Skip to main content
Log in

Continuous-Variable Nonlocality and Contextuality

  • Published:
Communications in Mathematical Physics Aims and scope Submit manuscript

Abstract

Contextuality is a non-classical behaviour that can be exhibited by quantum systems. It is increasingly studied for its relationship to quantum-over-classical advantages in informatic tasks. To date, it has largely been studied in discrete-variable scenarios, where observables take values in discrete and usually finite sets. Practically, on the other hand, continuous-variable scenarios offer some of the most promising candidates for implementing quantum computations and informatic protocols. Here we set out a framework for treating contextuality in continuous-variable scenarios. It is shown that the Fine–Abramsky–Brandenburger theorem extends to this setting, an important consequence of which is that Bell nonlocality can be viewed as a special case of contextuality, as in the discrete case. The contextual fraction, a quantifiable measure of contextuality that bears a precise relationship to Bell inequality violations and quantum advantages, is also defined in this setting. It is shown to be a non-increasing monotone with respect to classical operations that include binning to discretise data. Finally, we consider how the contextual fraction can be formulated as an infinite linear program. Through Lasserre relaxations, we are able to express this infinite linear program as a hierarchy of semi-definite programs that allow to calculate the contextual fraction with increasing accuracy.

This is a preview of subscription content, log in via an institution to check access.

Access this article

Price excludes VAT (USA)
Tax calculation will be finalised during checkout.

Instant access to the full article PDF.

Fig. 1

Similar content being viewed by others

Notes

  1. For a comprehensive treatment, we refer the reader to e.g. [21], or to [68] for a beautiful and more concise introduction aimed particularly at computer scientists.

  2. In fact, this holds more generally for \(\sigma \)-finite measures, i.e. when the space is a countable union of sets of finite measure.

  3. More concretely, it is the \(\sigma \)-algebra generated by the sets \(\mathsf {ev}_E^{-1}([0,r))=\left\{ \mu \in \mathbb {P}(X) \mid \mu (E) < r\right\} \) with \(E \in \mathcal {F}_X\) and \(r \in [0,1]\).

  4. The space [0, 1] is assumed to be equipped with its Borel \(\sigma \)-algebra.

  5. While it is more convenient to specify \(\mathcal {M}\), note that the set of contexts \(\mathcal {U}\) carries exactly the same information. It forms an abstract simplicial complex whose simplices are the contexts and whose facets are the maximal context. This combinatorial topological structure is emphasised in some presentations [4, 15, 16, 29, 49].

  6. For a comprehensive reference on sheaf theory see e.g. [59].

  7. In the language of [40], (non)contextuality here—and throughout this article—refers to global (non)contextuality as opposed to Bell (non)contextuality.

  8. Notions of partial extendability have also been considered in the discrete setting in [61, 79].

  9. The alternative term ontological model [81] has become widely used in quantum foundations in recent years. It indicates that the hidden variable, sometimes referred to as the ontic state, is supposed to provide an underlying description of the physical world at perhaps a more fundamental level than the empirical-level description via the quantum state for example.

  10. Recall from Sect. 2 that a probability kernel \(k_C :{\varvec{\varLambda }} \longrightarrow \mathcal {E}(C)\) is a function \(k_C :\varLambda \times \mathcal {F}_C \longrightarrow [0,1]\) which is a measurable function in the first argument and a probability measure in the second argument.

  11. Note that, due to the assumption of parameter independence (Eq. (6)), we can unambiguously write \(k_x(\lambda ,-)\) for \(k_C|_{\mathopen { \{ }x\mathclose { \} }}(\lambda ,-)\), as this marginal is independent of the context C from which one is restricting.

  12. Note that \(\mathbf {1}_D\) is just a simplified notation for the indicator function on the whole domain; i.e. \(\mathbf {1}_D = \chi _{_{D}} :D \longrightarrow \mathbb {R}\). Similarly, \(\mathbf {0}_D\) is the indicator function of the empty set; i.e. \(\mathbf {0}_D = \chi _{_\emptyset } :D \longrightarrow \mathbb {R}\).

  13. A function \(f :Y \longrightarrow \mathbb {R}\) on a locally compact space Y is said to vanish at infinity if the set \(\left\{ y \in Y \mid \Vert f(x)\Vert \ge \varepsilon \right\} \) is compact for all \(\varepsilon >0\).

  14. This theorem holds more generally for locally compact Hausdorff spaces if one considers only (finite signed) Radon measures, which are measures that play well with the underlying topology. However, second-countability, together with local compactness and Hausdorffness, guarantees that every Borel measure is Radon [37, Theorem 7.8].

  15. Here we will still use the form of the program (P-CF) though throughout this subsection one has to keep in mind that it is defined over finite-signed measures on a locally compact space rather than a compact space.

  16. The notation \(\langle \cdot ,\cdot \rangle _{_2}\) is further discussed and explained to be a canonical duality in Appendix A.

  17. Note that program (D-CF) might not have an optimal solution in which case it only has an optimal solution in the closure of the feasible set. In that case, we can always find a sequence of feasible solutions converging to an optimal solution.

  18. Categorically, this is a coproduct.

References

  1. Abramsky, S.: Relational databases and Bell’s theorem. In: Tannen, V., Wong, L., Libkin, L., Fan, W., Tan, W.C., Fourman, M. (eds.) In Search of Elegance in the Theory and Practice of Computation: Essays Dedicated to Peter Buneman, Lecture Notes in Computer Science, vol. 8000, pp. 13–35. Springer (2013). https://doi.org/10.1007/978-3-642-41660-6_2

  2. Abramsky, S.: Relational hidden variables and non-locality. Studia Logica 101(2), 411–452 (2013). https://doi.org/10.1007/s11225-013-9477-4

    Article  MathSciNet  MATH  Google Scholar 

  3. Abramsky, S.: Contextual semantics: From quantum mechanics to logic, databases, constraints, and complexity. In: Dzhafarov, E., Jordan, S., Zhang, R., Cervantes, V. (eds.) Contextuality from Quantum Physics to Psychology, Advanced Series on Mathematical Psychology, vol. 6, pp. 23–50. World Scientific (2015). https://doi.org/10.1142/9789814730617_0002

  4. Abramsky, S., Barbosa, R.S., Karvonen, M., Mansfield, S.: A comonadic view of simulation and quantum resources (2019). To appear in Proceedings of the 34th Annual ACM/IEEE Symposium on Logic in Computer Science (LICS 2019)

  5. Abramsky, S., Barbosa, R.S., Kishida, K., Lal, R., Mansfield, S.: Contextuality, cohomology and paradox. In: Kreutzer, S. (ed.) 24th EACSL Annual Conference on Computer Science Logic (CSL 2015), Leibniz International Proceedings in Informatics (LIPIcs), vol. 41, pp. 211–228. Schloss Dagstuhl–Leibniz-Zentrum fuer Informatik (2015). https://doi.org/10.4230/LIPIcs.CSL.2015.211

  6. Abramsky, S., Barbosa, R.S., Mansfield, S.: Contextual fraction as a measure of contextuality. Phys. Rev. Lett. 119(5), 050504 (2017). https://doi.org/10.1103/PhysRevLett.119.050504

    Article  ADS  Google Scholar 

  7. Abramsky, S., Barbosa, R.S., de Silva, N., Zapata, O.: The quantum monad on relational structures. In: Larsen, K.G., Bodlaender, H.L., Raskin, J.K. (eds.) 42nd International Symposium on Mathematical Foundations of Computer Science (MFCS 2017), Leibniz International Proceedings in Informatics (LIPIcs), vol. 83, pp. 35:1–35:19. Schloss Dagstuhl–Leibniz-Zentrum fuer Informatik (2017). https://doi.org/10.4230/LIPIcs.MFCS.2017.35

  8. Abramsky, S., Brandenburger, A.: The sheaf-theoretic structure of non-locality and contextuality. New J. Phys. 13(11), 113036 (2011). https://doi.org/10.1088/1367-2630/13/11/113036

    Article  ADS  MATH  Google Scholar 

  9. Abramsky, S., Gottlob, G., Kolaitis, P.G.: Robust constraint satisfaction and local hidden variables in quantum mechanics. In: Rossi, F. (ed.) 23rd International Joint Conference on Artificial Intelligence (IJCAI 2013), pp. 440–446. AAAI Press (2013)

  10. Abramsky, S., Hardy, L.: Logical Bell inequalities. Phys. Rev. A 85(6), 062114062114 (2012). https://doi.org/10.1103/PhysRevA.85.062114

    Article  ADS  Google Scholar 

  11. Abramsky, S., Mansfield, S., Barbosa, R.S.: The cohomology of non-locality and contextuality. In: Jacobs, B., Selinger, P., Spitters, B. (eds.) Proceedings of 8th International Workshop on Quantum Physics and Logic (QPL 2011), Electronic Proceedings in Theoretical Computer Science, vol. 95, pp. 1–14. Open Publishing Association (2012). https://doi.org/10.4204/EPTCS.95.1

