Skip to main content

Advertisement

Log in

Quantum Reverse Hypercontractivity: Its Tensorization and Application to Strong Converses

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

Abstract

In this paper we develop the theory of quantum reverse hypercontractivity inequalities and show how they can be derived from log-Sobolev inequalities. Next we prove a generalization of the Stroock–Varopoulos inequality in the non-commutative setting which allows us to derive quantum hypercontractivity and reverse hypercontractivity inequalities solely from 2-log-Sobolev and 1-log-Sobolev inequalities respectively. We then prove some tensorization-type results providing us with tools to prove hypercontractivity and reverse hypercontractivity not only for certain quantum superoperators but also for their tensor powers. Finally as an application of these results, we generalize a recent technique for proving strong converse bounds in information theory via reverse hypercontractivity inequalities to the quantum setting. We prove strong converse bounds for the problems of quantum hypothesis testing and classical-quantum channel coding based on the quantum reverse hypercontractivity inequalities that we derive.

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.

Institutional subscriptions

Fig. 1

Similar content being viewed by others

Notes

  1. For sake of brevity, we refrain from defining the phrases shown in italics throughout this introduction. Please refer to the main text and references therein for details.

  2. The test could be probabilistic, but for simplicity of presentation we restrict to deterministic tests.

  3. In the case \(p=0\), we define \(\hat{p}=0\) (see e.g., Definition 1.2 of [36]).

  4. Our entropy function here is different from the one in [26] by a factor of p. This modification ensures us that if X and \(\sigma \) commute, we get the usual entropy function in the classical case. Moreover, this extra factor makes the entropy function non-negative even for \(p<0\).

  5. By Brouwer’s fixed-point theorem, \(\Phi _1^*\), has a fixed point in \(\mathcal {D}(\mathcal {H})\) because it maps this compact convex set to itself. On the other hand, since \(\Phi _t^*= (\Phi _1^*)^t\), any fixed point of \(\Phi _1^*\) is an invariant of the whole semigroup. Thus \(\{\Phi _t^*:\, t\ge 0\}\) always has an invariant state in \(\mathcal {D}(\mathcal {H})\).

  6. Again, our definition of the Dirichlet form is different from that of [26] by a factor of p/2 and a negative sign.

  7. Note that this result was independently obtained recently in [9] by introducing the notion of a conditional log-Sobolev constant and finding a uniform lower bound on the latter. Moreover, a special case of the above theorem corresponding to \(\sigma \) being the completely mixed state was proved in [38].

  8. This 0-eigenvector is unique since \(\mathcal {L}\) is assumed to be primitive.

  9. What we really need is that the supports of \(\rho \) and \(\sigma \) being the same (and not being the whole \(\mathcal {H}\)) since in this case we may restrict everything to this support.

References

  1. Ahlswede, R., Gacs, P.: Spreading of sets in product spaces and hypercontraction of the Markov operator. Ann. Probab. 4(6), 925–939 (1976)

    MathSciNet  MATH  Google Scholar 

  2. Araki, H.: On an inequality of Lieb and Thirring. Lett. Math. Phys. 19(2), 167–170 (1990)

    ADS  MathSciNet  MATH  Google Scholar 

  3. Bardet, I.: Estimating the decoherence time using non-commutative Functional Inequalities. arXiv preprint arXiv:1710.01039 (2017)

  4. Bardet, I., Rouzé, C.: Hypercontractivity and logarithmic Sobolev inequality for non-primitive quantum Markov semigroups and estimation of decoherence rates. arXiv preprint: arXiv:1803.05379 (2018)

  5. Beigi, S.: Sandwiched Rényi divergence satisfies data processing inequality. J. Math. Phys. 54, 122202 (2013)

    ADS  MathSciNet  MATH  Google Scholar 

  6. Beigi, S., King, C.: Hypercontractivity and the logarithmic Sobolev inequality for the completely bounded norm. J. Math. Phys. 57(1), 015206 (2016)

    ADS  MathSciNet  MATH  Google Scholar 

  7. Bhatia, R.: Positive Definite Matrices. Princeton Series in Applied Mathematics. Princeton University Press, Princeton (2015)

    Google Scholar 

  8. Boucheron, S., Lugosi, G., Massart, P.: Concentration Inequalities: A Nonasymptotic Theory of Independence. Oxford University Press, Oxford (2013)

