Strong subadditivity of quantum entropy

Last updated

In quantum information theory, strong subadditivity of quantum entropy (SSA) is the relation among the von Neumann entropies of various quantum subsystems of a larger quantum system consisting of three subsystems (or of one quantum system with three degrees of freedom). It is a basic theorem in modern quantum information theory. It was conjectured by D. W. Robinson and D. Ruelle [1] in 1966 and O. E. Lanford III and D. W. Robinson [2] in 1968 and proved in 1973 by E.H. Lieb and M.B. Ruskai, [3] building on results obtained by Lieb in his proof of the Wigner-Yanase-Dyson conjecture. [4]

Contents

The classical version of SSA was long known and appreciated in classical probability theory and information theory. The proof of this relation in the classical case is quite easy, but the quantum case is difficult because of the non-commutativity of the reduced density matrices describing the quantum subsystems.

Some useful references here include:

Definitions

We use the following notation throughout the following: A Hilbert space is denoted by , and denotes the bounded linear operators on . Tensor products are denoted by superscripts, e.g., . The trace is denoted by .

Density matrix

A density matrix is a Hermitian, positive semi-definite matrix of trace one. It allows for the description of a quantum system in a mixed state. Density matrices on a tensor product are denoted by superscripts, e.g., is a density matrix on .

Entropy

The von Neumann quantum entropy of a density matrix is

.

Relative entropy

Umegaki's [8] quantum relative entropy of two density matrices and is

.

Joint concavity

A function of two variables is said to be jointly concave if for any the following holds

Subadditivity of entropy

Ordinary subadditivity [9] concerns only two spaces and a density matrix . It states that

This inequality is true, of course, in classical probability theory, but the latter also contains the theorem that the conditional entropies and are both non-negative. In the quantum case, however, both can be negative, e.g. can be zero while . Nevertheless, the subadditivity upper bound on continues to hold. The closest thing one has to is the Araki–Lieb triangle inequality [9]

which is derived in [9] from subadditivity by a mathematical technique known as purification.

Strong subadditivity (SSA)

Suppose that the Hilbert space of the system is a tensor product of three spaces: . Physically, these three spaces can be interpreted as the space of three different systems, or else as three parts or three degrees of freedom of one physical system.

Given a density matrix on , we define a density matrix on as a partial trace: . Similarly, we can define density matrices: , , , , .

Statement

For any tri-partite state the following holds

,

where , for example.

Equivalently, the statement can be recast in terms of conditional entropies to show that for tripartite state ,

.

This can also be restated in terms of quantum mutual information,

.

These statements run parallel to classical intuition, except that quantum conditional entropies can be negative, and quantum mutual informations can exceed the classical bound of the marginal entropy.

The strong subadditivity inequality was improved in the following way by Carlen and Lieb [10]

,

with the optimal constant .

J. Kiefer [11] [12] proved a peripherally related convexity result in 1959, which is a corollary of an operator Schwarz inequality proved by E.H.Lieb and M.B.Ruskai. [3] However, these results are comparatively simple, and the proofs do not use the results of Lieb's 1973 paper on convex and concave trace functionals. [4] It was this paper that provided the mathematical basis of the proof of SSA by Lieb and Ruskai. The extension from a Hilbert space setting to a von Neumann algebra setting, where states are not given by density matrices, was done by Narnhofer and Thirring . [13]

The theorem can also be obtained by proving numerous equivalent statements, some of which are summarized below.

Wigner–Yanase–Dyson conjecture

E. P. Wigner and M. M. Yanase [14] proposed a different definition of entropy, which was generalized by Freeman Dyson.

The Wigner–Yanase–Dyson p-skew information

The Wigner–Yanase–Dyson -skew information of a density matrix . with respect to an operator is

where is a commutator, is the adjoint of and is fixed.

Concavity of p-skew information

It was conjectured by E. P. Wigner and M. M. Yanase in [15] that - skew information is concave as a function of a density matrix for a fixed .

Since the term is concave (it is linear), the conjecture reduces to the problem of concavity of . As noted in, [4] this conjecture (for all ) implies SSA, and was proved for in, [15] and for all in [4] in the following more general form: The function of two matrix variables

 

 

 

 

(1)

is jointly concave in and when and .

