Skip to main content
Log in

Ensemble Steering, Weak Self-Duality, and the Structure of Probabilistic Theories

  • Published:
Foundations of Physics Aims and scope Submit manuscript

Abstract

In any probabilistic theory, we say that a bipartite state ω on a composite system AB steers its marginal state ω B if, for any decomposition of ω B as a mixture ω B=∑ i p i β i of states β i on B, there exists an observable {a i } on A such that the conditional states \(\omega_{B|a_{i}}\) are exactly the states β i . This is always so for pure bipartite states in quantum mechanics, a fact first observed by Schrödinger in 1935. Here, we show that, for weakly self-dual state spaces (those isomorphic, but perhaps not canonically isomorphic, to their dual spaces), the assumption that every state of a system A is steered by some bipartite state of a composite AA consisting of two copies of A, amounts to the homogeneity of the state cone. If the state space is actually self-dual, and not just weakly so, this implies (via the Koecher-Vinberg Theorem) that it is the self-adjoint part of a formally real Jordan algebra, and hence, quite close to being quantum mechanical.

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.

Similar content being viewed by others

Notes

  1. In the scheme, the two possible values to which Alice can commit are represented by two distinct ensembles for the same density matrix: Alice is to send samples from the ensemble to Bob in order to commit, and later reveal which states she drew so that Bob can check that she used the claimed ensemble. However, by sending, not a draw from the ensemble but one system of a pure bipartite entangled state with the specified density matrix, and keeping the other system, she can realize either ensemble after she’s already sent the systems to Bob by making measurements on her entangled system, enabling her to perfectly mimic commitment to either bit.

  2. A convex cone in a real vector space V is a convex set KV closed under multiplication by non-negative scalars. If K∩−K={0}, the cone is said to be pointed.

  3. The set Ω A is a base for the positive cone A +, i.e., every non-zero αA + is a positive scalar multiple of a unique vector—namely, α/u A (α)—belonging to Ω A . Indeed, what we are calling an abstract state space is essentially the same thing as a ordered vector space with a distinguished cone-base, i.e., a (finite-dimensional) cone-base space.

  4. We might also consider still larger spaces, of which AB is only a quotient—indeed, this is necessary to accommodate the states on the usual tensor product of real Hilbert spaces, as a composite state space, as discussed below. However, we shall not pursue this point any further here.

  5. Similarly, an unnormalized effect e on \(A \otimes_{{\rm min}}B\) amounts to a positive linear maps \(\hat{e} : A \rightarrow B^{*}\) (or, dually, to \(\hat{e}^{\ast}: B \rightarrow A^{*}\)), with e normalized iff \(\hat{e}(\alpha) \le u_{B}\) for normalized states αΩ A .

  6. In [11], it is assumed that physical processes are “completely positive”, meaning that they can be tensored with the identity to yield positive transformations. This property depends, not on the geometry of the two state spaces involved, but on the entire probabilistic theory at hand.

  7. It is easily seen that direct sums of homogeneous or weakly self-dual cones are, respectively, homogeneous or weakly self-dual. For weak self-duality, one just proves that any sum of isomorphisms α i :A i B i of direct summands, is an isomorphism of the direct sums ⨁ i A i and ⨁ i B i . For homogeneity, one uses the fact that the interior of a direct sum consists precisely of sums of interior points of all summands. Then to get from any such interior point x=∑ i x i to any other y=∑ i y i , one uses a sum of automorphisms α i of the summands \(A^{i}_{+}\), chosen (as homogeneity of each \(A^{i}_{+}\) ensures is possible) such that α i (x i )=y i .

  8. This last result also requires that all composites AB, AB′ of isomorphic systems AA′, BB′, contained in the theory, be isomorphic.

References

  1. Araki, H.: On a characterization of the state space of quantum mechanics. Commun. Math. Phys. 75, 1–24 (1980)

    Article  MathSciNet  ADS  MATH  Google Scholar 

  2. Barnum, H., Barrett, J., Leifer, M., Wilce, A.: Cloning and broadcasting in generic probabilistic models (2006). arXiv:quant-ph/0611295

  3. Barnum, H., Barrett, J., Leifer, M., Wilce, A.: A general no-cloning theorem. Phys. Rev. Lett. 99, 240501 (2007). arXiv:0707.0620

    Article  ADS  Google Scholar 

  4. Barnum, H., Barrett, J., Leifer, M., Wilce, A.: Teleportation in general probabilistic theories. In: Abramsky, S., Milsove, M. (eds.) Mathematical Foundations of Information Flow. Proceedings of Symposia in Applied Mathematics, vol. 71. Am. Math. Soc., Providence (2012). arXiv:0805.3553