    MATH  Google Scholar 

  9. Capel, A., Lucia, A., Pérez-García, D.: Quantum conditional relative entropy and quasi-factorization of the relative entropy. J. Phys. A Math. Theor. 51(48), 484001 (2018)

    MathSciNet  MATH  Google Scholar 

  10. Carbone, R., Martinelli, A.: Logarithmic Sobolev inequalities in non-commutative algebras. Infinite Dimens. Anal. Quantum Probab. Relat. Top. 18(02), 1550011 (2015)

    ADS  MathSciNet  MATH  Google Scholar 

  11. Carlen, E.A., Maas, J.: Gradient flow and entropy inequalities for quantum markov semigroups with detailed balance. J. Funct. Anal. 273(5), 1810–1869 (2017)

    MathSciNet  MATH  Google Scholar 

  12. Cheng, H.-C., Datta, N., Rouzé, C.: Strong converse bounds in quantum network information theory: distributed hypothesis testing and source coding. arXiv preprint: arXiv:1905.00873 (2019)

  13. Cheng, H.-C., Datta, N., Rouzé, C.: Strong converse for classical-quantum degraded broadcast channels. arXiv preprint arXiv:1905.00874 (2019)

  14. Cubitt, T., Kastoryano, M., Montanaro, A., Temme, K.: Quantum reverse hypercontractivity. J. Math. Phys. 56(10), 102204 (2015)

    ADS  MathSciNet  MATH  Google Scholar 

  15. de Wolf, R.: A brief introduction to Fourier analysis on the Boolean cube. Theory Comput. 1, 1–20 (2008)

    Google Scholar 

  16. Delgosha, P., Beigi, S.: Impossibility of local state transformation via hypercontractivity. Commun. Math. Phys. 332(1), 449–476 (2014)

    ADS  MathSciNet  MATH  Google Scholar 

  17. Devetak, I., Junge, M., King, C., Ruskai, M.B.: Multiplicativity of completely bounded \(p\)-norms implies a new additivity result. Commun. Math. Phys. 266(1), 37–63 (2006)

    ADS  MathSciNet  MATH  Google Scholar 

  18. Diaconis, P., Saloff-Coste, L.: Logarithmic Sobolev inequalities for finite Markov chains. Ann. Appl. Probab. 6(3), 695–750 (1996)

    MathSciNet  MATH  Google Scholar 

  19. Frank, R.L., Lieb, E.H.: Monotonicity of a relative Rényi entropy. J. Math. Phys. 54, 122201 (2013)

    ADS  MathSciNet  MATH  Google Scholar 

  20. Gorini, V., Kossakowski, A., Sudarshan, E.C.G.: Complete positive dynamical semigroups of N-level systems. J. Math. Phys. 17(1976), 821 (1976)

    ADS  Google Scholar 

  21. Gozlan, N., Leonard, C.: Transport inequalities. A survey. Markov Process. Relat. Fields 16, 635–736 (2010)

    MathSciNet  MATH  Google Scholar 

  22. Gross, L.: Logarithmic Sobolev inequalities. Am. J. Math. 97(4), 1061–1083 (1975)

    MathSciNet  MATH  Google Scholar 

  23. Hiai, F., Petz, D.: The proper formula for relative entropy and its asymptotics in quantum probability. Commun. Math. Phys. 143(1), 99–114 (1991)

    ADS  MathSciNet  MATH  Google Scholar 

  24. Holevo, A.S.: The capacity of the quantum channel with general signal states. IEEE Trans. Inf. Theory 44(1), 269–273 (1998)

    MathSciNet  MATH  Google Scholar 

  25. Kamath, S., Anantharam, V.: Non-interactive simulation of joint distributions: the Hirschfeld–Gebelein–Rényi maximal correlation and the hypercontractivity ribbon. In: Proceedings of 50th Annual Allerton Conference on Communication, Control, and Computing, pp. 1057–1064 (2012)

  26. Kastoryano, M.J., Temme, K.: Quantum logarithmic Sobolev inequalities and rapid mixing. J. Math. Phys. 54(5), 052202 (2013)

