle A_{1}^{(k)} =A1(l)=cos(θ)σX+sin(θ)σY.\displaystyle=A_{1}^{(l)}=\cos(\theta)\,\sigma_{X}+\sin(\theta)\,\sigma_{Y}.

Note that when ϕ=0\phi=0 Lemma˜8 reduces to Lemma˜4. Moreover, this Bell expression retains the same symmetry properties of the IθI_{\theta} family, namely that Jϕ,θ=ηϕ,θQ\langle J_{\phi,\theta}\rangle=-\eta^{\mathrm{Q}}_{\phi,\theta} for the state |ψ|\psi^{\prime}\rangle and the same measurements, and is hence a self-test. Lemma˜8 can be obtained as a corollary of the self-testing results from Refs [49, 50, 51] (see Section˜B.7 for details).

For future convenience, we define the set =[(π/4,π/4)×𝒢][(π/4,π/4){0}×{π/2,3π/2}]2\mathcal{F}=\Big{[}(-\pi/4,\pi/4)\times\mathcal{G}\Big{]}\cup\Big{[}(-\pi/4,\pi/4)\setminus\{0\}\times\{\pi/2,3\pi/2\}\Big{]}\subset\mathbb{R}^{2}. One can verify that points (ϕ,θ)(\phi,\theta)\in\mathcal{F} satisfy cos(2θ)cos(2ϕ)<0\cos(2\theta)\cos(2\phi)<0 and cos(θϕ)0\cos(\theta-\phi)\neq 0, and therefore define a valid self-test according to Lemma˜8.

5.1.2 Target strategy

For the NN-partite case, we will consider the following strategy:

ρ𝑸=|ψGHZψGHZ|A0(k)=cosϕσXsinϕσY,A1(k)=cosθσX+sinθσY,k{1,,N}.\begin{gathered}\rho_{\bm{Q}}=|\psi_{\mathrm{GHZ}}\rangle\!\langle\psi_{\mathrm{GHZ}}|\\ A_{0}^{(k)}=\cos\phi\,\sigma_{X}-\sin\phi\,\sigma_{Y},\\ A_{1}^{(k)}=\cos\theta\,\sigma_{X}+\sin\theta\,\sigma_{Y},\ k\in\{1,...,N\}.\end{gathered} (31)

Using the fact that A𝒙=sin(nθ(Nn)ϕ)\langle A_{\bm{x}}\rangle=\sin(n\theta-(N-n)\phi), where n=kxkn=\sum_{k}x_{k}, the MABK value, defined in Eq.˜1, of the above strategy is given by

MN(ϕ,θ)=2N12(cosN[(θ+ϕ)/2+π/4]sin[N(θϕ)/2+π/4]+cosN[(θ+ϕ)/2π/4]sin[N(θϕ)/2π/4]).\langle M_{N}(\phi,\theta)\rangle=2^{\frac{N-1}{2}}\Big{(}\cos^{N}\big{[}(\theta+\phi)/2+\pi/4\big{]}\\ \cdot\sin\big{[}N(\theta-\phi)/2+\pi/4\big{]}\\ +\cos^{N}\big{[}(\theta+\phi)/2-\pi/4\big{]}\sin\big{[}N(\theta-\phi)/2-\pi/4\big{]}\Big{)}. (32)

5.1.3 Constructing an NN-partite Bell inequality

Using the previous two building blocks, we construct the following Bell inequality using the Jϕ,θJ_{\phi,\theta} expressions as the seed.

Lemma 9.

Let 𝛍\bm{\mu} be a tuple of N2N-2 measurement outcomes for all parties excluding k,lk,l, and n𝛍{0,1}n_{\bm{\mu}}\in\{0,1\} be the parity of 𝛍\bm{\mu}. Let (ϕ,θ)2(\phi,\theta)\in\mathbb{R}^{2},

ϕ:=ϕN2,θ:=θN22ϕ,\phi^{\prime}:=\frac{\phi N}{2},\ \theta^{\prime}:=\theta-\frac{N-2}{2}\phi, (33)

and

I𝝁(k,l)=(1)n𝝁Jϕ,θ(k,l).I_{\bm{\mu}}^{(k,l)}=(-1)^{n_{\bm{\mu}}}J_{\phi^{\prime},\theta^{\prime}}^{(k,l)}. (34)

Define the following Bell polynomial

Iϕ,θ:=k=1N1(𝝁P~𝝁|𝟎(k,N)¯I𝝁(k,N)).I_{\phi^{\prime},\theta^{\prime}}:=\sum_{k=1}^{N-1}\Bigg{(}\sum_{\bm{\mu}}\tilde{P}^{\overline{(k,N)}}_{\bm{\mu}|\bm{0}}I_{\bm{\mu}}^{(k,N)}\Bigg{)}. (35)

If (ϕ,θ)(\phi^{\prime},\theta^{\prime})\in\mathcal{F}, Iϕ,θI_{\phi^{\prime},\theta^{\prime}} is an expanded Bell expression, and has quantum bounds ±ηN,ϕ,θQ\pm\eta_{N,\phi^{\prime},\theta^{\prime}}^{\mathrm{Q}} where ηN,ϕ,θQ=2(N1)sin2(θ+ϕ)sin(θϕ)\eta_{N,\phi^{\prime},\theta^{\prime}}^{\mathrm{Q}}=2(N-1)\sin^{2}(\theta^{\prime}+\phi^{\prime})\sin(\theta^{\prime}-\phi^{\prime}), which can be achieved up to relabelings by the strategy

ρ𝑸=|ψGHZψGHZ|,A0(k)=cosϕσXsinϕσY,A1(k)=cosθσX+sinθσY,k{1,,N}.\begin{gathered}\rho_{\bm{Q}}=|\psi_{\mathrm{GHZ}}\rangle\!\langle\psi_{\mathrm{GHZ}}|,\\ A_{0}^{(k)}=\cos\phi\,\sigma_{X}-\sin\phi\,\sigma_{Y},\\ A_{1}^{(k)}=\cos\theta\,\sigma_{X}+\sin\theta\,\sigma_{Y},\ k\in\{1,...,N\}.\end{gathered} (36)

In addition, this quantum bound cannot be achieved classically.

The proof is given in Section˜B.8. Note that the Bell expressions (35) are written in terms of the shifted parameters (ϕ,θ)(\phi^{\prime},\theta^{\prime}), instead of the measurement angles (ϕ,θ)(\phi,\theta). This is because the N2N-2 parties performing the projection no longer use σX\sigma_{X}, but use cos(ϕ)σXsin(ϕ)σY\cos(\phi)\,\sigma_{X}-\sin(\phi)\,\sigma_{Y}. This accumulates a phase factor on the state of the remaining parties, (|00+i(1)n𝝁ei(N2)ϕ|11)/2=:|Φ𝝁,ϕ(|00\rangle+i(-1)^{n_{\bm{\mu}}}e^{i(N-2)\phi}|11\rangle)/\sqrt{2}=:|\Phi_{\bm{\mu},\phi}\rangle, which is equivalent to the action of some local unitary UU on (|00+i(1)n𝝁|11)/2(|00\rangle+i(-1)^{n_{\bm{\mu}}}|11\rangle)/\sqrt{2}. To correct for this, we use the Bell expression Jϕ,θ(k,l)J_{\phi^{\prime},\theta^{\prime}}^{(k,l)} as the seed. See Section˜B.8 for the full details.

5.2 Randomness versus MABK value

Using the previously derived Bell inequality, we consider the trade-off between maximum device-independent randomness and MABK violation. Let us define

θ(ϕ)=N1N+1ϕ+θN,\theta(\phi)=\frac{N-1}{N+1}\phi+\theta^{*}_{N}, (37)

where θN\theta_{N}^{*} is defined in Eq.˜25 and let

ϕN=sgn[sin(2θN)]π4N.\phi_{N}^{*}=\mathrm{sgn}[\sin(2\theta_{N}^{*})]\frac{\pi}{4N}. (38)

For the strategy in Eq.˜31, direct calculation yields

MN(ϕN,θ(ϕN))=2(N1)/2,\langle M_{N}(\phi_{N}^{*},\theta(\phi_{N}^{*}))\rangle=2^{(N-1)/2}, (39)

and MN(0,θ(0))=mN\langle M_{N}(0,\theta(0))\rangle=m_{N}^{*}. Hence by varying ϕ[0,ϕN]\phi\in[0,\phi_{N}^{*}]555To ease notation, we write [0,ϕN][0,\phi_{N}^{*}] when ϕN>0\phi_{N}^{*}>0 and ϕN<0\phi_{N}^{*}<0, interpreting the latter case as [ϕN,0][\phi_{N}^{*},0]. and choosing θ=θ(ϕ)\theta=\theta(\phi) one can obtain the desired range of MABK values [mN,2(N1)/2][m_{N}^{*},2^{(N-1)/2}], from the conjectured maximum MABK value with maximum randomness to the maximum quantum value. Choosing this parameterization, the raw randomness, H(𝑹|𝑿=𝒙)=H({pϕ(𝒂|𝟎)})H(\bm{R}|\bm{X}=\bm{x}^{*})=H(\{p_{\phi}(\bm{a}|\bm{0})\}), of the strategy in Eq.˜31 (with θ=θ(ϕ)\theta=\theta(\phi)), as a function of ϕ\phi, is given by

H({pϕ(𝒂|𝟎)})r(ϕ)=N1+Hbin[1sinNϕ2],H(\{p_{\phi}(\bm{a}|\bm{0})\})\equiv r(\phi)=N-1+H_{\mathrm{bin}}\Big{[}\frac{1-\sin N\phi}{2}\Big{]}, (40)

where HbinH_{\mathrm{bin}} is the binary entropy function and r(ϕ)r(\phi) is a smooth, monotonically decreasing function of ϕ\phi in the range [0,ϕN][0,\phi_{N}^{*}].

As shown in Section˜4 (by combining Lemma˜5 with Proposition˜3), when ϕ=0\phi=0 we can certify maximum DI randomness. It also follows from the self-testing properties of the MABK family that when ϕ=ϕN\phi=\phi^{*}_{N} we obtain N1+3/2log2(1+2)/2N0.4N-1+3/2-\log_{2}(1+\sqrt{2})/\sqrt{2}\approx N-0.4 bits of global DI randomness when all parties use measurement 0. What remains then, is to apply the new Bell expression constructed in Lemma˜9 to certify r(ϕ)r(\phi) bits of DI randomness for ϕ(0,ϕN)\phi\in(0,\phi^{*}_{N}). To do so, the following proposition shows that for every ϕ[0,ϕN]\phi\in[0,\phi_{N}^{*}], there exists a valid Bell expression given by Lemma˜9.

Proposition 7.

Let ϕ[0,ϕN]\phi\in[0,\phi_{N}^{*}], and θ=θ(ϕ)\theta=\theta(\phi) as defined in Eqs.˜37 and 38. Then (ϕ,θ)(\phi^{\prime},\theta^{\prime})\in\mathcal{F}, where ϕ\phi^{\prime} and θ\theta^{\prime} are defined in Lemma˜9.

This is proven in Section˜B.9. Proposition˜7 implies that, for every ϕ\phi in the range we are interested in, the bipartite expression (1)n𝝁Jϕ,θ(-1)^{n_{\bm{\mu}}}J_{\phi^{\prime},\theta^{\prime}} is a valid self-test of the strategy

ρQkQl\displaystyle\rho_{Q_{k}Q_{l}} =|Φ𝝁Φ𝝁|,\displaystyle=|\Phi_{\bm{\mu}}\rangle\!\langle\Phi_{\bm{\mu}}|, (41)
A0(k)\displaystyle A_{0}^{(k)} =A0(l)=cosϕσXsinϕσY\displaystyle=A_{0}^{(l)}=\cos\phi^{\prime}\,\sigma_{X}-\sin\phi^{\prime}\,\sigma_{Y}
A1(k)\displaystyle A_{1}^{(k)} =A1(l)=cosθσX+sinθσY.\displaystyle=A_{1}^{(l)}=\cos\theta^{\prime}\,\sigma_{X}+\sin\theta^{\prime}\,\sigma_{Y}.

After applying the local unitary UϕUϕU_{\phi}\otimes U_{\phi}, where Uϕ=|00|+ei(N2)ϕ/2|11|U_{\phi}=|0\rangle\!\langle 0|+e^{i(N-2)\phi/2}|1\rangle\!\langle 1|, this is equivalent to

ρQkQl\displaystyle\rho_{Q_{k}Q_{l}} =|Φ𝝁,ϕΦ𝝁,ϕ|,\displaystyle=|\Phi_{\bm{\mu},\phi}\rangle\!\langle\Phi_{\bm{\mu},\phi}|, (42)
A0(k)\displaystyle A_{0}^{(k)} =A0(l)=cosϕσXsinϕσY\displaystyle=A_{0}^{(l)}=\cos\phi\,\sigma_{X}-\sin\phi\,\sigma_{Y}
A1(k)\displaystyle A_{1}^{(k)} =A1(l)=cosθσX+sinθσY,\displaystyle=A_{1}^{(l)}=\cos\theta\,\sigma_{X}+\sin\theta\,\sigma_{Y},

which is exactly the bipartite strategy we wanted to self-test, since it is the one held by parties kk and NN after the projector P𝝁|𝟎(k,N)¯P_{\bm{\mu}|\bm{0}}^{\overline{(k,N)}} is applied to the global state |ψGHZ|\psi_{\mathrm{GHZ}}\rangle. As a result, the correlations generated by the strategy in Eq.˜31, by choosing θ=θ(ϕ)\theta=\theta(\phi) and varying ϕ[0,ϕN]\phi\in[0,\phi_{N}^{*}], maximally violate the Bell inequality Jϕ,θJ_{\phi^{\prime},\theta^{\prime}} with ϕ\phi^{\prime} and θ\theta^{\prime} given in Eq.˜33. We can therefore employ the decoupling lemma to make the rate unconditioned on Eve, r(ϕ)r(\phi), device-independent.

Proposition 8.

Achieving the maximum quantum value of the Bell inequality in Lemma˜9 certifies r(ϕ)r(\phi) bits of randomness, i.e.,

RIϕ,θ(ηN,ϕ,θQ)=r(ϕ).R_{I_{\phi^{\prime},\theta^{\prime}}}(\eta^{\mathrm{Q}}_{N,\phi^{\prime},\theta^{\prime}})=r(\phi). (43)

We have established that, for every MABK value s[mN,2(N1)/2]s\in[m_{N}^{*},2^{(N-1)/2}], there exists a ϕs[0,ϕN]\phi_{s}\in[0,\phi_{N}^{*}] which defines a quantum strategy achieving ss, for which its generated randomness r(ϕs)r(\phi_{s}) is device-independent. We have therefore related every MABK value with a DI rate. We now conjecture the curve (s,r(ϕs))(s,r(\phi_{s})) is optimal in terms of raw randomness, which can then be made device-independent following the above discussion — see Lemma˜10.

Conjecture 2.

For NN even, the maximum randomness unconditioned on Eve, rr, that can be generated by quantum strategies achieving an MABK value, ss, is given by

r(s)={N,s(1,mN],r(ϕs),s(mN,2(N1)/2],r(s)=\begin{cases}N,\ s\in(1,m_{N}^{*}],\\ r(\phi_{s}),\ s\in(m_{N}^{*},2^{(N-1)/2}],\end{cases} (44)

where r(ϕ)r(\phi) is defined in Eq.˜40, and

ϕs=argmin{|ϕ|:ϕ[0,ϕN],MN(ϕ,θ(ϕ))=s}.\phi_{s}\!=\!\mathrm{arg\,min}\big{\{}|\phi|:\phi\in[0,\phi_{N}^{*}],\langle M_{N}(\phi,\theta(\phi))\rangle=s\big{\}}. (45)

For the range s(1,mN]s\in(1,m_{N}^{*}], and s(mN,2(N1)/2]s\in(m_{N}^{*},2^{(N-1)/2}], the rate r(s)r(s) and MABK value ss is achieved by the family of quantum strategies in Eq.˜20 and Eq.˜31, and certified device-independently by the Bell expressions in Lemma˜5 and Lemma˜9, respectively. Similarly to ˜1, Eq.˜45 is known to hold for the case N=2N=2 [23].

The minimization in the definition of ϕs\phi_{s} is included since we have not shown the set {ϕ[0,ϕN]:MN(ϕ,θ(ϕ))=s}\{\ \phi\in[0,\phi_{N}^{*}]\ :\ \langle M_{N}(\phi,\theta(\phi))\rangle=s\} is unique. Intuitively, the closer ϕs\phi_{s} is to zero the closer we are to maximum randomness; hence we expect r(ϕ)r(\phi) to be monotonically decreasing with |ϕ||\phi| for ϕ[0,ϕN]\phi\in[0,\phi_{N}^{*}], and minimization guarantees the best rate at a given ss. For the examples we have computed, the minimization turns out to be trivial, i.e., ϕs\phi_{s} is the unique solution to MN(ϕs,θ(ϕs))=s\langle M_{N}(\phi_{s},\theta(\phi_{s}))\rangle=s.

Lemma 10.

Suppose Eq.˜45 holds. Then the maximum amount of device-independent randomness that can be certified for the range of MABK values (1,2(N1)/2](1,2^{(N-1)/2}] is given by Eq.˜44.

The above lemma tells us that if the rate in Eq.˜44 is optimal without conditioning on Eve, it will also be optimal conditioned on Eve, i.e., the rate can be made device-independent. This is because Propositions˜3 and 7 ensure that, for every MABK value, the corresponding quantum strategy achieving it, given in Eq.˜20 or Eq.˜31, can generate certifiable randomness using the corresponding Bell expression IθI_{\theta} or Iϕ,θI_{\phi^{\prime},\theta^{\prime}}. A rewriting of Lemma˜10 without Conjecture 45, would say that (44) corresponds to a guaranteed lower bound on the maximum DI randomness as a function of MABK value. Our results are summarized in Fig.˜4.

Refer to caption
Figure 4: The conjectured curves of maximum device-independent randomness versus MABK value, where the MABK value is normalized by its maximum quantum value 2(N1)/22^{(N-1)/2}. When NN is odd, maximum global randomness can be achieved for any MABK value between the local and quantum bound, indicated by solid lines. The blue crosses indicate the conjectured maximum MABK value for which maximum randomness can be achieved when NN is even, mNm_{N}^{*}, which tends to the maximum quantum value as NN\rightarrow\infty. To the left of the blue crosses, NN bits of randomness can be achieved for MABK values between the local bound (not shown on this plot) and mNm_{N}^{*}; since this is the global maximum, it is optimal, indicated by solid lines. To the right of the blue crosses, dashed lines indicate lower bounds on the trade-off between maximum randomness and MABK value from mNm_{N}^{*} to the maximum quantum value, which are conjectured to be tight. The case of N=2N=2 was proven in Ref [23], which we reproduce with the results of the present paper.

We provide evidence for Conjectures 1 and 45 in Section˜B.10, where we derive a numerical technique for upper bounding the maximum randomness that can be generated by quantum strategies achieving a given MABK value ss. By studying the numerical results, we find our analytical lower bound closely agrees with the numerical upper bound.

5.3 Other extremal Bell inequalities

In the NN-partite 2-input 2-output scenario, the MABK family is just one class of extremal Bell inequalities, and the techniques developed in this section can be readily applied to others. For example, when N=3N=3 and restricting to Bell inequalities containing only three party correlators, there are a total of 3 non-trivial, new classes, one being the MABK inequality M3M_{3}, and two being of the form [41]

S1\displaystyle S_{1} =14x,y,zAxByCzA1B1C1,\displaystyle=\frac{1}{4}\sum_{x,y,z}A_{x}B_{y}C_{z}-A_{1}B_{1}C_{1}, (46)
S2\displaystyle S_{2} =A0B0(C0+C1)A1B1(C0C1),\displaystyle=A_{0}B_{0}(C_{0}+C_{1})-A_{1}B_{1}(C_{0}-C_{1}), (47)

where we used AxA_{x} for Ax1(1)A_{x_{1}}^{(1)}, ByB_{y} for Ax2(2)A_{x_{2}}^{(2)} etc. for legibility. S1S_{1} and S2S_{2} have local bounds of 11 and 22, and maximum quantum values 5/35/3 and 222\sqrt{2}, respectively. We applied our numerical technique (see Section˜B.10) to find upper bounds on the maximum amount of randomness certifiable whilst achieving a given violation of S1S_{1} and S2S_{2}. A comparison can be found in Fig.˜5, and we provide additional figures and details in Section˜B.11.

Refer to caption
Figure 5: Upper bounds on the trade-off between asymptotic global DI randomness and violation of non-trivial extremal Bell inequalities, for the tripartite scenario with two binary measurements per party. The violation has been normalized by the maximum quantum value. S1S_{1} and S2S_{2} are given by Eq.˜46 and Eq.˜47 respectively, and M3M_{3} is the MABK expression.

We find both S1S_{1} and S2S_{2} exhibit a trade-off for violations close to the quantum maximum. Due to its structure, S2S_{2} exhibits identical trade-off characteristics to the CHSH inequality found in [23], with the maximum CHSH value achievable with maximum randomness being numerically close to 33/23\sqrt{3}/2. On the other hand, we find more randomness can be generated from maximum violation of S1S_{1}, which is still strictly less than the maximum certified by MABK. Exact values are given in Section˜B.11.

6 Discussion

In this work, we studied how MABK violation constrains the generation of global DI randomness for an arbitrary number of parties. Whilst there is no trade-off in the odd case, we conjectured the precise trade-off for the even case. This conjecture is supported by an analytical lower bound and a numerical upper bound that agree up to several decimal places and have been checked for up to N=12N=12. We additionally showed that, in the asymptotic limit, this trade-off vanishes. Our main technical contribution was the extension of a recent tool, originally introduced to witness genuinely multi-partite nonlocality, to randomness certification.

The difference between the odd and even cases is explained by the odd MABK expressions containing half the full party correlators, allowing for input combinations not included in the Bell expression to be uniform. For even NN, all correlators are included and must be non-zero to obtain maximum violation, resulting in a trade-off between randomness and MABK value. The number of correlators is 2N2^{N}, whereas the quantum bound is 2(N1)/22^{(N-1)/2}. Maximum violation is achieved when the product of all correlators and their MABK coefficients have the same value, which in this case will be 2(N+1)/22^{-(N+1)/2}. Hence we find the penalty for fixing one correlator to zero (which is necessary for maximum randomness) diminishes as NN grows, and vanishes in the asymptotic limit.

We also quantified the nonlocality (measured using the dilution) of our constructions for low values of NN, calculated via a linear program. Interestingly we found our correlations are always separated from the local boundary (except for the case N=2N=2). It would be interesting to find correlations that lie arbitrarily close to the local boundary whilst still certifying maximum randomness, as was found in the N=2N=2 case [22, 23]. This could entail breaking the symmetry between the second measurements of each party, and the techniques presented here can be applied. Moreover, one could hope to go further and bound the trade-off between randomness and nonlocality in this setting; however, the challenge becomes finding a suitable measure of nonlocality that is efficient to work with. Specifically, for the case N=2N=2, constraining the amount of nonlocality of a quantum correlation is equivalent to constraining its CHSH value, which is linear in the correlators. When used as a constraint for maximizing randomness, this can then be handled using the techniques presented in Section˜B.10.1, which rely on the Navascués-Pironio-Acín hierarchy [52, 53]. For N>2N>2, the dilution measure is calculated via a linear program, therefore requiring a different approach.

We hope that the results found here will inform future experiments in DIRE; initial works have already made progress in this direction [54]. In such experiments, one should consider the optimal protocol for DIRE. In this work, we have proposed protocols that rely on spot-checking [7, 8, 45], where randomness is generated by a single input combination. An alternative approach has been explored in reference [45], where it was found that generating randomness from all inputs and averaging the result boosts the rate when the CHSH inequality is used. In the case of maximally violating a self-testing Bell inequality, spot-checking will always be optimal when a specific input combination leads to maximum randomness; having one maximally random setting necessarily implies that at least one other setting combination will not be maximally random for the correlations to be nonlocal, hence averaging will only decrease the overall rate. The noisy case remains an open question however, and the trade-off between using spot-checking and weighted averaging for our constructions deserves future investigation, with the aim of further improving practical rates.

Another interesting direction for future research is to connect expanded Bell inequalities to multi-partite self-testing. Our decoupling result allows the structure of the post-measurement state to be certified following the maximum violation of an expanded Bell expression; the open question we pose is whether an expanded Bell inequality can constitute a self-test of the NN-party state, opening up a new range of useful applications beyond randomness certification. Intuitively, if each pair of parties can self-test all of their measurements along with a projected state, the marginal information could be sufficient to make a statement about the global state. Such a strategy has already been adopted for self-testing certain multi-partite strategies [36]. However, there are subtleties to consider. For example, expanded Bell expressions tailored to some quantum strategies can result in trivial inequalities (i.e., ones that exhibit no classical-quantum gap), despite the underlying strategy being nonlocal, and furthermore self-testable. We therefore believe an additional ingredient is needed to self-test generic multi-partite states using the techniques in this work.

Finally, whilst we studied the MABK family of inequalities, one could consider a similar analysis for other families of extremal multi-partite Bell expressions [41]. For example, we explored upper bounds on the trade-offs for all extremal inequalities when N=3N=3, and found the MABK inequality to be optimal for global randomness; it would be interesting to find out if the upper bounds presented for the other inequalities are achievable. One might try to find explicit constructions that match these bounds, and use our techniques to make them device-independent. Analysing other extremal inequalities will build a more complete picture of how DI randomness can be generated in the multi-partite scenario, better informing the way forward for future multi-partite experiments.


Acknowledgements

This work was supported by the UK’s Engineering and Physical Sciences Research Council (EPSRC) via the Quantum Communications Hub (Grant No. EP/T001011/1) and Grant No. EP/SO23607/1.