  12. Abramsky, S., Sadrzadeh, M.: Semantic unification: A sheaf-theoretic approach to natural language. In: Casadio, C., Coecke, B., Moortgat, M., Scott, P. (eds.) Categories and Types in Logic, Language, and Physics: Essays dedicated to Jim Lambek on the occasion of his 90th birthday, pp. 1–13. Springer (2014). https://doi.org/10.1007/978-3-642-54789-8_1

  13. Acín, A., Fritz, T., Leverrier, A., Sainz, A.B.: A combinatorial approach to nonlocality and contextuality. Commun. Math. Phys. 334(2), 533–628 (2015). https://doi.org/10.1007/s00220-014-2260-1

    Article  ADS  MathSciNet  MATH  Google Scholar 

  14. Asadian, A., Budroni, C., Steinhoff, F.E., Rabl, P., Gühne, O.: Contextuality in phase space. Phys. Rev. Lett. 114(25), 250403 (2015). https://doi.org/10.1103/PhysRevLett.114.250403

    Article  ADS  Google Scholar 

  15. Barbosa, R.S.: On monogamy of non-locality and macroscopic averages: examples and preliminary results. In: Coecke, B., Hasuo, I., Panangaden, P. (eds.) 11th Workshop on Quantum Physics and Logic (QPL 2014), Electronic Proceedings in Theoretical Computer Science, vol. 172, pp. 36–55. Open Publishing Association (2014). https://doi.org/10.4204/EPTCS.172.4

  16. Barbosa, R.S.: Contextuality in quantum mechanics and beyond. Ph.D. thesis, University of Oxford (2015)

  17. Barvinok, A.: A Course in Convexity, Graduate Studies in Mathematics, vol. 54. American Mathematical Society (2002). https://doi.org/10.1090/gsm/054

  18. Bell, J.S.: On the Einstein Podolsky Rosen paradox. Physics Physique Fizika 1(3), 195 (1964). https://doi.org/10.1103/PhysicsPhysiqueFizika.1.195

    Article  MathSciNet  Google Scholar 

  19. Bell, J.S.: On the problem of hidden variables in quantum mechanics. Rev. Mod. Phys. 38(3), 447–452 (1966). https://doi.org/10.1103/RevModPhys.38.447

    Article  ADS  MathSciNet  MATH  Google Scholar 

  20. Bermejo-Vega, J., Delfosse, N., Browne, D.E., Okay, C., Raussendorf, R.: Contextuality as a resource for models of quantum computation with qubits. Phys. Rev. Lett. 119(12), 120505 (2017). https://doi.org/10.1103/PhysRevLett.119.120505

    Article  ADS  MathSciNet  Google Scholar 

  21. Billingsley, P.: Probability and Measure. Wiley Series in Probability and Mathematical Statistics. Wiley (1979)

  22. Bogachev, V.I.: Measure theory. Springer (2007). https://doi.org/10.1007/978-3-540-34514-5

  23. Brandenburger, A., Keisler, H.J.: Use of a canonical hidden-variable space in quantum mechanics. In: Coecke, B., Ong, L., Panangaden, P. (eds.) Computation, Logic, Games, and Quantum Foundations. The Many Facets of Samson Abramsky: Essays dedicated to Samson Abramsky on the occasion of his 60th birthday, pp. 1–6. Springer (2013). https://doi.org/10.1007/978-3-642-38164-5_1

  24. Brandenburger, A., Yanofsky, N.: A classification of hidden-variable properties. J. Phys. A Math. Theor. 41(42), 425302 (2008). https://doi.org/10.1088/1751-8113/41/42/425302

    Article  MathSciNet  MATH  Google Scholar 

  25. Braunstein, S.L., Van Loock, P.: Quantum information with continuous variables. Rev. Mod. Phys. 77(2), 513 (2005). https://doi.org/10.1103/RevModPhys.77.513

    Article  ADS  MathSciNet  MATH  Google Scholar 

  26. Cabello, A., Severini, S., Winter, A.: Graph-theoretic approach to quantum correlations. Phys. Rev. Lett. 112(4), 040401 (2014). https://doi.org/10.1103/PhysRevLett.112.040401

    Article  ADS  Google Scholar 

  27. Carù, G.: Detecting contextuality: Sheaf cohomology and all vs nothing arguments. Master’s thesis, University of Oxford (2015). Available at http://www.cs.ox.ac.uk/files/7608/Dissertation.pdf

  28. Carù, G.: On the cohomology of contextuality. In: Duncan, R., Heunen, C. (eds.) 13th International Conference on Quantum Physics and Logic (QPL 2016), Electronic Proceedings in Theoretical Computer Science, vol. 236, pp. 21–39. Open Publishing Association (2017). https://doi.org/10.4204/EPTCS.236.2

  29. Carù, G.: Towards a complete cohomology invariant for non-locality and contextuality. Preprint arXiv:1807.04203 [quant-ph] (2018)

  30. Cunha, M.T.: On measures and measurements: a fibre bundle approach to contextuality. Preprint arXiv:1903.08819 [math.PR] (2019)

  31. Dickson, W.M.: Quantum Chance and Non-locality: Probability and Non-locality in the Interpretations of Quantum Mechanics. Cambridge University Press (1998). https://doi.org/10.1017/CBO9780511524738

  32. Duarte, C., Amaral, B.: Resource theory of contextuality for arbitrary prepare-and-measure experiments. J. Math. Phys. 59(6), 062202 (2018). https://doi.org/10.1063/1.5018582

    Article  ADS  MathSciNet  MATH  Google Scholar 

  33. Dzhafarov, E.N., Kujala, J.V., Cervantes, V.H.: Contextuality-by-default: A brief overview of ideas, concepts, and terminology. In: Atmanspacher, H., Filk, T., Pothos, E. (eds.) 9th International Conference on Quantum Interaction (QI 2015), Lecture Notes in Computer Science, vol. 9535, pp. 12–23. Springer (2015). https://doi.org/10.1007/978-3-319-28675-4_2

  34. Dzhafarov, E.N., Zhang, R., Kujala, J.: Is there contextuality in behavioural and social systems? Theme issue on Quantum probability and the mathematical modelling of decision making. Philos. Trans. R. Soc. A Math. Phys. Eng. Sci. 374(2058), 20150099 (2016). https://doi.org/10.1098/rsta.2015.0099

    Article  ADS  MATH  Google Scholar 

  35. Einstein, A., Podolsky, B., Rosen, N.: Can quantum-mechanical description of physical reality be considered complete? Phys. Rev. 47(10), 777 (1935). https://doi.org/10.1103/PhysRev.47.777

    Article  ADS  MATH  Google Scholar 

  36. Fine, A.: Hidden variables, joint probability, and the bell inequalities. Phys. Rev. Lett. 48(5), 291 (1982). https://doi.org/10.1103/PhysRevLett.48.291

    Article  ADS  MathSciNet  Google Scholar 

  37. Folland, G.B.: Real Analysis: Modern Techniques and Their Applications. Pure and Applied Mathematics., 2nd edn. Wiley (1984) (1999)

  38. Fritz, T.: Possibilistic physics (2009). https://fqxi.org/community/forum/topic/569. FQXI Essay Contest 2009

  39. Giry, M.: A categorical approach to probability theory. In: Banaschewski, B. (ed.) Categorical aspects of topology and analysis. Lecture Notes in Mathematics, vol. 915, pp. 68–85. Springer (1982). https://doi.org/10.1007/BFb0092872

  40. Griffiths, R.B.: Quantum measurements and contextuality. Philos. Trans. R. Soc. A 377(2157), 20190033 (2019). https://doi.org/10.1098/rsta.2019.0033

    Article  ADS  MATH  Google Scholar 

  41. Haviland, E.: On the momentum problem for distribution functions in more than one dimension. ii. Am. J. Math. 58(1), 164–168 (1936). https://doi.org/10.2307/2371063

    Article  MathSciNet  MATH  Google Scholar 

  42. He, Q.Y., Cavalcanti, E.G., Reid, M.D., Drummond, P.D.: Bell inequalities for continuous-variable measurements. Phys. Rev. A 81(6), 062106 (2010). https://doi.org/10.1103/PhysRevA.81.062106

    Article  ADS  Google Scholar 

  43. Henaut, L., Catani, L., Browne, D.E., Mansfield, S., Pappa, A.: Tsirelson’s bound and Landauer’s principle in a single-system game. Phys. Rev. A 98(6), 060302 (2018). https://doi.org/10.1103/PhysRevA.98.060302

    Article  ADS  Google Scholar 

  44. Hilbert, D.: Über die darstellung definiter formen als summe von formenquadraten. Mathematische Annalen 32(3), 342–350 (1888). https://doi.org/10.1007/BF01443605

    Article  MathSciNet  MATH  Google Scholar 

  45. Hofmann, T., Schölkopf, B., Smola, A.J.: Kernel methods in machine learning. Ann. Stat. 36(3), 1171–1220 (2008). https://doi.org/10.1214/009053607000000677

    Article  MathSciNet  MATH  Google Scholar 

  46. Howard, M., Wallman, J., Veitch, V., Emerson, J.: Contextuality supplies the ‘magic’ for quantum computation. Nature 510(7505), 351–355 (2014). https://doi.org/10.1038/nature13460