    ADS  MathSciNet  MATH  Google Scholar 

  27. King, C.: Inequalities for trace norms of \(2\times 2\) block matrices. Commun. Math. Phys. 242(3), 531–545 (2003)

    ADS  MATH  Google Scholar 

  28. King, C.: Hypercontractivity for semigroups of unital qubit channels. Commun. Math. Phys. 328(1), 285–301 (2014)

    ADS  MathSciNet  MATH  Google Scholar 

  29. Lieb, E.H., Thirring, W.E.: Inequalities for the moments of the eigenvalues of the Schrodinger Hamiltonian and their relation to Sobolev inequalities. In: Thirring, W. (ed.) The Stability of Matter: From Atoms to Stars, pp. 135–169. Springer, Berlin (1991)

    Google Scholar 

  30. Lindblad, G.: On the generators of quantum dynamical semigroups. Commun. Math. Phys. 48(2), 119–130 (1976)

    ADS  MathSciNet  MATH  Google Scholar 

  31. Liu, J., van Handel, R., Verdú, S.: Beyond the blowing-up lemma: sharp converses via reverse hypercontractivity. In: 2017 IEEE International Symposium on Information Theory (ISIT), pp. 943–947 (June 2017)

  32. Montanaro, A.: Some applications of hypercontractive inequalities in quantum information theory. J. Math. Phys. 53(12), 1–18 (2012)

    MathSciNet  MATH  Google Scholar 

  33. Montanaro, A., Osborne, T.J.: Quantum boolean functions. Chic. J. Theor. Comput. Sci. (2010). https://doi.org/10.4086/cjtcs.2010.001

    Article  MathSciNet  MATH  Google Scholar 

  34. Mosonyi, M., Ogawa, T.: Quantum hypothesis testing and the operational interpretation of the quantum Rényi relative entropies. Commun. Math. Phys. 334(3), 1617–1648 (2015)

    ADS  MATH  Google Scholar 

  35. Mosonyi, M., Ogawa, T.: Strong converse exponent for classical-quantum channel coding. Commun. Math. Phys. 355(1), 373–426 (2017)

    ADS  MathSciNet  MATH  Google Scholar 

  36. Mossel, E., Oleszkiewicz, K., Sen, A.: On reverse hypercontractivity. Geom. Funct. Anal. 23(3), 1062–1097 (2013)

    MathSciNet  MATH  Google Scholar 

  37. Müler-Lennert, M., Dupuis, F., Szehr, O., Fehr, S., Tomamichel, M.: On quantum Rényi entropies: a new generalization and some properties. J. Math. Phys. 54(12), 122203 (2013)

    ADS  MathSciNet  MATH  Google Scholar 

  38. Müller-Hermes, A., França, D.S., Wolf, M.M.: Relative entropy convergence for depolarizing channels. J. Math. Phys. 57(2), 022202 (2016)

    ADS  MathSciNet  MATH  Google Scholar 

  39. Müller-Hermes, A., Stilck Franca, D., Wolf, M.M.: Entropy production of doubly stochastic quantum channels. J. Math. Phys. 57(2), 022203 (2016)

    ADS  MathSciNet  MATH  Google Scholar 

  40. Nelson, E.: A quartic interaction in two dimensions. In: Goodman, R. (ed.) Mathematical Theory of Elementary Particles, pp. 69–73. Springer, Berlin (1966)

    Google Scholar 

  41. Ogawa, T., Nagaoka, H.: Strong converse to the quantum channel coding theorem. IEEE Trans. Inf. Theory 45(7), 2486–2489 (1999)

    MathSciNet  MATH  Google Scholar 

  42. Ogawa, T., Nagaoka, H.: Strong converse and Stein’s lemma in quantum hypothesis testing. IEEE Trans. Inf. Theory 46(7), 2428–2433 (2000)

    MathSciNet  MATH  Google Scholar 

  43. Olkiewicz, R., Zegarlinski, B.: Hypercontractivity in noncommutative \(L_p\) spaces. J. Funct. Anal. 161(1), 246–285 (1999)

    MathSciNet  MATH  Google Scholar 

  44. Petz, D.: A variational expression for the relative entropy. Commun. Math. Phys. 114(2), 345–349 (1988)

    ADS  MathSciNet  MATH  Google Scholar 

  45. Raginsky, M., Sason, I.: Concentration of measure inequalities in information theory, communications, and coding. Found. Trends® Commun. Inf. Theory 10(1–2), 1–246 (2013)