This theorem is an essential part of the proof of SSA in. [3]

In their paper [15] E. P. Wigner and M. M. Yanase also conjectured the subadditivity of -skew information for , which was disproved by Hansen [16] by giving a counterexample.

First two statements equivalent to SSA

It was pointed out in [9] that the first statement below is equivalent to SSA and A. Ulhmann in [17] showed the equivalence between the second statement below and SSA.

Both of these statements were proved directly in. [3]

Joint convexity of relative entropy

As noted by Lindblad [18] and Uhlmann, [19] if, in equation ( 1 ), one takes and and and differentiates in at , one obtains the joint convexity of relative entropy: i.e., if , and , then

 

 

 

 

(2)

where with .

Monotonicity of quantum relative entropy

The relative entropy decreases monotonically under completely positive trace preserving (CPTP) operations on density matrices,

.

This inequality is called Monotonicity of quantum relative entropy. Owing to the Stinespring factorization theorem, this inequality is a consequence of a particular choice of the CPTP map - a partial trace map described below.

The most important and basic class of CPTP maps is a partial trace operation , given by . Then

 

 

 

 

(3)

which is called Monotonicity of quantum relative entropy under partial trace.

To see how this follows from the joint convexity of relative entropy, observe that can be written in Uhlmann's representation as

for some finite and some collection of unitary matrices on (alternatively, integrate over Haar measure). Since the trace (and hence the relative entropy) is unitarily invariant, inequality ( 3 ) now follows from ( 2 ). This theorem is due to Lindblad [18] and Uhlmann, [17] whose proof is the one given here.

SSA is obtained from ( 3 ) with replaced by and replaced . Take . Then ( 3 ) becomes

Therefore,

which is SSA. Thus, the monotonicity of quantum relative entropy (which follows from ( 1 ) implies SSA.

Relationship among inequalities

All of the above important inequalities are equivalent to each other, and can also be proved directly. The following are equivalent:

The following implications show the equivalence between these inequalities.

is convex. In [3] it was observed that this convexity yields MPT;

Moreover, if is pure, then and , so the equality holds in the above inequality. Since the extreme points of the convex set of density matrices are pure states, SSA follows from JC;

See, [21] [22] for a discussion.

The case of equality

Equality in monotonicity of quantum relative entropy inequality

In, [23] [24] D. Petz showed that the only case of equality in the monotonicity relation is to have a proper "recovery" channel:

For all states and on a Hilbert space and all quantum operators ,

if and only if there exists a quantum operator such that

and

Moreover, can be given explicitly by the formula

where is the adjoint map of .

D. Petz also gave another condition [23] when the equality holds in Monotonicity of quantum relative entropy: the first statement below. Differentiating it at we have the second condition. Moreover, M.B. Ruskai gave another proof of the second statement.

For all states and on and all quantum operators ,

if and only if the following equivalent conditions are satisfied:

where is the adjoint map of .

Equality in strong subadditivity inequality

P. Hayden, R. Jozsa, D. Petz and A. Winter described the states for which the equality holds in SSA. [25]

A state on a Hilbert space satisfies strong subadditivity with equality if and only if there is a decomposition of second system as

into a direct sum of tensor products, such that

with states on and on , and a probability distribution .

Carlen-Lieb Extension

E. H. Lieb and E.A. Carlen have found an explicit error term in the SSA inequality, [10] namely,

If and , as is always the case for the classical Shannon entropy, this inequality has nothing to say. For the quantum entropy, on the other hand, it is quite possible that the conditional entropies satisfy or (but never both!). Then, in this "highly quantum" regime, this inequality provides additional information.

The constant 2 is optimal, in the sense that for any constant larger than 2, one can find a state for which the inequality is violated with that constant.

Operator extension of strong subadditivity

In his paper [26] I. Kim studied an operator extension of strong subadditivity, proving the following inequality:

For a tri-partite state (density matrix) on ,

The proof of this inequality is based on Effros's theorem, [27] for which particular functions and operators are chosen to derive the inequality above. M. B. Ruskai describes this work in details in [28] and discusses how to prove a large class of new matrix inequalities in the tri-partite and bi-partite cases by taking a partial trace over all but one of the spaces.