    Article  ADS  Google Scholar 

  47. Jarrett, J.P.: On the physical significance of the locality conditions in the Bell arguments. Spec. Issue Found. Quant. Mech. Noûs 18(4), 569–589 (1984). https://doi.org/10.2307/2214878

    Article  MathSciNet  MATH  Google Scholar 

  48. Kakutani, S.: Concrete representation of abstract \((M)\)-spaces (a characterization of the space of continuous functions). Ann. Math. 42(4), 994–1024 (1941). https://doi.org/10.2307/1968778

    Article  MathSciNet  MATH  Google Scholar 

  49. Karvonen, M.: Categories of empirical models. In: Selinger, P., Chiribella, G. (eds.) 15th International Conference on Quantum Physics and Logic (QPL 2018), Electronic Proceedings in Theoretical Computer Science, vol. 287, pp. 239–252. Open Publishing Association (2019). https://doi.org/10.4204/EPTCS.287.14

  50. Ketterer, A., Laversanne-Finot, A., Aolita, L.: Continuous-variable supraquantum nonlocality. Phys. Rev. A 97(1), 012133 (2018). https://doi.org/10.1103/PhysRevA.97.012133

    Article  ADS  Google Scholar 

  51. Kishida, K.: Logic of local inference for contextuality in quantum physics and beyond. In: Chatzigiannakis, I., Mitzenmacher, M., Rabani, Y., Sangiorgi, D. (eds.) 43rd International Colloquium on Automata, Languages, and Programming (ICALP 2016), Leibniz International Proceedings in Informatics (LIPIcs), vol. 55, pp. 113:1–113:14. Schloss Dagstuhl–Leibniz-Zentrum fuer Informatik (2016). https://doi.org/10.4230/LIPIcs.ICALP.2016.113

  52. Kochen, S., Specker, E.P.: The problem of hidden variables in quantum mechanics. J. Math. Mech. 17(1), 59–87 (1967)

    MathSciNet  MATH  Google Scholar 

  53. Lasserre, J.B.: Global optimization with polynomials and the problem of moments. SIAM J. Optim. 11(3), 796–817 (2001). https://doi.org/10.1137/S1052623400366802

    Article  MathSciNet  MATH  Google Scholar 

  54. Lasserre, J.B.: Moments, Positive Polynomials and Their Applications, Series on Optmization and Its Applications, vol. 1. Imperial College Press (2009). https://doi.org/10.1142/p665

  55. Lasserre, J.B.: A new look at nonnegativity on closed sets and polynomial optimization. SIAM J. Optim. 21(3), 864–885 (2011). https://doi.org/10.1137/100806990

    Article  MathSciNet  MATH  Google Scholar 

  56. Laurent, M.: Sums of Squares, Moment Matrices and Optimization Over Polynomials, pp. 157–270. Springer New York, New York, NY (2009). https://doi.org/10.1007/978-0-387-09686-5_7

  57. Laversanne-Finot, A., Ketterer, A., Barros, M.R., Walborn, S.P., Coudreau, T., Keller, A., Milman, P.: General conditions for maximal violation of non-contextuality in discrete and continuous variables. J. Phys. A Math. Theor. 50(15), 155304 (2017). https://doi.org/10.1088/1751-8121/aa6016

    Article  ADS  MathSciNet  MATH  Google Scholar 

  58. Luenberger, D.G.: Optimization by Vector Space Methods. Wiley (1997)

  59. MacLane, S., Moerdijk, I.: Sheaves in Geometry and Logic: A First Introduction to Topos Theory. Universitext. Springer (1992). https://doi.org/10.1007/978-1-4612-0927-0

  60. Mansfield, S.: Consequences and applications of the completeness of hardy’s nonlocality. Phys. Rev. A (2017). https://doi.org/10.1103/PhysRevA.95.022122

    Article  Google Scholar 

  61. Mansfield, S., Barbosa, R.S.: Extendability in the sheaf-theoretic approach: Construction of Bell models from Kochen–Specker models. Preprint arXiv:1402.4827 [quant-ph] (2014)

  62. Mansfield, S., Fritz, T.: Hardy’s non-locality paradox and possibilistic conditions for non-locality. Found. Phys. 42(5), 709–719 (2012). https://doi.org/10.1007/s10701-012-9640-1

    Article  ADS  MathSciNet  MATH  Google Scholar 

  63. Mansfield, S., Kashefi, E.: Quantum advantage from sequential-transformation contextuality. Phys. Rev. Lett. 121(23), 230401 (2018). https://doi.org/10.1103/PhysRevLett.121.230401

    Article  ADS  Google Scholar 

  64. McKeown, G., Paris, M.G., Paternostro, M.: Testing quantum contextuality of continuous-variable states. Phys. Rev. A 83(6), 062105 (2011). https://doi.org/10.1103/PhysRevA.83.062105

    Article  ADS  Google Scholar 

  65. Okay, C., Roberts, S., Bartlett, S.D., Raussendorf, R.: Topological proofs of contextuality in quantum mechanics. Quant. Inf. Comput. 17(13–14), 1135–1166 (2017). https://doi.org/10.26421/QIC17.13-14

    Article  Google Scholar 

  66. Okay, C., Tyhurst, E., Raussendorf, R.: The cohomological and the resource-theoretic perspective on quantum contextuality: common ground through the contextual fraction. Preprint arXiv:1806.04657 [quant-ph] (2018)

  67. Orbanz, P.: Probability theory ii (2011)

  68. Panangaden, P.: Labelled Markov Processes. Imperial College Press (2009). https://doi.org/10.1142/p595

  69. Parrilo, P.A.: Semidefinite programming relaxations for semialgebraic problems. Math. Program. 96(2), 293–320 (2003). https://doi.org/10.1007/s10107-003-0387-5

    Article  MathSciNet  MATH  Google Scholar 

  70. Parthasarathy, K.: Probability Measures on Metric Spaces. Elsevier (1967). https://doi.org/10.1016/C2013-0-08107-8. https://linkinghub.elsevier.com/retrieve/pii/C20130081078

  71. Plastino, Á.R., Cabello, A.: State-independent quantum contextuality for continuous variables. Phys. Rev. A 82(2), 022114 (2010). https://doi.org/10.1103/PhysRevA.82.022114

    Article  ADS  Google Scholar 

  72. Popescu, S., Rohrlich, D.: Quantum nonlocality as an axiom. Found. Phys. 24(3), 379–385 (1994). https://doi.org/10.1007/BF02058098

    Article  ADS  MathSciNet  Google Scholar 

  73. Putinar, M.: Positive polynomials on compact semi-algebraic sets. Indiana Univ. Math. J. 42(3), 969–984 (1993)

    Article  MathSciNet  Google Scholar 

  74. Raussendorf, R.: Contextuality in measurement-based quantum computation. Phys. Rev. A 88(2), 022322 (2013). https://doi.org/10.1103/PhysRevA.88.022322

    Article  ADS  Google Scholar 

  75. Raussendorf, R.: Cohomological framework for contextual quantum computations. Preprint arXiv:1602.04155 [quant-ph] (2016)

  76. Raussendorf, R., Briegel, H.J.: A one-way quantum computer. Phys. Rev. Lett. 86(22), 5188 (2001). https://doi.org/10.1103/PhysRevLett.86.5188

    Article  ADS  Google Scholar 

  77. Roumen, F.: Cohomology of effect algebras. In: Duncan, R., Heunen, C. (eds.) 13th International Conference on Quantum Physics and Logic (QPL 2016), Electronic Proceedings in Theoretical Computer Science, vol. 236, pp. 174–201. Open Publishing Association(2017). https://doi.org/10.4204/EPTCS.236.12

  78. Shimony, A.: Events and processes in the quantum world. In: Penrose, R., Isham, C.J. (eds.) Quantum Concepts in Space and Time, pp. 182–203. Oxford University Press (1986). https://doi.org/10.1017/CBO9781139172196.011

  79. Simmons, A.W.: How (maximally) contextual is quantum mechanics? Preprint arXiv:1712.03766 [quant-ph] (2017)

  80. Slofstra, W.: Tsirelson’s problem and an embedding theorem for groups arising from non-local games. Preprint arXiv:1606.03140 [quant-ph] (2016)

  81. Spekkens, R.W.: Contextuality for preparations, transformations, and unsharp measurements. Phys. Rev. A 71(5), 052108 (2005). https://doi.org/10.1103/PhysRevA.71.052108

    Article  ADS  MathSciNet  Google Scholar 

  82. Su, H.Y., Chen, J.L., Wu, C., Yu, S., Oh, C.: Quantum contextuality in a one-dimensional quantum harmonic oscillator. Phys. Rev. A 85(5), 052126 (2012). https://doi.org/10.1103/PhysRevA.85.052126

    Article  ADS  Google Scholar 

  83. Vestrup, E.: The Theory of Measures and Integration. Wiley Series in Probability and Statistics. Wiley (2003)

  84. Weedbrook, C., Pirandola, S., García-Patrón, R., Cerf, N.J., Ralph, T.C., Shapiro, J.H., Lloyd, S.: Gaussian quantum information. Rev. Mod. Phys. 84(2), 621 (2012). https://doi.org/10.1103/RevModPhys.84.621

    Article  ADS  Google Scholar 

  85. Wester, L.: Classical and quantum structures of computation. Ph.D. thesis, University of Oxford (2018)

  86. Yoshikawa, J., Yokoyama, S., Kaji, T., Sornphiphatphong, C., Shiozawa, Y., Makino, K., Furusawa, A.: Invited article: generation of one-million-mode continuous-variable cluster state by unlimited time-domain multiplexing. APL Photon. 1(6), 060801 (2016). https://doi.org/10.1063/1.4962732

    Article  ADS  Google Scholar 

Download references

Acknowledgements

The authors thank Robert Booth, Ulysse Chabaud, Marcelo Terra Cunha, and Antoine Oustry for helpful comments and discussions. We also thank Robert Booth for pointing out that the proof of the FAB theorem could be extended straightforwardly to scenarios with an uncountable set of measurement labels.