References

  • [1] J. S. Bell. “Speakable and unspeakable in quantum mechanics”. Cambridge University Press.  (1987).
  • [2] Nicolas Brunner, Daniel Cavalcanti, Stefano Pironio, Valerio Scarani, and Stephanie Wehner. “Bell nonlocality”. Reviews of Modern Physics 86, 419–478 (2014).
  • [3] Jonathan Barrett, Noah Linden, Serge Massar, Stefano Pironio, Sandu Popescu, and David Roberts. “Nonlocal correlations as an information-theoretic resource”. Physical Review A 71, 022101 (2005).
  • [4] Roger Colbeck. “Quantum and relativistic protocols for secure multi-party computation”. PhD thesis. University of Cambridge.  (2007). arXiv:0911.3814.
  • [5] S. Pironio, A. Acin, S. Massar, A. Boyer de la Giroday, D. N. Matsukevich, P. Maunz, S. Olmschenk, D. Hayes, L. Luo, T. A. Manning, and C. Monroe. “Random numbers certified by Bell’s theorem”. Nature 464, 1021–1024 (2010).
  • [6] Roger Colbeck and Adrian Kent. “Private randomness expansion with untrusted devices”. Journal of Physics A 44, 095305 (2011).
  • [7] Carl A. Miller and Yaoyun Shi. “Robust protocols for securely expanding randomness and distributing keys using untrusted quantum devices”. Proceedings of the 46th Annual ACM Symposium on Theory of Computing (STOC ’14) (New York, USA) pages 417–426 (2014).
  • [8] Carl A. Miller and Yaoyun Shi. “Universal security for randomness expansion from the spot-checking protocol”. SIAM Journal of Computing 46, 1304–1335 (2017).
  • [9] Roger Colbeck and Renato Renner. “Free randomness can be amplified”. Nature Physics 8, 450–454 (2012).
  • [10] Artur K. Ekert. “Quantum cryptography based on Bell’s theorem”. Physical Review Letters 67, 661–663 (1991).
  • [11] John Barrett, Lucien Hardy, and Adrian Kent. “No signalling and quantum key distribution”. Physical Review Letters 95, 010503 (2005).
  • [12] Antonio Acin, Nicolas Brunner, Nicolas Gisin, Serge Massar, Stefano Pironio, and Valerio Scarani. “Device-independent security of quantum cryptography against collective attacks”. Physical Review Letters 98, 230501 (2007).
  • [13] Stefano Pironio, Antonio Acin, Nicolas Brunner, Nicolas Gisin, Serge Massar, and Valerio Scarani. “Device-independent quantum key distribution secure against collective attacks”. New Journal of Physics 11, 045021 (2009).
  • [14] Umesh Vazirani and Thomas Vidick. “Fully device-independent quantum key distribution”. Physical Review Letters 113, 140501 (2014).
  • [15] Rotem Arnon-Friedman, Frédéric Dupuis, Omar Fawzi, Renato Renner, and Thomas Vidick. “Practical device-independent quantum cryptography via entropy accumulation”. Nature Communications 9, 459 (2018).
  • [16] Dominic Mayers and Andrew Yao. “Self testing quantum apparatus” (2004). arXiv:quant-ph/0307205.
  • [17] M. McKague, T. H. Yang, and V. Scarani. “Robust self-testing of the singlet”. Journal of Physics A: Mathematical and Theoretical 45, 455304 (2012).
  • [18] Tzyh Haur Yang and Miguel Navascués. “Robust self-testing of unknown quantum systems into any entangled two-qubit states”. Physical Review A 87, 050102 (2013).
  • [19] Andrea Coladangelo, Koon Tong Goh, and Valerio Scarani. “All pure bipartite entangled states can be self-tested”. Nature Communications 8, 15485 (2017).
  • [20] Ivan Šupić, Joseph Bowles, Marc-Olivier Renou, Antonio Acín, and Matty J. Hoban. “Quantum networks self-test all entangled states”. Nature Physics 19, 670–675 (2023).
  • [21] Ivan Šupić and Joseph Bowles. “Self-testing of quantum systems: a review”. Quantum 4, 337 (2020).
  • [22] Antonio Acín, Serge Massar, and Stefano Pironio. “Randomness versus nonlocality and entanglement”. Physical Review Letters 108, 100402 (2012).
  • [23] Lewis Wooltorton, Peter Brown, and Roger Colbeck. “Tight analytic bound on the trade-off between device-independent randomness and nonlocality”. Physical Review Letters 129, 150403 (2022).
  • [24] N. David Mermin. “Extreme quantum entanglement in a superposition of macroscopically distinct states”. Physical Review Letters 65, 1838–1840 (1990).
  • [25] M. Ardehali. “Bell inequalities with a magnitude of violation that grows exponentially with the number of particles”. Physical Review A 46, 5375–5378 (1992).
  • [26] A. V. Belinskiĭ and D. N. Klyshko. “Interference of light and Bell’s theorem”. Physics-Uspekhi 36, 653 (1993).
  • [27] Erik Woodhead, Boris Bourdoncle, and Antonio Acín. “Randomness versus nonlocality in the Mermin-Bell experiment with three parties”. Quantum 2, 82 (2018).
  • [28] Federico Grasselli, Gláucia Murta, Hermann Kampermann, and Dagmar Bruß. “Entropy bounds for multiparty device-independent cryptography”. PRX Quantum 2, 010308 (2021).
  • [29] Federico Grasselli, Gláucia Murta, Hermann Kampermann, and Dagmar Bruß. “Boosting device-independent cryptography with tripartite nonlocality”. Quantum 7, 980 (2023).
  • [30] Cédric Bamps, Serge Massar, and Stefano Pironio. “Device-independent randomness generation with sublinear shared quantum resources”. Quantum 2, 86 (2018).
  • [31] Chirag Dhara, Giuseppe Prettico, and Antonio Acín. “Maximal quantum randomness in Bell tests”. Physical Review A 88, 052116 (2013).
  • [32] Gonzalo de la Torre, Matty J. Hoban, Chirag Dhara, Giuseppe Prettico, and Antonio Acín. “Maximally nonlocal theories cannot be maximally random”. Physical Review Letters 114, 160502 (2015).
  • [33] Samuel L. Braunstein and Carlton M. Caves. “Wringing out better Bell inequalities”. Annals of Physics 202, 22–56 (1990).
  • [34] I. Šupić, R. Augusiak, A. Salavrakos, and A. Acín. “Self-testing protocols based on the chained Bell inequalities”. New Journal of Physics 18, 035013 (2016).
  • [35] Florian J. Curchod, Mafalda L. Almeida, and Antonio Acín. “A versatile construction of Bell inequalities for the multipartite scenario”. New Journal of Physics 21, 023016 (2019).
  • [36] Ivan Šupić, A Coladangelo, R Augusiak, and A Acín. “Self-testing multipartite entangled states through projections onto two systems”. New Journal of Physics 20, 083041 (2018).
  • [37] M. Zwerger, W. Dür, J.-D. Bancal, and P. Sekatski. “Device-independent detection of genuine multipartite entanglement for all pure states”. Physical Review Letters 122, 060502 (2019).
  • [38] Xinhui Li, Yu Cai, Yunguang Han, Qiaoyan Wen, and Valerio Scarani. “Self-testing using only marginal information”. Physical Review A 98, 052331 (2018).
  • [39] T. Franz, F. Furrer, and R. F. Werner. “Extremal quantum correlations and cryptographic security”. Physical Review Letters 106, 250502 (2011).
  • [40] Vern Paulsen. “Completely bounded maps and operator algebras”. Cambridge Studies in Advanced Mathematics. Cambridge University Press.  (2003).
  • [41] R. F. Werner and M. M. Wolf. “All-multipartite Bell-correlation inequalities for two dichotomic observables per site”. Physical Review A 64, 032112 (2001).
  • [42] D. M. Greenberger, M. Horne, and A. Zeilinger. “Going beyond Bell’s theorem”. In M. Kafatos, editor, Bell’s Theorem, Quantum Mechanics and Conceptions of the Universe. Pages 69–72. Kluwer Academic, Dordrecht, The Netherlands (1989).
  • [43] Jȩdrzej Kaniewski. “Analytic and nearly optimal self-testing bounds for the Clauser-Horne-Shimony-Holt and Mermin inequalities”. Physical Review Letters 117, 070402 (2016).
  • [44] Jȩdrzej Kaniewski. “Self-testing of binary observables based on commutation”. Physical Review A 95, 062323 (2017).
  • [45] Rutvij Bhavsar, Sammy Ragy, and Roger Colbeck. “Improved device-independent randomness expansion rates using two sided randomness”. New Journal of Physics 25, 093035 (2023).
  • [46] Frédéric Dupuis, Omar Fawzi, and Renato Renner. “Entropy accumulation”. Communications in Mathematical Physics 379, 867–913 (2020).
  • [47] Frederic Dupuis and Omar Fawzi. “Entropy accumulation with improved second-order term”. IEEE Transactions on Information Theory 65, 7596–7612 (2019).
  • [48] Tony Metger, Omar Fawzi, David Sutter, and Renato Renner. “Generalised entropy accumulation”. Communications in Mathematical Physics 405, 261 (2024).
  • [49] Thinh P. Le, Chiara Meroni, Bernd Sturmfels, Reinhard F. Werner, and Timo Ziegler. “Quantum Correlations in the Minimal Scenario”. Quantum 7, 947 (2023).
  • [50] Victor Barizien, Pavel Sekatski, and Jean-Daniel Bancal. “Custom Bell inequalities from formal sums of squares”. Quantum 8, 1333 (2024).
  • [51] Lewis Wooltorton, Peter Brown, and Roger Colbeck. “Device-independent quantum key distribution with arbitrarily small nonlocality”. Physical Review Letters 132, 210802 (2024).
  • [52] Miguel Navascués, Stefano Pironio, and Antonio Acín. “Bounding the set of quantum correlations”. Physical Review Letters 98, 010401 (2007).
  • [53] Miguel Navascués, Stefano Pironio, and Antonio Acín. “A convergent hierarchy of semidefinite programs characterizing the set of quantum correlations”. New Journal of Physics 10, 073013 (2008).
  • [54] Alban Jean-Marie Seguinard, Amélie Piveteau, Piotr Mironowicz, and Mohamed Bourennane. “Experimental certification of more than one bit of quantum randomness in the two inputs and two outputs scenario”. New Journal of Physics 25, 113022 (2023).
  • [55] Marco Tomamichel. “Quantum information processing with finite resources”. Springer International Publishing.  (2016).
  • [56] C. Jordan. “Essai sur la géométrie à n dimensions”. Bulletin de la S. M. F. 3, 103–174 (1875).
  • [57] Armin Tavakoli, Alejandro Pozas-Kerstjens, Peter Brown, and Mateus Araújo. “Semidefinite programming relaxations for quantum correlations”. Reviews of Modern Physics 96, 045006 (2024).
  • [58] Koon Tong Goh, Jędrzej Kaniewski, Elie Wolfe, Tamás Vértesi, Xingyao Wu, Yu Cai, Yeong-Cherng Liang, and Valerio Scarani. “Geometry of the set of quantum correlations”. Physical Review A 97, 022104 (2018).
  • [59] Cédric Bamps and Stefano Pironio. “Sum-of-squares decompositions for a family of Clauser-Horne-Shimony-Holt-like inequalities and their application to self-testing”. Physical Review A 91, 052111 (2015).
  • [60] Cezary Śliwa. “Symmetries of the Bell correlation inequalities”. Physics Letters A 317, 165–168 (2003).
  • [61] Peter Brown, Hamza Fawzi, and Omar Fawzi. “Computing conditional entropies for quantum correlations”. Nature Communications 12, 575 (2021).

Appendix A Proofs for expanded Bell expressions

A.1 Proof of Lemma˜1

Lemma 1. Let II be an expanded Bell expression according to Footnote˜1. The maximum quantum value of I\langle I\rangle is upper bounded by ηNQ:=k,lck,lηQ\eta^{\mathrm{Q}}_{N}:=\sum_{k,l}c_{k,l}\eta^{\mathrm{Q}}.

Proof.

To prove the above, we show that the operator expression I¯=ηNQ𝕀I\bar{I}=\eta^{\mathrm{Q}}_{N}\mathbb{I}-I is non-negative. Since each bipartite expression satisfies I𝝁(k,l)ηQ\langle I_{\bm{\mu}}^{(k,l)}\rangle\leq\eta^{\mathrm{Q}} for all 𝝁,k,l\bm{\mu},k,l, we have positive expressions

I¯𝝁(k,l)=ηQ𝕀I𝝁(k,l)0,𝝁,k,l.\bar{I}_{\bm{\mu}}^{(k,l)}=\eta^{\mathrm{Q}}\mathbb{I}-I_{\bm{\mu}}^{(k,l)}\succeq 0,\ \forall\bm{\mu},k,l. (48)

We then have

I¯k,l\displaystyle\bar{I}_{k,l} :=ηQ𝕀𝝁P~𝝁|𝟎(k,l)¯I𝝁(k,l)\displaystyle:=\eta^{\mathrm{Q}}\mathbb{I}-\sum_{\bm{\mu}}\tilde{P}_{\bm{\mu}|\bm{0}}^{\overline{(k,l)}}I^{(k,l)}_{\bm{\mu}}
=𝝁P~𝝁|𝟎(k,l)¯(ηQ𝕀I𝝁(k,l))\displaystyle=\sum_{\bm{\mu}}\tilde{P}_{\bm{\mu}|\bm{0}}^{\overline{(k,l)}}\Big{(}\eta^{\mathrm{Q}}\mathbb{I}-I_{\bm{\mu}}^{(k,l)}\Big{)}
=𝝁P~𝝁|𝟎(k,l)¯I¯𝝁(k,l)0.\displaystyle=\sum_{\bm{\mu}}\tilde{P}_{\bm{\mu}|\bm{0}}^{\overline{(k,l)}}\bar{I}_{\bm{\mu}}^{(k,l)}\succeq 0. (49)

Finally it follows from the above that I¯0\bar{I}\succeq 0, which proves the upper bound on the maximum quantum value. We can also upper bound the maximum local value by k<lck,lηL<k<lck,lηQ=ηNQ\sum_{k<l}c_{k,l}\eta^{\mathrm{L}}<\sum_{k<l}c_{k,l}\eta^{\mathrm{Q}}=\eta_{N}^{\mathrm{Q}} since ηL<ηQ\eta^{\mathrm{L}}<\eta^{\mathrm{Q}}. ∎

A.2 Uniqueness of binary distributions with fixed marginals

Before proving Eq.˜13, we establish the following fact about classical distributions of bit strings when their conditional distributions are fixed. We will later use this result to justify that the distribution used for randomness achieving the quantum bound of an expanded Bell inequality is unique when the seed is a self-test.

Lemma 11.

Let 𝐀=A1AN\bm{A}=A_{1}...A_{N}, N3N\geq 3, be a random NN bit string which takes values 𝐚{0,1}N\bm{a}\in\{0,1\}^{N} according to the distribution p𝐀(𝐚)p_{\bm{A}}(\bm{a}). Let 𝐀k¯\bm{A}_{\overline{k}} be an N2N-2 bit partition of the string 𝐀\bm{A}, excluding bits NN and k{1,,N1}k\in\{1,...,N-1\}, taking values 𝐚k¯{0,1}N2\bm{a}_{\overline{k}}\in\{0,1\}^{N-2}. Then the distribution p𝐀(𝐚)p_{\bm{A}}(\bm{a}) is entirely determined by the set of conditional distributions {pAkAN|𝐀k¯(akaN|𝐚k¯)}k{1,,N1}\{p_{A_{k}A_{N}|\bm{A}_{\overline{k}}}(a_{k}a_{N}|\bm{a}_{\overline{k}})\}_{k\in\{1,...,N-1\}}, provided pAkAN|𝐀k¯(akaN|𝐚k¯)>0p_{A_{k}A_{N}|\bm{A}_{\overline{k}}}(a_{k}a_{N}|\bm{a}_{\overline{k}})>0 for all k,ak,aN,𝐚k¯k,a_{k},a_{N},\bm{a}_{\overline{k}}.

Proof.

Proof can be established using a recursive argument to solve for a set of marginal terms p𝑨l¯(𝒂l¯)p_{\bm{A}_{\overline{l}}}(\bm{a}_{\overline{l}}), for a fixed l{1,,N1}l\in\{1,...,N-1\}, from which the full distribution can be recovered. Let us choose another partition k{1,,N1}k\in\{1,...,N-1\} with klk\neq l. Then we have

p𝑨(𝒂)\displaystyle p_{\bm{A}}(\bm{a}) =p𝑨k¯(𝒂k¯)pAkAN|𝑨k¯(akaN|𝒂k¯)\displaystyle=p_{\bm{A}_{\overline{k}}}(\bm{a}_{\overline{k}})p_{A_{k}A_{N}|\bm{A}_{\overline{k}}}(a_{k}a_{N}|\bm{a}_{\overline{k}}) (50)
=p𝑨l¯(𝒂l¯)pAlAN|𝑨l¯(alaN|𝒂l¯).\displaystyle=p_{\bm{A}_{\overline{l}}}(\bm{a}_{\overline{l}})p_{A_{l}A_{N}|\bm{A}_{\overline{l}}}(a_{l}a_{N}|\bm{a}_{\overline{l}}).

By choosing different values of aka_{k}, this implies the two sets of equations

p𝑨k¯(𝒂k¯)\displaystyle p_{\bm{A}_{\overline{k}}}(\bm{a}_{\overline{k}}) pAkAN|𝑨k¯(ak=0,aN|𝒂k¯)\displaystyle p_{A_{k}A_{N}|\bm{A}_{\overline{k}}}(a_{k}=0,a_{N}|\bm{a}_{\overline{k}}) (51)
=p𝑨l¯(ak=0,𝒂kl¯)pAlAN|𝑨l¯(alaN|ak=0,𝒂kl¯),\displaystyle=p_{\bm{A}_{\overline{l}}}(a_{k}=0,\bm{a}_{\overline{kl}})\,p_{A_{l}A_{N}|\bm{A}_{\overline{l}}}(a_{l}a_{N}|a_{k}=0,\bm{a}_{\overline{kl}}),
p𝑨k¯(𝒂k¯)\displaystyle p_{\bm{A}_{\overline{k}}}(\bm{a}_{\overline{k}}) pAkAN|𝑨k¯(ak=1,aN|𝒂k¯)\displaystyle p_{A_{k}A_{N}|\bm{A}_{\overline{k}}}(a_{k}=1,a_{N}|\bm{a}_{\overline{k}})
=p𝑨l¯(ak=1,𝒂kl¯)pAlAN|𝑨l¯(alaN|ak=1,𝒂kl¯).\displaystyle=p_{\bm{A}_{\overline{l}}}(a_{k}=1,\bm{a}_{\overline{kl}})\,p_{A_{l}A_{N}|\bm{A}_{\overline{l}}}(a_{l}a_{N}|a_{k}=1,\bm{a}_{\overline{kl}}).

Since pAkAN|𝑨k¯(akaN|𝒂k¯)0p_{A_{k}A_{N}|\bm{A}_{\overline{k}}}(a_{k}a_{N}|\bm{a}_{\overline{k}})\neq 0, we can rearrange both for p𝑨k¯p_{\bm{A}_{\overline{k}}} and equate:

p𝑨l¯(ak=0,𝒂kl¯)pAlAN|𝑨l¯(alaN|ak=0,𝒂kl¯)pAkAN|𝑨k¯(ak=0,aN|𝒂k¯)=p𝑨l¯(ak=1,𝒂kl¯)pAlAN|𝑨l¯(alaN|ak=1,𝒂kl¯)pAkAN|𝑨k¯(ak=1,aN|𝒂k¯),\frac{p_{\bm{A}_{\overline{l}}}(a_{k}=0,\bm{a}_{\overline{kl}})\,p_{A_{l}A_{N}|\bm{A}_{\overline{l}}}(a_{l}a_{N}|a_{k}=0,\bm{a}_{\overline{kl}})}{p_{A_{k}A_{N}|\bm{A}_{\overline{k}}}(a_{k}=0,a_{N}|\bm{a}_{\overline{k}})}\\ =\frac{p_{\bm{A}_{\overline{l}}}(a_{k}=1,\bm{a}_{\overline{kl}})\,p_{A_{l}A_{N}|\bm{A}_{\overline{l}}}(a_{l}a_{N}|a_{k}=1,\bm{a}_{\overline{kl}})}{p_{A_{k}A_{N}|\bm{A}_{\overline{k}}}(a_{k}=1,a_{N}|\bm{a}_{\overline{k}})}, (52)

where we have explicitly written in the aka_{k} variable, and kl¯\overline{kl} denotes a tuple excluding parties kk and ll. The above equation tells us, for every choice of 𝒂kl¯{0,1}N3\bm{a}_{\overline{kl}}\in\{0,1\}^{N-3}, the unknowns p𝑨l¯(ak=0,𝒂kl¯)p_{\bm{A}_{\overline{l}}}(a_{k}=0,\bm{a}_{\overline{kl}}) and p𝑨l¯(ak=1,𝒂kl¯)p_{\bm{A}_{\overline{l}}}(a_{k}=1,\bm{a}_{\overline{kl}}) are linearly dependent. Hence we consider the 2N32^{N-3} unknowns p𝑨l¯(ak=0,𝒂kl¯)p_{\bm{A}_{\overline{l}}}(a_{k}=0,\bm{a}_{\overline{kl}}), since the ak=1a_{k}=1 terms can be computed via this linear dependence.

Consider choosing k{1,,N1}{k,l}k^{\prime}\in\{1,...,N-1\}\setminus\{k,l\}. We can find a new set of equations by repeating the above process with kk^{\prime} instead of kk, to find:

p𝑨l¯(ak=0,ak=0,𝒂kkl¯)pAkAN|𝑨k¯(ak=0,aN|ak=0,𝒂kk¯)\displaystyle\frac{p_{\bm{A}_{\overline{l}}}(a_{k}=0,a_{k^{\prime}}=0,\bm{a}_{\overline{kk^{\prime}l}})}{p_{A_{k^{\prime}}A_{N}|\bm{A}_{\bar{k^{\prime}}}}(a_{k^{\prime}}=0,a_{N}|a_{k}=0,\bm{a}_{\overline{kk^{\prime}}})} (53)
pAlAN|𝑨l¯(alaN|ak=0,ak=0,𝒂kkl¯)\displaystyle\hskip 22.76228pt\cdot p_{A_{l}A_{N}|\bm{A}_{\overline{l}}}(a_{l}a_{N}|a_{k}=0,a_{k^{\prime}}=0,\bm{a}_{\overline{kk^{\prime}l}})
=p𝑨l¯(ak=0,ak=1,𝒂kkl¯)pAkAN|𝑨k¯(ak=1,aN|ak=0,𝒂kk¯)\displaystyle=\frac{p_{\bm{A}_{\overline{l}}}(a_{k}=0,a_{k^{\prime}}=1,\bm{a}_{\overline{kk^{\prime}l}})}{p_{A_{k^{\prime}}A_{N}|\bm{A}_{\bar{k^{\prime}}}}(a_{k^{\prime}}=1,a_{N}|a_{k}=0,\bm{a}_{\overline{kk^{\prime}}})}
pAlAN|𝑨l¯(alaN|ak=0,ak=1,𝒂kkl¯),\displaystyle\hskip 22.76228pt\cdot p_{A_{l}A_{N}|\bm{A}_{\overline{l}}}(a_{l}a_{N}|a_{k}=0,a_{k^{\prime}}=1,\bm{a}_{\overline{kk^{\prime}l}}),

where we have only included the ak=0a_{k}=0 case, and written kkl¯\overline{kk^{\prime}l} to denote a tuple excluding parties k,k,lk,k^{\prime},l. Notice we have now identified linear dependence between p𝑨l¯(ak=0,ak=0,𝒂kkl¯)p_{\bm{A}_{\overline{l}}}(a_{k}=0,a_{k^{\prime}}=0,\bm{a}_{\overline{kk^{\prime}l}}) and p𝑨l¯(ak=0,ak=1,𝒂kkl¯)p_{\bm{A}_{\overline{l}}}(a_{k}=0,a_{k^{\prime}}=1,\bm{a}_{\overline{kk^{\prime}l}}). Neither of these equations contain a marginal with ak=1a_{k}=1, so they relate pairs of unknowns distinctly to the parings from the previous equations, and are hence linearly independent. As before, we can proceed with the 2N42^{N-4} unknowns p𝑨l¯(ak=0,ak=0,𝒂kkl¯)p_{\bm{A}_{\overline{l}}}(a_{k}=0,a_{k^{\prime}}=0,\bm{a}_{\overline{kk^{\prime}l}}) and recover the others by linear dependence.

We can apply the above procedure iteratively a total of N2N-2 times (for every k{1,,N1}{l}k\in\{1,...,N-1\}\setminus\{l\}), halving the number of unknowns on every iteration by establishing linear dependence. Starting with 2N22^{N-2} marginals p𝑨l¯(𝒂l¯)p_{\bm{A}_{\overline{l}}}(\bm{a}_{\overline{l}}), this leaves us with 1 unknown, which can be solved via normalization, completing the proof. ∎

Now we can present a corollary which will be needed for the proof of Eq.˜13, showing the conditional distribution used for randomness, p(𝒂|𝟎)p(\bm{a}|\bm{0}), achieving the quantum bound of an expanded Bell expression is unique.

Corollary 1.

Let II be an expanded Bell inequality according to Footnote˜1 with ck,N=1c_{k,N}=1 if k<Nk<N and zero otherwise, and ηNQ\eta_{N}^{\mathrm{Q}} be as defined in Lemma˜1. Let II be constructed using a seed with the self-testing properties described in Eq.˜13, and p(𝐚|𝐱)p(\bm{a}|\bm{x}) denote a quantum behaviour that achieves I=ηNQ\langle I\rangle=\eta^{\mathrm{Q}}_{N}. Then provided p(akaN|𝐛k¯,𝐱=𝟎)>0p(a_{k}a_{N}|\bm{b}_{\overline{k}},\bm{x}=\bm{0})>0 for all k,ak,aN,𝐛k¯k,a_{k},a_{N},\bm{b}_{\overline{k}}, the conditional the distribution p(𝐚|𝟎)p(\bm{a}|\bm{0}) is unique.

Proof.