    MATH  Google Scholar 

  46. Schumacher, B., Westmoreland, M.D.: Sending classical information via noisy quantum channels. Phys. Rev. A 56(1), 131 (1997)

    ADS  Google Scholar 

  47. Shor, P.W.: Additivity of the classical capacity of entanglement-breaking quantum channels. J. Math. Phys. 43(9), 4334–4340 (2002)

    ADS  MathSciNet  MATH  Google Scholar 

  48. Simon, B., Hoegh-Krohn, R.: Hypercontractive semigroups and two dimensional self-coupled bose fields. J. Funct. Anal. 9(2), 121–180 (1972)

    MathSciNet  MATH  Google Scholar 

  49. Temme, K., Pastawski, F., Kastoryano, M.J.: Hypercontractivity of quasi-free quantum semigroups. J. Phys. A Math. Gen. 47, 5303 (2014)

    MathSciNet  MATH  Google Scholar 

  50. Tomamichel, M., Berta, M., Hayashi, M.: Relating different quantum generalizations of the conditional Rényi entropy. J. Math. Phys. 55(8), 082206 (2014)

    ADS  MathSciNet  MATH  Google Scholar 

  51. Watrous, J.: The Theory of Quantum Information. Cambridge University Press, Cambridge (2018)

    MATH  Google Scholar 

  52. Wilde, M.M., Winter, A., Yang, D.: Strong converse for the classical capacity of entanglement-breaking channels. Commun. Math. Phys. 331(2), 593–622 (2014)

    ADS  MATH  Google Scholar 

  53. Winter, A.: Coding theorem and strong converse for quantum channels. IEEE Trans. Inf. Theory 45(7), 2481–2485 (1999)

    MathSciNet  MATH  Google Scholar 

  54. Wolf, M.M.: Quantum Channels & Operations: Guided Tour. http://www-m5.ma.tum.de/foswiki/pub/M5/Allgemeines/MichaelWolf/QChannelLecture.pdf (2012). Lecture Notes Based on a Course Given at the Niels-Bohr Institute

Download references

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Salman Beigi.

Additional information

Communicated by M. M. Wolf

Publisher's Note

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

Appendices

Appendix

Proof of Proposition 7

(i) As mentioned in [16] (and explicitly worked out in [5]) for \(p\ge 1\), contractivity can be proven using the Riesz–Thorin interpolation theorem. So we focus on \(p\in (-\infty , -1]\cup [1/2, 1)\). First let \(p=-q\in (-\infty , -1]\), and \(X> 0\). We note that

$$\begin{aligned} \Vert \Phi _t(X)\Vert _{p, \sigma } = \Vert \Phi _t(X)^{-1}\Vert _{q, \sigma }^{-1}. \end{aligned}$$

On the other hand, \(\Phi _t\) is completely positive and unital, and \(z\mapsto z^{-1}\) is operator convex. Therefore, by operator Jensen’s inequality \(\Phi _t(X^{-1})\ge \Phi _t(X)^{-1}\) and by the monotonicity of the norm we have \(\Vert \Phi _t(X)^{-1}\Vert _{q,\sigma } \le \Vert \Phi _t(X^{-1})\Vert _{q,\sigma }\). We conclude that

$$\begin{aligned} \Vert \Phi _t(X)\Vert _{p, \sigma } \ge \Vert \Phi _t(X^{-1})\Vert _{q,\sigma }^{-1} \ge \Vert X^{-1}\Vert _{q, \sigma }^{-1} = \Vert X\Vert _{p, \sigma }, \end{aligned}$$

where for the second inequality we use q-contractivity of \(\Phi _t\) for \(q\ge 1\).