This work was mostly carried out while RSB was based at the Department of Computer Science, University of Oxford and partly at the School of Informatics, University of Edinburgh and at their current affiliation; while TD was based at the School of Informatics, University of Edinburgh; and while SM was at LIP6, Sorbonne Université.

Financial support from the following is gratefully acknowledged: the Engineering and Physical Sciences Research Council, EP/N018745/1, ‘Contextuality as a Resource in Quantum Computation’ (RSB); EPSRC, EP/R044759/1, ‘Combining Viewpoints in Quantum Theory (Ext.)’ (RSB); the Portuguese Foundation for Science and Technology (FCT - Fundação para a Ciência e a Tecnologia), CEECINST/00062/2018 (RSB); and the European Union’s Horizon 2020 Research and Innovation Programme under the Marie Skłodowska-Curie Grant Agreement No. 750523, ‘Resource Sensitive Quantum Computing’ (SM).

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Pierre-Emmanuel Emeriau.

Additional information

Communicated by H-T. Yau.

Publisher's Note

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

Appendices

Appendices

Linear Programs in Standard Form

This appendix may be of particular interest to readers familiar with global optimisation. We express the problems (P-CF) and (D-CF) in the standard form of infinite linear programming [17, IV–(7.1)]. We recall them below:

$$\begin{aligned} \hbox {(P-CF)}\left\{ \begin{aligned}&\quad \text {Find } \mu \in \mathbb {M}_{\pm }({\varvec{O}}_X) \\&\quad \text {maximising } \mu (O_X) \\&\quad \text {subject to:} \\&\quad \begin{aligned}&\qquad \forall C \in \mathcal {M},\; \mu |_C \;\le \; e_C \\&\qquad \mu \;\ge \; 0 \text { .}\end{aligned} \end{aligned} \right.&\end{aligned}$$
$$\begin{aligned} \hbox {(D-CF)}\left\{ \begin{aligned}&\quad \text {Find } ({f_C})_{C \in \mathcal {M}} \in \prod _{C \in \mathcal {M}} C(O_C) \\&\quad \text {minimising } \int _{O_C}f_C\;\,\mathrm {d}\,e_C \\&\quad \text {subject to:} \\&\quad \begin{aligned}&\qquad \sum _{C \in \mathcal {M}} f_C \circ \rho ^X_C \;\ge \; \mathbf {1}_{O_X} \\&\qquad \forall C \in \mathcal {M},\; f_C \;\ge \; \mathbf {0}_{O_C} \text { .}\end{aligned} \end{aligned} \right.&\end{aligned}$$

Problems (P-CF) and (D-CF) are indeed infinite linear programs as both the objective and the constraints are linear with respect to the unknown measure \(\mu \in \mathbb {M}_{\pm }({{\varvec{O}}}_X)\). To write (P-CF) in the standard form [17], we introduce the following spaces:

  • \(\displaystyle E_1 :=\mathbb {M}_{\pm }({\varvec{O}}_X)\).

  • \(\displaystyle F_1 :=C(O_X)\), the dual space of \(E_1\).

  • \(\displaystyle E_2 :=\prod _{C \in \mathcal {M}} \mathbb {M}_{\pm }({\varvec{O}}_C)\).

  • \(\displaystyle F_2 :=\prod _{C \in \mathcal {M}} C(O_C) \), the dual space of \(E_2\).

The dualities \(\langle - , - \rangle _1 : E_1 \times F_1 \longrightarrow \mathbb {R}\) and \(\langle - , - \rangle _2 : E_2 \times F_2 \longrightarrow \mathbb {R}\) are defined as follows:

$$\begin{aligned}&\forall \mu \in E_1, \; f \in F_1,&\langle \mu ,f \rangle _1 :=\int _{O_X}f\;\,\mathrm {d}\,\mu \\&\forall (\nu _C) \in E_2, \; (f_C) \in F_2,&\langle (\nu _C),(f_C) \rangle _2 :=\sum _{C \in \mathcal {M}} \int _{O_C}f_C\;\,\mathrm {d}\,\nu _C\, , \end{aligned}$$

where, for simplicity, we have omitted \(C \in \mathcal {M}\) as a subscript for the families of functions. We fix \(K_1\) to be the convex cone of positive measures in \(E_1 = \mathbb {M}_{\pm }({\varvec{O}}_C)\) and \(K_2\) to be the convex cone of families of positive measures in \(E_2 = \prod _{C \in \mathcal {M}} \mathbb {M}_{\pm }({\varvec{O}}_C)\). Then \(K_1^*\) is the convex cone of positive function in \(F_1 = C(O_X)\) and \(K_2^*\) is the convex cone of families of positive functions in \(F_2 = \prod _{C \in \mathcal {M}} C(O_C)\).

Let \(A :E_1 \longrightarrow E_2\) be the following linear transformation: for \(\mu \in E_1\)

$$\begin{aligned} A(\mu ) := (\mu |_C)_{C \in \mathcal {M}} \in E_2 \text{. } \end{aligned}$$

We also define the linear transformation \(A^* :F_2 \longrightarrow F_1\) as follows: for \((f_C) \in F_2\),

$$\begin{aligned} A^*((f_C)) := \sum _{C \in \mathcal {M}} f_C \circ \rho ^X_C \in F_1 \text{. } \end{aligned}$$

We can verify that \(A^*\) is the dual transformation of A: given \(\mu \in E_1\) and \((f_C) \in F_2\), we have

$$\begin{aligned} \langle A(\mu ), (f_C) \rangle _2&= \langle (\mu |_C), (f_C) \rangle _2 \end{aligned}$$
(47)
$$\begin{aligned}&= \sum _{C \in \mathcal {M}} \int _{O_C}f_C\;\,\mathrm {d}\,\mu |_C \end{aligned}$$
(48)
$$\begin{aligned}&= \int _{O_X}\sum _{C \in \mathcal {M}} f_C \circ \rho ^X_C \;\,\mathrm {d}\,\mu \end{aligned}$$
(49)
$$\begin{aligned}&= \langle \mu , \sum _{C \in \mathcal {M}} f_C \circ \rho ^X_C \rangle _1 \end{aligned}$$
(50)
$$\begin{aligned}&= \langle \mu , A^*((f_C)) \rangle _1 \text { .}\end{aligned}$$
(51)

Now fixing the vector function in the objective to be \(c :=- {\varvec{1}}_{O_X} \in F_1\) and the vector in the constraints to be \(b := (-e_C)_{C \in \mathcal {M}} \in E_2\), the program (P-CF) (resp. (D-CF)) can be expressed as in the standard form given in [17]:

$$\begin{aligned} \hbox {LP}\left\{ \begin{aligned}&\quad \text {Find } e_1 \in E_1 \\&\quad \text {minimising } \langle e_1,c \rangle _1 \\&\quad \text {subject to:} \\&\quad \begin{aligned}&A(e_1) \ge _{K_2} b \\&e_1 \ge _{K_1} 0 \text { .}\end{aligned} \end{aligned} \right.&\end{aligned}$$
$$\begin{aligned} \hbox {(D-LP)}\left\{ \begin{aligned}&\quad \text {Find } f_2 \in F_2 \\&\quad \text {maximising } \langle b,f_2 \rangle _2 \\&\quad \text {subject to:} \\&\quad \begin{aligned}&A^*(f_2) \le _{K_1^*} c \\&f_2 \ge _{K_2} 0 \text { .}\end{aligned} \end{aligned} \right.&\end{aligned}$$

Note that the minus sign in the vectors c and b was added because we chose the primal program in the standard form to be a minimisation problem while the primal program (P-CF) at hand is a maximisation problem.

Proof of Proposition 2: Zero Duality Gap

In this appendix we give a full proof of Proposition 2; i.e. that strong duality holds between problems (P-CF) and (D-CF).