If the quantum value I=ηNQ\langle I\rangle=\eta_{N}^{\mathrm{Q}} is achieved, we must have

k=1N1ηQ𝕀𝝁P~𝝁|𝟎(k,N)¯I𝝁(k,N)=0,\displaystyle\sum_{k=1}^{N-1}\big{\langle}\eta^{\mathrm{Q}}\mathbb{I}-\sum_{\bm{\mu}}\tilde{P}_{\bm{\mu}|\bm{0}}^{\overline{(k,N)}}I_{\bm{\mu}}^{(k,N)}\big{\rangle}=0, (54)

which implies ηQ𝕀𝝁P~𝝁|𝟎(k,N)¯I𝝁(k,N)=0\big{\langle}\eta^{\mathrm{Q}}\mathbb{I}-\sum_{\bm{\mu}}\tilde{P}_{\bm{\mu}|\bm{0}}^{\overline{(k,N)}}I_{\bm{\mu}}^{(k,N)}\big{\rangle}=0 from Eq.˜49. Next, we observe

P~𝝁|𝟎(k,N)¯I𝝁(k,N)=Tr[P~𝝁|𝟎(k,N)¯ρ𝑸~EP~𝝁|𝟎(k,N)¯I𝝁(k,N)]|Tr[P~𝝁|𝟎(k,N)¯ρ𝑸~EP~𝝁|𝟎(k,N)¯I𝝁(k,N)]|P~𝝁|𝟎(k,N)¯ρ𝑸~EP~𝝁|𝟎(k,N)¯I𝝁(k,N)1,\langle\tilde{P}_{\bm{\mu}|\bm{0}}^{\overline{(k,N)}}I_{\bm{\mu}}^{(k,N)}\rangle=\mathrm{Tr}\Big{[}\tilde{P}_{\bm{\mu}|\bm{0}}^{\overline{(k,N)}}\rho_{\tilde{\bm{Q}}E}\tilde{P}_{\bm{\mu}|\bm{0}}^{\overline{(k,N)}}I_{\bm{\mu}}^{(k,N)}\Big{]}\\ \leq|\mathrm{Tr}\Big{[}\tilde{P}_{\bm{\mu}|\bm{0}}^{\overline{(k,N)}}\rho_{\tilde{\bm{Q}}E}\tilde{P}_{\bm{\mu}|\bm{0}}^{\overline{(k,N)}}I_{\bm{\mu}}^{(k,N)}\Big{]}|\\ \leq\|\tilde{P}_{\bm{\mu}|\bm{0}}^{\overline{(k,N)}}\rho_{\tilde{\bm{Q}}E}\tilde{P}_{\bm{\mu}|\bm{0}}^{\overline{(k,N)}}I_{\bm{\mu}}^{(k,N)}\|_{1}, (55)

where the second inequality follows from the fact |Tr[A]|A1|\mathrm{Tr}[A]|\leq\|A\|_{1} for any operator AA (see e.g. [55, Lemma 3.3]). Above, ρ𝑸~E=|ΨΨ|\rho_{\tilde{\bm{Q}}E}=|\Psi\rangle\!\langle\Psi|, where |Ψ|\Psi\rangle is the quantum state that realizes the correlations, and identities on EE are implicit. Next we combine Eq.˜55 with Hölder’s inequality to obtain

P~𝝁|𝟎(k,N)¯I𝝁(k,N)P~𝝁|𝟎(k,N)¯ρ𝑸~EP~𝝁|𝟎(k,N)¯I𝝁(k,N)1P~𝝁|𝟎(k,N)¯ρ𝑸~EP~𝝁|𝟎(k,N)¯1I𝝁(k,N)=P~𝝁|𝟎(k,N)¯ηQ.\langle\tilde{P}_{\bm{\mu}|\bm{0}}^{\overline{(k,N)}}I_{\bm{\mu}}^{(k,N)}\rangle\leq\|\tilde{P}_{\bm{\mu}|\bm{0}}^{\overline{(k,N)}}\rho_{\tilde{\bm{Q}}E}\tilde{P}_{\bm{\mu}|\bm{0}}^{\overline{(k,N)}}I_{\bm{\mu}}^{(k,N)}\|_{1}\\ \leq\|\tilde{P}_{\bm{\mu}|\bm{0}}^{\overline{(k,N)}}\rho_{\tilde{\bm{Q}}E}\tilde{P}_{\bm{\mu}|\bm{0}}^{\overline{(k,N)}}\|_{1}\|I_{\bm{\mu}}^{(k,N)}\|_{\infty}=\langle\tilde{P}_{\bm{\mu}|\bm{0}}^{\overline{(k,N)}}\rangle\eta^{\mathrm{Q}}. (56)

Since

𝝁(P~𝝁|𝟎(k,N)¯ηQP~𝝁|𝟎(k,N)¯I𝝁(k,N))=0\sum_{\bm{\mu}}\Big{(}\langle\tilde{P}_{\bm{\mu}|\bm{0}}^{\overline{(k,N)}}\rangle\eta^{\mathrm{Q}}-\langle\tilde{P}_{\bm{\mu}|\bm{0}}^{\overline{(k,N)}}I_{\bm{\mu}}^{(k,N)}\rangle\Big{)}=0 (57)

we conclude P~𝝁|𝟎(k,N)¯I𝝁(k,N)=P~𝝁|𝟎(k,N)¯ηQ\langle\tilde{P}_{\bm{\mu}|\bm{0}}^{\overline{(k,N)}}I_{\bm{\mu}}^{(k,N)}\rangle=\langle\tilde{P}_{\bm{\mu}|\bm{0}}^{\overline{(k,N)}}\rangle\eta^{\mathrm{Q}}. Now we can write

Ψ|I𝝁(k,N)|Ψ=ηQ,\langle\Psi^{\prime}|I_{\bm{\mu}}^{(k,N)}|\Psi^{\prime}\rangle=\eta^{\text{Q}}, (58)

where we define the normalized state |Ψ=P~𝝁|𝟎(k,N)¯|Ψ/P~𝝁|𝟎(k,N)¯|\Psi^{\prime}\rangle=\tilde{P}_{\bm{\mu}|\bm{0}}^{\overline{(k,N)}}|\Psi\rangle/\sqrt{\langle\tilde{P}_{\bm{\mu}|\bm{0}}^{\overline{(k,N)}}\rangle}. The self-testing properties of I𝝁(k,N)I_{\bm{\mu}}^{(k,N)} (see definition in Eq.˜13) then determine the unique correlations Ψ|P~ak|x(k)P~aN|y(N)|Ψ=p𝝁k,N(akaN|xy)\langle\Psi^{\prime}|\tilde{P}_{a_{k}|x}^{(k)}\tilde{P}_{a_{N}|y}^{(N)}|\Psi^{\prime}\rangle=p^{k,N}_{\bm{\mu}}(a_{k}a_{N}|xy). Using the definition of |Ψ|\Psi^{\prime}\rangle, we find

P~𝝁|𝟎(k,N)¯P~ak|0(k)P~aN|0(N)=P~𝝁|𝟎(k,N)¯p𝝁k,N(akaN|00).\big{\langle}\tilde{P}_{\bm{\mu}|\bm{0}}^{\overline{(k,N)}}\tilde{P}_{a_{k}|0}^{(k)}\tilde{P}_{a_{N}|0}^{(N)}\big{\rangle}=\langle\tilde{P}_{\bm{\mu}|\bm{0}}^{\overline{(k,N)}}\rangle p^{k,N}_{\bm{\mu}}(a_{k}a_{N}|00). (59)

We can now directly apply Lemma˜11. Explicitly we have p𝑨(𝒂)=P~𝝁|𝟎(k,N)¯P~ak|0(k)P~aN|0(N)p_{\bm{A}}(\bm{a})=\big{\langle}\tilde{P}_{\bm{\mu}|\bm{0}}^{\overline{(k,N)}}\tilde{P}_{a_{k}|0}^{(k)}\tilde{P}_{a_{N}|0}^{(N)}\big{\rangle}, and the N1N-1 fixed conditional distributions pAkAN|𝑨k¯(akaN|𝒂k¯)=p(akaN|𝒂k¯,𝒙=𝟎)=p𝝁k,N(akaN|00)p_{A_{k}A_{N}|\bm{A}_{\overline{k}}}(a_{k}a_{N}|\bm{a}_{\overline{k}})=p(a_{k}a_{N}|\bm{a}_{\overline{k}},\bm{x}=\bm{0})=p^{k,N}_{\bm{\mu}}(a_{k}a_{N}|00), where we have identified 𝝁=𝒂k¯\bm{\mu}=\bm{a}_{\overline{k}}.

A.3 Proof of Eq.˜13

We can now proceed to prove Eq.˜13 in the main text, which is restated below:

Lemma 2. Let II be an expanded NN-party Bell expression defined in Footnote˜1 with binary inputs and outputs, and ck,N=1c_{k,N}=1 if k<Nk<N and zero otherwise. Suppose for every I𝛍(k,N)I_{\bm{\mu}}^{(k,N)}, there exists an SOS decomposition which self-tests the same pure bipartite entangled state |Φ|\Phi\rangle between parties k,Nk,N, along with some ideal measurements Pak|xk(k),PaN|xN(N)P^{(k)}_{a_{k}|x_{k}},P^{(N)}_{a_{N}|x_{N}}, according to Eq.˜3, which satisfy Φ|Pak|0(k)PaN|0(N)|Φ>0\langle\Phi|P^{(k)}_{a_{k}|0}\otimes P^{(N)}_{a_{N}|0}|\Phi\rangle>0 for all ak,aNa_{k},a_{N}. Then for any strategy |Ψ𝐐~E,{{P~ak|xk(k)}xk}k|\Psi\rangle_{\tilde{\bm{Q}}E},\big{\{}\{\tilde{P}_{a_{k}|x_{k}}^{(k)}\}_{x_{k}}\big{\}}_{k} that achieves I=ηNQ\langle I\rangle=\eta^{\mathrm{Q}}_{N}, the post-measurement state ρ𝐑E\rho_{\bm{R}E}, for measurement settings 𝐱=𝟎\bm{x}=\bm{0}, admits the tensor product decomposition,

ρ𝑹E|𝟎=ρ𝑹|𝟎ρE.\rho_{\bm{R}E|\bm{0}}=\rho_{\bm{R}|\bm{0}}\otimes\rho_{E}. (60)
Proof.

We will break this proof up into four steps. First, we will provide an SOS decomposition for the Bell operator I¯\bar{I}, and derive the algebraic constraints satisfied by any state and measurements that achieves its maximum value. These turn out to be those implied by the bipartite Bell expressions that build up II. Next, we will employ Jordan’s lemma, reducing the problem to qubits, and show how these constraints self-test the bipartite states. Finally we write down the global post-measurement state, and show how these self-tests imply a tensor product with EE.

A.3.1 SOS decomposition

By assumption, there exist SOS decompositions for the bipartite Bell operators I¯𝝁(k,N)\bar{I}_{\bm{\mu}}^{(k,N)}, i.e., we can write

I¯𝝁(k,N)=iM𝝁,i(k,N)M𝝁,i(k,N),\bar{I}_{\bm{\mu}}^{(k,N)}=\sum_{i}M_{\bm{\mu},i}^{(k,N)\dagger}M_{\bm{\mu},i}^{(k,N)}, (61)

where M𝝁,i(k,N)M_{\bm{\mu},i}^{(k,N)} is a polynomial of the operators Axk(k)A_{x_{k}}^{(k)}, AxN(N)A_{x_{N}}^{(N)}. We now use this to build an SOS decomposition for I¯\bar{I},

I¯\displaystyle\bar{I} =k=1N1I¯k,N\displaystyle=\sum_{k=1}^{N-1}\bar{I}_{k,N}
=k=1N1𝝁P~𝝁|𝟎(k,N)¯I¯𝝁(k,N)\displaystyle=\sum_{k=1}^{N-1}\sum_{\bm{\mu}}\tilde{P}_{\bm{\mu}|\bm{0}}^{\overline{(k,N)}}\bar{I}_{\bm{\mu}}^{(k,N)}
=k=1N1𝝁i(P~𝝁|𝟎(k,N)¯M𝝁,i(k,N))(P~𝝁|𝟎(k,N)¯M𝝁,i(k,N)).\displaystyle=\sum_{k=1}^{N-1}\sum_{\bm{\mu}}\sum_{i}\Big{(}\tilde{P}_{\bm{\mu}|\bm{0}}^{\overline{(k,N)}}M_{\bm{\mu},i}^{(k,N)}\Big{)}^{\dagger}\Big{(}\tilde{P}_{\bm{\mu}|\bm{0}}^{\overline{(k,N)}}M_{\bm{\mu},i}^{(k,N)}\Big{)}. (62)

Now, for some physical state |Ψ|\Psi\rangle and observables, observation of the maximum quantum value, I¯=0\langle\bar{I}\rangle=0, immediately implies the algebraic constraints

P~𝝁|𝟎(k,N)¯M𝝁,i(k,N)|Ψ=0,k{1,,N1},𝝁,i.\tilde{P}_{\bm{\mu}|\bm{0}}^{\overline{(k,N)}}M_{\bm{\mu},i}^{(k,N)}|\Psi\rangle=0,\ \forall k\in\{1,...,N-1\},\bm{\mu},i. (63)

Since P~𝝁|𝟎(k,N)¯\tilde{P}_{\bm{\mu}|\bm{0}}^{\overline{(k,N)}} and M𝝁,i(k,N)M_{\bm{\mu},i}^{(k,N)} commute, we define the post-measurement states

|Ψ𝝁,k=P~𝝁|𝟎(k,N)¯|ΨΨ|P~𝝁|𝟎(k,N)¯|Ψ,|\Psi_{\bm{\mu},k}\rangle=\frac{\tilde{P}_{\bm{\mu}|\bm{0}}^{\overline{(k,N)}}|\Psi\rangle}{\sqrt{\langle\Psi|\tilde{P}_{\bm{\mu}|\bm{0}}^{\overline{(k,N)}}|\Psi\rangle}}, (64)

and arrive at the set of algebraic constraints, for a fixed k,𝝁k,\bm{\mu},

M𝝁,i(k,N)|Ψ𝝁,k=0,i.M_{\bm{\mu},i}^{(k,N)}|\Psi_{\bm{\mu},k}\rangle=0,\quad\forall i. (65)

A.3.2 Applying Jordan’s lemma

Next, we employ Jordan’s lemma [56], which allows us to reduce our analysis to a convex combination of NN-qubit systems. Specifically, we will use the version presented in [45, Lemma 4], and apply [45, Lemma 6] to consider (without loss of generality) the set of post-measurement states ρ𝑹E\rho_{\bm{R}E} arising from block diagonal measurement operators with block size 2×22\times 2, and a state which has support only on each 2×22\times 2 block. We write Qk\mathcal{H}_{Q_{k}} to denote the qubit Hilbert space of party kk, and introduce a flag system Fk\mathcal{H}_{F_{k}} which indicates the 2×22\times 2 block. Then for every party Q~k=QkFk\mathcal{H}_{\tilde{Q}_{k}}=\mathcal{H}_{Q_{k}}\otimes\mathcal{H}_{F_{k}}, and the purified state takes the form

|Ψ𝑸~E=𝒇p𝒇|𝒇𝑭|φ𝒇𝑸E|𝒇E,|\Psi\rangle_{\tilde{\bm{Q}}E}=\sum_{\bm{f}}\sqrt{p_{\bm{f}}}|\bm{f}\rangle_{\bm{F}}\otimes|\varphi^{\bm{f}}\rangle_{\bm{Q}E}\otimes|\bm{f}\rangle_{E^{\prime}}, (66)

where ρ𝒇=TrE[|φ𝒇φ𝒇|]\rho^{\bm{f}}=\mathrm{Tr}_{E}\big{[}|\varphi^{\bm{f}}\rangle\!\langle\varphi^{\bm{f}}|\big{]} is an NN-qubit state corresponding to block 𝒇=(f1,,fN)\bm{f}=(f_{1},...,f_{N}), where fkf_{k} indexes the block for party kk and p𝒇>0p_{\bm{f}}>0 form a probability distribution. Tracing out EEEE^{\prime}, the total state received by the parties is then a classical quantum state, where the classical register indexes which NN-qubit state is in the quantum system. Eve then holds a purification of the state in register EE along with a label in register EE^{\prime} telling her which block it pertains to. Similarly, the projectors decompose

P~ak|xk(k)=fk|fkfk|FkPak|xk(fk),\tilde{P}_{a_{k}|x_{k}}^{(k)}=\sum_{f_{k}}|f_{k}\rangle\!\langle f_{k}|_{F_{k}}\otimes P_{a_{k}|x_{k}}^{(f_{k})}, (67)

where Pak|xk(fk)P_{a_{k}|x_{k}}^{(f_{k})} is a rank one single qubit projector on Qk\mathcal{H}_{Q_{k}} corresponding to block fkf_{k} [56, 13, 45]. The projector performed on all systems except kk and NN for inputs 0 is

P~𝝁|𝟎(k,N)¯=𝒇(k,N¯)|𝒇(k,N¯)𝒇(k,N¯)|𝑭(k,N¯)P𝝁|𝟎(𝒇(k,N¯)),\tilde{P}_{\bm{\mu}|\bm{0}}^{\overline{(k,N)}}=\sum_{\bm{f}_{(\overline{k,N})}}|\bm{f}_{(\overline{k,N})}\rangle\!\langle\bm{f}_{(\overline{k,N})}|_{\bm{F}_{(\overline{k,N})}}\otimes P_{\bm{\mu}|\bm{0}}^{(\bm{f}_{(\overline{k,N})})}, (68)

where the notation (k,N¯)(\overline{k,N}) is understood to denote a tuple excluding entries k,Nk,N, e.g., 𝒇(k,N¯)\bm{f}_{(\overline{k,N})} is a tuple of indices fkf_{k^{\prime}} for kk,Nk^{\prime}\neq k,N. P𝝁|𝟎(𝒇(k,N¯))=kk,NPμk|0(fk)P_{\bm{\mu}|\bm{0}}^{(\bm{f}_{(\overline{k,N})})}=\bigotimes_{k^{\prime}\neq k,N}P^{(f_{k^{\prime}})}_{\mu_{k^{\prime}}|0} is then a rank one projector on the N2N-2 qubit Hilbert space shared by all parties excluding kk and NN, corresponding to the block 𝒇(k,N¯)\bm{f}_{(\overline{k,N})}.

We can now apply Jordan’s lemma to the projected states |Ψ𝝁,k:=P~𝝁|𝟎(k,N)¯|Ψ/P~𝝁|𝟎(k,N)¯|\Psi_{\bm{\mu},k}\rangle:=\tilde{P}_{\bm{\mu}|\bm{0}}^{\overline{(k,N)}}|\Psi\rangle/\sqrt{\langle\tilde{P}_{\bm{\mu}|\bm{0}}^{\overline{(k,N)}}\rangle}, by computing

|Ψ𝝁,k\displaystyle|\Psi_{\bm{\mu},k}\rangle =𝒇p𝒇|𝒇𝑭\displaystyle=\sum_{\bm{f}}\sqrt{p_{\bm{f}}}|\bm{f}\rangle_{\bm{F}}
(𝕀QkQNEP𝝁|𝟎(𝒇(k,N¯)))|φ𝒇𝑸EΨ|P~𝝁|𝟎(k,N)¯|Ψ|𝒇E\displaystyle\ \ \ \ \otimes\frac{\Big{(}\mathbb{I}_{Q_{k}Q_{N}E}\otimes P_{\bm{\mu}|\bm{0}}^{(\bm{f}_{(\overline{k,N})})}\Big{)}|\varphi^{\bm{f}}\rangle_{\bm{Q}E}}{\sqrt{\langle\Psi|\tilde{P}_{\bm{\mu}|\bm{0}}^{\overline{(k,N)}}|\Psi\rangle}}\otimes|\bm{f}\rangle_{E^{\prime}}
=𝒇S𝝁kp𝒇|𝒇𝑭|ϕ𝝁𝒇(k,N¯)𝑸(k,N¯)\displaystyle=\sum_{\bm{f}\in S_{\bm{\mu}}^{k}}\sqrt{p_{\bm{f}}}|\bm{f}\rangle_{\bm{F}}\otimes|\phi_{\bm{\mu}}^{\bm{f}_{(\overline{k,N})}}\rangle_{\bm{Q}_{(\overline{k,N})}}
|ψ𝝁,k𝒇QkQNE|𝒇E\displaystyle\ \ \ \ \ \ \ \ \otimes|\psi_{\bm{\mu},k}^{\bm{f}}\rangle_{Q_{k}Q_{N}E}\otimes|\bm{f}\rangle_{E^{\prime}} (69)

where S𝝁kS_{\bm{\mu}}^{k} is the set of blocks for which the projection is non-zero. Here 𝒇\bm{f} is understood to be the concatenation of fk,fN,𝒇(k,N¯)f_{k},f_{N},\bm{f}_{(\overline{k,N})}, and we used the fact that the qubit projectors are rank one to write the state as a tensor product,

(𝕀QkQNEP𝝁|𝟎(𝒇(k,N¯)))|φ𝒇𝑸EΨ|P~𝝁|𝟎(k,N)¯|Ψ={|ϕ𝝁𝒇(k,N¯)𝑸(k,N¯)|ψ𝝁,k𝒇QkQNEif𝒇S𝝁k,0otherwise,\frac{\Big{(}\mathbb{I}_{Q_{k}Q_{N}E}\otimes P_{\bm{\mu}|\bm{0}}^{(\bm{f}_{(\overline{k,N})})}\Big{)}|\varphi^{\bm{f}}\rangle_{\bm{Q}E}}{\sqrt{\langle\Psi|\tilde{P}_{\bm{\mu}|\bm{0}}^{\overline{(k,N)}}|\Psi\rangle}}\\ =\begin{cases}|\phi_{\bm{\mu}}^{\bm{f}_{(\overline{k,N})}}\rangle_{\bm{Q}_{(\overline{k,N})}}\otimes|\psi_{\bm{\mu},k}^{\bm{f}}\rangle_{Q_{k}Q_{N}E}\ \mathrm{if}\ \bm{f}\in S_{\bm{\mu}}^{k},\\ 0\ \mathrm{otherwise},\end{cases} (70)

where |ϕ𝝁𝒇(k,N¯)=kk,N|ϕμkfk|\phi_{\bm{\mu}}^{\bm{f}_{(\overline{k,N})}}\rangle=\bigotimes_{k^{\prime}\neq k,N}|\phi_{\mu_{k^{\prime}}}^{f_{k^{\prime}}}\rangle, and Pμk|0(fk)=|ϕμkfkϕμkfk|P^{(f_{k^{\prime}})}_{\mu_{k^{\prime}}|0}=|\phi_{\mu_{k^{\prime}}}^{f_{k^{\prime}}}\rangle\!\langle\phi_{\mu_{k^{\prime}}}^{f_{k^{\prime}}}|. |ψ𝝁,k𝒇QkQlE|\psi_{\bm{\mu},k}^{\bm{f}}\rangle_{Q_{k}Q_{l}E} is then the state held by parties k,Nk,N and Eve following the projection. By writing |ϕ~𝝁𝒇(k,N¯)𝑸~(k,N¯)=|𝒇(k,N¯)𝑭(k,N¯)|ϕ𝝁𝒇(k,N¯)𝑸(k,N¯)|\tilde{\phi}_{\bm{\mu}}^{\bm{f}_{(\overline{k,N})}}\rangle_{\tilde{\bm{Q}}_{{(\overline{k,N})}}}=|\bm{f}_{(\overline{k,N})}\rangle_{\bm{F}_{(\overline{k,N})}}\otimes|\phi_{\bm{\mu}}^{\bm{f}_{(\overline{k,N})}}\rangle_{\bm{Q}_{{(\overline{k,N})}}}, and |ψ~𝝁,k𝒇Q~kQ~NE=|fkfNFkFN|ψ𝝁,k𝒇QkQNE|\tilde{\psi}_{\bm{\mu},k}^{\bm{f}}\rangle_{\tilde{Q}_{k}\tilde{Q}_{N}E}=|f_{k}f_{N}\rangle_{F_{k}F_{N}}\otimes|\psi_{\bm{\mu},k}^{\bm{f}}\rangle_{Q_{k}Q_{N}E}, we can re-write the constraints in Eq.˜65 as (suppressing the identity operator on Eve’s systems)

𝒇S𝝁kp𝒇|ϕ~𝝁𝒇(k,N¯)𝑸~(k,N¯)M𝝁,i(k,N)|ψ~𝝁,k𝒇Q~kQ~NE|𝒇E=0,\sum_{\bm{f}\in S_{\bm{\mu}}^{k}}\sqrt{p_{\bm{f}}}|\tilde{\phi}_{\bm{\mu}}^{\bm{f}_{(\overline{k,N})}}\rangle_{\tilde{\bm{Q}}_{{(\overline{k,N})}}}\\ \otimes M_{\bm{\mu},i}^{(k,N)}|\tilde{\psi}_{\bm{\mu},k}^{\bm{f}}\rangle_{\tilde{Q}_{k}\tilde{Q}_{N}E}\otimes|\bm{f}\rangle_{E^{\prime}}=0, (71)

implying, for a fixed k,𝝁k,\bm{\mu},

M𝝁,i(k,N)|ψ~𝝁,k𝒇Q~kQ~NE=0,i,𝒇S𝝁k.M_{\bm{\mu},i}^{(k,N)}|\tilde{\psi}_{\bm{\mu},k}^{\bm{f}}\rangle_{\tilde{Q}_{k}\tilde{Q}_{N}E}=0,\ \forall i,\forall\bm{f}\in S_{\bm{\mu}}^{k}. (72)

This subsequently implies

M𝝁,i(k,N)|fk,fN|ψ𝝁,k𝒇QkQNE=0,i,𝒇S𝝁k,M_{\bm{\mu},i}^{(k,N)|f_{k},f_{N}}|\psi_{\bm{\mu},k}^{\bm{f}}\rangle_{Q_{k}Q_{N}E}=0,\ \forall i,\forall\bm{f}\in S_{\bm{\mu}}^{k}, (73)

where we wrote

M𝝁,i(k,N)|fk,fN=fk,fN|fk,fNfk,fN|FkFNM𝝁,i(k,N)|fk,fNM_{\bm{\mu},i}^{(k,N)|f_{k},f_{N}}=\sum_{f_{k},f_{N}}|f_{k},f_{N}\rangle\!\langle f_{k},f_{N}|_{F_{k}F_{N}}\otimes M_{\bm{\mu},i}^{(k,N)|f_{k},f_{N}} (74)

using the decomposition in Eq.˜67.