Now suppose that \(p\in [1/2, 1)\). We note that its Hölder conjugate \({\hat{p}}\in (-\infty , -1]\), and that \(\Phi _t\) is reverse \({\hat{p}}\)-contractive. Then using Hölder’s duality, for \(X>0\) we have

$$\begin{aligned} \Vert \Phi _t(X)\Vert _p&= \inf _{Y>0: \Vert Y\Vert _{{\hat{p}}, \sigma }\ge 1} \langle Y, \Phi _t(X)\rangle _\sigma \\&= \inf _{Y>0: \Vert Y\Vert _{{\hat{p}}, \sigma }\ge 1} \langle \widehat{\Phi }_t(Y), X\rangle _\sigma \\&\ge \inf _{Z>0: \Vert Z\Vert _{{\hat{p}}, \sigma }\ge 1} \langle Z, X\rangle _\sigma \\&= \Vert X\Vert _{p, \sigma }, \end{aligned}$$

where \(\widehat{\Phi }_t\) is the adjoint of \(\Phi _t\) with respect to \(\langle .,.\rangle _\sigma \), for each \(t\ge 0\). Here the first equality follows from Lemma 6, and the inequality follows from the \({\hat{p}}\)-contractivity of \(\Phi _t\), i.e, \(\Vert \Phi _t(Y)\Vert _{{\hat{p}}, \sigma } \ge \Vert Y\Vert _{{\hat{p}}, \sigma } \ge 1\).

(ii) As worked out in [14] this is an immediate consequence of the operator Jensen inequality.

Second Proof of Theorem 14

The proof is very similar to the one used in [3] to prove the strong \(L_p\)-regularity of the Dirichlet forms. Before stating the proof we need some definitions.

For a compact set I we let C(I) to be the Banach space of continuous, complex valued functions on I (equipped with the supremum norm). Then the Banach space \(C(I\times I)\) becomes a \(*\)-algebra when endowed with the natural involution \(f\mapsto f^*\) with \(f^*(x,y)=\overline{f(x,y)}\). Thus \(C(I\times I)\) is a \(C^*\)-algebra.

We endow \(\mathcal {B}(\mathcal {H})\) with a Hilbert space structure by equipping it with the Hilbert–Schmidt inner product:

$$\begin{aligned} \langle X, Y\rangle _{\text {HS}}:= \text {tr}(X^\dagger Y). \end{aligned}$$

Fix \(X,Y\in \mathcal {B}_{sa}(\mathcal {H})\), and let I be a compact interval containing the spectrum of both X and Y. We define a \(*\)-representation \(\pi _{X,Y}: C(I\times I)\rightarrow \mathcal {B}\big (\mathcal {B}(\mathcal {H})\big )\) that is uniquely determined by its action on tensor products of functions as follows. For \(f, g\in C(I)\) we define \(\pi _{X, Y}(f\otimes g)\in \mathcal {B}\big ( \mathcal {B}(\mathcal {H}) \big )\) by

$$\begin{aligned} \pi _{X,Y}(f\otimes g) (Z)=f(X) Zg(Y),\qquad Z\in \mathcal {B}(\mathcal {H}). \end{aligned}$$

The following lemma can be found in [3] (see Lemma 4.2):

Lemma 34

\(\pi _{XY}\) is a \(*\)-representation between \(C^*\)-algebras. That is,

  1. (i)

    \(\pi _{XY}(1)=\mathcal {I}\), where 1 is the constant function on \(I\times I\) equal to 1.

  2. (ii)

    \(\pi _{XY}(f^*g)=\pi _{XY}(f)^*\pi _{XY}(g)\) for all \(f,g\in C(I\times I)\).

  3. (iii)

    If \(f\in C(I\times I)\), is a non-negative function, then \(\pi _{XY}(f)\) is a positive semi-definite operator on \(\mathcal {B}(\mathcal {H})\) for the Hilbert–Schmidt inner product, i.e., \(\pi _{X, Y}(f)\in \mathcal {P}\big ( \mathcal {B}(\mathcal {H}) \big )\).

Now, for any function \(f\in C(I)\), define \({\tilde{f}}\) to be the function in \(C(I\times I)\) defined by

$$\begin{aligned} {\tilde{f}}(s,t)=\left\{ \begin{aligned}&\frac{f(s)-f(t)}{s-t}\qquad s\ne t\\&f'(s)\qquad ~~~~~~~~~ s=t. \end{aligned}\right. \end{aligned}$$
(55)

The following lemma, proved in [3] (see Lemma 4.2), gives a generalization of the chain rule formula to a derivation.

Lemma 35

Let \(X, Y\in \mathcal {B}_{sa}(\mathcal {H})\) and let I be a compact interval containing the spectrums of XY. Let \(f\in C(I)\) be a continuously differentiable function such that \(f(0)=0\). Then for all \(V\in \mathcal {B}(\mathcal {H})\) we have

$$\begin{aligned} Vf(Y)-f(X)V=\pi _{XY}({\tilde{f}})(VY-XV), \end{aligned}$$

where \({\tilde{f}}\) is defined by (55).