Proof

To show strong duality, we rely on [17, Theorem 7.2]. Because \(\mu _0= {\varvec{0}}_{{\varvec{O}}_X}\)—the measure that assigns 0 to every measurable set of \({\varvec{O}}_X\)— is a feasible solution for (P-CF) and the noncontextual fraction lies between 0 and 1, (P-CF) is consistent with finite value. Thus it suffices to show that the following cone

$$\begin{aligned} \mathcal {K}= \{ \left( A(\mu ), \langle \mu ,c \rangle _1 \right) \, \mid \, \mu \in K_1 \} = \{ \left( \, (\mu |_C)_C, \mu (O_X)\, \right) \, \mid \, \mu \in K_{1} \} \end{aligned}$$
(52)

is weakly closed in \(E_2 \oplus \mathbb {R}\) (i.e. closed in the weak topology of \(K_1\)) where we recall that \(K_1\) is the convex cone of positive measures in \(E_1 = \mathbb {M}_{\pm }({\varvec{O}}_X)\).

We first notice that the linear transformation A is a bounded linear operator and thus continuous. Boundedness comes from the fact that for all \(\mu \in K_1\),

$$\begin{aligned} \left\Vert A(\mu )\right\Vert _{E_2}&= \left\Vert (\mu |_C)_C\right\Vert _{E_2} \end{aligned}$$
(53)
$$\begin{aligned}&= \sum _{C \in \mathcal {M}} \left\Vert \mu |_C\right\Vert _{\mathbb {M}_{\pm }({\varvec{O}}_C)} \end{aligned}$$
(54)
$$\begin{aligned}&\le \sum _{C \in \mathcal {M}} \left\Vert \mu \right\Vert _{E_1} \end{aligned}$$
(55)
$$\begin{aligned}&= \vert \mathcal {M}\vert \left\Vert \mu \right\Vert _{E_1} \text { ,}\end{aligned}$$
(56)

where we take the strong topology—i.e. the norm induced by the total variation distance—on finite-signed measure spaces. It is defined as:

$$\begin{aligned} \left\Vert \mu \right\Vert _{\mathbb {M}_{\pm }({\varvec{U}})} = \vert \mu \vert (U)\text { .}\end{aligned}$$

We also equip the finite product space \(E_2 = \prod _{C \in \mathcal {M}} \mathbb {M}_{\pm }({\varvec{O}}_C)\) with the norm obtained by summingFootnote 18 the individual total variation norms. The inequality in Eq. (55) is due to the fact that \(\mu \in K_1\) so this is a positive measure and thus \(\left\| \mu |_C\right\| _{\mathbb {M}_{\pm }({\varvec{O}}_C)} \le \left\| \mu \right\| _{E_1}\). This, of course, extends to the weak topology.

Secondly, we consider a sequence \((\mu ^k)_{k \in \mathbb {N}}\) in \(K_1\) and we want to show that the accumulation point \(((\varTheta _C)_C,\lambda ) = \lim _{k \rightarrow \infty } \left( A(\mu ^k), \langle \mu ^k,c \rangle _1 \right) \) belongs to \(\mathcal {K}\), where \(\varTheta = (\varTheta _C)_C \in E_2\) and \(\lambda \in \mathbb {R}\). If we consider the product of indicator functions \((\mathbf {1}_{O_C})_C \in F_2\) then \(\langle A(\mu ^k), (\mathbf {1}_{O_C}) \rangle _2 = \sum _{C \in \mathcal {M}} \mu |_C^k(O_C) \longrightarrow _{k} \sum _{C \in \mathcal {M}} \varTheta _C(O_C) < \infty \) as \(\varTheta _C\) is a finite measure for all maximal contexts \(C \in \mathcal {M}\). Then, because \(\mathcal {M}\) is a covering family of X, we have that, for all \(k \in \mathbb {N} \ \mu ^k(O_X) \le \sum _{C \in \mathcal {M}} \mu ^k|_C(O_C) < \infty \). Since \((\mu ^k) \in K_1^{\mathbb {N}^{}}\) is a sequence of positive measures, this implies that \((\mu ^k)\) is bounded. Next, by weak-\(*\) compactness of the unit ball (Alaoglu’s theorem [58]), there exists a subsequence \((\mu ^{k_i})_{i}\) that converges weakly to an element \(\omega \in K_1\). By continuity of A, we obtain that the accumulation point is such that \(((\varTheta _C)_C,\lambda ) = \left( A(\omega ), \langle \omega ,c \rangle _1 \right) \in \mathcal {K}\). \(\quad \square \)

The Lasserre–Parrilo Hierarchy

Below we introduce the Lasserre–Parrilo hierarchy for relaxing infinite-dimensional linear programs known as Generalised Moment Problems [53, 54, 69]. We start by giving insightful results: Sect. C.1 provides results concerning the representation of positive polynomials while Sect. C.2 provides results to understand when a sequence can be represented by a Borel measure.

Notation, terminology

We work in \(\mathbb {R}^d\) for \(d \in \mathbb {N}^*\). We fix \({\varvec{K}}\) to be a generic Borel measurable subspace of \({\varvec{\mathbb {R}}}^d\). Below we fix some multi-index notations. Let \(\mathbb {R}[1]{[}{\varvec{x}}[1]{]}\) denote the ring of real polynomials in the variables \({\varvec{x}} \in \mathbb {R}^d\), and let \(\mathbb {R}[1]{[}{\varvec{x}} [1]{]}_{k}\subset \mathbb {R}[1]{[}{\varvec{x}}[1]{]}\) contain those polynomials of total degree at most k. The latter forms a vector space of dimension \(s(k) :=\left( {\begin{array}{c}d+k\\ k\end{array}}\right) \), with a canonical basis consisting of monomials \({\varvec{x}}^{{\varvec{\alpha }}} :=x_1^{\alpha _1}\cdots x_d^{\alpha _d}\) indexed by the set \(\mathbb {N}^d_k :=\left\{ {\varvec{\alpha }}\in \mathbb {N}^d \mid |{\varvec{\alpha }}|\le k\right\} \) where \(|{\varvec{\alpha }}|:=\sum _{i=1}^d\alpha _i\). For \(k \in \mathbb {N}\), \({\varvec{x}} \in \mathbb {R}^d\), we define \({\varvec{\mathrm {v}}}_k({\varvec{x}}) :=({\varvec{x}}^{{\varvec{\alpha }}})_{\vert {\varvec{\alpha }}\vert \le d} = (1,x_1,\dots ,x_n,x_1^2,x_1x_2,\dots ,x_n^k)^T \) the vector of monomials of total degree less or equal than k.

Any \( p \in \mathbb {R}[1]{[}{\varvec{x}} [1]{]}_{k}\) is associated with a vector of coefficients \({\varvec{p}} :=(p_{{\varvec{\alpha }}}) \in \mathbb {R}^{s(k)}\) via expansion in the canonical basis as \(p({\varvec{x}}) = \sum _{{\varvec{\alpha }}\in \mathbb {N}^d_k} p_{{\varvec{\alpha }}}{\varvec{x}}^{{\varvec{\alpha }}}\).

Moment problem in probability

Given a finite set of indices \(\varGamma \), a set of reals \(\mathopen { \{ }\gamma _j : j \in \varGamma \mathclose { \} }\) and functions \(h_j :K \longrightarrow \mathbb {R}\), \(j \in \varGamma \), that are integrable with respect to every measure \(\mu \in \mathbb {M}_{\pm }({\varvec{K}})\), the corresponding Global Moment Problem (GMP) can be expressed as:

$$\begin{aligned} \hbox {(GMP)}\left\{ \begin{aligned}&\quad \text {Find } \mu \in \mathbb {M}_{\pm }({\varvec{K}}) \\&\quad \text {maximising } \mu ({\varvec{K}}) \\&\quad \text {subject to:} \\&\qquad \qquad \begin{aligned}&\forall j \in \varGamma ,\quad \int _{K}h_j\;\,\mathrm {d}\,\mu \le \gamma _j \text { .}\end{aligned} \end{aligned} \right.&\end{aligned}$$

It dual program can be expressed as:

$$\begin{aligned} \hbox {(D-GMP)}\left\{ \begin{aligned}&\quad \text {Find } {\varvec{\lambda }}\in \mathbb {R}^\varGamma \\&\quad \text {minimising } \sum _{j \in \varGamma } \gamma _j \lambda _j \\&\quad \text {subject to:} \\&\quad \begin{aligned}&\qquad \forall {\varvec{x}} \in K,\quad \sum _{j \in \varGamma } \lambda _j h_j({\varvec{x}}) - {\varvec{x}} \ge 0 \\&\qquad \forall j \in \varGamma , \quad \lambda _j \ge 0 \text { .}\end{aligned} \end{aligned} \right.&\end{aligned}$$

1.1 Positive polynomials and sum-of-squares

Here we present the link between positive polynomials and sum-of-squares representation so that we can derive a converging hierarchy of restriction problems for program (D-GMP).

Definition 14

A polynomial \(p \in \mathbb {R}[1]{[}{\varvec{x}}[1]{]}\) is a sum-of-squares (SOS) polynomial if there exists a finite family of polynomials \(({q_i})_{i \in I}\) such that \(p = \sum _{i\in I} q_i^2\).