A.3.3 Applying the bipartite self-tests

For the next part of the proof, we use the self-testing properties of the bipartite Bell expression I𝝁(k,l)I^{(k,l)}_{\bm{\mu}}, which are in turn derived from the SOS polynomials, and the constraints they impose on any state which achieves the maximum quantum value. Explicitly, for any state which satisfies the polynomial constraints, there exists a local isometry which transforms that state into an ideal state in tensor product with some junk. Due to Jordan’s lemma, the constraints in Eq.˜73 certify the the existence of a pair of local unitaries on the qubit registers of parties kk and NN, such that, for every measurement outcome 𝝁\bm{\mu} and block combination 𝒇S𝝁k\bm{f}\in S_{\bm{\mu}}^{k}, U𝝁,k𝒇U_{\bm{\mu},k}^{\bm{f}}666Note that the notation U𝝁,k𝒇U_{\bm{\mu},k}^{\bm{f}} suppresses the tensor product of local unitaries to ease notation, that is, U𝝁,k𝒇=VWU_{\bm{\mu},k}^{\bm{f}}=V\otimes W where VV is a unitary on system QkQ_{k} and WW is a unitary on system QNQ_{N}., there exists a fixed two qubit state |Φ|\Phi\rangle and projectors Pak|xk(k)P_{a_{k}|x_{k}}^{(k)}, PaN|xN(N)P_{a_{N}|x_{N}}^{(N)} satisfying

(𝕀FkFNU𝝁,k𝒇𝕀E)(P~ak|xk(k)P~aN|xN(N))|ψ~𝝁,k𝒇Q~kQ~NE=|fkfNFkFN(Pak|xk(k)PaN|xN(N))|ΦQkQN|λ𝝁,k𝒇E,\Big{(}\mathbb{I}_{F_{k}F_{N}}\otimes U_{\bm{\mu},k}^{\bm{f}}\otimes\mathbb{I}_{E}\Big{)}\Big{(}\tilde{P}_{a_{k}|x_{k}}^{(k)}\otimes\tilde{P}_{a_{N}|x_{N}}^{(N)}\Big{)}|\tilde{\psi}_{\bm{\mu},k}^{\bm{f}}\rangle_{\tilde{Q}_{k}\tilde{Q}_{N}E}\\ =|f_{k}f_{N}\rangle_{F_{k}F_{N}}\otimes\Big{(}P_{a_{k}|x_{k}}^{(k)}\otimes P_{a_{N}|x_{N}}^{(N)}\Big{)}|\Phi\rangle_{Q_{k}Q_{N}}\otimes|\lambda_{\bm{\mu},k}^{\bm{f}}\rangle_{E}, (75)

where |λ𝝁,k𝒇E|\lambda_{\bm{\mu},k}^{\bm{f}}\rangle_{E} is the junk system left over in the Eve register, which a priori may depend on 𝝁,k\bm{\mu},k and 𝒇\bm{f}. Since we required that all bipartite expressions self-test the same state up to local unitaries, the above equation holds for all 𝝁\bm{\mu}.

A.3.4 Global post-measurement state

Finally, we use the existence of the local unitaries U𝝁,k𝒇U_{\bm{\mu},k}^{\bm{f}} to analyze the global post-measurement state held by all parties following their zero measurement, and storing their outcomes in the register 𝑹\bm{R}, which takes the form

ρ𝑹EE|𝟎=ak,aN,𝝁|akaN𝝁akaN𝝁|𝑹Tr𝑸~[(P~ak|0(k)P~aN|0(N)P~𝝁|𝟎(k,N)¯𝕀EE)|ΨΨ|].\rho_{\bm{R}EE^{\prime}|\bm{0}}=\sum_{a_{k},a_{N},\bm{\mu}}|a_{k}a_{N}\bm{\mu}\rangle\!\langle a_{k}a_{N}\bm{\mu}|_{\bm{R}}\otimes\\ \mathrm{Tr}_{\tilde{\bm{Q}}}\Big{[}\Big{(}\tilde{P}_{a_{k}|0}^{(k)}\otimes\tilde{P}_{a_{N}|0}^{(N)}\otimes\tilde{P}_{\bm{\mu}|\bm{0}}^{\overline{(k,N)}}\otimes\mathbb{I}_{EE^{\prime}}\Big{)}|\Psi\rangle\!\langle\Psi|\Big{]}. (76)

We can rewrite the term inside the partial trace

(P~ak|0(k)P~aN|0(N)P~𝝁|𝟎(k,N)¯𝕀EE)|Ψ\displaystyle\Big{(}\tilde{P}_{a_{k}|0}^{(k)}\otimes\tilde{P}_{a_{N}|0}^{(N)}\otimes\tilde{P}_{\bm{\mu}|\bm{0}}^{\overline{(k,N)}}\otimes\mathbb{I}_{EE^{\prime}}\Big{)}|\Psi\rangle
=Ψ|P~𝝁|𝟎(k,N)¯|Ψ(P~ak|0(k)P~aN|0(N)𝕀𝑸~(k,N¯)𝕀EE)|Ψ𝝁,k\displaystyle=\sqrt{\langle\Psi|\tilde{P}_{\bm{\mu}|\bm{0}}^{\overline{(k,N)}}|\Psi\rangle}\Big{(}\tilde{P}_{a_{k}|0}^{(k)}\otimes\tilde{P}_{a_{N}|0}^{(N)}\otimes\mathbb{I}_{\tilde{\bm{Q}}_{(\overline{k,N})}}\otimes\mathbb{I}_{EE^{\prime}}\Big{)}|\Psi_{\bm{\mu},k}\rangle
=Ψ|P~𝝁|𝟎(k,N)¯|Ψ𝒇S𝝁kp𝒇|ϕ~𝝁𝒇(k,N¯)𝑸~(k,N¯)\displaystyle=\sqrt{\langle\Psi|\tilde{P}_{\bm{\mu}|\bm{0}}^{\overline{(k,N)}}|\Psi\rangle}\sum_{\bm{f}\in S_{\bm{\mu}}^{k}}\sqrt{p_{\bm{f}}}|\tilde{\phi}_{\bm{\mu}}^{\bm{f}_{(\overline{k,N})}}\rangle_{\tilde{\bm{Q}}_{{(\overline{k,N})}}}
(P~ak|0(k)P~aN|0(N))|ψ~𝝁,k𝒇Q~kQ~NE|𝒇E\displaystyle\ \ \ \ \ \ \ \ \otimes\Big{(}\tilde{P}_{a_{k}|0}^{(k)}\otimes\tilde{P}_{a_{N}|0}^{(N)}\Big{)}|\tilde{\psi}_{\bm{\mu},k}^{\bm{f}}\rangle_{\tilde{Q}_{k}\tilde{Q}_{N}E}\otimes|\bm{f}\rangle_{E^{\prime}}
=Ψ|P~𝝁|𝟎(k,N)¯|Ψ𝒇S𝝁kp𝒇|ϕ~𝝁𝒇(k,N¯)𝑸~(k,N¯)\displaystyle=\sqrt{\langle\Psi|\tilde{P}_{\bm{\mu}|\bm{0}}^{\overline{(k,N)}}|\Psi\rangle}\sum_{\bm{f}\in S_{\bm{\mu}}^{k}}\sqrt{p_{\bm{f}}}|\tilde{\phi}_{\bm{\mu}}^{\bm{f}_{(\overline{k,N})}}\rangle_{\tilde{\bm{Q}}_{{(\overline{k,N})}}}
|fkfNFkFNU𝝁,k𝒇(Pak|0(k)PaN|0(N))|ΦQkQN\displaystyle\otimes|f_{k}f_{N}\rangle_{F_{k}F_{N}}\otimes U_{\bm{\mu},k}^{\bm{f}\dagger}\Big{(}P_{a_{k}|0}^{(k)}\otimes P_{a_{N}|0}^{(N)}\Big{)}|\Phi\rangle_{Q_{k}Q_{N}}
|λ𝝁,k𝒇E|𝒇E.\displaystyle\ \ \ \ \ \ \ \ \otimes|\lambda_{\bm{\mu},k}^{\bm{f}}\rangle_{E}\otimes|\bm{f}\rangle_{E^{\prime}}. (77)

Now, tracing out 𝑸~\tilde{\bm{Q}} equates to tracing out systems 𝑭𝑸\bm{F}\bm{Q}, which yields

Tr𝑸~[(P~ak|0(k)P~aN|0(N)P~𝝁|𝟎(k,N)¯𝕀EE)|ΨΨ|]\displaystyle\mathrm{Tr}_{\tilde{\bm{Q}}}\Big{[}\Big{(}\tilde{P}_{a_{k}|0}^{(k)}\otimes\tilde{P}_{a_{N}|0}^{(N)}\otimes\tilde{P}_{\bm{\mu}|\bm{0}}^{\overline{(k,N)}}\otimes\mathbb{I}_{EE^{\prime}}\Big{)}|\Psi\rangle\!\langle\Psi|\Big{]}
=Ψ|P~𝝁|𝟎(k,N)¯|ΨΦ|(Pak|0(k)PaN|0(N))|Φ\displaystyle=\langle\Psi|\tilde{P}_{\bm{\mu}|\bm{0}}^{\overline{(k,N)}}|\Psi\rangle\langle\Phi|\Big{(}P_{a_{k}|0}^{(k)}\otimes P_{a_{N}|0}^{(N)}\Big{)}|\Phi\rangle
𝒇S𝝁kp𝒇|λ𝝁,k𝒇λ𝝁,k𝒇|E|𝒇𝒇|E.\displaystyle\ \ \ \ \ \ \ \ \cdot\sum_{\bm{f}\in S_{\bm{\mu}}^{k}}p_{\bm{f}}|\lambda_{\bm{\mu},k}^{\bm{f}}\rangle\!\langle\lambda_{\bm{\mu},k}^{\bm{f}}|_{E}\otimes|\bm{f}\rangle\!\langle\bm{f}|_{E^{\prime}}. (78)

By employing Corollary˜1, we know that, following observation of I=ηQ\langle I\rangle=\eta^{\mathrm{Q}}, the marginals Ψ|P~𝝁|𝟎(k,l)¯|Ψ\langle\Psi|\tilde{P}_{\bm{\mu}|\bm{0}}^{\overline{(k,l)}}|\Psi\rangle are unique; hence the following equality holds:

Ψ|P~𝝁|𝟎(k,N)¯|ΨΦ|(Pak|0(k)PaN|0(N))|Φ=p(ak,aN,𝝁|𝟎),\langle\Psi|\tilde{P}_{\bm{\mu}|\bm{0}}^{\overline{(k,N)}}|\Psi\rangle\langle\Phi|\Big{(}P_{a_{k}|0}^{(k)}\otimes P_{a_{N}|0}^{(N)}\Big{)}|\Phi\rangle=p(a_{k},a_{N},\bm{\mu}|\bm{0}), (79)

where {p(ak,aN,𝝁|𝟎)}ak,aN,𝝁\{p(a_{k},a_{N},\bm{\mu}|\bm{0})\}_{a_{k},a_{N},\bm{\mu}} are those unique correlations needed to achieve I=ηQ\langle I\rangle=\eta^{\mathrm{Q}}. We then find

ρ𝑹EE|𝟎=ak,aN,𝝁p(ak,aN,𝝁|𝟎)|akaN𝝁akaN𝝁|𝑹𝒇S𝝁kp𝒇|λ𝝁,k𝒇λ𝝁,k𝒇|E|𝒇𝒇|E.\rho_{\bm{R}EE^{\prime}|\bm{0}}=\sum_{a_{k},a_{N},\bm{\mu}}p(a_{k},a_{N},\bm{\mu}|\bm{0})|a_{k}a_{N}\bm{\mu}\rangle\!\langle a_{k}a_{N}\bm{\mu}|_{\bm{R}}\\ \otimes\sum_{\bm{f}\in S_{\bm{\mu}}^{k}}p_{\bm{f}}|\lambda^{\bm{f}}_{\bm{\mu},k}\rangle\!\langle\lambda^{\bm{f}}_{\bm{\mu},k}|_{E}\otimes|\bm{f}\rangle\!\langle\bm{f}|_{E^{\prime}}. (80)

For a fixed k{1,,N1}k\in\{1,...,N-1\}, Eve can still establish correlations with the NN parties excluding kk and NN, by choosing the sets {S𝝁k}𝝁\{S_{\bm{\mu}}^{k}\}_{\bm{\mu}} to be disjoint. She could then distinguish the different outcomes 𝝁\bm{\mu} by a projective measurement on EE^{\prime}. She can also establish correlations by making the set {|λ𝝁,k𝒇}𝝁\{|\lambda^{\bm{f}}_{\bm{\mu},k}\rangle\}_{\bm{\mu}} distinguishable and measuring EE. To show Eve is in fact uncorrelated with all parties, we use the fact that there are N1N-1 choices of kk, i.e., one can choose different combinations of parties to maximally violate the bipartite self-tests.

We now include extra notation kN¯\overline{kN}, 𝝁𝝁kN¯\bm{\mu}\rightarrow\bm{\mu}_{\overline{kN}}, which denotes a tuple of N2N-2 outcomes for all parties excluding kk and NN, and define the following state, which is just a rewriting of Eq.˜80 making the choice kk explicit:

ρ𝑹EE|𝟎k:=ak,aN,𝝁kN¯p(ak,aN,𝝁kN¯|𝟎)|akaN𝝁kN¯akaN𝝁kN¯|𝑹𝒇S𝝁kN¯kp𝒇|λ𝝁kN¯,k𝒇λ𝝁kN¯,k𝒇|E|𝒇𝒇|E.\rho_{\bm{R}EE^{\prime}|\bm{0}}^{k}:=\sum_{a_{k},a_{N},\bm{\mu}_{\overline{kN}}}p(a_{k},a_{N},\bm{\mu}_{\overline{kN}}|\bm{0})\\ \cdot|a_{k}a_{N}\bm{\mu}_{\overline{kN}}\rangle\!\langle a_{k}a_{N}\bm{\mu}_{\overline{kN}}|_{\bm{R}}\\ \otimes\sum_{\bm{f}\in S_{\bm{\mu}_{\overline{kN}}}^{k}}p_{\bm{f}}|\lambda^{\bm{f}}_{\bm{\mu}_{\overline{kN}},k}\rangle\!\langle\lambda^{\bm{f}}_{\bm{\mu}_{\overline{kN}},k}|_{E}\otimes|\bm{f}\rangle\!\langle\bm{f}|_{E^{\prime}}. (81)

Now, we could equally define a second state ρ𝑹EE|𝟎k\rho_{\bm{R}EE^{\prime}|\bm{0}}^{k^{\prime}} where kk,Nk^{\prime}\neq k,N in the exact same way, except we chose kk^{\prime} instead of kk as the self-testing party. Since both states are a rewriting of the same post measurement state ρ𝑹EE|𝟎\rho_{\bm{R}EE|\bm{0}} given by Eq.˜80, they must be equal, and we establish the equalities

ρ𝑹EE|𝟎k=ρ𝑹EE|𝟎k,k,k{1,,N1},kk.\rho_{\bm{R}EE^{\prime}|\bm{0}}^{k}=\rho_{\bm{R}EE^{\prime}|\bm{0}}^{k^{\prime}},\ \forall k,k^{\prime}\in\{1,...,N-1\},\ k^{\prime}\neq k. (82)

These equalities will allow us to place constraints on the sets S𝝁kN¯kS_{\bm{\mu}_{\overline{kN}}}^{k}, and ultimately show they are all equal. Specifically, Eq.˜82 implies777Note that we have implicitly equated the rewriting of the same outcome string 𝒂=akaN𝝁kN¯=akaN𝝁kN¯\bm{a}=a_{k}a_{N}\bm{\mu}_{\overline{kN}}=a_{k^{\prime}}a_{N}\bm{\mu}_{\overline{k^{\prime}N}}, and only considered cases which appear in the summation, i.e., strings with nonzero probability, p(𝒂|𝟎)>0p(\bm{a}|\bm{0})>0.

𝒇S𝝁kN¯kp𝒇|λ𝝁kN¯,k𝒇λ𝝁kN¯,k𝒇|E|𝒇𝒇|E=𝒇S𝝁kN¯kp𝒇|λ𝝁kN¯,k𝒇λ𝝁kN¯,k𝒇|E|𝒇𝒇|E\sum_{\bm{f}\in S_{\bm{\mu}_{\overline{kN}}}^{k}}p_{\bm{f}}|\lambda^{\bm{f}}_{\bm{\mu}_{\overline{kN}},k}\rangle\!\langle\lambda^{\bm{f}}_{\bm{\mu}_{\overline{kN}},k}|_{E}\otimes|\bm{f}\rangle\!\langle\bm{f}|_{E^{\prime}}\\ =\sum_{\bm{f}\in S_{\bm{\mu}_{\overline{k^{\prime}N}}}^{k^{\prime}}}p_{\bm{f}}|\lambda^{\bm{f}}_{\bm{\mu}_{\overline{k^{\prime}N}},k^{\prime}}\rangle\!\langle\lambda^{\bm{f}}_{\bm{\mu}_{\overline{k^{\prime}N}},k^{\prime}}|_{E}\otimes|\bm{f}\rangle\!\langle\bm{f}|_{E^{\prime}} (83)

By tracing out system EE, and using the fact that p𝒇0p_{\bm{f}}\neq 0, Eq.˜83 can only hold if

S𝝁kN¯k=S𝝁kN¯k,a1,,aN1.S_{\bm{\mu}_{\overline{kN}}}^{k}=S_{\bm{\mu}_{\overline{k^{\prime}N}}}^{k^{\prime}},\ \forall a_{1},...,a_{N-1}. (84)

Consider the case k=1k=1 and k=2k^{\prime}=2. Then we have for a fixed a3,,aN1a_{3},...,a_{N-1},

S(a2,a3,,aN1)1=S(a1,a3,a4,,aN1)2,a2.S_{(a_{2},a_{3},...,a_{N-1})}^{1}=S_{(a_{1},a_{3},a_{4},...,a_{N-1})}^{2},\ \forall a_{2}. (85)

Since the RHS is independent of a2a_{2}, we must have that S(a2,a3,,aN1)1S(a3,,aN1)1S^{1}_{(a_{2},a_{3},...,a_{N-1})}\equiv S^{1}_{(a_{3},...,a_{N-1})}, i.e., the LHS is also independent of a2a_{2}. We can apply the same argument for k=3,4,,N1k^{\prime}=3,4,...,N-1, deducing that

S(a2,a3,,aN1)1S,S_{(a_{2},a_{3},...,a_{N-1})}^{1}\equiv S, (86)

which is independent of the outcome string 𝝁1N¯=(a2,a3,,aN1)\bm{\mu}_{\overline{1N}}=(a_{2},a_{3},...,a_{N-1}). There is nothing unique about choosing k=1k=1, allowing us to write

ρ𝑹EE|𝟎k=ak,aN,𝝁kN¯p(ak,aN,𝝁kN¯|𝟎)|akaN𝝁kN¯akaN𝝁kN¯|𝑹𝒇Sp𝒇|λ𝝁kN¯,k𝒇λ𝝁kN¯,k𝒇|E|𝒇𝒇|E,\rho_{\bm{R}EE^{\prime}|\bm{0}}^{k}=\sum_{a_{k},a_{N},\bm{\mu}_{\overline{kN}}}p(a_{k},a_{N},\bm{\mu}_{\overline{kN}}|\bm{0})\\ \cdot|a_{k}a_{N}\bm{\mu}_{\overline{kN}}\rangle\!\langle a_{k}a_{N}\bm{\mu}_{\overline{kN}}|_{\bm{R}}\\ \otimes\sum_{\bm{f}\in S}p_{\bm{f}}|\lambda^{\bm{f}}_{\bm{\mu}_{\overline{kN}},k}\rangle\!\langle\lambda^{\bm{f}}_{\bm{\mu}_{\overline{kN}},k}|_{E}\otimes|\bm{f}\rangle\!\langle\bm{f}|_{E^{\prime}}, (87)

Hence Eve can learn nothing about any of the outcomes by measuring her register EE^{\prime}. What remains is to deal with the vectors |λ𝝁kN¯,k𝒇|\lambda^{\bm{f}}_{\bm{\mu}_{\overline{kN}},k}\rangle. We will show that the set {|λ𝝁kN¯,k𝒇}𝝁kN¯\{|\lambda^{\bm{f}}_{\bm{\mu}_{\overline{kN}},k}\rangle\}_{\bm{\mu}_{\overline{kN}}} contains one linearly independent vector, hence Eve can learn nothing about the outcome 𝝁kN¯\bm{\mu}_{\overline{kN}} from measuring her register EE.

Using the same approach as before, Eq.˜82 now implies, for every 𝒇S\bm{f}\in S,

akaN𝝁kN¯Ψ|P~𝝁kN¯|𝟎(k,N)¯|ΨΦ|(Pak|0(k)PaN|0(N))|Φ|akaN𝝁kN¯akaN𝝁kN¯||λ𝝁kN¯,k𝒇λ𝝁kN¯,k𝒇|=akaN𝝁kN¯Ψ|P~𝝁kN¯|𝟎(k,N)¯|ΨΦ|(Pak|0(k)PaN|0(N))|Φ|akaN𝝁kN¯akaN𝝁kN¯||λ𝝁kN¯,k𝒇λ𝝁kN¯,k𝒇|.\sum_{a_{k}a_{N}\bm{\mu}_{\overline{kN}}}\langle\Psi|\tilde{P}_{\bm{\mu}_{\overline{kN}}|\bm{0}}^{\overline{(k,N)}}|\Psi\rangle\langle\Phi|\Big{(}P_{a_{k}|0}^{(k)}\otimes P_{a_{N}|0}^{(N)}\Big{)}|\Phi\rangle\\ \cdot|a_{k}a_{N}\bm{\mu}_{\overline{kN}}\rangle\!\langle a_{k}a_{N}\bm{\mu}_{\overline{kN}}|\otimes|\lambda^{\bm{f}}_{\bm{\mu}_{\overline{kN}},k}\rangle\!\langle\lambda^{\bm{f}}_{\bm{\mu}_{\overline{kN}},k}|\\ =\sum_{a_{k^{\prime}}a_{N}\bm{\mu}_{\overline{k^{\prime}N}}}\langle\Psi|\tilde{P}_{\bm{\mu}_{\overline{k^{\prime}N}}|\bm{0}}^{\overline{(k^{\prime},N)}}|\Psi\rangle\langle\Phi|\Big{(}P_{a_{k^{\prime}}|0}^{(k^{\prime})}\otimes P_{a_{N}|0}^{(N)}\Big{)}|\Phi\rangle\\ \cdot|a_{k^{\prime}}a_{N}\bm{\mu}_{\overline{k^{\prime}N}}\rangle\!\langle a_{k^{\prime}}a_{N}\bm{\mu}_{\overline{k^{\prime}N}}|\otimes|\lambda^{\bm{f}}_{\bm{\mu}_{\overline{k^{\prime}N}},k^{\prime}}\rangle\!\langle\lambda^{\bm{f}}_{\bm{\mu}_{\overline{k^{\prime}N}},k^{\prime}}|. (88)

Now equating for a given value of 𝒂=(a1,,ak,,ak,,aN)\bm{a}=(a_{1},...,a_{k},...,a_{k^{\prime}},...,a_{N}), we obtain

Ψ|P~𝝁kN¯|𝟎(k,N)¯|ΨΦ|(Pak|0(k)PaN|0(N))|Φ|λ𝝁kN¯,k𝒇λ𝝁kN¯,k𝒇|=Ψ|P~𝝁kN¯|𝟎(k,N)¯|ΨΦ|(Pak|0(k)PaN|0(N))|Φ|λ𝝁kN¯,k𝒇λ𝝁kN¯,k𝒇|.\langle\Psi|\tilde{P}_{\bm{\mu}_{\overline{kN}}|\bm{0}}^{\overline{(k,N)}}|\Psi\rangle\langle\Phi|\Big{(}P_{a_{k}|0}^{(k)}\otimes P_{a_{N}|0}^{(N)}\Big{)}|\Phi\rangle\\ \cdot|\lambda^{\bm{f}}_{\bm{\mu}_{\overline{kN}},k}\rangle\!\langle\lambda^{\bm{f}}_{\bm{\mu}_{\overline{kN}},k}|\\ =\langle\Psi|\tilde{P}_{\bm{\mu}_{\overline{k^{\prime}N}}|\bm{0}}^{\overline{(k^{\prime},N)}}|\Psi\rangle\langle\Phi|\Big{(}P_{a_{k^{\prime}}|0}^{(k)}\otimes P_{a_{N}|0}^{(N)}\Big{)}|\Phi\rangle\\ \cdot|\lambda^{\bm{f}}_{\bm{\mu}_{\overline{k^{\prime}N}},k^{\prime}}\rangle\!\langle\lambda^{\bm{f}}_{\bm{\mu}_{\overline{k^{\prime}N}},k^{\prime}}|. (89)

We can hence conclude the two vectors are linearly dependent,

|λ𝝁kN¯,k𝒇|λ𝝁kN¯,k𝒇,|\lambda^{\bm{f}}_{\bm{\mu}_{\overline{kN}},k}\rangle\sim|\lambda^{\bm{f}}_{\bm{\mu}_{\overline{k^{\prime}N}},k^{\prime}}\rangle, (90)

where \sim denotes equality up to a constant888Note it is sufficient to just consider the cases for which the probabilities Ψ|P~𝝁kN¯|𝟎(k,N)¯|ΨΦ|(Pak|0(k)PaN|0(N))|Φ\langle\Psi|\tilde{P}_{\bm{\mu}_{\overline{kN}}|\bm{0}}^{\overline{(k,N)}}|\Psi\rangle\langle\Phi|\Big{(}P_{a_{k}|0}^{(k)}\otimes P_{a_{N}|0}^{(N)}\Big{)}|\Phi\rangle are nonzero. If they are zero, then the vector |λ𝝁kN¯,k𝒇|\lambda^{\bm{f}}_{\bm{\mu}_{\overline{kN}},k}\rangle will not appear in the summation, and Eve will never observe it.. Notice on the left hand side of the above equation, one has the freedom to choose any value of aka_{k^{\prime}} without changing the right hand side up to a constant. Therefore

|λ𝝁kN¯|ak=0,k𝒇|λ𝝁kN¯|ak=1,k𝒇.|\lambda^{\bm{f}}_{\bm{\mu}_{\overline{kN}}|a_{k^{\prime}}=0,k}\rangle\sim|\lambda^{\bm{f}}_{\bm{\mu}_{\overline{kN}}|a_{k^{\prime}}=1,k}\rangle. (91)

There are 2N32^{N-3} such equations written above, corresponding to the different choices of 𝝁kN¯\bm{\mu}_{\overline{kN}} once aka_{k^{\prime}} is fixed. We have 2N22^{N-2} unknowns that we want to fix, hence this is sufficient to show linear dependence for the N=3N=3 case. If N>3N>3, we can select another party k′′k^{\prime\prime} not equal to k,k,Nk,k^{\prime},N and add another 2N32^{N-3} equations,

|λ𝝁kN¯,k𝒇|λ𝝁k′′N¯,k′′𝒇,|\lambda^{\bm{f}}_{\bm{\mu}_{\overline{kN}},k}\rangle\sim|\lambda^{\bm{f}}_{\bm{\mu}_{\overline{k^{\prime\prime}N}},k^{\prime\prime}}\rangle, (92)

allowing us to write

|λ𝝁kN¯|ak=0,k𝒇\displaystyle|\lambda^{\bm{f}}_{\bm{\mu}_{\overline{kN}}|a_{k^{\prime}}=0,k}\rangle |λ𝝁kN¯|ak=1,k𝒇\displaystyle\sim|\lambda^{\bm{f}}_{\bm{\mu}_{\overline{kN}}|a_{k^{\prime}}=1,k}\rangle
|λ𝝁kN¯|ak′′=0,k𝒇\displaystyle|\lambda^{\bm{f}}_{\bm{\mu}_{\overline{kN}}|a_{k^{\prime\prime}}=0,k}\rangle |λ𝝁kN¯|ak′′=1,k𝒇.\displaystyle\sim|\lambda^{\bm{f}}_{\bm{\mu}_{\overline{kN}}|a_{k^{\prime\prime}}=1,k}\rangle. (93)

The combination of the two will solve the N=4N=4 case. We can see that for the first choice of kk^{\prime}, we get 2N32^{N-3} equations which is half the unknowns, reducing the number of unknowns to 2N32^{N-3}. Each new choice of kk^{\prime}, of which there are N2N-2 in total, halves the number of unknowns, until we are left with one, which shows that every vector can be written as a linear combination of a fixed vector, i.e.

|λ𝝁kN¯,k𝒇=β𝝁kN¯k|λk𝒇,|\lambda^{\bm{f}}_{\bm{\mu}_{\overline{kN}},k}\rangle=\beta_{\bm{\mu}_{\overline{kN}}}^{k}|\lambda^{\bm{f}}_{k}\rangle, (94)

for some constants β𝝁kN¯k\beta_{\bm{\mu}_{\overline{kN}}}^{k}, which due to normalization must be equal to 1. Putting the above into Eq.˜88 we also notice |λk𝒇=|λk𝒇|\lambda^{\bm{f}}_{k}\rangle=|\lambda^{\bm{f}}_{k^{\prime}}\rangle, hence we can drop the kk label.