We can now prove the theorem. By the result of [11] (an extension of Lemma 13), there are superoperators \(\partial _j:\mathcal {B}(\mathcal {H})\rightarrow \mathcal {B}(\mathcal {H})\) of the form

$$\begin{aligned} \partial _j(X)= [V_j, X] = V_j X- XV_j, \end{aligned}$$

where \(V_j\in \mathcal {B}(\mathcal {H})\), such that

$$\begin{aligned} \langle X,\mathcal {L}(Y)\rangle _\sigma = \sum _{j}\langle \partial _jX,\partial _j Y\rangle _{\sigma }. \end{aligned}$$
(56)

Moreover, \(V_j\)’s are such that there are \(\omega _j\ge 0\) with

$$\begin{aligned} \sigma V_j = \omega _j V_j \sigma . \end{aligned}$$

Using the above equation one can show [3] that

$$\begin{aligned}&\partial _j\big (I_{q,p}(X)\big )=\Gamma _\sigma ^{-\frac{1}{q}}\bigg (V_j \Big (\Gamma _{\sigma }^{\frac{1}{p}}\big (\omega _j^{-\frac{1}{2p}}X\big )\Big )^{\frac{p}{q}}-\Big (\Gamma ^{\frac{1}{p}}_\sigma \big (\omega _j^{\frac{1}{2p}}X\big )\Big )^{\frac{p}{q}}V_j\bigg ). \end{aligned}$$
(57)

For arbitrary \(X> 0\) define \(Y_j:= \omega _j^{-1/4}\, \Gamma _{\sigma }^{\frac{1}{2}}(X)\) and \({Z}_j:= \omega _j^{1/4}\,\Gamma _{\sigma }^{\frac{1}{2}}(X)\). Using (57) we compute

$$\begin{aligned} \mathcal {E}_{q,\mathcal {L}}\big (I_{q,2}(X)\big )&=\frac{q{\hat{q}}}{4}\big \langle I_{\hat{q},q}\big (I_{q,2}(X)\big ),\mathcal {L}\big (I_{q,2}(X)\big )\big \rangle _\sigma \nonumber \\&=\frac{q{\hat{q}}}{4}\big \langle I_{\hat{q},2}(X),\mathcal {L}\big (I_{q,2}(X)\big )\big \rangle _\sigma \nonumber \\&=\frac{q{\hat{q}}}{4} \sum _{j} \langle \partial _j I_{\hat{q},2}(X),\partial _j I_{q,2}(X)\rangle _\sigma \end{aligned}$$
(58)
$$\begin{aligned}&=\frac{q{\hat{q}}}{4} \sum _{j} \Big \langle \Gamma _\sigma ^{-\frac{1}{\hat{q}}} \Big (V_j {Y_j}^{2/\hat{q}}-{Z}_j^{2/\hat{q}}V_j\Big ), \Gamma ^{-\frac{1}{q}} \Big (V_jY_j^{2/q}-{Z}_j^{2/q}V_j\Big )\Big \rangle _\sigma \end{aligned}$$
(59)
$$\begin{aligned}&=\frac{q{\hat{q}}}{4} \sum _{j} \Big \langle V_j {Y_j}^{2/\hat{q}}-{Z}_j^{2/\hat{q}}V_j, V_j{Y}_j^{2/q}-{Z}_j^{2/q}V_j\Big \rangle _{\text {HS}}\nonumber \\&=\frac{q{\hat{q}}}{4}\sum _{j} \Big \langle \pi _{{Z}_j,{Y}_j}\big ({\tilde{f}}_{2/\hat{q}}\big )(V_j{Y}_j-{Z}_jV_j),\pi _{{Z}_j,{Y}_j}\big ({\tilde{f}}_{2/q}\big )(V_j{Y}_j-{Z}_jV_j)\Big \rangle _{\text {HS}}\end{aligned}$$
(60)
$$\begin{aligned}&=\frac{q{\hat{q}}}{4}\sum _{j} \Big \langle V_j{Y}_j-{Z}_jV_j, \pi _{{Z}_j,{Y}_j}\big ({\tilde{f}}_{2/\hat{q}}\big )^*\pi _{{Z}_j,{Y}_j}\big ({\tilde{f}}_{2/q}\big )(V_j{Y}_j-{Z}_jV_j)\Big \rangle _{\text {HS}}\nonumber \\&= \frac{q{\hat{q}}}{4} \sum _{j} \Big \langle V_j {Y}_j-Z_jV_j,\pi _{Z_j,Y_j}\big ({\tilde{f}}_{2/\hat{q}}^*{\tilde{f}}_{2/q}\big )(V_jY_j-Z_jV_j)\Big \rangle _{\text {HS}}, \end{aligned}$$
(61)