SOS polynomials are widely used in convex optimisation. We will denote by \(\varSigma ^2 \mathbb {R}[1]{[}{\varvec{x}} [1]{]}\subset \mathbb {R}[1]{[}{\varvec{x}}[1]{]}\) the set of (multivariate) SOS polynomials, and \(\varSigma ^2 \mathbb {R}[1]{[}{\varvec{x}} [1]{]}_{k}\subset \varSigma ^2 \mathbb {R}[1]{[}{\varvec{x}} [1]{]}\) the set of SOS polynomials of degree at most 2k. The following proposition hints towards the reason why it is desirable to be able to look for a sum-of-squares decomposition: it can be cast as a semidefinite optimisation problem.

Proposition 3

(Prop. 2.1, [54]). A polynomial \(p \in \mathbb {R}[1]{[}{\varvec{x}}[1]{]}_{2k}\) has a sum-of-squares decomposition if and only if there exists a real symmetric positive semidefinite matrix \(Q \in \textsf {Sym}_{s(k)}\) such that \(\forall {\varvec{x}} \in \mathbb {R}^d\), \(p({\varvec{x}}) = {\varvec{\mathrm {v}}}_k({\varvec{x}})^T Q {\varvec{\mathrm {v}}}_k({\varvec{x}})\).

Then we will be looking at conditions under which a nonnegative polynomial can be expressed as a sum-of-squares polynomial. This is in essence the question raised by Hilbert in his 17\(^\text {th}\) conjecture [44].

Definition 15

For a family \(q = (q_j)_{j\in \mathopen { \{ }1, \ldots , m\mathclose { \} }}\) of polynomials, the set:

$$\begin{aligned} Q(q) :=\left\{ \sum _{j=0}^m \sigma _j q_j \mid (\sigma _j)_{j\in \mathopen { \{ }0,\ldots , m\mathclose { \} }} \subset \varSigma ^2 \mathbb {R}[1]{[}{\varvec{x}} [1]{]}\right\} \end{aligned}$$
(57)

is a convex cone in \(\mathbb {R}[1]{[}{\varvec{x}}[1]{]}\) called the quadratic module generated by the family q with, for convenience, \(q_0 = 1\) added. For \(k \in \mathbb {N}\), we define \(Q_k(q)\) to be the quadratic module Q(q) where we further impose that \((\sigma _j)_{j\in \mathopen { \{ }0,\ldots , m\mathclose { \} }} \subset \varSigma ^2 \mathbb {R}[1]{[}{\varvec{x}} [1]{]}_{k}\) i.e. we limit the degree of SOS polynomials.

In the family of polynomials \((g_j)_{j\in \mathopen { \{ }1, \ldots , m\mathclose { \} }}\) we add \(g_0 = 1\) for convenience.

Assumptions 1

Let \(K \subset \mathbb {R}^d\). We make the following three assumptions on K.

  1. (i)

    Suppose K is a basic semi-algebraic set i.e. there exists a family of polynomials \(g = (g_j)_{j\in \mathopen { \{ }1, \ldots , m\mathclose { \} }} \in \mathbb {R}[1]{[}{\varvec{x}}[1]{]}^m\) of degrees deg(\(g_j\)) respectively such that:

    $$\begin{aligned} K :=\left\{ {\varvec{x}} \in \mathbb {R}^d \mid \forall j = 1,\dots ,m, \; g_j({\varvec{x}}) \ge 0\right\} . \end{aligned}$$
    (58)
  2. (ii)

    Further suppose that K is compact.

  3. (iii)

    Finally suppose that there exists \(u \in Q(q)\) such that the level set \(\left\{ {\varvec{x}} \in \mathbb {R}^d \mid u({\varvec{x}}) \ge 0\right\} \) is compact.

The following theorem is the key result that we will exploit for deriving the hierarchy of SDP restrictions for the dual program (GMP).

Theorem 6

(Putinar’s Positivellensatz [73]). Let \(K \subset \mathbb {R}^d\) satisfy Assumptions 1. If \(p \in \mathbb {R}[1]{[}{\varvec{x}}[1]{]}\) is strictly positive on K then \(p \in Q(g)\), that is

$$\begin{aligned} p = \sum _{j=0}^m \sigma _j g_j \end{aligned}$$
(59)

for some sum-of-squares polynomials \(\sigma _j \in \varSigma ^2 \mathbb {R}[1]{[}{\varvec{x}} [1]{]}\) for \(j=0,1,\dots ,m\).

A proof can also be found in [56].

Using the results above and Assumption 1, one can derive a hierarchy of SDPs [54] which provide a converging sequence of optimal values towards the value of program (D-GMP):

figure e

1.2 Moment sequences and moment matrices

In this subsection, we want to understand why the program (P-CF) can be relaxed so that a converging hierarchy of SDPs can be derived. The program (P-CF) is essentially a maximisation problem on finite-signed Borel measures with additional constraints such as the fact that these are proper measures (i.e. they are nonnegative). We will represent a measure by its moment sequence and find conditions for which this moment sequence has a (unique) representing Borel measure.

Definition 16

Given a sequence \({\varvec{y}} = (y_{{\varvec{\alpha }}})_{{\varvec{\alpha }}\in \mathbb {N}^d}\in \mathbb {R}^{\mathbb {N}^d}\), we define the linear functional \(L_{{\varvec{y}}} :\mathbb {R}[1]{[}{\varvec{x}}[1]{]} \longrightarrow \mathbb {R}\) by

$$\begin{aligned} L_{{\varvec{y}}}(p) :=\sum _{{\varvec{\alpha }}\in \mathbb {N}^d}p_{{\varvec{\alpha }}} y_{{\varvec{\alpha }}}. \end{aligned}$$
(60)

Definition 17

Given a measure \(\mu \in \mathbb {M}({\varvec{K}})\), its moment sequence \({\varvec{y}} = (y_{{\varvec{\alpha }}})_{{\varvec{\alpha }}\in \mathbb {N}^d} \in \mathbb {R}^{\mathbb {N}^d}\) is given by

$$\begin{aligned} y_{{\varvec{\alpha }}} :=\int _{{\varvec{K}}}{\varvec{x}}^{{\varvec{\alpha }}}\;\,\mathrm {d}\,\mu ({\varvec{x}}) \text { .} \end{aligned}$$
(61)

We say that \({\varvec{y}}\) has a unique representing measure \(\mu \) when there exists a unique \(\mu \) such that Eq. (61) holds. If \(\mu \) is unique then we say it is determinate (i.e. determined by its moments).

The linear functional \(L_{{\varvec{y}}}\) then gives integration of polynomials with respect to \(\mu \) i.e. for any \(p \in \mathbb {R}[1]{[}{\varvec{x}}[1]{]}\):

$$\begin{aligned} \begin{aligned} L_{{\varvec{y}}}(p)&= \sum _{{\varvec{\alpha }}\in \mathbb {N}^d}p_{{\varvec{\alpha }}} y_{{\varvec{\alpha }}} = \sum _{{\varvec{\alpha }}\in \mathbb {N}^d}p_{{\varvec{\alpha }}} \int _{{\varvec{K}}}{\varvec{x}}^{{\varvec{\alpha }}}\;\,\mathrm {d}\,\mu ({\varvec{x}}) = \int _{{\varvec{K}}}\sum _{{\varvec{\alpha }}\in \mathbb {N}^d}p_{{\varvec{\alpha }}} {\varvec{x}}^{{\varvec{\alpha }}}\;\,\mathrm {d}\,\mu ({\varvec{x}}) \\&= \int _{{\varvec{K}}}p({\varvec{x}})\;\,\mathrm {d}\,\mu ({\varvec{x}}) \\&= \int _{{\varvec{K}}}p\;\,\mathrm {d}\,\mu , \end{aligned} \end{aligned}$$
(62)

where we reversed summation and integration because the sum is finite since p is a polynomial.

The following theorem is often used in optimisation theory over measures as it provides a necessary and sufficient condition for a sequence to have a representing measure.

Theorem 7

(Riesz-Haviland [41]). Let \({\varvec{y}} = (y_{{\varvec{\alpha }}})_{{\varvec{\alpha }}\in \mathbb {N}^d} \in \mathbb {R}^{\mathbb {N}^d}\) and suppose that \(K \subseteq \mathbb {R}^d\) is closed. Then \({\varvec{y}}\) has a representation (nonnegative) measure i.e. there exists \(\mu \) a measure on K such that:

$$\begin{aligned} \forall {\varvec{\alpha }}\in \mathbb {N}^d, \; \int _{K}{\varvec{x}}^{{\varvec{\alpha }}}\;\,\mathrm {d}\,\mu = y_{{\varvec{\alpha }}} \end{aligned}$$

if and only if \(L_{{\varvec{y}}}(p) \ge 0\) for all polynomials \(p \in \mathbb {R}[1]{[}{\varvec{x}}[1]{]}\) nonnegative on K.

We recall that for \(k \in \mathbb {N}\), \(s(k) = \left( {\begin{array}{c}d+k\\ k\end{array}}\right) \).