Now we finally conclude

ρ𝑹EE|𝟎ρ𝑹EE|𝟎k=(ak,aN,𝝁kN¯p(ak,aN,𝝁kN¯|𝟎)|akaN𝝁kN¯akaN𝝁kN¯|𝑹)(𝒇Sp𝒇|λ𝒇λ𝒇|E|𝒇𝒇|E),\rho_{\bm{R}EE^{\prime}|\bm{0}}\equiv\rho_{\bm{R}EE^{\prime}|\bm{0}}^{k}=\Bigg{(}\sum_{a_{k},a_{N},\bm{\mu}_{\overline{kN}}}p(a_{k},a_{N},\bm{\mu}_{\overline{kN}}|\bm{0})\\ \cdot|a_{k}a_{N}\bm{\mu}_{\overline{kN}}\rangle\!\langle a_{k}a_{N}\bm{\mu}_{\overline{kN}}|_{\bm{R}}\Bigg{)}\\ \otimes\Bigg{(}\sum_{\bm{f}\in S}p_{\bm{f}}|\lambda^{\bm{f}}\rangle\!\langle\lambda^{\bm{f}}|_{E}\otimes|\bm{f}\rangle\!\langle\bm{f}|_{E^{\prime}}\Bigg{)}, (95)

which is of the desired tensor product form, and completes the proof. ∎

A.4 Proof of Lemma˜3

Lemma 3. Let I,ηNQI,\eta_{N}^{\mathrm{Q}} be defined as in Eq.˜13. Then

RI(ηNQ)=H({p(𝒂|𝟎)}),R_{I}(\eta_{N}^{\mathrm{Q}})=H(\{p(\bm{a}|\bm{0})\}), (96)

where H({pi})H(\{p_{i}\}) is the Shannon entropy of a distribution {pi}i\{p_{i}\}_{i}.

Proof.

Lemma˜3 directly follows from Eq.˜13 since the state for which then entropy is evaluated on has Eve’s part in tensor product. This implies, for all compatible post measurement states, H(𝑹|𝑿=𝟎,E)ρ𝑹E|𝟎=H(𝑹|𝑿=𝟎)ρ𝑹|𝟎H(\bm{R}|\bm{X}=\bm{0},E)_{\rho_{\bm{R}E|\bm{0}}}=H(\bm{R}|\bm{X}=\bm{0})_{\rho_{\bm{R}|\bm{0}}}. Since ρ𝑹|𝟎\rho_{\bm{R}|\bm{0}}, derived explicitly in the proof of Eq.˜13, is the same for all compatible states and measurements, we obtain the claim:

inf|Ψ𝑸~E,{{P~ak|xk(k)}ak}kcompatiblewithI=ηNQH(𝑹|𝑿=𝟎,E)ρ𝑹E|𝟎=H({p(ak,al,𝝁|𝟎)}).\inf_{\begin{subarray}{c}|\Psi\rangle_{\tilde{\bm{Q}}E},\big{\{}\{\tilde{P}_{a_{k}|x_{k}}^{(k)}\}_{a_{k}}\big{\}}_{k}\\ \mathrm{compatible\ with}\ \langle I\rangle\\ =\eta^{\mathrm{Q}}_{N}\end{subarray}}H(\bm{R}|\bm{X}=\bm{0},E)_{\rho_{\bm{R}E|\bm{0}}}\\ =H(\{p(a_{k},a_{l},\bm{\mu}|\bm{0})\}). (97)

Appendix B Proofs and details for DI randomness certification

Recall the bipartite Bell-inequalities IθI_{\theta} which are defined as

Iθ(k,l)=cosθcos2θA0(k)A0(l)cos2θ(A0(k)A1(l)+A1(k)A0(l))+cosθA1(k)A1(l)\langle I_{\theta}^{(k,l)}\rangle=\cos\theta\cos 2\theta\langle A_{0}^{(k)}A_{0}^{(l)}\rangle-\\ \cos 2\theta\big{(}\langle A_{0}^{(k)}A_{1}^{(l)}\rangle+\langle A_{1}^{(k)}A_{0}^{(l)}\rangle\big{)}+\cos\theta\langle A_{1}^{(k)}A_{1}^{(l)}\rangle (98)

for θ𝒢\theta\in\mathcal{G} where

𝒢=(π/4,π/2)(π/2,3π/4)(5π/4,3π/2)(3π/2,7π/4).\mathcal{G}=(\pi/4,\pi/2)\cup(\pi/2,3\pi/4)\cup(5\pi/4,3\pi/2)\cup(3\pi/2,7\pi/4). (99)

We now prove a few results about these Bell-inequalities that were stated in the main text.

B.1 Proof of Lemma˜5

Lemma 5. Let 𝛍\bm{\mu} be a tuple of N2N-2 measurement outcomes for all parties excluding k,lk,l, and n𝛍{0,1}n_{\bm{\mu}}\in\{0,1\} be the parity of 𝛍\bm{\mu}. Let θ𝒢\theta\in\mathcal{G} and {I𝛍(k,l)}𝛍\{I_{\bm{\mu}}^{(k,l)}\}_{\bm{\mu}} be a set of bipartite Bell expressions between parties k,lk,l, where

I𝝁(k,l)=(1)n𝝁Iθ(k,l).I_{\bm{\mu}}^{(k,l)}=(-1)^{n_{\bm{\mu}}}I_{\theta}^{(k,l)}. (100)

Then the expanded Bell expression given by

Iθ=k=1N1(𝝁P~𝝁|𝟎(k,N)¯I𝝁(k,N))I_{\theta}=\sum_{k=1}^{N-1}\Bigg{(}\sum_{\bm{\mu}}\tilde{P}^{\overline{(k,N)}}_{\bm{\mu}|\bm{0}}I_{\bm{\mu}}^{(k,N)}\Bigg{)} (101)

has quantum bounds ±ηN,θQ\pm\eta_{N,\theta}^{\mathrm{Q}} where ηN,θQ=2(N1)sin3θ\eta_{N,\theta}^{\mathrm{Q}}=2(N-1)\sin^{3}\theta. Moreover, Iθ=ηN,θQ\langle I_{\theta}\rangle=\eta_{N,\theta}^{\mathrm{Q}} is achieved up to relabelings by the strategy in Eq.˜20, and cannot be achieved classically.

Proof.

The above is an example of Footnote˜1, where we used the bipartite expression in Lemma˜4 as a seed, and chose ck,l=1c_{k,l}=1 if l=Nl=N and k{1,..,N1}k\in\{1,..,N-1\}, and 0 otherwise. First we notice that if the bound ηN,θQ\eta_{N,\theta}^{\mathrm{Q}} is achievable, then IθI_{\theta} must define a nontrivial Bell inequality. This is because the maximum local value is upper bounded by (N1)ηθL<ηN,θQ(N-1)\eta_{\theta}^{\mathrm{L}}<\eta_{N,\theta}^{\mathrm{Q}}, since ηθL<ηθQ\eta_{\theta}^{\mathrm{L}}<\eta_{\theta}^{\mathrm{Q}}. A similar argument holds for the minimum local value.

We now proceed to show I=ηN,θQ\langle I\rangle=\eta_{N,\theta}^{\mathrm{Q}} is achieved by the strategy in Eq.˜20. Notice that, for the ideal operators Pak|0(k)=|++|P_{a_{k^{\prime}}|0}^{(k^{\prime})}=|+\rangle\!\langle+| (|||-\rangle\!\langle-|) if ak=0a_{k^{\prime}}=0 (ak=1a_{k^{\prime}}=1),

P𝝁|𝟎(k,N)¯|ψGHZ=ψGHZ|P𝝁|𝟎(k,N)¯|ψGHZ|ϕ𝝁|Φ𝝁,P^{\overline{(k,N)}}_{\bm{\mu}|\bm{0}}|\psi_{\mathrm{GHZ}}\rangle=\sqrt{\langle\psi_{\mathrm{GHZ}}|P^{\overline{(k,N)}}_{\bm{\mu}|\bm{0}}|\psi_{\mathrm{GHZ}}\rangle}|\phi_{\bm{\mu}}\rangle\otimes|\Phi_{\bm{\mu}}\rangle, (102)

where |ϕ𝝁|\phi_{\bm{\mu}}\rangle spans the support of P𝝁|𝟎(k,N)¯P^{\overline{(k,N)}}_{\bm{\mu}|\bm{0}}, and |Φ𝝁=(|00+i(1)n𝝁|11)/2|\Phi_{\bm{\mu}}\rangle=(|00\rangle+i(-1)^{n_{\bm{\mu}}}|11\rangle)/\sqrt{2}. We then have

𝝁\displaystyle\sum_{\bm{\mu}} ψGHZ|P𝝁|𝟎(k,N)¯I𝝁(k,N)|ψGHZ\displaystyle\langle\psi_{\mathrm{GHZ}}|P^{\overline{(k,N)}}_{\bm{\mu}|\bm{0}}\otimes I_{\bm{\mu}}^{(k,N)}|\psi_{\mathrm{GHZ}}\rangle
=𝝁ψGHZ|P𝝁|𝟎(k,N)¯|ψGHZΦ𝝁|I𝝁(k,N)|Φ𝝁\displaystyle=\sum_{\bm{\mu}}\langle\psi_{\mathrm{GHZ}}|P_{\bm{\mu}|\bm{0}}^{\overline{(k,N)}}|\psi_{\mathrm{GHZ}}\rangle\langle\Phi_{\bm{\mu}}|I_{\bm{\mu}}^{(k,N)}|\Phi_{\bm{\mu}}\rangle
=2sin3θ,k{1,,N1},\displaystyle=2\sin^{3}\theta,\ \forall k\in\{1,...,N-1\}, (103)

where the final equality comes from the fact that when n𝝁=0n_{\bm{\mu}}=0, we recover the state and measurements in Eq.˜19 which yields the maximum quantum value of Iθ(k,N)I^{(k,N)}_{\theta}, and when n𝝁=1n_{\bm{\mu}}=1, we obtain a strategy equivalent to Eq.˜19 up to a relabeling, which yields the maximum quantum value of Iθ(k,N)-I^{(k,N)}_{\theta}.

Since the maximum quantum bound is achievable, a nontrivial Bell inequality follows. The minimum quantum value ηN,θQ-\eta_{N,\theta}^{\mathrm{Q}} is simply achieved by relabeling the outcomes. ∎

B.2 Proof of Proposition˜3

Proposition 3. For even NN, let

θN=2πtNN+1,\theta^{*}_{N}=\frac{2\pi t_{N}}{N+1}, (104)

where tNt_{N} is the (N/2)th(N/2)^{\mathrm{th}} element of the sequence 1,1,5,7,3,3,11,13,5,5,1,1,5,7,3,3,11,13,5,5,... given by

tN={N/4+1/2,ifN=8n+2,N/4,ifN=8n+4,3N/4+1/2,ifN=8n+6,3N/4+1,ifN=8n+8,n0.t_{N}=\begin{cases}N/4+1/2,\ \mathrm{if}\ N=8n+2,\\ N/4,\ \mathrm{if}\ N=8n+4,\\ 3N/4+1/2,\ \mathrm{if}\ N=8n+6,\\ 3N/4+1,\ \mathrm{if}\ N=8n+8,\ n\in\mathbb{N}_{0}.\end{cases} (105)

Then θN𝒢\theta^{*}_{N}\in\mathcal{G}.

Proof.

The sequence θN\theta_{N}^{*} splits naturally into four subsequences for n0n\in\mathbb{N}_{0}

θ8n+2\displaystyle\theta_{8n+2}^{*} =2π(2n+1)8n+3\displaystyle=\frac{2\pi(2n+1)}{8n+3}\qquad θ8n+4=2π(2n+1)8n+5\displaystyle\theta_{8n+4}^{*}=\frac{2\pi(2n+1)}{8n+5} (106)
θ8n+6\displaystyle\theta_{8n+6}^{*} =2π(6n+5)8n+7\displaystyle=\frac{2\pi(6n+5)}{8n+7}\qquad θ8n+8=2π(6n+7)8n+9.\displaystyle\theta_{8n+8}^{*}=\frac{2\pi(6n+7)}{8n+9}\,.

First notice that the first terms in each of these subsequences are contained in one of the subintervals of 𝒢\mathcal{G}

θ2\displaystyle\theta_{2}^{*} (π/2,3π/4)\displaystyle\in(\pi/2,3\pi/4)\qquad θ4(π/4,π/2)\displaystyle\theta_{4}^{*}\in(\pi/4,\pi/2) (107)
θ6\displaystyle\theta_{6}^{*} (5π/4,3π/2)\displaystyle\in(5\pi/4,3\pi/2)\qquad θ8(3π/2,7π/4).\displaystyle\theta_{8}^{*}\in(3\pi/2,7\pi/4)\,.

Looking at the ratio of subsequent terms in each subsequence

θ8n+2θ8n+10\displaystyle\frac{\theta_{8n+2}^{*}}{\theta_{8n+10}^{*}} =16n2+30n+1116n2+30n+9θ8n+4θ8n+12=16n2+34n+1316n2+34n+15\displaystyle=\frac{16n^{2}+30n+11}{16n^{2}+30n+9}\quad\frac{\theta_{8n+4}^{*}}{\theta_{8n+12}^{*}}=\frac{16n^{2}+34n+13}{16n^{2}+34n+15}
θ8n+6θ8n+14\displaystyle\frac{\theta_{8n+6}^{*}}{\theta_{8n+14}^{*}} =48n2+130n+7548n2+130n+77θ8n+8θ8n+16=48n2+158n+11948n2+158n+117\displaystyle=\frac{48n^{2}+130n+75}{48n^{2}+130n+77}\quad\frac{\theta_{8n+8}^{*}}{\theta_{8n+16}^{*}}=\frac{48n^{2}+158n+119}{48n^{2}+158n+117}\,

we see that the subsequences θ8n+2\theta_{8n+2}^{*} and θ8n+8\theta_{8n+8}^{*} are strictly monotonically decreasing sequences whereas θ8n+4\theta_{8n+4}^{*} and θ8n+6\theta_{8n+6}^{*} are strictly monotonically increasing sequences. Furthermore their limits are

limnθ8n+2\displaystyle\lim_{n\to\infty}\theta_{8n+2}^{*} =limnθ8n+4=π/2\displaystyle=\lim_{n\to\infty}\theta_{8n+4}^{*}=\pi/2 (108)
limnθ8n+6\displaystyle\lim_{n\to\infty}\theta_{8n+6}^{*} =limnθ8n+8=3π/2.\displaystyle=\lim_{n\to\infty}\theta_{8n+8}^{*}=3\pi/2\,.

The limits correspond to the boundaries of the respective subinterval of 𝒢\mathcal{G} that the first terms of the subsequences are contained in. As they converge strictly monotonically to these limits this implies that every term in the subsequences in contained within their respective subintervals and hence θN𝒢\theta_{N}^{*}\in\mathcal{G} for all even NN. ∎

B.3 Proof of Propositions˜4 and 5

Proposition 4. Let NN be an even integer. For every MABK value ss in the range (1,mN](1,m_{N}^{*}], there exists a θs𝒢\theta_{s}\in\mathcal{G} that satisfies s=MN(θs)s=\langle M_{N}(\theta_{s})\rangle.

Proof.

We examine each subsequence independently.

Case 1. For N=8n+2N=8n+2, recall that θN(π/2,3π/4)\theta_{N}^{*}\in(\pi/2,3\pi/4). When n=0n=0, the claim can be seen from Fig.˜1, which indicates MABK values below 11 for θ(π/2,3π/4)\theta\in(\pi/2,3\pi/4), and the fact that MN(θ)M_{N}(\theta) is continuous. We hence consider n1n\geq 1 in the following. First note

M8n+2(π/2)=24n.\langle M_{8n+2}(\pi/2)\rangle=2^{4n}. (109)

Consider the point θ~n=π/2π/(4n+1)>π/4\tilde{\theta}_{n}=\pi/2-\pi/(4n+1)>\pi/4. We have

M8n+2(θ~n)=24n(sin8n+2(π8n+2)+cos8n+2(π8n+2))<0.\langle M_{8n+2}(\tilde{\theta}_{n})\rangle=-2^{4n}\Big{(}\sin^{8n+2}\Big{(}\frac{\pi}{8n+2}\Big{)}\\ +\cos^{8n+2}\Big{(}\frac{\pi}{8n+2}\Big{)}\Big{)}<0. (110)

Therefore, by continuity of the function MN(θ)\langle M_{N}(\theta)\rangle in θ\theta, the range of achievable MABK values corresponding to θ(θ~n,π/2)𝒢\theta\in(\tilde{\theta}_{n},\pi/2)\subset\mathcal{G} contains the sub-interval (1,24n)(1,2^{4n}). By a similar argument, the sub-interval (24n,m8n+2)(2^{4n},m_{8n+2}^{*}) must also be achievable for θ(π/2,θ8n+2)𝒢\theta\in(\pi/2,\theta_{8n+2}^{*})\subset\mathcal{G}. For the value 24n2^{4n}, notice

M8n+2(3π/4)=2(8n+1)/2(1)nsin8n+2(3π/8),\langle M_{8n+2}(3\pi/4)\rangle=2^{(8n+1)/2}(-1)^{n}\sin^{8n+2}(3\pi/8), (111)

and sin8n+2(3π/8)<1/2\sin^{8n+2}(3\pi/8)<1/\sqrt{2}. Hence there must exist a θ(θ8n+2,3π/4)𝒢\theta\in(\theta_{8n+2}^{*},3\pi/4)\subset\mathcal{G} such that M8n+2(θ)=24n\langle M_{8n+2}(\theta)\rangle=2^{4n}.

Case 2. For N=8n+4N=8n+4, recall θN(π/4,π/2)\theta_{N}^{*}\in(\pi/4,\pi/2). As before, we only need to consider n1n\geq 1 by analyzing Fig.˜1. Let θ~n=π/2+π/(4n+2)<3π/4\tilde{\theta}_{n}=\pi/2+\pi/(4n+2)<3\pi/4. Then

M8n+4(θ~n)=24n+1(sin8n+4(π8n+4)cos8n+4(π8n+4))<0.\langle M_{8n+4}(\tilde{\theta}_{n})\rangle=2^{4n+1}\Big{(}\sin^{8n+4}\Big{(}\frac{\pi}{8n+4}\Big{)}\\ -\cos^{8n+4}\Big{(}\frac{\pi}{8n+4}\Big{)}\Big{)}<0. (112)

We also note

M8n+4(π/2)=24n+1,\langle M_{8n+4}(\pi/2)\rangle=2^{4n+1}, (113)

and

M8n+4(π/4)=24n+1(1)n(cos8n+4(3π/8)+sin8n+4(3π/8))<24n+1.\langle M_{8n+4}(\pi/4)\rangle=2^{4n+1}(-1)^{n}\big{(}\cos^{8n+4}(3\pi/8)\\ +\sin^{8n+4}(3\pi/8)\big{)}<2^{4n+1}. (114)

We can therefore apply the same arguments as the previous case.

Case 3. For N=8n+6N=8n+6, recall θN(5π/4,3π/2)\theta_{N}^{*}\in(5\pi/4,3\pi/2). We observe the claim holds for n=0n=0 by examining Fig.˜1, and consider n1n\geq 1. Let θ~n=3π/2+π/(4n+3)\tilde{\theta}_{n}=3\pi/2+\pi/(4n+3), and note

M8n+6(θ~n)=24n+2(sin8n+6(π8n+6)+cos8n+6(π8n+6))<0.\langle M_{8n+6}(\tilde{\theta}_{n})\rangle=-2^{4n+2}\Big{(}\sin^{8n+6}\Big{(}\frac{\pi}{8n+6}\Big{)}\\ +\cos^{8n+6}\Big{(}\frac{\pi}{8n+6}\Big{)}\Big{)}<0. (115)

By checking

M8n+6(3π/2)=24n+2,\langle M_{8n+6}(3\pi/2)\rangle=2^{4n+2}, (116)

and

M8n+6(5π/4)=2(8n+5)/2(1)nsin8n+6(π/8)<24n+2,\langle M_{8n+6}(5\pi/4)\rangle\\ =-2^{(8n+5)/2}(-1)^{n}\sin^{8n+6}(\pi/8)<2^{4n+2}, (117)

we can apply the same arguments as case 1.

Case 4. Finally, for N=8n+8N=8n+8, we have θN(3π/2,7π/4)\theta_{N}^{*}\in(3\pi/2,7\pi/4). The case n=0n=0 can be seen by examining Fig.˜1, and for n1n\geq 1 we define θ~n=3π/2π/(4n+4)>5π/4\tilde{\theta}_{n}=3\pi/2-\pi/(4n+4)>5\pi/4. Then

M8n+8(θ~n)=24n+3(sin8n+8(π8n+8)cos8n+8(π8n+8))<0.\langle M_{8n+8}(\tilde{\theta}_{n})\rangle=2^{4n+3}\Big{(}\sin^{8n+8}\Big{(}\frac{\pi}{8n+8}\Big{)}\\ -\cos^{8n+8}\Big{(}\frac{\pi}{8n+8}\Big{)}\Big{)}<0. (118)

We also have

M8n+8(3π/2)=24n+3,\langle M_{8n+8}(3\pi/2)\rangle=2^{4n+3}, (119)

and

M8n+8(7π/4)=24n+3(1)n+1(cos8n+8(π/8)sin8n+8(π/8))<24n+3.\langle M_{8n+8}(7\pi/4)\rangle=2^{4n+3}(-1)^{n+1}\big{(}\cos^{8n+8}(\pi/8)\\ -\sin^{8n+8}(\pi/8)\big{)}<2^{4n+3}. (120)

By the arguments in case 1, this completes the proof. ∎

Proposition 5. Let NN be an odd integer. For every MABK value ss in the range (1,2(N1)/2)(1,2^{(N-1)/2}), there exists a θs𝒢\theta_{s}\in\mathcal{G} that satisfies s=MN(θs)s=\langle M_{N}(\theta_{s})\rangle.

Proof.

As was done for the even case, we consider four subsequences.

Case 1. For N=8n+3N=8n+3, we note M8n+3(π/2)=24n+1\langle M_{8n+3}(\pi/2)\rangle=2^{4n+1}. When n=0n=0, the claim can be verified from Fig.˜2, hence we consider n1n\geq 1. Consider the point θ~n=π/2+π/(4n+2)(π/2,3π/4)𝒢\tilde{\theta}_{n}=\pi/2+\pi/(4n+2)\in(\pi/2,3\pi/4)\subset\mathcal{G}. Then

MN(θ~n)=24n+1(sin8n+3(π8n+4)sin(πf(n))+cos8n+3(π8n+4)cos(πf(n))),\langle M_{N}(\tilde{\theta}_{n})\rangle=-2^{4n+1}\Big{(}\sin^{8n+3}\Big{(}\frac{\pi}{8n+4}\Big{)}\sin\big{(}\pi f(n)\big{)}\\ +\cos^{8n+3}\Big{(}\frac{\pi}{8n+4}\Big{)}\cos\big{(}\pi f(n)\big{)}\Big{)}, (121)

where f(n)=16n2+24n+78n+4f(n)=\frac{16n^{2}+24n+7}{8n+4}. Notice that

sin(πf(n))+sin(π8n+4)=2sin(π(n+1))cos(π28n2+12n+34n+2)=0.\sin\big{(}\pi f(n)\big{)}+\sin\Big{(}\frac{\pi}{8n+4}\Big{)}\\ =2\sin\big{(}\pi(n+1)\big{)}\cos\Big{(}\frac{\pi}{2}\frac{8n^{2}+12n+3}{4n+2}\Big{)}=0. (122)

Similarly

cos(πf(n))cos(π8n+4)=2sin(π(n+1))sin(π28n2+12n+34n+2)=0.\cos\big{(}\pi f(n)\big{)}-\cos\Big{(}\frac{\pi}{8n+4}\Big{)}\\ =-2\sin\big{(}\pi(n+1)\big{)}\sin\Big{(}\frac{\pi}{2}\frac{8n^{2}+12n+3}{4n+2}\Big{)}=0. (123)

Hence

MN(θ~n)=24n+1(sin8n+4(π8n+4)cos8n+4(π8n+4))<0.\langle M_{N}(\tilde{\theta}_{n})\rangle=2^{4n+1}\Big{(}\sin^{8n+4}\Big{(}\frac{\pi}{8n+4}\Big{)}\\ -\cos^{8n+4}\Big{(}\frac{\pi}{8n+4}\Big{)}\Big{)}<0. (124)

Therefore, the achievable MABK values from the interval (π/2,θ~n)𝒢(\pi/2,\tilde{\theta}_{n})\subset\mathcal{G} include the sub-interval (24n+1,1)(2^{4n+1},1) by continuity of the function MN(θ)\langle M_{N}(\theta)\rangle in θ\theta.

