Skip to main content
Log in

Local Tomography and the Jordan Structure of Quantum Theory

  • Published:
Foundations of Physics Aims and scope Submit manuscript

Abstract

Using a result of H. Hanche-Olsen, we show that (subject to fairly natural constraints on what constitutes a system, and on what constitutes a composite system), orthodox finite-dimensional complex quantum mechanics with superselection rules is the only non-signaling probabilistic theory in which (i) individual systems are Jordan algebras (equivalently, their cones of unnormalized states are homogeneous and self-dual), (ii) composites are locally tomographic (meaning that states are determined by the joint probabilities they assign to measurement outcomes on the component systems) and (iii) at least one system has the structure of a qubit. Using this result, we also characterize finite dimensional quantum theory among probabilistic theories having the structure of a dagger-monoidal category.

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. State-completeness is assumed in the derivation of an homogeneous, self-dual cone in [32]; on the other hand, it is a consequence of the somewhat simpler assumptions in [30].

  2. The reader may have noticed that the only explicit use we’ve made of local tomography in the derivation of Corollary 3.7 is in the proof of part (a) of Lemma 3.4. However, to apply Hanche-Olsen’s theorem together with Corollary 3.7, we need to know that the ordinary vector space tensor product \(\mathbf{E}(A) \otimes \mathbf{E}(B)\), where \(B\) is a qubit, has an HSD structure; this only follows if the composite \(AB\) is locally tomographic.

References

  1. Abramsky, S., Coecke, B.: A categorical semantics of quantum protocols. In: Proceedings of the 19th Annual IEEE Symposium on Logic in Computer Science (LICS ’04), pp. 415–425 (2004)

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

    Article  ADS  MATH  MathSciNet  Google Scholar 

  3. Baez, J.: Quantum quandaries: a category-theoretic perspective. In: Rickles, D., French, S., Saatsi, J. (eds.) The Structural Foundations of Quantum Gravity. Oxford University Press, Oxford (2006). (Preprint arxiv.org/abs/quant-ph/0404040v2, 2004)

    Google Scholar 

  4. Barnum, H., Duncan, R., Wilce, A.: Symmetry, compact closure, and dagger compactness for categories of convex operational models. arxiv:1004.2920 (2010)

  5. Barnum, H., Fuchs, C., Renes, J., Wilce, A.: Influence-free states on compound quantum systems. arXiv:quant-ph/0507108 (2005)

  6. Barnum, H., Gaebler, P., Wilce, A.: Ensemble steering, weak self-duality, and the structure of probabilistic theories. arXiv:0912.5532 (2009)

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

  8. Barnum, H., Barrett, J., Leifer, M., Wilce, A.: A generalized no-broadcasting theorem. Phys. Rev. Lett. 99, 240501–240504 (2007). (Preprint arxiv:0707.0620, 2007)

    Article  ADS  Google Scholar 

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

  10. Dakić, B., Brukner, Č.: Quantum theory and beyond: is entanglement special? arXiv:0911.0695 (2009)

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

    Article  ADS  MATH  MathSciNet  Google Scholar 

  12. de la Torre, G., Masanes, Ll., Short, A., Müller, M.: Deriving quantum theory from its local structure and reversibility, Preprint arXiv:1110:5482 (2011)

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

    MATH  Google Scholar 

  14. Foulis, D., Randall, C.: Empirical logic and tensor products. In: Neumann, H. (ed.) Interpretations and Foundations of Quantum Theory. B. I. Wisssencshaft, Mannheim (1981)

    Google Scholar 

  15. Goyal, P.: From information geometry to to quantum theory. New J. Phys. 12, 023012 (2010)

    Article  ADS  MathSciNet  Google Scholar 

  16. Hanche-Olsen, H.: JB-algebras with tensor products are \(C^{\ast }\)-algebras. In: Araki, H. (ed.) Operator Algebras and their Connections with Topology and Ergodic Theory, Lecture Notes in Mathematics 1132. Springer, Berlin (1985)

    Google Scholar 

  17. Hardy, L.: Quantum theory from five reasonable axioms. arXiv:quant-ph/0101012 (2000)

  18. Holevo, A.: Probabilistic and Statistical Aspects of Quantum Mechanics, 2nd edn. North-Holland, Amsterdam (1982). (Edizioni della Normale, Pisa, 2011)