    Google Scholar 

  5. Barnum, H., Barrett, J., Clark, L., Leifer, M., Spekkens, R., Stepanik, N., Wilce, A., Wilke, R.: Entropy and information causality in non-signaling theories. New J. Phys. 12, 038824 (2010). arXiv:0909.5075

    Article  MathSciNet  Google Scholar 

  6. Barnum, H., Dahlsten, O., Leifer, M., Toner, B.: Nonclassicality without entanglement enables bit committment. In: IEEE Information Theory Workshop, Porto (2008). arXiv:0803.1264

    Google Scholar 

  7. Barnum, H., Wilce, A.: Information processing in convex operational theories. Electron. Notes Theor. Comput. Sci. 270, 3–15 (2011). arXiv:0908.2352

    Article  Google Scholar 

  8. Barnum, H., Wilce, A.: Ordered linear spaces and categories as frameworks for information-processing characterizations of quantum and classical theory (2009). arXiv:0908.2354

  9. Barnum, H., Wilce, A.: Local tomography and the Jordan structure of quantum theory (2012). arXiv:1202.4513

  10. Barrett, J.: Information processing in general probabilistic theories. Phys. Rev. A 75, 032304 (2007). arXiv:quant-ph/0508211

    Article  ADS  Google Scholar 

  11. Chiribella, G., D’Ariano, G.M., Perinotti, P.: Probabilistic theories with purification. Phys. Rev. A 81, 062348 (2010). arXiv:0908.1583

    Article  ADS  Google Scholar 

  12. Bennett, C.H., Brassard, G.: Quantum cryptography: public key distribution and coin tossing. In: Proceedings of IEEE International Conference on Computers Systems and Signal Processing, Bangalore India, pp. 175–179 (1984)

    Google Scholar 

  13. Dakic, B., Brukner, C.: Quantum theory and beyond: is entanglement special? (2009). arXiv:0911.0695

  14. Davies, E.B., Lewis, J.T.: An operational approach to quantum probability. Commun. Math. Phys. 17, 239–260 (1970)

    Article  MathSciNet  ADS  MATH  Google Scholar 

  15. Edwards, C.M.: The operational approach to quantum probability I. Commun. Math. Phys. 17, 207–230 (1971)

    Google Scholar 

  16. Faraut, J., Korányi, A.: Analysis on Symmetric Cones. Oxford University Press, London (1994)

    MATH  Google Scholar 

  17. Foulis, D.J., Randall, C.H.: Empirical logic and tensor products. In: Neumann, H. (ed.) Interpretations and Foundations of Quantum Mechanics. Wissenschaftsverlag, Mannheim (1981)

    Google Scholar 

  18. Hanche-Olsen, H.: In: Jordan Algebras with Tensor Products Are C -Algebras. Springer Lecture Notes in Mathematics, vol. 1132 (1985)

    Google Scholar 

  19. Hadjisaavas, N.: Properties of mixtures of non-orthogonal quantum states. Lett. Math. Phys. 5, 327–332 (1981)

    Article  MathSciNet  ADS  Google Scholar 

  20. Hardy, L.: Quantum theory from five reasonable axioms. J. Phys. A 40, 3081 (2007). arXiv:quant-ph/0101012. See also, Hardy, L.: A framework for probabilistic theories with non-fixed causal structure. J. Phys. A. 40, 3081 (2007)

    Article  MathSciNet  ADS  MATH  Google Scholar 

  21. Hughston, L.P., Jozsa, R., Wootters, W.K.: A complete classification of quantum ensembles having a given density matrix. Phys. Lett. A 183, 14–18 (1993)

    Article  MathSciNet  ADS  Google Scholar 

  22. Kläy, M.: Einstein-Podolski-Rosen Experiments: the structure of the sample space. Found. Phys. Lett. 1, 205–244 (1988)

    Article  MathSciNet  Google Scholar 

  23. Kläy, M., Randall, C.H., Foulis, D.J.: Tensor products and probability weights. Int. J. Theor. Phys. 26, 199–219 (1987)

    Article  MATH  Google Scholar 

  24. Koecher, M.: Die geoodätischen von positivitaätsbereichen. Math. Ann. 135, 192–202 (1958)

    Article  MathSciNet  MATH  Google Scholar 

  25. Koecher, M.: Jordan Algebras and Their Applications Univ. Minnesota Lecture Notes, Minneapolis (1962) Reprinted as Krieger, A., Walcher, S.: The Minnesota Notes on Jordan Algebras and their Applications. Lecture Notes in Mathematics, vol. 1710. Springer, Berlin (1999)