Case 2. For N=8n+7N=8n+7, note M8n+7(3π/2)=24n+3\langle M_{8n+7}(3\pi/2)\rangle=2^{4n+3}. The claim for n=0n=0 can be verified using Fig.˜2, so we consider n1n\geq 1. Let θ~n=3π/2+π/(4n+4)<7π/4\tilde{\theta}_{n}=3\pi/2+\pi/(4n+4)<7\pi/4, and note

MN(θ~n)=24n+3(sin8n+7(π8n+8)cos(πf(n))cos8n+7(π8n+8)sin(πf(n))),\langle M_{N}(\tilde{\theta}_{n})\rangle=2^{4n+3}\Big{(}\sin^{8n+7}\Big{(}\frac{\pi}{8n+8}\Big{)}\cos\big{(}\pi f(n)\big{)}\\ -\cos^{8n+7}\Big{(}\frac{\pi}{8n+8}\Big{)}\sin\big{(}\pi f(n)\big{)}\Big{)}, (125)

where f(n)=48n2+100n+518n+8f(n)=\frac{48n^{2}+100n+51}{8n+8}. Following the same steps as before, we have

sin(πf(n))cos(π8n+8)=0,\sin\big{(}\pi f(n)\big{)}-\cos\Big{(}\frac{\pi}{8n+8}\Big{)}=0, (126)

and

cos(πf(n))sin(π8n+8)=0,\cos\big{(}\pi f(n)\big{)}-\sin\Big{(}\frac{\pi}{8n+8}\Big{)}=0, (127)

implying

MN(θ~n)=24n+3(sin8n+8(π8n+8)cos8n+8(π8n+8))<0.\langle M_{N}(\tilde{\theta}_{n})\rangle=2^{4n+3}\Big{(}\sin^{8n+8}\Big{(}\frac{\pi}{8n+8}\Big{)}\\ -\cos^{8n+8}\Big{(}\frac{\pi}{8n+8}\Big{)}\Big{)}<0. (128)

The same arguments used in case 1 prove the claim.

Case 3. For N=8n+5N=8n+5, we consider the MABK expression obtained by relabelling the inputs of every party, followed by the output of every party’s first measurement, i.e., A~0(k)A~1(k)\tilde{A}_{0}^{(k)}\mapsto\tilde{A}_{1}^{(k)} and A~1(k)A~0(k)\tilde{A}_{1}^{(k)}\mapsto-\tilde{A}_{0}^{(k)}. The resulting MABK value of the strategy in Eq.˜20 is given by999Rather than relabelling the MABK expression, we could equivalently consider the strategy in Eq. 31, which achieves the MABK value in Eq. 32 of the original expression (Eq. 1), with the substitution ϕθ,θ0\phi\mapsto\theta,\ \theta\mapsto 0. This also results in Eq. 129.

M~N(θ)=2N12(cosN(θ/2+π/4)sin(Nθ/2+π/4)cosN(θ/2π/4)sin(Nθ/2+π/4)).\langle\tilde{M}_{N}(\theta)\rangle=2^{\frac{N-1}{2}}\Big{(}\cos^{N}\big{(}\theta/2+\pi/4\big{)}\sin\big{(}-N\theta/2+\pi/4\big{)}\\ -\cos^{N}\big{(}\theta/2-\pi/4\big{)}\sin\big{(}N\theta/2+\pi/4\big{)}\Big{)}. (129)

We then have M~8n+5(π/2)=24n+2\langle\tilde{M}_{8n+5}(\pi/2)\rangle=2^{4n+2}. We can verify the claim for n=0n=0 from Fig.˜2, hence we consider n1n\geq 1. Let θ~=π/2+π/(4n+3)<3π/4\tilde{\theta}=\pi/2+\pi/(4n+3)<3\pi/4, and note

M~N(θ~n)=24n+2(sin8n+5(π8n+6)cos(πf(n))+8cos8n+5(π8n+6)sin(πf(n))),\langle\tilde{M}_{N}(\tilde{\theta}_{n})\rangle=-2^{4n+2}\Big{(}\sin^{8n+5}\Big{(}\frac{\pi}{8n+6}\Big{)}\cos\big{(}\pi f(n)\big{)}\\ +8\cos^{8n+5}\Big{(}\frac{\pi}{8n+6}\Big{)}\sin\big{(}\pi f(n)\big{)}\Big{)}, (130)

where f(n)=8n2+16n+74n+3f(n)=\frac{8n^{2}+16n+7}{4n+3}. We have

sin(π8n+6)cos(πf(n))=0,\sin\Big{(}\frac{\pi}{8n+6}\Big{)}-\cos(\pi f(n))=0, (131)

and

cos(π8n+6)sin(πf(n))=0,\cos\Big{(}\frac{\pi}{8n+6}\Big{)}-\sin(\pi f(n))=0, (132)

which implies

M~N(θ~n)=24n+2(sin8n+6(π8n+6)+cos8n+6(π8n+6))<0.\langle\tilde{M}_{N}(\tilde{\theta}_{n})\rangle=-2^{4n+2}\Big{(}\sin^{8n+6}\Big{(}\frac{\pi}{8n+6}\Big{)}\\ +\cos^{8n+6}\Big{(}\frac{\pi}{8n+6}\Big{)}\Big{)}<0. (133)

The claim then follows from the same arguments as the previous cases.

Case 4. For N=8n+9N=8n+9, we also consider the relabelled MABK expression in Eq.˜129, and note M~8n+5(3π/2)=24n+4\langle\tilde{M}_{8n+5}(3\pi/2)\rangle=2^{4n+4}. Let θ~n=3π/2+π/(4n+5)<7π/4\tilde{\theta}_{n}=3\pi/2+\pi/(4n+5)<7\pi/4, and consider

M~N(θ~n)=24n+4(sin8n+9(π8n+10)sin(πf(n))cos8n+9(π8n+10)cos(πf(n))),\langle\tilde{M}_{N}(\tilde{\theta}_{n})\rangle=2^{4n+4}\Big{(}\sin^{8n+9}\Big{(}\frac{\pi}{8n+10}\Big{)}\sin\big{(}\pi f(n)\big{)}\\ -\cos^{8n+9}\Big{(}\frac{\pi}{8n+10}\Big{)}\cos\big{(}\pi f(n)\big{)}\Big{)}, (134)

where f(n)=48n2+124n+798n+10f(n)=\frac{48n^{2}+124n+79}{8n+10}. We have

cos(π8n+10)cos(πf(n))=0,\cos\Big{(}\frac{\pi}{8n+10}\Big{)}-\cos(\pi f(n))=0, (135)

and

sin(π8n+10)+sin(πf(n))=0,\sin\Big{(}\frac{\pi}{8n+10}\Big{)}+\sin(\pi f(n))=0, (136)

implying

M~N(θ~n)=24n+4(sin8n+10(π8n+10)+cos8n+10(π8n+10))<0.\langle\tilde{M}_{N}(\tilde{\theta}_{n})\rangle=-2^{4n+4}\Big{(}\sin^{8n+10}\Big{(}\frac{\pi}{8n+10}\Big{)}\\ +\cos^{8n+10}\Big{(}\frac{\pi}{8n+10}\Big{)}\Big{)}<0. (137)

By following the same steps as before this proves the claim. ∎

B.4 Evidence for ˜1

Conjecture 1. The maximum MABK value achievable by quantum strategies that generate maximum randomness, i.e., strategies with p(𝐚|𝟎)=1/2N,𝐚p(\bm{a}|\bm{0})=1/2^{N},\ \forall\bm{a}, is mNm_{N}^{*}.

The goal of ˜1 is to claim the analytical lower bound, mNm_{N}^{*}, on the maximum MABK value achievable whilst generating maximum randomness, is optimal, in the sense there exists no other set of quantum correlations with a higher MABK value achieving maximum randomness. To justify this claim, we developed a numerical technique to upper bound the maximum MABK value of quantum correlations achieving maximum randomness, and checked this against the achievable lower bounds mNm_{N}^{*}. We found close agreement between our numerical calculations and the analytic lower bounds, modulo a small discrepancy which pushed the numerical values slightly below the analytical ones. Since the SDP being solved is a minimization, we judge this to be caused by numerical precision, and the accumulations of errors pushing the global optimum below its true value. A comparison between our analytical lower bound and the numerics can be found in Table˜1, and details of the numerical method can be found in Section˜B.10.1. We also remark that this conjecture was proven analytically for the N=2N=2 case in [23].

NN Analytical lower bound, mNm_{N}^{*} Numerical upper bound
2 33/41.299038113\sqrt{3}/4\approx 1.29903811 1.29903810
4 (5/8)(5/2)(5+5)2.65828378(5/8)\sqrt{(5/2)(5+\sqrt{5})}\approx 2.65828378 2.65828370
6
(7/8)(cos(π/14)+32cos(3π/28)3cos(π/7)(7/8)\big{(}\cos(\pi/14)+3\sqrt{2}\cos(3\pi/28)-3\cos(\pi/7)
+5cos(3π/14))5.41251947+5\cos(3\pi/14)\big{)}\approx 5.41251947
5.41251940
8
45/16+333/4+(452/16)(3+3)cos(π/36)45/16+33\sqrt{3}/4+(45\sqrt{2}/16)(3+\sqrt{3})\cos(\pi/36)
(1/16)(1803+90)cos2(π/36)10.93208548-(1/16)(180\sqrt{3}+90)\cos^{2}(\pi/36)\approx 10.93208548
10.93208548
10
(11/32)(cos(π/22)+302cos(3π/44)5cos(π/11)+15cos(3π/22)(11/32)\big{(}\cos(\pi/22)+30\sqrt{2}\cos(3\pi/44)-5\cos(\pi/11)+15\cos(3\pi/22)
+52cos(7π/44)30cos(2π/11)+42cos(5π/22)22.00126184+5\sqrt{2}\cos(7\pi/44)-30\cos(2\pi/11)+42\cos(5\pi/22)\approx 22.00126184
22.00125885
12
(4292/16)cos(π/52)+(392/32)cos(9π/52)+(7152/64)cos(5π/52)(429\sqrt{2}/16)\cos(\pi/52)+(39\sqrt{2}/32)\cos(9\pi/52)+(715\sqrt{2}/64)\cos(5\pi/52)
+(13/64)cos(π/26)+(1287/64)cos(5π/26)(715/64)cos(2π/13)+(143/32)cos(3π/26)+(13/64)\cos(\pi/26)+(1287/64)\cos(5\pi/26)-(715/64)\cos(2\pi/13)+(143/32)\cos(3\pi/26)
(39/32)cos(π/13)(429/16)cos(3π/13)44.19316043-(39/32)\cos(\pi/13)-(429/16)\cos(3\pi/13)\approx 44.19316043
44.19316040
Table 1: Comparison between upper and lower bounds on the maximum MABK value which can be achieved by quantum correlations certifying maximum randomness, for different numbers of parties NN, rounded to 8 d.p. The analytical lower bound is provided by the family of strategies Eq.˜20 in the main text, and the numerical upper bound is calculated with an SOS approach, which amounts to solving an SDP. We observe instances where the numerical upper bound is seemingly below the analytical value. Since the SDP being solved is a minimization, we put this down to the accumulation of errors in the solver resulting in a numerical value below the true minimum. Data in this table supports the conjecture that our analytical lower bound is tight.

B.5 Proof of Proposition˜6

Proposition 6. In the limit of large even NN, one can achieve arbitrarily close to the maximum quantum violation of the NN party MABK inequality, 2(N1)/22^{(N-1)/2}, whilst certifying maximum device-independent randomness.

Proof.

We consider the asymptotic behaviour of mNm_{N}^{*}, which gives an achievable lower bound on the maximum MABK value compatible with maximum randomness. Our aim is to show that as NN becomes very large, whilst remaining an even integer, the MABK violation tends to the maximum quantum violation 2(N1)/22^{(N-1)/2}, i.e.,

limNNevenmN2N12=1.\lim_{\begin{subarray}{c}N\to\infty\\ N\ \mathrm{even}\end{subarray}}\frac{m_{N}^{*}}{2^{\frac{N-1}{2}}}=1. (138)

This can be established directly from the asymptotic behaviour of θN\theta^{*}_{N} given in Eq.˜108, as follows.

We first insert the expression for mN=MN(θN)m_{N}^{*}=\langle M_{N}(\theta^{*}_{N})\rangle in Eq.˜24 into the left hand side of (138) to give

limN\displaystyle\lim_{\begin{subarray}{c}N\to\infty\end{subarray}} (cosN(θN/2+π/4)sin(NθN/2+π/4)\displaystyle\Big{(}\cos^{N}\big{(}\theta^{*}_{N}/2+\pi/4\big{)}\sin\big{(}N\theta^{*}_{N}/2+\pi/4\big{)}
+cosN(θN/2π/4)sin(NθN/2π/4)).\displaystyle+\cos^{N}\big{(}\theta^{*}_{N}/2-\pi/4\big{)}\sin\big{(}N\theta^{*}_{N}/2-\pi/4\big{)}\Big{)}. (139)

We then consider two cases:

Case 1: N=8n+2N=8n+2 or N=8n+4N=8n+4. For these values of NN, we know limNθN=π/2\lim_{N\to\infty}\theta_{N}^{*}=\pi/2. One can then verify

limNcosN(θN/2+π/4)=0,\displaystyle\lim_{\begin{subarray}{c}N\to\infty\end{subarray}}\cos^{N}\big{(}\theta_{N}^{*}/2+\pi/4\big{)}=0, (140)
limNcosN(θN/2π/4)=1.\displaystyle\lim_{\begin{subarray}{c}N\to\infty\end{subarray}}\cos^{N}\big{(}\theta_{N}^{*}/2-\pi/4\big{)}=1.

Now we split into two sub-cases to evaluate NθNN\theta^{*}_{N} in the limit (B.5).

1a: N=8n+2N=8n+2. Here we find

NθN2=π(8n+2)(2n+1)8n+3,\frac{N\theta^{*}_{N}}{2}=\frac{\pi(8n+2)(2n+1)}{8n+3}, (141)

which has the oblique asymptote 2πn+3π/42\pi n+3\pi/4. We hence find

limNsin(NθN/2π/4)=sin(π/2)=1.\lim_{\begin{subarray}{c}N\to\infty\end{subarray}}\sin\big{(}N\theta^{*}_{N}/2-\pi/4\big{)}=\sin(\pi/2)=1. (142)

1b: N=8n+4N=8n+4. Here we find

NθN2=4π(2n+1)28n+5,\frac{N\theta^{*}_{N}}{2}=\frac{4\pi(2n+1)^{2}}{8n+5}, (143)

which also has the oblique asymptote 2πn+3π/42\pi n+3\pi/4, implying Eq.˜142.

Case 2: N=8n+6N=8n+6 or N=8n+8N=8n+8. For these values of NN, we know limNθN=3π/2\lim_{N\to\infty}\theta_{N}^{*}=3\pi/2. One can then verify

limNcosN(θN/2+π/4)=1,\displaystyle\lim_{\begin{subarray}{c}N\to\infty\end{subarray}}\cos^{N}\big{(}\theta_{N}^{*}/2+\pi/4\big{)}=1, (144)
limNcosN(θN/2π/4)=0.\displaystyle\lim_{\begin{subarray}{c}N\to\infty\end{subarray}}\cos^{N}\big{(}\theta_{N}^{*}/2-\pi/4\big{)}=0.

Splitting into two sub-cases:

2a, N=8n+6N=8n+6: Here

NθN2=π(8n+6)(6n+5)8n+7,\frac{N\theta^{*}_{N}}{2}=\frac{\pi(8n+6)(6n+5)}{8n+7}, (145)

which has the oblique asymptote 6πn+17π/46\pi n+17\pi/4. Inserting into the limit we find

limNsin(NθN/2+π/4)=sin(9π/2)=1.\lim_{\begin{subarray}{c}N\to\infty\end{subarray}}\sin\big{(}N\theta^{*}_{N}/2+\pi/4\big{)}=\sin(9\pi/2)=1. (146)

2b, N=8n+8N=8n+8: For the final case

NθN2=π(8n+8)(6n+7)8n+9,\frac{N\theta^{*}_{N}}{2}=\frac{\pi(8n+8)(6n+7)}{8n+9}, (147)

which has the oblique asymptote 6πn+25π/46\pi n+25\pi/4. Inserting into the limit we find

limNsin(NθN/2+π/4)=sin(13π/2)=1.\lim_{\begin{subarray}{c}N\to\infty\end{subarray}}\sin\big{(}N\theta^{*}_{N}/2+\pi/4\big{)}=\sin(13\pi/2)=1. (148)

This shows that the MABK violation mNm_{N}^{*} tends to the maximum quantum value in the regime of large NN. Note however, in the limit θ\theta tends to π/2\pi/2 or 3π/23\pi/2; these values are not self-testable, since θ\theta does not lie in a valid interval for there to be a gap between the local and quantum bound for the two party Bell inequality. However, the points either side of both angles are self-testable. We hence obtain the result that one can achieve arbitrarily close to maximum MABK violation whilst certifying maximum DI randomness, for an arbitrarily large finite NN. ∎

B.6 Local dilution measure of nonlocality

In this section we briefly outline the linear program used to generate Fig.˜3 in the main text (see also [57, Section I.B.1] for a description of this technique). Given a distribution {p(𝒂|𝒙)}\{p(\bm{a}|\bm{x})\}, we can represent it as a vector in a high dimensional space, p22Np\in\mathbb{R}^{2^{2N}}. We denote the set of 22N2^{2N} local deterministic distributions as {dλ}λ\{d_{\lambda}\}_{\lambda}, and the local polytope can be constructed by taking its convex hull,

={λqλdλ|λqλ=1,qλ0}.\mathcal{L}=\Big{\{}\sum_{\lambda}q_{\lambda}d_{\lambda}\ |\ \sum_{\lambda}q_{\lambda}=1,\ q_{\lambda}\geq 0\Big{\}}. (149)

To check wether pp admits a local description, we can search for a set {qλ}λ\{q_{\lambda}\}_{\lambda}, where λqλ=1\sum_{\lambda}q_{\lambda}=1, qλ0q_{\lambda}\geq 0, using a standard linear program.

To quantify the distance of a given pp from \mathcal{L}, we define the diluted sets, for ϵ0\epsilon\geq 0,

ϵ={λqλdλ|λqλ=1,qλϵ}.\mathcal{L}_{\epsilon}=\Big{\{}\sum_{\lambda}q_{\lambda}d_{\lambda}\ |\ \sum_{\lambda}q_{\lambda}=1,\ q_{\lambda}\geq-\epsilon\Big{\}}. (150)

As ϵ\epsilon\rightarrow\infty, ϵ\mathcal{L}_{\epsilon} tends to the affine hull of the points {dλ}λ\{d_{\lambda}\}_{\lambda}, and we have the inclusion: ϵ1ϵ2\mathcal{L}\subset\mathcal{L}_{\epsilon_{1}}\subset\mathcal{L}_{\epsilon_{2}}\subset\mathcal{L}_{\infty}, where 0<ϵ1<ϵ20<\epsilon_{1}<\epsilon_{2}. We then define the following measure, which we call the local dilution:

𝒟[p]:=min{ϵ|pϵ}.\mathcal{D}[p]:=\min\{\epsilon\ |\ p\in\mathcal{L}_{\epsilon}\}. (151)

𝒟[p]\mathcal{D}[p] can be interpreted as the smallest dilution of the local polytope that contains pp. If 𝒟[p1]<𝒟[p2]\mathcal{D}[p_{1}]<\mathcal{D}[p_{2}], one has to enlarge the local set more to include p2p_{2} than to include p1p_{1}. We can hence understand p2p_{2} to be more nonlocal than p1p_{1}.

B.7 Proof of Lemma˜8

In the main text, we detailed a two parameter family of bipartite Bell expressions which are used to build our NN-partite expressions. These are generalizations of the two families introduced in [23], and can be recovered as sub-families of the self-tests studied in [50, 51, 49].

Lemma 8. Let (ϕ,θ)2(\phi,\theta)\in\mathbb{R}^{2} such that cos(2θ)cos(2ϕ)<0\cos(2\theta)\cos(2\phi)<0 and cos(θϕ)0\cos(\theta-\phi)\neq 0. Define the family of Bell expressions parameterized by ϕ\phi and θ\theta,

Jϕ,θ(k,l)=cos2θcos(θϕ)A0(k)A0(l)cos2θcos2ϕ(A0(k)A1(l)+A1(k)A0(l))+cos2ϕcos(θϕ)A1(k)A1(l).\langle J_{\phi,\theta}^{(k,l)}\rangle=\cos 2\theta\cos(\theta-\phi)\langle A_{0}^{(k)}A_{0}^{(l)}\rangle\\ -\cos 2\theta\cos 2\phi\big{(}\langle A_{0}^{(k)}A_{1}^{(l)}\rangle+\langle A_{1}^{(k)}A_{0}^{(l)}\rangle\big{)}\\ +\cos 2\phi\cos(\theta-\phi)\langle A_{1}^{(k)}A_{1}^{(l)}\rangle. (152)

Then we have the following:

  1. (i)

    The local bounds are given by ±ηϕ,θL\pm\eta_{\phi,\theta}^{\mathrm{L}}, where

    ηϕ,θL=max{|cos(θϕ)(cos2θcos2ϕ)|,|cos(θϕ)(cos2θ+cos2ϕ)|+|2cos2ϕcos2θ|}.\eta_{\phi,\theta}^{\mathrm{L}}=\max\big{\{}|\cos(\theta-\phi)(\cos 2\theta-\cos 2\phi)|,\\ |\cos(\theta-\phi)\big{(}\cos 2\theta+\cos 2\phi\big{)}|+|2\cos 2\phi\cos 2\theta|\big{\}}. (153)
  2. (ii)

    The quantum bounds are given by ±ηϕ,θQ\pm\eta_{\phi,\theta}^{\mathrm{Q}}, where ηϕ,θQ=2sin2(θ+ϕ)sin(θϕ)\eta_{\phi,\theta}^{\mathrm{Q}}=2\sin^{2}(\theta+\phi)\sin(\theta-\phi).

  3. (iii)

    |ηϕ,θQ|>ηϕ,θL|\eta^{\mathrm{Q}}_{\phi,\theta}|>\eta^{\mathrm{L}}_{\phi,\theta}.

  4. (iv)

    Up to local isometries, there exists a unique strategy that achieves Jϕ,θ(k,l)=ηϕ,θQ\langle J_{\phi,\theta}^{(k,l)}\rangle=\eta_{\phi,\theta}^{\mathrm{Q}}:

    ρQkQl\displaystyle\rho_{Q_{k}Q_{l}} =|ψψ|,where|ψ=12(|00+i|11)\displaystyle=|\psi\rangle\!\langle\psi|,\ \mathrm{where}\ |\psi\rangle=\frac{1}{\sqrt{2}}\big{(}|0\rangle+i|1\rangle\big{)} (154)
    A0(k)\displaystyle A_{0}^{(k)} =A0(l)=cosϕσXsinϕσY\displaystyle=A_{0}^{(l)}=\cos\phi\,\sigma_{X}-\sin\phi\,\sigma_{Y}
    A1(k)\displaystyle A_{1}^{(k)} =A1(l)=cosθσX+sinθσY.\displaystyle=A_{1}^{(l)}=\cos\theta\,\sigma_{X}+\sin\theta\,\sigma_{Y}.
Proof.

The above is a direct corollary of [51, Proposition 1]. Specifically, let β=2ϕ\beta=-2\phi, γ=θϕ\gamma=\theta-\phi and α=θ+ϕ\alpha=\theta+\phi. Then the Bell expression Bα,β,γB_{\alpha,\beta,\gamma} of [51, Proposition 1] is equal to the expression Jϕ,θJ_{\phi,\theta} up to a factor of cos(θϕ)\cos(\theta-\phi). The condition for ηϕ,θL<|ηϕ,θQ|\eta^{\mathrm{L}}_{\phi,\theta}<|\eta^{\mathrm{Q}}_{\phi,\theta}| is given by [51, Equation 4], which reads

cos(α+β)cosβcos(α+γ)cosγ=cos2(θϕ)cos2θcos2ϕ<0,\cos(\alpha+\beta)\cos\beta\cos(\alpha+\gamma)\cos\gamma=\\ \cos^{2}(\theta-\phi)\cos 2\theta\cos 2\phi<0, (155)

and is satisfied if and only if cos(2θ)cos(2ϕ)<0\cos(2\theta)\cos(2\phi)<0 and cos(θϕ)0\cos(\theta-\phi)\neq 0. ∎

Lemma˜4 can be proven as a special case of the above, by setting ϕ=0\phi=0. This follows from 𝒢\mathcal{G}\subset\mathcal{F}, where \mathcal{F} is a set of valid self-testing points defined below Lemma˜8.