    Google Scholar 

  19. Jordan, P.: Ueber verallgemeinerungsmöglichkeiten des formalismus der quantenmechanik. Nachr. Akad. Wiss. Göttingen Math. Phys. Kl. 41, 209–217 (1933)

    Google Scholar 

  20. Jordan, P., von Neumann, J., Wigner, E.P.: On an algebraic generalization of the quantum-mechanical formalism. Ann. Math. 35, 29–64 (1934)

    Article  Google Scholar 

  21. Knapp, A.: Lie Groups Beyond an Introduction, 2nd edn. Birkhauser, Basel (2002)

    MATH  Google Scholar 

  22. Koecher, M.: Die geodätischen von positivitätsbereichen. Math. Annalen. 135, 192–202 (1958)

    Article  MATH  MathSciNet  Google Scholar 

  23. Ludwig, G.: Foundations of Quantum Mechanics. Springer, Berlin (1985)

    Book  MATH  Google Scholar 

  24. Mackey, G.: Mathematical Foundations of Quantum Mechanics. Addison Wesley, Boston (1963)

    MATH  Google Scholar 

  25. Masanes, L.I., Müller, M.: A derivation of quantum theory from physical requirements. New J. Phys. 13, 063001 (2011). (arXiv:1004.1483, 2011)

  26. Müller, M., Ududec, C.: The computational power of quantum mechanics determines its self-duality. Phys. Rev. Lett. 108, 130401 (2012). (Preprint arXiv:1110:3516, 2011)

    Article  Google Scholar 

  27. Rau, J.: On quantum vs. classical probability. Ann. Phys. 324, 2622–2637 (2009)

    Article  ADS  MATH  MathSciNet  Google Scholar 

  28. Selinger, P.: Towards a semantics for higher-order quantum computation. In: Proceedings of the 2nd International Workshop on Quantum Programming Languages, Turku, pp. 127–143. Turku Center for Computer Science, Publication No. 33 (2004)

  29. Vinberg, E.B.: Homogeneous cones. Dokl. Acad. Nauk. SSSR 141, 270–273 (1961). (English trans. Soviet Math. Dokl. 2(1961), 1416–1619)

    MathSciNet  Google Scholar 

  30. Wilce, A.: Conjugates, correlation and quantum mechanics, arXiv:1206.2897 (2012)

  31. Wilce, A.: Four and a half axioms for finite-dimensional quantum theory. In: Ben-Menahem, Y., Hemmo, M. (eds.) Probability in Physics: Essays in Honor of Itamar Pitowsky. Springer, Berlin (2012). (Preprint arXiv:0912.5530, 2009)

  32. Wilce, A.: Symmetry, self-duality, and the Jordan structure of quantum theory. Preprint arXiv:1110.6607 (2011)

  33. Wilce, A.: The tensor product in generalized measure theory. Int. J. Theor. Phys. 31, 1915–1928 (1992)

    Article  MATH  MathSciNet  Google Scholar 

Download references

Acknowledgments

We thank C. M. Edwards for drawing our attention to Hanche-Olsen’s paper. Part of this work was done while the authors were guests of the Oxford University Computing Laboratory, whose hospitality is also gratefully acknowledged. H. B. thanks the Foundational Questions Institute (FQXi) for travel support for the visit. Additional work was done at the 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.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Alexander Wilce.

Appendix: The Koecher-Vinberg Theorem

Appendix: The Koecher-Vinberg Theorem

This appendix contains a detailed sketch of the proof of the Koecher-Vinberg Theorem. This is almost entirely extracted from Faraut and Koranyi [13], to whom we refer for many of the details, but with a few minor modifications in order to obtain the precise form of the theorem (our Theorem 2.3) that we required above.