Extensions of strong subadditivity in terms of recoverability

A significant strengthening of strong subadditivity was proved in 2014, [29] which was subsequently improved in [30] and. [31] In 2017, [32] it was shown that the recovery channel can be taken to be the original Petz recovery map. These improvements of strong subadditivity have physical interpretations in terms of recoverability, meaning that if the conditional mutual information of a tripartite quantum state is nearly equal to zero, then it is possible to perform a recovery channel (from system E to AE) such that . These results thus generalize the exact equality conditions mentioned above.

See also

Related Research Articles

In physics, the CHSH inequality can be used in the proof of Bell's theorem, which states that certain consequences of entanglement in quantum mechanics cannot be reproduced by local hidden-variable theories. Experimental verification of the inequality being violated is seen as confirmation that nature cannot be described by such theories. CHSH stands for John Clauser, Michael Horne, Abner Shimony, and Richard Holt, who described it in a much-cited paper published in 1969. They derived the CHSH inequality, which, as with John Stewart Bell's original inequality, is a constraint—on the statistical occurrence of "coincidences" in a Bell test—which is necessarily true if an underlying local hidden-variable theory exists. In practice, the inequality is routinely violated by modern experiments in quantum mechanics.

In quantum physics, a measurement is the testing or manipulation of a physical system to yield a numerical result. A fundamental feature of quantum theory is that the predictions it makes are probabilistic. The procedure for finding a probability involves combining a quantum state, which mathematically describes a quantum system, with a mathematical representation of the measurement to be performed on that system. The formula for this calculation is known as the Born rule. For example, a quantum particle like an electron can be described by a quantum state that associates to each point in space a complex number called a probability amplitude. Applying the Born rule to these amplitudes gives the probabilities that the electron will be found in one region or another when an experiment is performed to locate it. This is the best the theory can do; it cannot say for certain where the electron will be found. The same quantum state can also be used to make a prediction of how the electron will be moving, if an experiment is performed to measure its momentum instead of its position. The uncertainty principle implies that, whatever the quantum state, the range of predictions for the electron's position and the range of predictions for its momentum cannot both be narrow. Some quantum states imply a near-certain prediction of the result of a position measurement, but the result of a momentum measurement will be highly unpredictable, and vice versa. Furthermore, the fact that nature violates the statistical conditions known as Bell inequalities indicates that the unpredictability of quantum measurement results cannot be explained away as due to ignorance about "local hidden variables" within quantum systems.

In mathematics, subadditivity is a property of a function that states, roughly, that evaluating the function for the sum of two elements of the domain always returns something less than or equal to the sum of the function's values at each element. There are numerous examples of subadditive functions in various areas of mathematics, particularly norms and square roots. Additive maps are special cases of subadditive functions.

In mathematical statistics, the Kullback–Leibler (KL) divergence, denoted , is a type of statistical distance: a measure of how one probability distribution P is different from a second, reference probability distribution Q. A simple interpretation of the KL divergence of P from Q is the expected excess surprise from using Q as a model when the actual distribution is P. While it is a measure of how different two distributions are, and in some sense is thus a "distance", it is not actually a metric, which is the most familiar and formal type of distance. In particular, it is not symmetric in the two distributions, and does not satisfy the triangle inequality. Instead, in terms of information geometry, it is a type of divergence, a generalization of squared distance, and for certain classes of distributions, it satisfies a generalized Pythagorean theorem.