One can also recover the IδI_{\delta} and JγJ_{\gamma} family of Bell expressions from [23], as special cases of the Jϕ,θJ_{\phi,\theta} expressions. By choosing ϕ=0\phi=0, and θ=δ+π/2\theta=\delta+\pi/2, δ(0,π/6]\delta\in(0,\pi/6], we obtain

J0,δ+π/2=sin(δ)cos(2δ)A0(k)A0(l)+cos(2δ)(A0(k)A1(l)+A1(k)A0(l))sin(δ)A1(k)A1(l)2cos3(δ),\langle J_{0,\delta+\pi/2}\rangle=\sin(\delta)\cos(2\delta)\langle A_{0}^{(k)}A_{0}^{(l)}\rangle\\ +\cos(2\delta)\big{(}\langle A_{0}^{(k)}A_{1}^{(l)}\rangle+\langle A_{1}^{(k)}A_{0}^{(l)}\rangle\big{)}\\ -\sin(\delta)\langle A_{1}^{(k)}A_{1}^{(l)}\rangle\leq 2\cos^{3}(\delta), (156)

which is identical to the expression found in [23]. One can obtain the JγJ_{\gamma} family by choosing ϕ=3γ/2\phi=-3\gamma/2 and θ=ϕ/3+2π/3\theta=\phi/3+2\pi/3 using Eq.˜37, and multiplying though by 4cos2(γ+π/6)/sin2(2γ+π/3)4\cos^{2}(\gamma+\pi/6)/\sin^{2}(2\gamma+\pi/3):

4cos2(γ+π/6)sin2(2γ+π/3)J3γ/2,γ/2+2π/3=A0(k)A0(l)+(4cos2(γ+π/6)1)(A0(k)A1(l)+A1(k)A0(l)A1(k)A1(l))8cos3(γ+π/6).\frac{4\cos^{2}(\gamma+\pi/6)}{\sin^{2}(2\gamma+\pi/3)}\langle J_{-3\gamma/2,-\gamma/2+2\pi/3}\rangle=\langle A_{0}^{(k)}A_{0}^{(l)}\rangle\\ +\Big{(}4\cos^{2}(\gamma+\pi/6)-1\Big{)}\Big{(}\langle A_{0}^{(k)}A_{1}^{(l)}\rangle+\langle A_{1}^{(k)}A_{0}^{(l)}\rangle\\ -\langle A_{1}^{(k)}A_{1}^{(l)}\rangle\Big{)}\leq 8\cos^{3}(\gamma+\pi/6). (157)

B.8 Proof of Lemma˜9

Lemma 9. Let 𝛍\bm{\mu} be a tuple of N2N-2 measurement outcomes for all parties excluding k,lk,l, and n𝛍{0,1}n_{\bm{\mu}}\in\{0,1\} be the parity of 𝛍\bm{\mu}. Let (ϕ,θ)2(\phi,\theta)\in\mathbb{R}^{2},

ϕ:=ϕN2,θ:=θN22ϕ,\phi^{\prime}:=\frac{\phi N}{2},\ \theta^{\prime}:=\theta-\frac{N-2}{2}\phi, (158)

and

I𝝁(k,l)=(1)n𝝁Jϕ,θ(k,l).I_{\bm{\mu}}^{(k,l)}=(-1)^{n_{\bm{\mu}}}J_{\phi^{\prime},\theta^{\prime}}^{(k,l)}. (159)

Define the following Bell polynomial

Iϕ,θ:=k=1N1(𝝁P~𝝁|𝟎(k,N)¯I𝝁(k,N)).I_{\phi^{\prime},\theta^{\prime}}:=\sum_{k=1}^{N-1}\Bigg{(}\sum_{\bm{\mu}}\tilde{P}^{\overline{(k,N)}}_{\bm{\mu}|\bm{0}}I_{\bm{\mu}}^{(k,N)}\Bigg{)}. (160)

If (ϕ,θ)(\phi^{\prime},\theta^{\prime})\in\mathcal{F}, Iϕ,θI_{\phi^{\prime},\theta^{\prime}} is an expanded Bell expression, and has quantum bounds ±ηN,ϕ,θQ\pm\eta_{N,\phi^{\prime},\theta^{\prime}}^{\mathrm{Q}} where ηN,ϕ,θQ=2(N1)sin2(θ+ϕ)sin(θϕ)\eta_{N,\phi^{\prime},\theta^{\prime}}^{\mathrm{Q}}=2(N-1)\sin^{2}(\theta^{\prime}+\phi^{\prime})\sin(\theta^{\prime}-\phi^{\prime}), which can be achieved up to relabelings by the strategy

ρ𝑸=|ψGHZψGHZ|,A0(k)=cosϕσXsinϕσY,A1(k)=cosθσX+sinθσY,k{1,,N}.\begin{gathered}\rho_{\bm{Q}}=|\psi_{\mathrm{GHZ}}\rangle\!\langle\psi_{\mathrm{GHZ}}|,\\ A_{0}^{(k)}=\cos\phi\,\sigma_{X}-\sin\phi\,\sigma_{Y},\\ A_{1}^{(k)}=\cos\theta\,\sigma_{X}+\sin\theta\,\sigma_{Y},\ k\in\{1,...,N\}.\end{gathered} (161)

In addition, this quantum bound cannot be achieved classically.

Proof.

Suppose (ϕ,θ)(\phi^{\prime},\theta^{\prime})\in\mathcal{F}, and consider the ideal projection operators Pak|0(k)=|φakφak|P_{a_{k^{\prime}}|0}^{(k^{\prime})}=|\varphi_{a_{k^{\prime}}}\rangle\!\langle\varphi_{a_{k^{\prime}}}| derived from the observables in Eq.˜31, i.e., cosϕσXsinϕσY=|φ0φ0||φ1φ1|\cos\phi\,\sigma_{X}-\sin\phi\,\sigma_{Y}=|\varphi_{0}\rangle\!\langle\varphi_{0}|-|\varphi_{1}\rangle\!\langle\varphi_{1}| where

|φak=|0+(1)akeiϕ|12.|\varphi_{a_{k^{\prime}}}\rangle=\frac{|0\rangle+(-1)^{a_{k^{\prime}}}e^{-i\phi}|1\rangle}{\sqrt{2}}. (162)

We then have

P𝝁|𝟎(k,N)¯|ψGHZ=ψGHZ|P𝝁|𝟎(k,N)¯|ψGHZ|φ𝝁|Φ𝝁,ϕQkQN,P_{\bm{\mu}|\bm{0}}^{\overline{(k,N)}}|\psi_{\mathrm{GHZ}}\rangle=\\ \sqrt{\langle\psi_{\mathrm{GHZ}}|P_{\bm{\mu}|\bm{0}}^{\overline{(k,N)}}|\psi_{\mathrm{GHZ}}\rangle}|\varphi_{\bm{\mu}}\rangle\otimes|\Phi_{\bm{\mu},\phi}\rangle_{Q_{k}Q_{N}}, (163)

where |φ𝝁|\varphi_{\bm{\mu}}\rangle spans the support of P𝝁|𝟎(k,N)¯P_{\bm{\mu}|\bm{0}}^{\overline{(k,N)}} and

|Φ𝝁,ϕ=|00+i(1)n𝝁ei(N2)ϕ|112.|\Phi_{\bm{\mu},\phi}\rangle=\frac{|00\rangle+i(-1)^{n_{\bm{\mu}}}e^{i(N-2)\phi}|11\rangle}{\sqrt{2}}. (164)

Notice |Φ𝝁,ϕ=(UϕUϕ)|Φ𝝁|\Phi_{\bm{\mu},\phi}\rangle=(U_{\phi}\otimes U_{\phi})|\Phi_{\bm{\mu}}\rangle where Uϕ=|00|+ei(N2)ϕ/2|11|U_{\phi}=|0\rangle\!\langle 0|+e^{i(N-2)\phi/2}|1\rangle\!\langle 1|. We now calculate

Φ𝝁,ϕ|I𝝁(k,l)|Φ𝝁,ϕ=(1)n𝝁Φ𝝁|(UϕUϕ)Jϕ,θ(k,N)(UϕUϕ)|Φ𝝁.\langle\Phi_{\bm{\mu},\phi}|I_{\bm{\mu}}^{(k,l)}|\Phi_{\bm{\mu},\phi}\rangle\\ =(-1)^{n_{\bm{\mu}}}\langle\Phi_{\bm{\mu}}|(U_{\phi}\otimes U_{\phi})^{\dagger}J_{\phi^{\prime},\theta^{\prime}}^{(k,N)}(U_{\phi}\otimes U_{\phi})|\Phi_{\bm{\mu}}\rangle. (165)

We can now calculate how I𝝁(k,N)I_{\bm{\mu}}^{(k,N)} transforms under (UϕUϕ)(U_{\phi}\otimes U_{\phi})^{\dagger}. Let O=cos(t)σX+sin(t)σYO=\cos(t)\,\sigma_{X}+\sin(t)\,\sigma_{Y} for tt\in\mathbb{R} be an arbitrary observable. Then

UϕOUϕ=cos(N22ϕ+t)σX+sin(N22ϕ+t)σY.U_{\phi}^{\dagger}OU_{\phi}=\cos\Big{(}-\frac{N-2}{2}\phi+t\Big{)}\,\sigma_{X}+\sin\Big{(}-\frac{N-2}{2}\phi+t\Big{)}\,\sigma_{Y}. (166)

Substituting t=ϕt=-\phi and t=θt=\theta into the above, we find

UϕA0(k)Uϕ\displaystyle U_{\phi}^{\dagger}A_{0}^{(k)}U_{\phi} =cosϕσXsinϕσY=:A0(k)\displaystyle=\cos\phi^{\prime}\,\sigma_{X}-\sin\phi^{\prime}\sigma_{Y}=:A_{0}^{{}^{\prime}(k)}
UϕA1(k)Uϕ\displaystyle U_{\phi}^{\dagger}A_{1}^{(k)}U_{\phi} =cosθσX+sinθσY=:A1(k),\displaystyle=\cos\theta^{\prime}\,\sigma_{X}+\sin\theta^{\prime}\sigma_{Y}=:A_{1}^{{}^{\prime}(k)}, (167)

and similarly for party NN, where we used the definitions of Ax(k)A_{x}^{(k)} in Eq.˜154. We therefore find

J:=(UϕUϕ)Jϕ,θ(k,N)(UϕUϕ)=cos2θcos(θϕ)A0(k)A0(N)cos2θcos2ϕ(A0(k)A1(N)+A1(k)A0(N))+cos2ϕcos(θϕ)A1(k)A1(N),J^{\prime}:=(U_{\phi}\otimes U_{\phi})^{\dagger}J_{\phi^{\prime},\theta^{\prime}}^{(k,N)}(U_{\phi}\otimes U_{\phi})\\ =\cos 2\theta^{\prime}\cos(\theta^{\prime}-\phi^{\prime})A_{0}^{{}^{\prime}(k)}\otimes A_{0}^{{}^{\prime}(N)}\\ -\cos 2\theta^{\prime}\cos 2\phi^{\prime}\big{(}A_{0}^{{}^{\prime}(k)}\otimes A_{1}^{{}^{\prime}(N)}+A_{1}^{{}^{\prime}(k)}\otimes A_{0}^{{}^{\prime}(N)}\big{)}\\ +\cos 2\phi^{\prime}\cos(\theta^{\prime}-\phi^{\prime})A_{1}^{{}^{\prime}(k)}\otimes A_{1}^{{}^{\prime}(N)}, (168)

The largest and smallest eigenvalues of JJ^{\prime} are ±ηθ,ϕQ\pm\eta^{\mathrm{Q}}_{\theta^{\prime},\phi^{\prime}}, and are associated with the eigenvectors (|00±i|11)/2(|00\rangle\pm i|11\rangle)/\sqrt{2}. We therefore have

Φ𝝁,ϕ|I𝝁(k,N)|Φ𝝁,ϕ=(1)n𝝁Φ𝝁|J|Φ𝝁=ηϕ,θQ.\langle\Phi_{\bm{\mu},\phi}|I_{\bm{\mu}}^{(k,N)}|\Phi_{\bm{\mu},\phi}\rangle=(-1)^{n_{\bm{\mu}}}\langle\Phi_{\bm{\mu}}|J^{\prime}|\Phi_{\bm{\mu}}\rangle=\eta_{\phi^{\prime},\theta^{\prime}}^{\mathrm{Q}}. (169)

Putting everything together, we obtain

𝝁\displaystyle\sum_{\bm{\mu}} ψGHZ|P𝝁|𝟎(k,N)¯I𝝁(k,N)|ψGHZ\displaystyle\langle\psi_{\mathrm{GHZ}}|P^{\overline{(k,N)}}_{\bm{\mu}|\bm{0}}\otimes I_{\bm{\mu}}^{(k,N)}|\psi_{\mathrm{GHZ}}\rangle
=𝝁ψGHZ|P𝝁|𝟎(k,N)¯|ψGHZΦ𝝁,ϕ|I𝝁(k,N)|Φ𝝁,ϕ\displaystyle=\sum_{\bm{\mu}}\langle\psi_{\mathrm{GHZ}}|P_{\bm{\mu}|\bm{0}}^{\overline{(k,N)}}|\psi_{\mathrm{GHZ}}\rangle\langle\Phi_{\bm{\mu},\phi}|I_{\bm{\mu}}^{(k,N)}|\Phi_{\bm{\mu},\phi}\rangle
=ηϕ,θQ,k{1,,N1}.\displaystyle=\eta_{\phi^{\prime},\theta^{\prime}}^{\mathrm{Q}},\ \forall k\in\{1,...,N-1\}. (170)

The quantum bound ηN,ϕ,θQ\eta_{N,\phi^{\prime},\theta^{\prime}}^{\mathrm{Q}} then follows by summing over kk, which shows achievability. The minimum quantum value ηN,ϕ,θQ-\eta_{N,\phi^{\prime},\theta^{\prime}}^{\mathrm{Q}} is achieved by relabeling the outcomes. Moreover, the quantum bound cannot be achieved classically since Iϕ,θI_{\phi^{\prime},\theta^{\prime}} corresponds to an expanded Bell inequality when (ϕ,θ)(\phi^{\prime},\theta^{\prime})\in\mathcal{F} (cf. the discussion below Eq.˜49). ∎

B.9 Proof of Proposition˜7

Proposition 7. Let ϕ[0,ϕN]\phi\in[0,\phi_{N}^{*}], and θ=θ(ϕ)\theta=\theta(\phi) as defined in Eqs.˜37 and 38. Then (ϕ,θ)(\phi^{\prime},\theta^{\prime})\in\mathcal{F}, where ϕ\phi^{\prime} and θ\theta^{\prime} are defined in Lemma˜9.

Proof.

For convenience, we restate the set \mathcal{F} below,

=[(π/4,π/4)×𝒢][(π/4,π/4){0}×{π/2,3π/2}].\mathcal{F}=\Big{[}(-\pi/4,\pi/4)\times\mathcal{G}\Big{]}\cup\Big{[}(-\pi/4,\pi/4)\setminus\{0\}\times\{\pi/2,3\pi/2\}\Big{]}. (171)

First we will consider the range of ϕ\phi^{\prime} as we vary ϕ\phi. By definition, ϕ=ϕN/2\phi^{\prime}=\phi N/2, which takes values over the interval [0,ϕNN/2][0,\phi^{*}_{N}N/2]. Then

ϕNN2=sgn[sin(2θN)]π8,\frac{\phi^{*}_{N}N}{2}=\mathrm{sgn}[\sin(2\theta_{N}^{*})]\frac{\pi}{8}, (172)

which implies π/4<ϕ<π/4-\pi/4<\phi^{\prime}<\pi/4.

Next we consider θ\theta^{\prime} as a function of ϕ\phi,

θ=θ(ϕ)N22ϕ=N(N3)2(N+1)ϕ+θN.\theta^{\prime}=\theta(\phi)-\frac{N-2}{2}\phi=-\frac{N(N-3)}{2(N+1)}\phi+\theta_{N}^{*}. (173)

When ϕ=0\phi=0, θ=θN\theta^{\prime}=\theta_{N}^{*}, and when ϕ=ϕN\phi=\phi_{N}^{*},

θ\displaystyle\theta^{\prime} =sgn[sin(2θN)]π8N3N+1+θN=:θ~.\displaystyle=-\mathrm{sgn}[\sin(2\theta_{N}^{*})]\frac{\pi}{8}\frac{N-3}{N+1}+\theta_{N}^{*}=:\tilde{\theta}^{\prime}. (174)

We first consider N=8n+2N=8n+2, which satisfies sgn[sin(2θN)]=1\mathrm{sgn}[\sin(2\theta_{N}^{*})]=-1 and π/2<θN<3π/4\pi/2<\theta_{N}^{*}<3\pi/4. By inserting the definition of θN\theta^{*}_{N} from Proposition˜3,

θ~=π88n18n+3+2π8n+3(8n+24+12)=5π8(π/2,3π/4).\tilde{\theta}^{\prime}=\frac{\pi}{8}\frac{8n-1}{8n+3}+\frac{2\pi}{8n+3}\big{(}\frac{8n+2}{4}+\frac{1}{2}\big{)}=\frac{5\pi}{8}\in(\pi/2,3\pi/4). (175)

We therefore see θ(π/2,3π/4)\theta^{\prime}\in(\pi/2,3\pi/4) (since θ\theta^{\prime} is a linear function of ϕ\phi and lies in that range for the extremal values of ϕ\phi).

For N=8n+8N=8n+8, sgn[sin(2θN)]=1\mathrm{sgn}[\sin(2\theta_{N}^{*})]=-1 and 3π/2<θN<7π/43\pi/2<\theta_{N}^{*}<7\pi/4. We have

θ~=π88n+58n+9+2π8n+9(3(8n+8)4+1)=13π8(3π/2,7π/4),\tilde{\theta}^{\prime}=\frac{\pi}{8}\frac{8n+5}{8n+9}+\frac{2\pi}{8n+9}\big{(}\frac{3(8n+8)}{4}+1\big{)}\\ =\frac{13\pi}{8}\in(3\pi/2,7\pi/4), (176)

implying 3π/2<θ<7π/43\pi/2<\theta^{\prime}<7\pi/4. For N=8n+4N=8n+4, we have sgn[sin(2θN)]=+1\mathrm{sgn}[\sin(2\theta_{N}^{*})]=+1, π/4<θN<π/2\pi/4<\theta_{N}^{*}<\pi/2 and

θ~=π88n+18n+5+2π8n+58n+44=3π8(π/4,π/2),\tilde{\theta}^{\prime}=-\frac{\pi}{8}\frac{8n+1}{8n+5}+\frac{2\pi}{8n+5}\frac{8n+4}{4}=\frac{3\pi}{8}\in(\pi/4,\pi/2), (177)

implying π/4<θ<π/2\pi/4<\theta^{\prime}<\pi/2. Finally, for N=8n+6N=8n+6, sgn[sin(2θN)]=+1\mathrm{sgn}[\sin(2\theta_{N}^{*})]=+1, 5π/4<θN<3π/25\pi/4<\theta_{N}^{*}<3\pi/2 and

θ~=π88n+38n+7+2π8n+7(3(8n+6)4+12)=11π8(5π/4,3π/2),\tilde{\theta}^{\prime}=-\frac{\pi}{8}\frac{8n+3}{8n+7}+\frac{2\pi}{8n+7}\Big{(}\frac{3(8n+6)}{4}+\frac{1}{2}\Big{)}\\ =\frac{11\pi}{8}\in(5\pi/4,3\pi/2), (178)

implying π/4<θ<π/2\pi/4<\theta^{\prime}<\pi/2. For every NN, we have shown θ𝒢\theta^{\prime}\in\mathcal{G} and ϕ(π/4,π/4)\phi^{\prime}\in(-\pi/4,\pi/4) for ϕ[0,ϕN]\phi\in[0,\phi^{*}_{N}], hence (ϕ,θ)(\phi^{\prime},\theta^{\prime})\in\mathcal{F}. ∎

B.10 Evidence for Eq.˜45

Conjecture 2. For NN even, the maximum randomness unconditioned on Eve, rr, that can be generated by quantum strategies achieving an MABK value, ss, is given by

r(s)={N,s(1,mN],r(ϕs),s(mN,2(N1)/2],r(s)=\begin{cases}N,\ s\in(1,m_{N}^{*}],\\ r(\phi_{s}),\ s\in(m_{N}^{*},2^{(N-1)/2}],\end{cases} (179)

where r(ϕ)r(\phi) is defined in Eq.˜40, and

ϕs=argmin{|ϕ||ϕ[0,ϕN],MN(ϕ,θ(ϕ))=s}.\phi_{s}=\mathrm{arg\,min}\big{\{}|\phi|\ |\ \phi\in[0,\phi_{N}^{*}],\langle M_{N}(\phi,\theta(\phi))\rangle=s\big{\}}. (180)

B.10.1 Numerical approach for upper bounding maximum randomness versus MABK value

The objective of this section is to generalize the technique introduced in [23], which found upper bounds on the maximum randomness versus CHSH value using an SOS technique. We begin by defining the problem in detail.

Let 𝒬\mathcal{Q} denote the set of quantum distributions, and (P)\mathcal{M}(P) denote the MABK value of a distribution P𝒬P\in\mathcal{Q}. Moreover, let H(𝑹|𝑿=𝟎,E)PH(\bm{R}|\bm{X}=\bm{0},E)_{P} be the conditional von Neumann entropy of the outputs 𝑹\bm{R} for inputs 𝑿=𝟎\bm{X}=\bm{0} given the observed distribution PP, taking the infimum over all quantum strategies that could give rise to PP, i.e.,

H(𝑹|𝑿=𝟎,E)P:=inf|Ψ𝑸~E,{{P~ak|xk(k)}ak}k,compatible withPH(𝑹|𝑿=𝟎,E)ρ𝑹E|𝟎.H(\bm{R}|\bm{X}=\bm{0},E)_{P}:=\inf_{\begin{subarray}{c}|\Psi\rangle_{\tilde{\bm{Q}}E},\\ \big{\{}\{\tilde{P}_{a_{k}|x_{k}}^{(k)}\}_{a_{k}}\big{\}}_{k},\\ \text{compatible with}\ P\end{subarray}}H(\bm{R}|\bm{X}=\bm{0},E)_{\rho_{\bm{R}E|\bm{0}}}. (181)

Similarly let H(𝑹|𝑿=𝟎)PH(\bm{R}|\bm{X}=\bm{0})_{P} be the Shannon entropy of the distribution on 𝑹\bm{R} for inputs 𝑿=𝟎\bm{X}=\bm{0}. Then the curve R:[mN,2(N1)/2][0,N],sR(s)R:[m_{N}^{*},2^{(N-1)/2}]\to[0,N],\ s\mapsto R(s) we want to find is defined by the optimization

R(s)=maxP\displaystyle R(s)=\max_{P}\ H(𝑹|𝑿=𝟎,E)P\displaystyle H(\bm{R}|\bm{X}=\bm{0},E)_{P}
s.t. (P)=s,\displaystyle\mathcal{M}(P)=s,
P𝒬.\displaystyle P\in\mathcal{Q}. (182)

Our technique proceeds by defining a sequence of upper bounds on Graph[R(s)]={(s,r)|r=R(s)}\text{Graph}[R(s)]=\{(s,r)\ |\ r=R(s)\}. [We speak of upper bounds on Graph[R(s)]\text{Graph}[R(s)] because we either upper bound R(s)R(s) or R1(r)R^{-1}(r), and because the functions turn out to be monotonic either gives an upper bound.]

We begin by removing the dependence on Eve, which results in an upper bound R(s)R¯(s)R(s)\leq\bar{R}(s) on the previous problem via strong subadditivity of the von Neumann entropy. Explicitly,

R(s)R¯(s)=maxP\displaystyle R(s)\leq\bar{R}(s)=\max_{P}\ H(𝑹|𝑿=𝟎)P\displaystyle H(\bm{R}|\bm{X}=\bm{0})_{P}
s.t. (P)=s,\displaystyle\mathcal{M}(P)=s,
P𝒬.\displaystyle P\in\mathcal{Q}. (183)

Next we state the following two lemmas, adapted from Ref. [23].

Lemma 12.

[Monotonicity of R¯(s)\bar{R}(s)] Suppose mNm_{N}^{*} is the largest MABK value compatible with p(𝐚|𝟎)=2Np(\bm{a}|\bm{0})=2^{-N} (cf. Conjecture 1). Then the function R¯(s)\bar{R}(s) is strictly decreasing on [mN,2(N1)/2][m_{N}^{*},2^{(N-1)/2}].

We will prove Lemma 12 as a corollary of the following.

Lemma 13.

Let 𝒬\mathcal{Q}^{\prime} be a convex set, H:𝒬H:\mathcal{Q}^{\prime}\to\mathbb{R} be concave, :𝒬\mathcal{M}^{\prime}:\mathcal{Q}^{\prime}\to\mathbb{R} be linear and R¯(s)=maxPH(P)\bar{R}^{\prime}(s)=\max_{P}H(P) subject to (P)=s\mathcal{M}^{\prime}(P)=s, P𝒬P\in\mathcal{Q}^{\prime}. Then R¯(s)\bar{R}^{\prime}(s) is convex.

Proof.

We directly calculate

R¯[λs1+(1λ)s2]=maxP\displaystyle\bar{R}^{\prime}[\lambda s_{1}+(1-\lambda)s_{2}]=\max_{P}\ H(P)\displaystyle H(P)
s.t. (P)=λs1+(1λ)s2,\displaystyle\mathcal{M}^{\prime}(P)=\lambda s_{1}+(1-\lambda)s_{2},
P𝒬,\displaystyle P\in\mathcal{Q}^{\prime},
maxP1,P2\displaystyle\geq\max_{P_{1},P_{2}}\ H(λP1+(1λ)P2)\displaystyle H(\lambda P_{1}+(1-\lambda)P_{2})
s.t. (P1)=s1,(P2)=s2,\displaystyle\mathcal{M}^{\prime}(P_{1})=s_{1},\ \mathcal{M}^{\prime}(P_{2})=s_{2},
P1,P2𝒬,\displaystyle P_{1},P_{2}\in\mathcal{Q}^{\prime},
maxP1,P2\displaystyle\geq\max_{P_{1},P_{2}}\ λH(P1)+(1λ)H(P2)\displaystyle\lambda H(P_{1})+(1-\lambda)H(P_{2})
s.t. (P1)=s1,(P2)=s2,\displaystyle\mathcal{M}^{\prime}(P_{1})=s_{1},\ \mathcal{M}^{\prime}(P_{2})=s_{2},
P1,P2𝒬,\displaystyle P_{1},P_{2}\in\mathcal{Q}^{\prime},
=λR¯(s1\displaystyle=\lambda\bar{R}^{\prime}(s_{1} )+(1λ)R¯(s2),\displaystyle)+(1-\lambda)\bar{R}^{\prime}(s_{2}), (184)

where the first inequality follows due to the linearity of \mathcal{M}^{\prime} and convexity of 𝒬\mathcal{Q}^{\prime}, and the second inequality due to the concavity of HH. ∎

Proof of Lemma 12.

Note that H(𝑹|𝑿=𝟎)PH(\bm{R}|\bm{X}=\bm{0})_{P} is concave in PP, the quantum set, 𝒬\mathcal{Q}, is convex and the MABK value, \mathcal{M}, is linear. Lemma 13 hence implies that R¯(s)\bar{R}(s) is concave in ss.