Definition 18

For each \(k \in \mathbb {N}\), the moment matrix of order k, \(M_k({\varvec{y}}) \in \textsf {Sym}_{s(k)}\), of a truncated sequence \((y_{{\varvec{\alpha }}})_{{\varvec{\alpha }}\in \mathbb {N}^d_{2k}}\) is the \(s(k) \times s(k)\) symmetric matrix with rows and columns indexed by \(\mathbb {N}^d_k\) (i.e. by the canonical basis for \(\mathbb {R}[1]{[}{\varvec{x}} [1]{]}_{k}\)) defined as follows: for any \({\varvec{\alpha }}, {\varvec{\beta }}\in \mathbb {N}^d_k\),

$$\begin{aligned} \left( M_k({\varvec{y}})\right) _{{\varvec{\alpha }},{\varvec{\beta }}} :=L_{{\varvec{y}}}({\varvec{x}}^{{\varvec{\alpha }}+ {\varvec{\beta }}}) = y_{{\varvec{\alpha }}+ {\varvec{\beta }}} \text { .}\end{aligned}$$
(63)

Definition 19

Given a polynomial \(p \in \mathbb {R}[1]{[}{\varvec{x}}[1]{]}\), the localising matrix \(M_k(p {\varvec{y}}) \in \textsf {Mat}_{s(k)}(\mathbb {R})\) of a moment sequence \((y_{{\varvec{\alpha }}})_{{\varvec{\alpha }}\in \mathbb {N}^d} \in \mathbb {R}^{\mathbb {N}^d}\) is defined by: for all \({\varvec{\alpha }}, {\varvec{\beta }}\in \mathbb {N}^d_{k}\),

$$\begin{aligned} \left( M_k(p {\varvec{y}})\right) _{{\varvec{\alpha }}, {\varvec{\beta }}} :=L_{{\varvec{y}}}(p(x) x^{{\varvec{\alpha }}+ {\varvec{\beta }}}) = \sum _{\gamma \in \mathbb {N}^d} p_{\gamma } y_{{\varvec{\alpha }}+ {\varvec{\beta }}+ \gamma } \text { .}\end{aligned}$$
(64)

The localising matrix reduces to the moment matrix for \(p=1\). For well-defined moment sequences, i.e. sequences that have a representing finite Borel measure, moment matrices and localising matrices are positive semidefinite, which provides insight on the reason why problem (P-CF) can be relaxed to a problem with positive semidefiniteness constraints.

Proposition 4

Let \({\varvec{y}} = (y_{{\varvec{\alpha }}})_{{\varvec{\alpha }}\in \mathbb {N}^d} \in \mathbb {R}^{\mathbb {N}^d}\) be a sequence of moments for some finite Borel measure \(\mu \) on \({\varvec{K}}\). Then for all \(k\in \mathbb {N}\), \(M_k({\varvec{y}}) \succeq 0 \). If \(\mu \) has support contained in the set \(\left\{ {\varvec{x}} \in K \mid g({\varvec{x}}) \ge 0\right\} \) for some polynomial \(g \in \mathbb {R}[1]{[}{\varvec{x}}[1]{]}\) then, for all \(k \in \mathbb {N}\), \(M_k(g {\varvec{y}}) \succeq 0\).

Proof

Let \({\varvec{y}} = (y_{{\varvec{\alpha }}})\) be the moment sequence of a given Borel measure \(\mu \) on \({\varvec{K}}\). Fix \(k \in \mathbb {N}\). For any vector \({\varvec{v}} \in \mathbb {R}^{s(k)}\) (noting that \({\varvec{v}}\) is canonically associated with a polynomial \(v \in \mathbb {R}[1]{[}{\varvec{x}} [1]{]}_{k}\) in the basis \(({\varvec{x}}^{{\varvec{\alpha }}})\)):

$$\begin{aligned} {\varvec{v}}^T M_k({\varvec{y}}) {\varvec{v}}&= \sum _{{\varvec{\alpha }},{\varvec{\beta }}\in \mathbb {N}^d_k} v_{{\varvec{\alpha }}} y_{{\varvec{\alpha }}+ {\varvec{\beta }}} v_{{\varvec{\beta }}} \end{aligned}$$
(65)
$$\begin{aligned}&= \sum _{{\varvec{\alpha }}, {\varvec{\beta }}\in \mathbb {N}^d_k} v_{{\varvec{\alpha }}} v_{{\varvec{\beta }}} \int _{K}{\varvec{x}}^{{\varvec{\alpha }}+ {\varvec{\beta }}}\;\,\mathrm {d}\,\mu \end{aligned}$$
(66)
$$\begin{aligned}&= \int _{K}\left( \sum _{{\varvec{\alpha }}\in \mathbb {N}^d_k} v_{{\varvec{\alpha }}} {\varvec{x}}^{{\varvec{\alpha }}} \right) ^2\;\,\mathrm {d}\,\mu \end{aligned}$$
(67)
$$\begin{aligned}&= \int _{K}v^2({\varvec{x}})\;\,\mathrm {d}\,\mu \ge 0 \text { .} \end{aligned}$$
(68)

Thus \(M_{k}({\varvec{y}}) \succeq 0 \).

Similarly we can prove that the localising matrix \(M_k(g {\varvec{y}})\) is positive semidefinite when g is a nonnegative polynomial on the support of \(\mu \). Indeed for all \({\varvec{v}} \in \mathbb {R}^{s(k)}\):

$$\begin{aligned} {\varvec{v}}^T M_k(g {\varvec{y}}) {\varvec{v}} = \int _{K}v^2({\varvec{x}}) g({\varvec{x}})\;\,\mathrm {d}\,\mu \ge 0 \text { ,}\end{aligned}$$
(69)

which concludes the proof. \(\quad \square \)

The following theorem, which is the dual facet of Theorem 6, is the key result for deriving the hierarchy of SDP relaxations for the primal problem (P-CF). It provides a necessary and sufficient condition for a sequence to have a representing measure.

Theorem 8

(Theorem 3.8 [54]). Let \({\varvec{y}} = (y_{{\varvec{\alpha }}})_{{\varvec{\alpha }}\in \mathbb {N}^d} \in \mathbb {R}^{\mathbb {N}^d}\) be a given infinite sequence in \(\mathbb {R}\). Let \(K \subset \mathbb {R}^d\) satisfy Assumptions 1. Then \({\varvec{y}}\) has a finite Borel representing measure with support contained in K if and only if:

$$\begin{aligned}&M_k({\varvec{y}}) \succeq 0, \; \forall k \in \mathbb {N}, \end{aligned}$$
(70)
$$\begin{aligned}&M_k(g_j {\varvec{y}}) \succeq 0, \; \forall j=1,\dots ,m , \; \forall k \in \mathbb {N}. \end{aligned}$$
(71)

Using the results above and Assumption 1, one can derive a hierarchy of SDPs [54] which provide a converging sequence of optimal values towards the value of program (GMP):

figure f

We refer readers to [54] for the proof of convergence of the hierarchies given by programs (\(\hbox {GMP}^k\)) and (D-\(\hbox {GMP}^k\)).

Duality between programs (SDP-\(\hbox {CF}^k\)) and (DSDP-\(\hbox {CF}^k\))

As mentioned above, we chose to derive programs (SDP-\(\hbox {CF}^k\)) and (DSDP-\(\hbox {CF}^{k}\)) using dual arguments. These programs should therefore be dual to one another, which will immediately provide weak duality. We prove this for completeness.

Proposition 5

The program (DSDP-\(\hbox {CF}^{k}\)) is the dual formulation of the program (SDP-\(\hbox {CF}^k\)).

Proof

We start by rewriting \(M_k({\varvec{y}})\) as \(\sum _{{\varvec{\alpha }}\in \mathbb {N}^d_k} {\varvec{y}}_{{\varvec{\alpha }}} A_{{\varvec{\alpha }}}\) and \(M_{k-1}(g_j {\varvec{y}})\) as \(\sum _{{\varvec{\alpha }}\in \mathbb {N}^d_k} y_{{\varvec{\alpha }}} B^j_{{\varvec{\alpha }}}\) for \(1 \le j \le m\) and for appropriate real symmetric matrices \(A_{{\varvec{\alpha }}}\) and \((B^j_{{\varvec{\alpha }}})_j\). For instance, in the basis \(({\varvec{x}}^{{\varvec{\alpha }}})\):

$$\begin{aligned} ( A_{{\varvec{\alpha }}} )_{{\varvec{s}}, {\varvec{t}}} = \bigg ( {\left\{ \begin{array}{ll} 1 \text { if } {\varvec{s}}+{\varvec{t}} = {\varvec{\alpha }}\\ 0 \text { otherwise} \end{array}\right. } \bigg )_{{\varvec{s}}, {\varvec{t}}} \text { .}\end{aligned}$$