The Peres–Horodecki criterion is a necessary condition, for the joint density matrix of two quantum mechanical systems and , to be separable. It is also called the PPT criterion, for positive partial transpose. In the 2×2 and 2×3 dimensional cases the condition is also sufficient. It is used to decide the separability of mixed states, where the Schmidt decomposition does not apply. The theorem was discovered in 1996 by Asher Peres and the Horodecki family

In physics, the von Neumann entropy, named after John von Neumann, is an extension of the concept of Gibbs entropy from classical statistical mechanics to quantum statistical mechanics. For a quantum-mechanical system described by a density matrix ρ, the von Neumann entropy is

The quantum Heisenberg model, developed by Werner Heisenberg, is a statistical mechanical model used in the study of critical points and phase transitions of magnetic systems, in which the spins of the magnetic systems are treated quantum mechanically. It is related to the prototypical Ising model, where at each site of a lattice, a spin represents a microscopic magnetic dipole to which the magnetic moment is either up or down. Except the coupling between magnetic dipole moments, there is also a multipolar version of Heisenberg model called the multipolar exchange interaction.

In linear algebra, the Schmidt decomposition refers to a particular way of expressing a vector in the tensor product of two inner product spaces. It has numerous applications in quantum information theory, for example in entanglement characterization and in state purification, and plasticity.

In quantum mechanics, notably in quantum information theory, fidelity is a measure of the "closeness" of two quantum states. It expresses the probability that one state will pass a test to identify as the other. The fidelity is not a metric on the space of density matrices, but it can be used to define the Bures metric on this space.

In quantum information theory, quantum relative entropy is a measure of distinguishability between two quantum states. It is the quantum mechanical analog of relative entropy.

In quantum information science, the concurrence is a state invariant involving qubits.

In quantum mechanics, and especially quantum information theory, the purity of a normalized quantum state is a scalar defined as

In quantum mechanics, and especially quantum information and the study of open quantum systems, the trace distanceT is a metric on the space of density matrices and gives a measure of the distinguishability between two states. It is the quantum generalization of the Kolmogorov distance for classical probability distributions.

The Fannes–Audenaert inequality is a mathematical bound on the difference between the von Neumann entropies of two density matrices as a function of their trace distance. It was proved by Koenraad M. R. Audenaert in 2007 as an optimal refinement of Mark Fannes' original inequality, which was published in 1973. Mark Fannes is a Belgian physicist specialised in mathematical quantum mechanics. He works at the KU Leuven. Koenraad M. R. Audenaert is a Belgian physicist and civil engineer. He currently works at Royal Holloway, University of London.

In the theory of quantum communication, an amplitude damping channel is a quantum channel that models physical processes such as spontaneous emission. A natural process by which this channel can occur is a spin chain through which a number of spin states, coupled by a time independent Hamiltonian, can be used to send a quantum state from one location to another. The resulting quantum channel ends up being identical to an amplitude damping channel, for which the quantum capacity, the classical capacity and the entanglement assisted classical capacity of the quantum channel can be evaluated.

In quantum mechanics, negativity is a measure of quantum entanglement which is easy to compute. It is a measure deriving from the PPT criterion for separability. It has shown to be an entanglement monotone and hence a proper measure of entanglement.

In quantum information theory, the classical capacity of a quantum channel is the maximum rate at which classical data can be sent over it error-free in the limit of many uses of the channel. Holevo, Schumacher, and Westmoreland proved the following least upper bound on the classical capacity of any quantum channel :

In mathematics, there are many kinds of inequalities involving matrices and linear operators on Hilbert spaces. This article covers some important operator inequalities connected with traces of matrices.

The min-entropy, in information theory, is the smallest of the Rényi family of entropies, corresponding to the most conservative way of measuring the unpredictability of a set of outcomes, as the negative logarithm of the probability of the most likely outcome. The various Rényi entropies are all equal for a uniform distribution, but measure the unpredictability of a nonuniform distribution in different ways. The min-entropy is never greater than the ordinary or Shannon entropy and that in turn is never greater than the Hartley or max-entropy, defined as the logarithm of the number of outcomes with nonzero probability.

Generalized relative entropy is a measure of dissimilarity between two quantum states. It is a "one-shot" analogue of quantum relative entropy and shares many properties of the latter quantity.