    MATH  Google Scholar 

  26. Lo, H.K., Chau, H.F.: Is quantum bit commitment really possible? Phys. Rev. Lett. 78, 3410–3413 (1997)

    Article  ADS  Google Scholar 

  27. Mackey, G.: Mathematical Foundations of Quantum Mechanics. Addison-Wesley, Reading (1963)

    MATH  Google Scholar 

  28. Mayers, D.: Unconditionally secure quantum bit commitment is impossible. Phys. Rev. Lett. 78, 3414 (1997). arXiv:quant-ph/9605044

    Article  ADS  Google Scholar 

  29. Pawlowski, M., Paterek, T., Kazlikowski, D., Scarani, V., Winter, A., Żukowski, M., et al.: A new physical principle: information causality. Nature 461, 1101–1104 (2009). arXiv:0905.2292

    Article  ADS  Google Scholar 

  30. Satake, I.: Algebraic Structures of Symmetric Domains. Publications of the Mathematical Society of Japan, vol. 14. Iwanami Shoten/Princeton University Press, Tokyo/Princeton (1980)

    MATH  Google Scholar 

  31. Schrödinger, E.: Probability relations between separated systems. Proc. Camb. Philos. Soc. 32, 446–452 (1936)

    Article  ADS  Google Scholar 

  32. Vinberg, E.B.: Homogeneous cones. Dokl. Akad. Nauk SSSR 141, 270–273 (1960). English transl., Sov. Math. Dokl. 2, 1416–1619 (1961)

    MathSciNet  Google Scholar 

  33. Vinberg, E.B.: The theory of convex homogeneous cones. Tr. Mosk. Mat. Obŝ. 12, 303–358 (1963). English transl., Trans. Mosc. Math. Soc. 340–403 (1963)

    MathSciNet  Google Scholar 

  34. Wilce, A.: Tensor products in generalized measure theory. Int. J. Theor. Phys. 31, 1915 (1992)

    Article  MathSciNet  MATH  Google Scholar 

  35. Wilce, A.: Four and a half axioms for quantum mechanics. In: Ben-Menahem, Y., Hemmo, M. (eds.) Probability in Physics. Springer, Berlin (2012)

    Google Scholar 

Download references

Acknowledgements

This research was supported by the United States Government through grant OUR-0754079 from the National Science Foundation. It was also supported by Perimeter Institute for Theoretical Physics; work at Perimeter Institute is supported in part by the Government of Canada through Industry Canada and by the Province of Ontario through the Ministry of Research and Innovation. H. Barnum and A. Wilce also wish to thank: Samson Abramsky and Bob Coecke for extending them the hospitality of the Oxford University Computing Laboratory during November of 2009, when parts of this paper were written; C. Martin Edwards for referring us to the work of H. Hanche-Olsen; the Foundational Questions Institute and the University of Cambridge’s DAMTP for sponsoring the workshop Operational Probabilistic Theories as Foils to Quantum Theory, Cambridge (UK), July 2007, where some of these ideas were initially presented and their development into the present paper begun. Further work was done at other workshops and conferences: we thank Renato Renner, Oscar Dahlsten, and the Pauli Center for Theoretical Studies and the initiative “Quantum Science and Technology” at ETH Zürich, for sponsoring the workshop “Information Primitives and Laws of Physics”, March, 2008; Jeffrey Bub, Robert Rynasiewicz, and the University of Maryland, College Park and Georgetown University for organizing and sponsoring New Directions in the Foundations of Physics, May 2008; Hans Briegel, Bob Coecke, and the other organizers and the European network QICS, the EPSRC, and the IQOQI of the Austrian Academy of Sciences for the workshop “Foundational Structures for Quantum Information and Computation”, September 2008.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Alexander Wilce.

Examples for Sect. 5

Examples for Sect. 5

Example A.1

(An example in which there is a section that enables steering)

Consider the following state in the maximal tensor product of two state spaces with square base. We’ll view this as the state space of two two-outcome tests {a,a′} and {b,b′}, and also identify a,a′,b,b′ with the atomic effects in the dual cone. We’ll label each of the four vertices of the normalized state space by the two atomic effects it makes certain: ab,ab′,ab,ab′.