In what follows, \(\mathbf{E}\) is a finite-dimensional order-unit space with a self-dual positive cone \(\mathbf{E}_+\). By this we mean that there exists a self-dualizing inner product on \(\mathbf{E}\). Let \(\text{ Aut }(\mathbf{E})\) denote the group of order-automorphisms of \(\mathbf{E}\). This is a Lie group, as is any closed subgroup \(G \le \text{ Aut }(\mathbf{E})\). When \(G\) is connected and acts homogeneously on \(\mathbf{E}_+\) (that is, transitively on the interior of \(\mathbf{E}_+\)), we can use this action to construct a Jordan product on \(\mathbf{E}\), as per Theorem 2.3, which, for convenience, we now restate:

Theorem 2.3

Let \(G\) be a closed, connected subgroup of \(\text{ Aut }(\mathbf{E})\) and let \({\mathfrak g}_u\) denote the Lie algebra of \(G_u\), the stabilizer of \(u\) in \(G\). Then

  1. (a)

    It is possible to choose a self-dualizing inner product on \(\mathbf{E}_+\) in such a way that \(G_u = G \cap O(\mathbf{E})\), where \(O(\mathbf{E})\) is the orthogonal group with respect to the chosen inner product;

  2. (b)

    If \(G = G^{\dagger }\) with respect to this inner product, then

    $$\begin{aligned} {\mathfrak g}_u = \{ X \in {\mathfrak g}| X^{\dagger } = -X\} = \{ X \in {\mathfrak g}| Xu = 0\}, \end{aligned}$$

    and \({\mathfrak g}= {\mathfrak p}\oplus {\mathfrak g}_u\), where \({\mathfrak p}= \{ X \in {\mathfrak g}| X^{\dagger } = X\}\).

  3. (c)

    In this case the mapping \({\mathfrak p}\rightarrow \mathbf{E}\), given by \(X \mapsto Xu\), is an isomorphism of linear spaces. Letting \(L_a\) denote the unique element \(X \in {\mathfrak p}\) with \(X u = a\), define

    $$\begin{aligned} a \bullet b = L_{a} b \end{aligned}$$

    for all \(a, b \in \mathbf{E}\). Then \(\bullet \) makes \(\mathbf{E}\) a formally real Jordan algebra, with identity element \(u\).

We break the proof into a series of Lemmas. Throughout, \(G\) is a connected, closed subgroup of \(\text{ Aut }(\mathbf{E})\), acting homogeneously on the interior, \(\mathbf{E}^{\circ }_{+}\), of the cone \(\mathbf{E}_+\).

Lemma A

If \(g \in \text{ Aut }(\mathbf{E})\), then \(g^{*} \in \text{ Aut }(\mathbf{E})\), where \(g^{*}\) is the adjoint with respect to any self-dualizing inner product.

Proof

If \(g \in \text{ Aut }(\mathbf{E})\) preserves \(\mathbf{E}_+\), then \(g^{\dagger }\) preserves \(\mathbf{E}^{+} = \mathbf{E}_+\). \(\square \)

Lemma B

Any compact subgroup of \(\text{ Aut }(\mathbf{E})\) fixes some point \(a\) in the interior of \(\mathbf{E}_+\). In particular, a maximal compact subgroup is a stabilizer, and vice versa. Thus, all maximal compact subgroups of \(\text{ Aut }(\mathbf{E})\) are conjugate.

For a proof, see [13], Proposition I.1.8ff.

Lemma C

For a suitable choice of self-dualizing inner product, \(O(\mathbf{E}) \cap G \le G_u\), where \(O(\mathbf{E})\) is the orthogonal group relative to the chosen inner product.

Proof

If \(\langle , \rangle \) is any inner product on \(\mathbf{E}\) with respect to which \(\mathbf{E}_{+} = \mathbf{E}^{+}\), one can show that \(O(\mathbf{E}) \cap \text{ Aut }(\mathbf{E})\) is the stabilizer of some \(a \in \mathbf{E}_{+}^{\circ }\) ([13], Proposition I.1.9). It follows that \(O(\mathbf{E}) \cap G \le G_{a}\). Since \(G\) acts transitively on \(\mathbf{E}_{+}^{\circ }\), we can find some \(g \in G\) with \(ga = u\); replacing \(\langle , \rangle \), if necessary, by the inner product \(\langle x, y \rangle _{g} : = \langle g x, g y \rangle \)—which is also self-dualizing, by [13], Proposition I.1.7—we can assume that \(a = u\), whence, that \(O(\mathbf{E}) \cap G \le G_u\). \(\square \)

This gives us part (a) of the Theorem. Now let \(K = G \cap O(\mathbf{E})\), and let \({\mathfrak k}\) denote \(K\)’s Lie algebra. Notice that \({\mathfrak k}= {\mathfrak g}\cap \mathfrak {o}(\mathbf{E}) = \{ X \in {\mathfrak g}| X^{\dagger } = -X\}\).