References

  1. Robinson, Derek W.; Ruelle, David (1967). "Mean entropy of states in classical statistical mechanics". Communications in Mathematical Physics. Springer Science and Business Media LLC. 5 (4): 288–300. Bibcode:1967CMaPh...5..288R. doi:10.1007/bf01646480. ISSN   0010-3616. S2CID   115134613.
  2. Lanford, Oscar E.; Robinson, Derek W. (1968). "Mean Entropy of States in Quantum‐Statistical Mechanics". Journal of Mathematical Physics. AIP Publishing. 9 (7): 1120–1125. Bibcode:1968JMP.....9.1120L. doi:10.1063/1.1664685. ISSN   0022-2488.
  3. 1 2 3 4 5 Lieb, Elliott H.; Ruskai, Mary Beth (1973). "Proof of the strong subadditivity of quantum‐mechanical entropy" (PDF). Journal of Mathematical Physics. AIP Publishing. 14 (12): 1938–1941. Bibcode:1973JMP....14.1938L. doi:10.1063/1.1666274. ISSN   0022-2488.
  4. 1 2 3 4 Lieb, Elliott H (1973). "Convex trace functions and the Wigner-Yanase-Dyson conjecture". Advances in Mathematics . 11 (3): 267–288. doi: 10.1016/0001-8708(73)90011-X . ISSN   0001-8708.
  5. M. Nielsen, I. Chuang, Quantum Computation and Quantum Information, Cambr. U. Press, (2000)
  6. M. Ohya, D. Petz, Quantum Entropy and Its Use, Springer (1993)
  7. E. Carlen, Trace Inequalities and Quantum Entropy: An Introductory Course, Contemp. Math. 529 (2009).
  8. Umegaki, Hisaharu (1962). "Conditional expectation in an operator algebra. IV. Entropy and information". Kodai Mathematical Seminar Reports. Tokyo Institute of Technology, Department of Mathematics. 14 (2): 59–85. doi: 10.2996/kmj/1138844604 . ISSN   0023-2599.
  9. 1 2 3 4 5 Araki, Huzihiro; Lieb, Elliott H. (1970). "Entropy inequalities". Communications in Mathematical Physics. 18 (2): 160–170. Bibcode:1970CMaPh..18..160A. doi:10.1007/BF01646092. ISSN   0010-3616. S2CID   189832417.
  10. 1 2 Carlen, Eric A.; Lieb, Elliott H. (2012). "Bounds for Entanglement via an Extension of Strong Subadditivity of Entropy". Letters in Mathematical Physics. 101 (1): 1–11. arXiv: 1203.4719 . Bibcode:2012LMaPh.101....1C. doi:10.1007/s11005-012-0565-6. S2CID   119317605.
  11. Kiefer, J. (July 1959). "Optimum Experimental Designs". Journal of the Royal Statistical Society, Series B (Methodological). 21 (2): 272–310. doi:10.1111/j.2517-6161.1959.tb00338.x.
  12. Ruskai, Mary Beth. "Evolution of a Fundemental [sic] Theorem on Quantum Entropy". youtube.com. World Scientific. Retrieved 20 August 2020. Invited talk at the Conference in Honour of the 90th Birthday of Freeman Dyson, Institute of Advanced Studies, Nanyang Technological University, Singapore, 26–29 August 2013. The note on Kiefer (1959) is at the 26:40 mark.
  13. Narnhofer, H. (1985). "From Relative Entropy to Entropy". Fizika. 17: 258–262.
  14. Wigner, E. P.; Yanase, M. M. (1 May 1963). "Information Content of Distributions". Proceedings of the National Academy of Sciences. 49 (6): 910–918. Bibcode:1963PNAS...49..910W. doi: 10.1073/pnas.49.6.910 . ISSN   0027-8424. PMC   300031 . PMID   16591109.
  15. 1 2 3 Wigner, Eugene P.; Yanase, Mutsuo M. (1964). "On the Positive Semidefinite Nature of a Certain Matrix Expression". Canadian Journal of Mathematics. Canadian Mathematical Society. 16: 397–406. doi:10.4153/cjm-1964-041-x. ISSN   0008-414X.