Writing, e.g. abab′ for a bipartite product state, the state:

$$ \omega= {\frac{1}{2}} \bigl(ab \otimes ab + a'b \otimes a'b \bigr) $$
(5)

has

$$ \hat{\omega}(a) = {\frac{1}{2}}ab, ~~ \hat{\omega} \bigl(a' \bigr) = {\frac{1}{2}}a'b , $$
(6)

so measuring {a,a′} on the first system gives the ensemble \(\{{\frac{1}{2}}ab, {\frac{1}{2}}a'b\}\) for the second-system marginal \(\omega^{B} = {\frac{1}{2}}(ab + a'b)\). \({{\rm Face}}(\omega^{B})\) is generated as nonnegative linear combinations of ab and ab, and is thus a two-dimensional ordered subspace of the three-dimensional state space of the second system. Its unit interval is the square that is the convex hull of (0,0),ab/2,ab/2,(ab+ab)/2. The quotient of the first system space, A, by the kernel of the linear map \(\hat{\omega}\), can be represented by setting up the state space to have 90 opening angle between opposite extremal rays, and taking the orthogonal (90) projection onto the plane normal to the ray b′, i.e. the projection along the ray generated by b′. This indeed gives a two-dimensional classical state space, i.e. one isomorphic to \({{\rm Face}}(\omega^{B})\). Moreover, there is an affine section σ of this quotient map into \(A^{*}_{+}\), given by inverting the relations (6), thus allowing us to map the unit interval [0,ω B] in \({{\rm Face}}(\omega^{B})\) to the diagonal cross section, in the a,a′ plane, of the unit interval of \(A^{*}_{+}\). σπ is indeed a positive projection on \(A^{*}_{+}\), namely the orthogonal projection onto this plane.

For the state (5), the mentioned section is the only affine section over \({{\rm Face}}(\omega^{B})\). A slight modification, however, gives an example in which many affine sections exist.

Example A.2

(An example in which there are many affine sections over the image of the order unit)

Let

$$ \hat{\omega}= {\frac{1}{2}} \bigl( x \otimes ab + y \otimes a'b \bigr) $$
(7)

with x any state in the line segment [ab,ab′] and y any state in the segment [ab,ab′] respectively. This still gives a state steering for the same marginal \(\omega^{B} = {\frac{1}{2}}(ab + a'b)\) as in the preceding example. If we choose x=ab,y=ab′, that is,

$$ \hat{\omega}= {\frac {1}{2}} \bigl( ab \otimes ab + a'b' \otimes a'b \bigr) , $$
(8)

we can steer the marginal using either of the two observables {a,a′},{b,b′}. In this case, there are two distinct sections over \({{\rm Face}}(\omega^{B})\) that enable steering.

In the example of Eq. (8), the kernel of \(\hat{\omega}\) is generated, not by b′ as before, but by abab′, and the natural way of attempting to represent the quotient map, namely by projection onto a subspace complementary to the kernel in the original space, if we project orthogonally in the natural geometry, projects onto the subspace spanned by ab,ab′, and the order unit, giving a slice of the order interval that includes the diagonal of the square.

Although the kernel of this map is of course one-dimensional, its positive kernel, which we’ve claimed must be an exposed face of the state cone, is the trivial such exposed face: the zero subspace. This example shows that quotients whose positive kernel is trivial are not necessarily uninteresting or ill-behaved.

Example A.3

(A positive section of the order-quotient map is not necessary for steering)

Let \(A \simeq {\mathbb{R}}^{4}\) be an abstract state space whose normalized states form a cube; its dual cone is a regular polyhedral cone in \({\mathbb{R}} ^{4}\) with octahedral base. The atomic effects (largest effects on extremal rays of the effect cone) are the vertices of such an octahedral base. The example has BR 3 a cone with regular hexagonal base; the center of the hexagon is a state having three distinct two-state extremal ensembles, each composed of a pair of opposite vertices of the hexagon (with weights 1/2 on each one). \(\hat{\omega}^{AB}\) is chosen to take the six vertices of the octahedron of atomic effects in A to the six vertices of the hexagon of states normalized to 1/2, in such a way that opposite vertex-pairs are mapped to opposite vertex-pairs. This is clearly positive and linear, and maps the order-unit (twice the center of the octahedron of atomic effects) to the center of the hexagon of normalized states (which is twice the center of the hexagon of states normalized to 1/2). \({{\rm Face}}(\hat{\omega}^{B})\) is the entire hexagonal cone in \({\mathbb{R}}^{3}\), so \(\hat{\omega}\) is itself (technically, is a representative of) the quotient π. All three ensembles for ω B consist of extremal states, and because they have unique \(\hat{\omega}\)-preimages, a section of \(\hat{\omega}\), we know that a section of π must take these, the vertices of a hexagon, to the vertices of the octahedron, matching opposites to opposites. But no affine map can do this. The affine span of the hexagonal points has affine dimension 2, while that of the octagonal ones has affine dimension 3.

Rights and permissions

Reprints and permissions

About this article

Cite this article

Barnum, H., Gaebler, C.P. & Wilce, A. Ensemble Steering, Weak Self-Duality, and the Structure of Probabilistic Theories. Found Phys 43, 1411–1427 (2013). https://doi.org/10.1007/s10701-013-9752-2

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s10701-013-9752-2

Keywords

Navigation