Corollary 1

Let the inner product on \(\mathbf{E}\) be chosen as per Lemma C. Let \(K = G \cap O(\mathbf{E})\), and let \({\mathfrak k}\) be the Lie algebra of \(K\), and let \({\mathfrak g}_u\) denote the Lie algebra of \(G_u \le G\). Then

  1. (a)

    \(G_{u}\) is connected;

  2. (b)

    \(G_{u} = K\);

  3. (c)

    \({\mathfrak g}_u = {\mathfrak k}\);

Proof

(a) \(G/G_u\) is homeomorphic to the simply connected space \(\mathbf{E}^{\circ }_{+}\); hence, as \(G\) is connected, so is \(G_{u}\) (see, e.g., [21], Proposition 1.94). It follows that if \(G_u\) and \(K\) have the same Lieq algebra, they coincide—in other words, (b) follows from (c). To prove (c), note that since \(K \le G_u\), by the choice of inner product, we have \({\mathfrak k}\le {\mathfrak g}_u\). For every \(X \in {\mathfrak g}_u \le {\mathfrak g}\), we have the decomposition \(X = X_1 + X_2\) with \(X_1\) self-adjoint and \(X_2\), skew-adjoint. Since \(X_2 \in {\mathfrak k}\subseteq {\mathfrak g}_{u}\), it follows that \(X_1 = X - X_2 \in {\mathfrak g}_u\) as well, whence, \(e^{tX_1} \in G_u\) for all \(t\). However, \(G_u\) is compact, so this implies that \(e^{tX_1}\) is bounded as a function of \(t\). Since \(X_1\) is self-adjoint, this is possible only if \(X_1 = 0\). Hence, \(X = X_2 \in {\mathfrak k}\), and we have \({\mathfrak g}_u \le {\mathfrak k}\). \(\square \)

Proof of Part (b) of Theorem 2.3: Suppose that \(G\) is self-adjoint with respect to the self-dualizing inner product of Lemma C. Then, for every \(X \in {\mathfrak g}\), \(X^{*} \in {\mathfrak g}\). To see this, let \(X = \gamma '(0)\) where \(\gamma : {\mathbb R} \rightarrow G\) is a smooth path with \(\gamma (0) = \text{1 }\), and note that \(\gamma ^{*} : t \mapsto \gamma (t)^{*}\) is another such path, with \({\gamma ^{*}}'(0) = X^{*}\). Thus, for every \(X \in {\mathfrak g}\), \(X_1 := \left( X + X^{\dagger }\right) /2\) and \(X_2 = \left( X - X^{\dagger }\right) /2\) also lie in \({\mathfrak g}\); \(X = X_1 + X_2\), so \({\mathfrak g}\) decomposes as the direct sum \({\mathfrak g}= {\mathfrak p}+ {\mathfrak k}\), where \({\mathfrak p}\) and \({\mathfrak k}\) are respectively the spaces of self-adjoint and skew-adjoint elements of \({\mathfrak g}\).\(\square \)

Remark

Note also, for later reference, that if \(X, Y \in {\mathfrak p}\), we also have \([X,Y] \in {\mathfrak g}\) and \([X,Y]^{\dagger } = [Y,X] = -[X,Y]\), i.e., \([X,Y] \in {\mathfrak k}\).

The interior, \(\mathbf{E}^{\circ }_{+}\), of the cone \(\mathbf{E}_+\) is a smooth manifold, on which \(G\) acts smoothly. Thus, we have a canonical smooth mapping \(\phi : G \rightarrow \mathbf{E}^{\circ }_{+}\) given by \(g \mapsto gu\). Differentiating this, we obtain a linear mapping \(d\phi (\text{1 }) : {\mathfrak g}\rightarrow T_{u}\left( \mathbf{E}^{\circ }_{+}\right) = \mathbf{E}\). Explicitly, if \(X = \gamma '(0) \in {\mathfrak g}\), where \(\gamma \) is a smooth path in \(G\) with \(\gamma (0) = \text{1 }\), then

$$\begin{aligned} d\phi (\text{1 })(X) = \frac{d}{dt} \phi (\gamma (t))_{t = 0} = \gamma '(0)u = Xu. \end{aligned}$$
(3)

Lemma D

Let \(X \in {\mathfrak g}\). Then \(X \in {\mathfrak k}\) iff \(Xu = 0\).