  16. Hansen, Frank (18 January 2007). "The Wigner-Yanase Entropy is not Subadditive". Journal of Statistical Physics. Springer Nature. 126 (3): 643–648. arXiv: math-ph/0609019 . Bibcode:2007JSP...126..643H. doi:10.1007/s10955-006-9265-x. ISSN   0022-4715. S2CID   119667187.
  17. 1 2 A. Ulhmann, Endlich Dimensionale Dichtmatrizen, II, Wiss. Z. Karl-Marx-University Leipzig 22 Jg. H. 2., 139 (1973).
  18. 1 2 Lindblad, Göran (1974). "Expectations and entropy inequalities for finite quantum systems". Communications in Mathematical Physics. 39 (2): 111–119. Bibcode:1974CMaPh..39..111L. doi:10.1007/BF01608390. ISSN   0010-3616. S2CID   120760667.
  19. Uhlmann, A. (1977). "Relative entropy and the Wigner-Yanase-Dyson-Lieb concavity in an interpolation theory". Communications in Mathematical Physics. 54 (1): 21–32. Bibcode:1977CMaPh..54...21U. doi:10.1007/BF01609834. ISSN   0010-3616. S2CID   15800519.
  20. Lindblad, Göran (1975). "Completely positive maps and entropy inequalities". Communications in Mathematical Physics. Springer Science and Business Media LLC. 40 (2): 147–151. Bibcode:1975CMaPh..40..147L. doi:10.1007/bf01609396. ISSN   0010-3616. S2CID   121650206.
  21. 1 2 Lieb, E. H. (1975). "Some Convexity and Subadditivity Properties of Entropy". Bull. Am. Math. Soc. 81: 1–13. doi: 10.1090/s0002-9904-1975-13621-4 .
  22. Ruskai, Mary Beth (2002). "Inequalities for quantum entropy: A review with conditions for equality". Journal of Mathematical Physics. AIP Publishing. 43 (9): 4358–4375. arXiv: quant-ph/0205064 . Bibcode:2002JMP....43.4358R. doi:10.1063/1.1497701. ISSN   0022-2488. S2CID   3051292. erratum 46, 019901 (2005)
  23. 1 2 Petz, Dénes (1986). "Sufficient subalgebras and the relative entropy of states of a von Neumann algebra". Communications in Mathematical Physics. Springer Science and Business Media LLC. 105 (1): 123–131. Bibcode:1986CMaPh.105..123P. doi:10.1007/bf01212345. ISSN   0010-3616. S2CID   18836173.
  24. D. Petz, Sufficiency of Channels over von Neumann Algebras, Quart. J. Math. Oxford 35, 475–483 (1986).
  25. P. Hayden, R. Jozsa, D. Petz, A. Winter, Structure of States which Satisfy Strong Subadditivity of Quantum Entropy with Equality, Comm. Math. Phys. 246, 359–374 (2003).
  26. I. Kim, Operator Extension of Strong Subadditivity of Entropy, arXiv : 1210.5190 (2012).
  27. Effros, E. G. (2009). "A Matrix Convexity Approach to Some Celebrated Quantum Inequalities". Proc. Natl. Acad. Sci. USA. 106 (4): 1006–1008. arXiv: 0802.1234 . Bibcode:2009PNAS..106.1006E. doi: 10.1073/pnas.0807965106 . PMC   2633548 . PMID   19164582.
  28. M. B. Ruskai, Remarks on Kim’s Strong Subadditivity Matrix Inequality: Extensions and Equality Conditions, arXiv : 1211.0049 (2012).
  29. O. Fawzi, R. Renner. Quantum conditional mutual information and approximate Markov chains. Communications in Mathematical Physics: 340, 2 (2015)
  30. M. M. Wilde. Recoverability in quantum information theory. Proceedings of the Royal Society A, vol. 471, no. 2182, page 20150338 October 2015
  31. Marius Junge, Renato Renner, David Sutter, Mark M. Wilde, Andreas Winter. Universal recovery maps and approximate sufficiency of quantum relative entropy. Annales Henri Poincare, vol. 19, no. 10, pages 2955--2978, October 2018 arXiv : 1509.07127
  32. Carlen, Eric A.; Vershynina, Anna (2017-10-06). "Recovery map stability for the Data Processing Inequality". arXiv: 1710.02409 [math.OA].