From \(A_{{\varvec{\alpha }}}\), we also extract \(A_{{\varvec{\alpha }}}^C\) for \(C \in \mathcal {M}\) in order to rewrite \(M_k({\varvec{y}}|_C)\) as \(\sum _{\alpha \in \mathbb {N}^d_k} {\varvec{y}}_{{\varvec{\alpha }}} A_{{\varvec{\alpha }}}^C\). This amounts to identifying which matrices \((A_{{\varvec{\alpha }}})\) contribute to a given context \(C \in \mathcal {M}\). Then (SDP-\(\hbox {CF}^k\)) can be rewritten as:

figure g

Via the Lagrangian method, this is equivalent to:

$$\begin{aligned} \sup _{{\varvec{y}} \in \mathbb {R}^{s(k)}} \; \inf _{ \begin{array}{c} (X^C), (Y_j), Z \\ \text { SDP matrices} \end{array}} \; \mathcal {L}({\varvec{y}},(X_C),Y,(Z_j)), \end{aligned}$$
(72)

with

$$\begin{aligned} \begin{aligned} \mathcal {L}({\varvec{y}},(X^C),(Y_j),Z)&= y_{{\varvec{0}}} \\&\quad + \sum _{C \in \mathcal {M}} \text {Tr}(M_k({\varvec{y}}^{e,C}) X^C) - \sum _{C \in \mathcal {M}} \sum _{{\varvec{\alpha }}\in \mathbb {N}^d_k} y_{{\varvec{\alpha }}} \text {Tr}(A_{{\varvec{\alpha }}}^C X^C) \\&\quad + \sum _{j=1}^m \sum _{{\varvec{\alpha }}\in \mathbb {N}^d_k} y_{{\varvec{\alpha }}} \text {Tr}(B_{{\varvec{\alpha }}}^j Y_j) \\&\quad + \sum _{{\varvec{\alpha }}\in \mathbb {N}^d_k} y_{{\varvec{\alpha }}} \text {Tr}(A_{{\varvec{\alpha }}} Z) \text { .}\end{aligned} \end{aligned}$$
(73)

The dual program corresponds to

$$\begin{aligned} \inf _{ \begin{array}{c} (X^C), (Y_j), Z \\ \text { SDP matrices} \end{array}} \; \sup _{{\varvec{y}} \in \mathbb {R}^{s(k)}} \; \mathcal {L}({\varvec{y}},(X^C),(Y_j),Z) \text { .}\end{aligned}$$
(74)

We rewrite the Lagrangian as:

$$\begin{aligned} \mathcal {L}({\varvec{y}},(X_C),Y,(Z_j))= & {} \sum _{C \in \mathcal {M}} \text {Tr}(M_k({\varvec{y}}^{e,C}) X^C) \nonumber \\&+ \sum _{{\varvec{\alpha }}\in \mathbb {N}^d_k} y_{{\varvec{\alpha }}} \left( \delta _{{\varvec{\alpha }}, {\varvec{0}}} - \sum _{C \in \mathcal {M}} \text {Tr}(A_{{\varvec{\alpha }}}^C X^C) + \sum _{j=1}^m \text {Tr}(B_{{\varvec{\alpha }}}^j Y_j) + \text {Tr}(A_{{\varvec{\alpha }}} Z) \right) \text { .}\quad \quad \end{aligned}$$
(75)

Then the dual program of (SDP-\(\hbox {CF}^k\)) reads:

$$\begin{aligned} \left\{ \begin{aligned}&\quad \text {Find } ({X^C})_{C \in \mathcal {M}}, ({Y_j})_{j=1,\dots ,m} \text { and } Z \text { SDP matrices } \\&\quad \text {maximising } \sum _{C \in \mathcal {M}} \text {Tr}(M_k({\varvec{y}}^{e,C}) X^C) \\&\quad \text {subject to:} \\&\quad \begin{aligned}&\forall {\varvec{\alpha }}\in \mathbb {N}^d_k, \; \sum _{C \in \mathcal {M}} \text {Tr}(A_{{\varvec{\alpha }}}^C X^C) - \sum _{j=1}^m \text {Tr}(B_{{\varvec{\alpha }}}^j Y_j) - \text {Tr}(A_{{\varvec{\alpha }}} Z) = \delta _{{\varvec{\alpha }}, {\varvec{0}}} \text { .}\end{aligned} \end{aligned} \right.&\end{aligned}$$

We finally show that the above program exactly corresponds to (DSDP-\(\hbox {CF}^{k}\)). We start with the objective. For all \(C \in \mathcal {M}\), with \(X^C\) a positive semidefinite matrix:

$$\begin{aligned} \text {Tr}(M_k({\varvec{y}}^{e,C}) X^C)&= \sum _{{\varvec{\alpha }}} \sum _{{\varvec{\beta }}} (M_k({\varvec{y}}^{e,C}))_{{\varvec{\alpha }}{\varvec{\beta }}} (X^C)_{{\varvec{\beta }}{\varvec{\alpha }}} \end{aligned}$$
(76)
$$\begin{aligned}&= \sum _{{\varvec{\alpha }}} \sum _{{\varvec{\beta }}} {\varvec{y}}^{e,C}_{{\varvec{\alpha }}+ {\varvec{\beta }}} (X^C)_{{\varvec{\alpha }}{\varvec{\beta }}} \end{aligned}$$
(77)
$$\begin{aligned}&= \sum _{{\varvec{\alpha }}} \sum _{{\varvec{\beta }}} \int _{\mathcal {O}_C}{\varvec{x}}^{{\varvec{\alpha }}+ {\varvec{\beta }}}\;\,\mathrm {d}\,e_C (X^C)_{{\varvec{\alpha }}{\varvec{\beta }}} \end{aligned}$$
(78)
$$\begin{aligned}&= \int _{\mathcal {O}_C} {\varvec{\mathrm {v}}}_k({\varvec{x}})^T X^C {\varvec{\mathrm {v}}}_k({\varvec{x}})\;\,\mathrm {d}\,e_C \end{aligned}$$
(79)
$$\begin{aligned}&= \int _{\mathcal {O}_C} p_C \;\,\mathrm {d}\,e_C, \end{aligned}$$
(80)

with for all \(C \in \mathcal {M}\), \(p_C \in \varSigma ^2 \mathbb {R}[1]{[}{\varvec{x}} [1]{]}_{k}\) a sum-of-squares polynomial via Proposition 3 and where we used \({\varvec{\mathrm {v}}}_k({\varvec{x}})\) the vectors of monomials of maximal total degree k.

Now, to retrieve the constraint, we multiply each side by \({\varvec{x}}^{{\varvec{\alpha }}}\) and we sum for all \({\varvec{\alpha }}\):

$$\begin{aligned} \sum _{C \in \mathcal {M}} \text {Tr}(\sum _{{\varvec{\alpha }}} {\varvec{x}}^{{\varvec{\alpha }}} A^C_{{\varvec{\alpha }}} X^C) - {\varvec{1}} = \sum _{j=1}^m \text {Tr}(\sum _{{\varvec{\alpha }}} {\varvec{x}}^{{\varvec{\alpha }}} B_{{\varvec{\alpha }}}^j Y_j) + \text {Tr}(\sum _{{\varvec{\alpha }}} {\varvec{x}}^{{\varvec{\alpha }}} A_{{\varvec{\alpha }}} Z) \end{aligned}$$
(81)

Recalling the definition of moment and localising matrices:

$$\begin{aligned}&\sum _{{\varvec{\alpha }}} A^C_{{\varvec{\alpha }}} {\varvec{x}}^{{\varvec{\alpha }}} = {\varvec{\mathrm {v}}}_k({\varvec{x}}) {\varvec{\mathrm {v}}}_k({\varvec{x}})^T \end{aligned}$$
(82)
$$\begin{aligned}&\sum _{{\varvec{\alpha }}} B_{{\varvec{\alpha }}}^j \mathrm {v}_k({\varvec{x}}) \mathrm {v}_k({\varvec{x}})^T = g_j({\varvec{x}}) {\varvec{\mathrm {v}}}_{k-1}({\varvec{x}}) {\varvec{\mathrm {v}}}_{k-1}({\varvec{x}})^T, \; \forall j=1,\dots ,m \end{aligned}$$
(83)

Thus, by Proposition 3, for appropriate sum-of-squares polynomials \((\sigma _j)_{j=0,1,\dots ,m} \subset \mathbb {R}[1]{[}{\varvec{x}}[1]{]}_{k-1}\):

$$\begin{aligned}&\text {Tr}(\sum _{{\varvec{\alpha }}} {\varvec{x}}^{{\varvec{\alpha }}} A^C_{{\varvec{\alpha }}} X^C) = p_C \circ \rho _C^X ({\varvec{x}}) \end{aligned}$$
(84)
$$\begin{aligned}&\text {Tr}(\sum _{{\varvec{\alpha }}} {\varvec{x}}^{{\varvec{\alpha }}} B_{{\varvec{\alpha }}}^j Y_j) = g_j({\varvec{x}}) \sigma _j({\varvec{x}}) \end{aligned}$$
(85)
$$\begin{aligned}&\text {Tr}(\sum _{{\varvec{\alpha }}} {\varvec{x}}^{{\varvec{\alpha }}} A_{{\varvec{\alpha }}} Z) = \sigma _0({\varvec{x}}) \end{aligned}$$
(86)

This is exactly the constraint in (DSDP-\(\hbox {CF}^{k}\)) with, for convenience, \(g_0 = 1\). \(\quad \square \)

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Barbosa, R.S., Douce, T., Emeriau, PE. et al. Continuous-Variable Nonlocality and Contextuality. Commun. Math. Phys. 391, 1047–1089 (2022). https://doi.org/10.1007/s00220-021-04285-7

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s00220-021-04285-7

Navigation