Proof

Suppose \(d\phi (\text{1 })(X) = Xu = 0\), where \(X = \gamma '(0)\). The vector-valued function \(v(t) = e^{tX}u\) then satisfies

$$\begin{aligned} v' = Xv = Xe^{tX}u = e^{tX}Xu = 0. \end{aligned}$$

It follows that \(v\) is constant, i.e, that \(e^{tX}u = u\) for all \(t\). But then \(e^{tX} \in G_u\), so that \(X = \frac{d}{dt} e^{tX}|_{t =0} \in {\mathfrak k}\). Conversely, if \(X \in {\mathfrak k}\), then \(X = \gamma '(0)\) where \(\gamma (t) \in K\) and \(\gamma (0) = \text{1 }\), so that \(Xu = \gamma '(0) u = \left[ \frac{d}{dt}(\gamma (t)u)\right] _{t = 0} - \left[ \gamma (t) \frac{d}{dt} u\right] _{t=0} = \left[ \frac{d}{dt}(\gamma (t)u)\right] _{t = 0} = \frac{d}{dt} u |_{t = 0} = 0\) (the last equality uses \(\gamma (t) \in K\) and \(Ku=u\)). \(\square \)

Corollary 2

\(d\phi (\text{1 }) : {\mathfrak p}\simeq \text{ ran }(d\phi (\text{1 })) \le \mathbf{E}\).

Proof

By Lemma D and (0), \({\mathfrak k}\) is the kernel of \(d\phi (\text{1 })\); as established above (in the proof of part (b) of Lemma C, \({\mathfrak g}= {\mathfrak p}\oplus {\mathfrak k}\). \(\square \)

Lemma E

\(\text{ ran }(d\phi (\text{1 }))\) is all of \(\mathbf{E}\), i.e., \(d\phi (\text{1 }) : {\mathfrak p}\simeq \mathbf{E}\).

Proof

Note, first, that \(T_{g}(G) = gT_{1}(G) = g{\mathfrak g}\) for any \(g \in G\). We also have

$$\begin{aligned} (d\phi (g))(gX) = gXu \end{aligned}$$

for all \(X \in {\mathfrak g}\). Hence, \(\text{ ran }(d\phi (g)) = g \text{ ran }(d\phi (1))\). If the latter is not all of \(\mathbf{E}\), then every \(g\) is a critical point of \(\phi \), whence, every point \(\phi (g) = gu \in \mathbf{E}\) is a critical value. Sard’s Theorem now tells us that \(Gu = \mathbf{E}^{\circ }_{+}\) has measure zero, a contradiction. \(\square \)

Construction of the Jordan product Now define \(L_x \in {\mathfrak p}\) to be the unique self-adjoint element of \({\mathfrak g}\) with \(L_x u = x\). Set \(x \bullet y = L_x y\) for all \(x, y \in \mathbf{E}\). This is evidently bilinear. A series of computations (see [13], pp. 49-50) shows that it makes \(\mathbf{E}\) a formally real Jordan algebra with identity element \(u\). Specifically,

  1. (1)

    \(x \bullet y = y \bullet x\) since \(x \bullet y - y \bullet x = x \bullet (y \bullet u)u - y \bullet (x \bullet u) = [L_x,L_y]u = 0\) (as remarked above following the proof of Corollary 1, \(X, Y \in {\mathfrak p}\Rightarrow [X,Y] \in {\mathfrak k}\), and, by Lemma D, \({\mathfrak k}u = 0\));

  2. (2)

    \(x \bullet u = L_x u = x\) by definition of \(L_x\), so \(u\) serves as the identity;

  3. (3)

    The product satisfies the Jacobi identity. This is proved exactly as in [13]. Note that the argument uses the fact that the inner product is associative, which follows from \(L_x\) being self-adjoint. This also then gives us that the Jordan algebra \(\mathbf{E}\) is formally real.

This completes the proof of the Koecher-Vinberg Theorem. \(\square \)

Rights and permissions

Reprints and permissions

About this article

Cite this article

Barnum, H., Wilce, A. Local Tomography and the Jordan Structure of Quantum Theory. Found Phys 44, 192–212 (2014). https://doi.org/10.1007/s10701-014-9777-1

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s10701-014-9777-1

Keywords

Navigation