Then note that R¯(mN)>R¯(s)s(mN,2(N1)/2]\bar{R}(m_{N}^{*})>\bar{R}(s)\ \forall s\in(m_{N}^{*},2^{(N-1)/2}]. This follows from the assumption that the largest MABK value achievable when p(𝒂|𝟎)=2Np(\bm{a}|\bm{0})=2^{-N} is mNm_{N}^{*}, and hence R¯(s)\bar{R}(s) is initially decreasing, which implies the claim. ∎

In the case N=2N=2, it is known m2m_{2}^{*} is the largest MABK value compatible with p(a,b|0,0)=1/4p(a,b|0,0)=1/4; see [23, Corollary 1]. For N=4,6,8,10,12N=4,6,8,10,12, we have verified tightness numerically, as shown in Table˜1101010Monotonicity of R¯(s)\bar{R}(s) is not needed to establish [23, Corollary 1] or the upper bounds presented in Table 1.

Lemma 14 (Inverse function of R¯(s)\bar{R}(s)).

Suppose r=R¯(s)r=\bar{R}(s). The function R¯(s)\bar{R}(s) has the following inverse, denoted R¯1\bar{R}^{-1}, that satisfies s=R¯1(r)s=\bar{R}^{-1}(r), given by

R¯1(r)=maxP\displaystyle\bar{R}^{-1}(r)=\max_{P}\ (P)\displaystyle\mathcal{M}(P)
s.t.\displaystyle\mathrm{s.t.}\ \ H(𝑹|𝑿=𝟎)P=r,\displaystyle H(\bm{R}|\bm{X}=\bm{0})_{P}=r,
P𝒬.\displaystyle P\in\mathcal{Q}. (185)
Proof.

We prove Lemma˜14 by showing R¯1(R¯(s))=s\bar{R}^{-1}(\bar{R}(s))=s, and R¯(R¯1(r))=r\bar{R}(\bar{R}^{-1}(r))=r, and using Lemma˜12. First consider R¯1(R¯(s))=s\bar{R}^{-1}(\bar{R}(s))=s,

R¯1(R¯(s))=maxP\displaystyle\bar{R}^{-1}(\bar{R}(s))=\max_{P}\ (P)\displaystyle\mathcal{M}(P)
s.t. H(𝑹|𝑿=𝟎)P=R¯(s),\displaystyle H(\bm{R}|\bm{X}=\bm{0})_{P}=\bar{R}(s),
P𝒬.\displaystyle P\in\mathcal{Q}. (186)

The constraint H(𝑹|𝑿=𝟎)P=R¯(s)H(\bm{R}|\bm{X}=\bm{0})_{P}=\bar{R}(s) implies the achievable MABK values for the distribution PP must lie to the left of ss, i.e., (P)s\mathcal{M}(P)\leq s, since the curve R¯(s)\bar{R}(s) is strictly decreasing (cf. Lemma˜12). To see this, suppose s=(P)>ss^{\prime}=\mathcal{M}(P)>s. Then R¯(s)<R¯(s)\bar{R}(s^{\prime})<\bar{R}(s) by Lemma˜12. But, R¯(s)H(𝑹|𝑿=𝟎)P\bar{R}(s^{\prime})\geq H(\bm{R}|\bm{X}=\bm{0})_{P}, implying H(𝑹|𝑿=𝟎)P<R¯(s)H(\bm{R}|\bm{X}=\bm{0})_{P}<\bar{R}(s), which by assumption cannot hold, giving a contradiction. We therefore have that R¯1(R¯(s))max{P𝒬s.t.(P)s}(P)=s\bar{R}^{-1}(\bar{R}(s))\leq\max_{\{P\in\mathcal{Q}\ \text{s.t.}\ \mathcal{M}(P)\leq s\}}\mathcal{M}(P)=s. Moreover, this bound is saturated by the behaviour PP^{*} which achieves the maximum defining R¯(s)\bar{R}(s)111111Note that the function PH(𝑹|𝑿=𝟎)PP\mapsto H(\bm{R}|\bm{X}=\bm{0})_{P} is continuous, and the quantum set 𝒬\mathcal{Q} is compact (see the discussion in [58, Appendix B]), from which it follows that 𝒬s={P𝒬:(P)=s}\mathcal{Q}_{s}=\{P\in\mathcal{Q}\ :\ \mathcal{M}(P)=s\} is also compact. This implies the objective function in (183) attains its maximum over its domain., i.e., satisfies R¯(s)=H(AB|X=0,Y=0)P\bar{R}(s)=H(AB|X=0,Y=0)_{P^{*}} and (P)=s\mathcal{M}(P^{*})=s. This establishes R¯1(R¯(s))=s\bar{R}^{-1}(\bar{R}(s))=s.

For the other direction R¯(R¯1(r))\bar{R}(\bar{R}^{-1}(r)), the same reasoning holds. The constraint (P)=R¯1(r)\mathcal{M}(P)=\bar{R}^{-1}(r) implies that H(𝑹|𝑿=𝟎)PrH(\bm{R}|\bm{X}=\bm{0})_{P}\leq r since any distribution that achieves a CHSH value of R¯1(r)\bar{R}^{-1}(r) can generate no more than rr bits of randomness. Hence R¯(R¯1(r))r\bar{R}(\bar{R}^{-1}(r))\leq r. This bound is saturated by the behaviour PP^{*} achieving the maximum defining R¯1(s)\bar{R}^{-1}(s), completing the proof. ∎

From the above lemma, we can solve for upper bounds on the points (s,R(s))Graph[R(s)](s,R(s))\in\text{Graph}[R(s)] using the inverse function, i.e., (s,R(s))(s,R¯(s))=(R¯1(r),r)(s,R(s))\leq(s,\bar{R}(s))=(\bar{R}^{-1}(r),r) where R¯(s)=r\bar{R}(s)=r. What remains is to upper bound R¯1(r)\bar{R}^{-1}(r), which will correspond to an upper bound on R(s)R(s) due to the monotonicity argument. Next we obtain a upper bound by applying a symmetrization step to the distribution {p(𝒂|𝟎)}\{p(\bm{a}|\bm{0})\}:

Lemma 15.

Let \mathcal{E} be the local channel that maps all even (odd) parity outcome strings to a uniform mixture of all even (odd) parity outcome strings,

[p(𝒂|𝟎)]={12N1𝒃evenp(𝒃|𝟎),if𝒂hasevenparity,12N1𝒃oddp(𝒃|𝟎),if𝒂hasoddparity.\mathcal{E}[p(\bm{a}|\bm{0})]=\begin{cases}\frac{1}{2^{N-1}}\sum_{\bm{b}\ \mathrm{even}}p(\bm{b}|\bm{0}),\ \mathrm{if}\ \bm{a}\ \mathrm{has\ even\ parity,}\\ \frac{1}{2^{N-1}}\sum_{\bm{b}\ \mathrm{odd}}p(\bm{b}|\bm{0}),\ \mathrm{if}\ \bm{a}\ \mathrm{has\ odd\ parity.}\end{cases} (187)

The entropy after applying \mathcal{E} is non-decreasing, and the MABK value is invariant under \mathcal{E}.

Proof.

The first claim follows from the data processing inequality, H(𝑹|𝑿=𝟎)PH(𝑹|𝑿=𝟎)(P)H(\bm{R}|\bm{X}=\bm{0})_{P}\leq H(\bm{R}|\bm{X}=\bm{0})_{\mathcal{E}(P)}. The second claim comes from the fact that the correlators A𝒙\langle A_{\bm{x}}\rangle are invariant under \mathcal{E}. ∎

After the map \mathcal{E} is applied, the probabilities are symmeterized, i.e., p(𝒂|𝟎)=ϵp(\bm{a}|\bm{0})=\epsilon if 𝒂\bm{a} is even and p(𝒂|𝟎)=(12N1ϵ)/2N1=(22Nϵ)/2Np(\bm{a}|\bm{0})=(1-2^{N-1}\epsilon)/2^{N-1}=(2-2^{N}\epsilon)/2^{N} if 𝒂\bm{a} is odd, for ϵ[0,21N]\epsilon\in[0,2^{1-N}]. We can hence define the following upper bound on R¯(s)R¯¯(s)\bar{R}(s)\leq\bar{\bar{R}}(s) where

R¯¯(s)=max\displaystyle\bar{\bar{R}}(s)=\max\ H(𝑹|𝑿=𝟎)(P)\displaystyle H(\bm{R}|\bm{X}=\bm{0})_{\mathcal{E}(P)}
s.t. ((P))=s,\displaystyle\mathcal{M}(\mathcal{E}(P))=s,
P𝒬\displaystyle P\in\mathcal{Q}
=max\displaystyle=\max\ H(𝑹|𝑿=𝟎)P\displaystyle H(\bm{R}|\bm{X}=\bm{0})_{P}
s.t. (P)=s,\displaystyle\mathcal{M}(P)=s,
p(𝒂|𝟎)={ϵif𝒂iseven,(22Nϵ)/2Nif𝒂isodd,\displaystyle p(\bm{a}|\bm{0})=\begin{cases}\epsilon\ \mathrm{if}\ \bm{a}\ \mathrm{is\ even,}\\ (2-2^{N}\epsilon)/2^{N}\ \mathrm{if}\ \bm{a}\ \mathrm{is\ odd,}\end{cases}
P𝒬.\displaystyle P\in\mathcal{Q}. (188)

We can now apply Lemma˜14 (which still holds after applying \mathcal{E}, since it preserves the convexity of 𝒬\mathcal{Q}), and define the function

R¯¯1(r)=max\displaystyle\bar{\bar{R}}^{-1}(r)=\max\ (P)\displaystyle\mathcal{M}(P)
s.t. H(𝑹|𝑿=𝟎)P=r,\displaystyle H(\bm{R}|\bm{X}=\bm{0})_{P}=r,
p(𝒂|𝟎)={ϵif𝒂iseven,(22Nϵ)/2Nif𝒂isodd,\displaystyle p(\bm{a}|\bm{0})=\begin{cases}\epsilon\ \mathrm{if}\ \bm{a}\ \mathrm{is\ even,}\\ (2-2^{N}\epsilon)/2^{N}\ \mathrm{if}\ \bm{a}\ \mathrm{is\ odd,}\end{cases}
P𝒬.\displaystyle P\in\mathcal{Q}. (189)

We can further relax this by considering the correlator after symmetrization, A𝟎=2N1ϵ2N1(22Nϵ)/2N=2Nϵ1\langle A_{\bm{0}}\rangle=2^{N-1}\epsilon-2^{N-1}(2-2^{N}\epsilon)/2^{N}=2^{N}\epsilon-1. Then we have R¯¯1(r)R¯¯¯1(r)\bar{\bar{R}}^{-1}(r)\leq\bar{\bar{\bar{R}}}^{-1}(r) where

R¯¯¯1(r)=max\displaystyle\bar{\bar{\bar{R}}}^{-1}(r)=\max\ (P)\displaystyle\mathcal{M}(P)
s.t. H(𝑹|𝑿=𝟎)P=r,\displaystyle H(\bm{R}|\bm{X}=\bm{0})_{P}=r,
A𝟎=2Nϵ1,\displaystyle\langle A_{\bm{0}}\rangle=2^{N}\epsilon-1,
P𝒬.\displaystyle P\in\mathcal{Q}. (190)

This follows from the fact that the feasible region of the former optimization is a subset of the latter.

Next, consider the expression for the entropy H(𝑹|𝑿=𝟎)PH(\bm{R}|\bm{X}=\bm{0})_{P} of a symmetrized distribution PP. By direct computation we find

H(𝑹|𝑿=𝟎)(P)\displaystyle H(\bm{R}|\bm{X}=\bm{0})_{\mathcal{E}(P)} =𝒂evenϵlogϵ\displaystyle=-\sum_{\bm{a}\ \mathrm{even}}\epsilon\log\epsilon
𝒂odd22Nϵ2Nlog22Nϵ2N\displaystyle\ \ \ \ \ \ \ \ \ \ -\sum_{\bm{a}\ \mathrm{odd}}\frac{2-2^{N}\epsilon}{2^{N}}\log\frac{2-2^{N}\epsilon}{2^{N}}
=2N1(ϵlogϵ+22Nϵ2Nlog22Nϵ2N)\displaystyle=-2^{N-1}\Big{(}\epsilon\log\epsilon+\frac{2-2^{N}\epsilon}{2^{N}}\log\frac{2-2^{N}\epsilon}{2^{N}}\Big{)}
=2N1(ϵlog2N1ϵϵ(N1)\displaystyle=-2^{N-1}\Big{(}\epsilon\log 2^{N-1}\epsilon-\epsilon(N-1)
+12N1(12N1ϵ)\displaystyle+\frac{1}{2^{N-1}}(1-2^{N-1}\epsilon)
(log(12N1ϵ)(N1)))\displaystyle\ \ \ \ \ \ \ \ \ \ \cdot(\log(1-2^{N-1}\epsilon)-(N-1))\Big{)}
=N1+Hbin(2N1ϵ).\displaystyle=N-1+H_{\mathrm{bin}}(2^{N-1}\epsilon). (191)

Since 2N1ϵ[0,1/2]2^{N-1}\epsilon\in[0,1/2] for ϵ[0,1/2N]\epsilon\in[0,1/2^{N}], there is a one to one correspondence between H(𝑹|𝑿=𝟎)(P)H(\bm{R}|\bm{X}=\bm{0})_{\mathcal{E}(P)} and ϵ\epsilon. Hence if H(𝑹|𝑿=𝟎)(P)=rH(\bm{R}|\bm{X}=\bm{0})_{\mathcal{E}(P)}=r, there exists a unique ϵr[0,1/2N]\epsilon_{r}\in[0,1/2^{N}] that satisfies r=N1+Hbin(2N1ϵr)r=N-1+H_{\mathrm{bin}}(2^{N-1}\epsilon_{r}). We can therefore write

R¯¯¯1(r)=max\displaystyle\bar{\bar{\bar{R}}}^{-1}(r)=\max\ (P)\displaystyle\mathcal{M}(P)
s.t. A𝟎=2Nϵr1,\displaystyle\langle A_{\bm{0}}\rangle=2^{N}\epsilon_{r}-1,
P𝒬.\displaystyle P\in\mathcal{Q}. (192)

Next we introduce the function

R~1(r)=mint,z,S\displaystyle\tilde{R}^{-1}(r)=\min_{t,z,S}\ t\displaystyle t
s.t.\displaystyle\mathrm{s.t.}\ t𝕀MN=S+z(A𝟎\displaystyle t\mathbb{I}-M_{N}=S+z\Big{(}A_{\bm{0}} (2Nϵr1)𝕀)\displaystyle-\big{(}2^{N}\epsilon_{r}-1\big{)}\mathbb{I}\Big{)}
S0.\displaystyle S\succeq 0. (193)

Here SS is an SOS decomposition for the operator expression

S=(t+z(2Nϵr1))𝕀(MN+zA𝟎)0,S=\Big{(}t+z\big{(}2^{N}\epsilon_{r}-1\big{)}\Big{)}\mathbb{I}-\Big{(}M_{N}+zA_{\bm{0}}\Big{)}\succeq 0, (194)

hence for any feasible point (t,z,S)(t,z,S), all quantum correlations satisfying A𝟎=2Nϵr1\langle A_{\bm{0}}\rangle=2^{N}\epsilon_{r}-1 also satisfy tMNt\geq\langle M_{N}\rangle. This implies R¯¯¯1(r)R~1(r)\bar{\bar{\bar{R}}}^{-1}(r)\leq\tilde{R}^{-1}(r). Now, by choosing a basis of monomials for SS, one can evaluate the function R~1(r)\tilde{R}^{-1}(r) numerically using a standard SDP solver. We refer the reader to [59] for details of how to find such formulation, and provide a self-contained description below.

If the monomial basis is chosen to be {Rμ}μ\{R_{\mu}\}_{\mu}, where each RμR_{\mu} is a product of operators from the set {Axk(k)}xk,k\{A_{x_{k}}^{(k)}\}_{x_{k},k}, then one can construct polynomials of the form Mi=μqiμRμM_{i}=\sum_{\mu}q^{\mu}_{i}R_{\mu}, qiμq_{i}^{\mu}\in\mathbb{C}. SS then admits an SOS decomposition if qiμ,Rμq^{\mu}_{i},R_{\mu} are chosen such that

S=iMiMi=μνMμνRμRν,S=\sum_{i}M_{i}^{\dagger}M_{i}=\sum_{\mu\nu}M^{\mu\nu}R_{\mu}^{\dagger}R_{\nu}, (195)

where Mμν=i(qiμ)qiνM^{\mu\nu}=\sum_{i}(q_{i}^{\mu})^{*}q_{i}^{\nu} are matrix elements 𝑴\bm{M}, the Gram matrix of the set of vectors {𝒒μ}μ\{\bm{q}^{\mu}\}_{\mu}, where 𝒒μ=[q1μ,,qiμ,]T\bm{q}^{\mu}=[q_{1}^{\mu},...,q_{i}^{\mu},...]^{\mathrm{T}}. One now arrives at the following optimization problem

mint,z,𝑴\displaystyle\min_{t,z,\bm{M}}\ t\displaystyle t
s.t.\displaystyle\mathrm{s.t.}\ μνMμνRμRν=(t+z(2Nϵr1))𝕀\displaystyle\sum_{\mu\nu}M^{\mu\nu}R_{\mu}^{\dagger}R_{\nu}=\Big{(}t+z\big{(}2^{N}\epsilon_{r}-1\big{)}\Big{)}\mathbb{I}
(MN+zA𝟎),\displaystyle\ \ \ \ \ \ \ \ -\Big{(}M_{N}+zA_{\bm{0}}\Big{)},
𝑴0.\displaystyle\bm{M}\succeq 0. (196)

Now one can write RμRν=iFμνiEiR_{\mu}^{\dagger}R_{\nu}=\sum_{i}F^{i}_{\mu\nu}E_{i}, where 𝑭i\bm{F}^{i} is some complex matrix with elements FμνiF^{i}_{\mu\nu}, for EiCanonical[{RμRν}μν]E_{i}\in\mathrm{Canonical}\big{[}\{R_{\mu}^{\dagger}R_{\nu}\}_{\mu\nu}\big{]}, which denotes the canonical reduction of the set {RμRν}μν\{R_{\mu}^{\dagger}R_{\nu}\}_{\mu\nu} using the commutation and projective relations (see Ref. [59] for details). Similarly, one can express the operator expression in this basis, (t+z(2Nϵr1))𝕀(MN+zA𝟎)=isiEi\Big{(}t+z\big{(}2^{N}\epsilon_{r}-1\big{)}\Big{)}\mathbb{I}-\Big{(}M_{N}+zA_{\bm{0}}\Big{)}=\sum_{i}s^{i}E_{i}, and arrive at an SDP of the form

mint,z,𝑴\displaystyle\min_{t,z,\bm{M}}\ t\displaystyle t
s.t.\displaystyle\mathrm{s.t.}\ μνFμνiMμν=sii,\displaystyle\sum_{\mu\nu}F_{\mu\nu}^{i}M^{\mu\nu}=s^{i}\ \forall i,
𝑴0.\displaystyle\bm{M}\succeq 0. (197)

For our numerical calculations, we choose the monomial basis {k=1N/2Axk(k),k=N/2+1NAxk(k)}xk\Big{\{}\prod_{k=1}^{N/2}A_{x_{k}}^{(k)},\prod_{k=N/2+1}^{N}A_{x_{k}}^{(k)}\Big{\}}_{x_{k}}, consisting of 2N/2+12^{N/2+1} elements.

B.10.2 Numerical results

We present the numerical results which support our conjecture in Fig.˜6, for N=4,6,8,10,12N=4,6,8,10,12. We find the two curves coincide very well, suggesting the lower bounds are indeed tight. The case of N=2N=2 was proven in [23]. The technique also generates the numerical results in Table˜1.

B.11 Other Bell inequalities for N=3N=3

The above technique can be directly applied to find bounds on other Bell inequalities. We applied this to the tripartite expressions in Eqs.˜46 and 47, which, along with the MABK expression, trivial extensions of CHSH and positivity cover all inequivalent classes of facets for the local polytope with uniform marginals (see [60] for the general case). The exact trade-offs we found are detailed in Fig.˜7, and we summarize various bounds of each class in Table˜2. We restate the Bell expressions studied below [41]:

M3\displaystyle M_{3} =A0(B0C1+B1C0)/2+A1(B0C0B1C1)/2,\displaystyle=A_{0}(B_{0}C_{1}+B_{1}C_{0})/2+A_{1}(B_{0}C_{0}-B_{1}C_{1})/2,
S1\displaystyle S_{1} =14x,y,zAxByCzA1B1C1,\displaystyle=\frac{1}{4}\sum_{x,y,z}A_{x}B_{y}C_{z}-A_{1}B_{1}C_{1},
S2\displaystyle S_{2} =A0B0(C0+C1)A1B1(C0C1),\displaystyle=A_{0}B_{0}(C_{0}+C_{1})-A_{1}B_{1}(C_{0}-C_{1}),
S3\displaystyle S_{3} =A0(B0C0+B0C1+B1C0B1C1),\displaystyle=A_{0}(B_{0}C_{0}+B_{0}C_{1}+B_{1}C_{0}-B_{1}C_{1}),
S4\displaystyle S_{4} =A0B0C0,\displaystyle=A_{0}B_{0}C_{0}, (198)

where we have made the substitution Ax0(0)AxA_{x_{0}}^{(0)}\rightarrow A_{x}, Ax1(1)ByA_{x_{1}}^{(1)}\rightarrow B_{y} etc. for readability. For S1S_{1} and S2S_{2}, we applied our numerical technique by substituting them in for the MABK expression. When ϵ=1/8\epsilon=1/8, the SDP returns an upper bound on the maximum violation with maximum randomness, which we label ss^{*} (the facet class in question will be clear from context). For S1S_{1}, we numerically identified that ϵ=1/6\epsilon=1/6 achieves the maximum quantum value of 5/35/3, for which the corresponding entropy is given by Eq.˜191, which evaluates to 2+Hbin(2/3)2+H_{\mathrm{bin}}(2/3); by varying ϵ[1/8,1/6]\epsilon\in[1/8,1/6] we achieve the trade off in Fig.˜7, which corresponds to an upper bound. Similarly, for S2S_{2} we found ϵ=(2+2)/16\epsilon=(2+\sqrt{2})/16 achieves the maximum quantum value 222\sqrt{2}, so by varying ϵ[1/8,(2+2)/16]\epsilon\in[1/8,(2+\sqrt{2})/16] we find the curve in Fig.˜7. The corresponding entropy is 2+Hbin((2+2)/4)2+H_{\mathrm{bin}}((2+\sqrt{2})/4) which is exactly 1 plus the amount certified by maximum CHSH violation [45, 61, 23].

We checked the value of our constructions in Eq.˜20, with N=3N=3, for the expressions S1\langle S_{1}\rangle and S2\langle S_{2}\rangle. We find

S1\displaystyle\langle S_{1}\rangle =2sin(3θ/2)cos3(θ/2),\displaystyle=2\sin(3\theta/2)\cos^{3}(\theta/2),
S2\displaystyle\langle S_{2}\rangle =sinθsin2θ+sin3θ.\displaystyle=\sin\theta-\sin 2\theta+\sin 3\theta. (199)

By varying θ\theta, we find one can achieve a wide range of values from the local bound, however the values of ss^{*} are inaccessible using this parameterization; one will need to consider a larger family of strategies to further investigate if ss^{*} is achievable with maximum randomness.

\onecolumngrid
Refer to caption
Refer to caption
Refer to caption
Refer to caption
Refer to caption
Figure 6: Lower bounds on the maximum DI randomness versus MABK value, normalized by the maximum quantum value 2(N1)/22^{(N-1)/2}, for even NN. These are compared to numerical upper bounds computed via the SOS technique. Each plot starts at the conjectured maximum MABK value achievable with maximum randomness, mNm_{N}^{*}, and ends with the maximum quantum MABK value. The points coincide closely with each curve, supporting our conjecture that the analytical lower bound provided in the text is tight.
Facet ηL\eta^{\mathrm{L}} ss^{*} ηQ\eta^{\mathrm{Q}} r(S3=ηQ)r(\langle S_{3}\rangle=\eta^{\mathrm{Q}})
M3M_{3} 1 2 2 3
S1S_{1} 1 1.64621108 5/31.666666675/3\approx 1.66666667 2+Hbin(2/3)2.918295832+H_{\mathrm{bin}}(2/3)\approx 2.91829583
S2S_{2} 2 2.59807617 222.828427122\sqrt{2}\approx 2.82842712 7/2log2(1+2)/22.600876047/2-\log_{2}(1+\sqrt{2})/\sqrt{2}\approx 2.60087604
S3S_{3} 2 33/22.598076213\sqrt{3}/2\approx 2.59807621 222.828427122\sqrt{2}\approx 2.82842712 5/2log2(1+2)/21.600876045/2-\log_{2}(1+\sqrt{2})/\sqrt{2}\approx 1.60087604
S4S_{4} 1 0 1 0
Table 2: Relevant bounds on violations of all inequivalent facet classes of the local polytope in the N=3N=3 scenario, with two binary measurement per party, given in Eq.˜198. M3M_{3} is the MABK expression, S1,S2S_{1},S_{2} are the other non trivial facets for this scenario, S3S_{3} is the trivial extension of CHSH and S4S_{4} is a positivity facet. ηL\eta^{\mathrm{L}} and ηQ\eta^{\mathrm{Q}} are the local and quantum bounds respectively, and ss^{*} is the computed upper bound on the maximum value achievable with maximum randomness. r(S3=ηQ)r(\langle S_{3}\rangle=\eta^{\mathrm{Q}}) is an upper bound on the asymptotic global DI randomness when the maximum quantum violation is achieved. The cases M3M_{3} and S3S_{3} are known to be tight [31, 23]. Due to its structure, S2S_{2} retains the same characteristics as CHSH [23], whereas S1S_{1} exhibits an unexplored trade-off. Since S4S_{4} cannot be violated, it cannot be used for DI randomness certification.
Refer to caption
Refer to caption
Figure 7: Upper bounds on the trade-off between global DI randomness generation and violation of non-trivial facets in the tripartite scenario with two binary measurements per party. The bounds are numerical and generated by our SOS technique. S1S_{1} and S2S_{2} are two non-trivial extremal Bell expressions given in Eq.˜198. Details of the exact numerical values are given in Table˜2.
\twocolumngrid