where in (58) we used (56), in (59) we used (57), and in (60) we used the chain rule formula of Lemma 35 for the functions \(f_\alpha \) with \(f_\alpha (x)=x^{\alpha }\). Finally, in (61) we used part (ii) of Lemma 34.

Now, using the proofs of Theorem 2.1 and Lemma 2.4 of [36], for any \(x,y\ge 0\) and \(0\le p\le q\le 2\) we have

$$\begin{aligned} q\hat{q}(x^{1/\hat{q}}-y^{1/\hat{q}})(x^{1/q}-y^{1/q})\le p\hat{p}(x^{1/\hat{p}}-y^{1/\hat{p}})(x^{1/p}-y^{1/p}). \end{aligned}$$
(62)

This means that for all xy we have

$$\begin{aligned} q{\hat{q}} \big ({\tilde{f}}_{2/\hat{q}}^*{\tilde{f}}_{2/q}\big )(x, y) \le p{\hat{p}} \big ({\tilde{f}}_{2/\hat{p}}^*{\tilde{f}}_{2/p}\big )(x, y). \end{aligned}$$

Hence, by part (iii) of Lemma 34 we have

$$\begin{aligned} \mathcal {E}_{q,\mathcal {L}}(I_{q,2}(X))&\le \frac{p \hat{p}}{4} \sum _{j} \Big \langle V_j Y_j-Z_jV_j,\pi _{Z_j,Y_j}({\tilde{f}}_{2/\hat{p}}^*{\tilde{f}}_{2/p})(V_jY_j-Z_jV_j)\Big \rangle _{\text {HS}}\\&=\mathcal {E}_{p,\mathcal {L}}(I_{p,2}(X)). \end{aligned}$$

Remark 9

The difference with the proof of \(L_p\)-regularity of [3] lies in the choice of the inequality (62) used at the end of the proof.

Proof of Theorem 25

Since both \(\text {Ent}_{2, \sigma }(X)\) and \(\mathcal {E}_{2, \mathcal {L}}(X)\) are homogenous of degree two in X, to prove a log-Sobolev inequality, without loss of generality we can assume that X is of the form \(X=\Gamma _\sigma ^{-1/2}(\sqrt{\rho })\) where \(\rho \) is a density matrix. In this case

$$\begin{aligned} \text {Ent}_{2, \sigma }(X) = D(\rho \Vert \sigma ), \qquad \langle X, \mathcal {L}X\rangle _\sigma = 1- \big [\text {tr}\big (\sqrt{\sigma }\sqrt{\rho }\big )\big ]^2. \end{aligned}$$

Let \(\sigma = \sum _{i=1}^d s_i|i\rangle \langle i|\) and \(\rho =\sum _{k=1}^d r_k |{{\tilde{k}}}\rangle \langle {{\tilde{k}}}|\) be the eigen-decompositions of \(\sigma \) and \(\rho \). Then

$$\begin{aligned} \text {Ent}_{2, \sigma }(X) = \sum _{k=1}^d r_k \log r_k - \sum _{i, k=1}^d |\langle i| {{\tilde{k}}}\rangle |^2r_k \log s_i, \end{aligned}$$

and

$$\begin{aligned} \langle X, \mathcal {L}X\rangle _\sigma = 1- \Big ( \sum _{i, k=1}^d |\langle i| {{\tilde{k}}}\rangle |^2 \sqrt{s_i r_k} \Big )^2. \end{aligned}$$

Let \(A=(a_{ik})_{d\times d}\) be a \(d\times d\) matrix whose entries are given by

$$\begin{aligned} a_{ik} = |\langle i| {{\tilde{k}}}\rangle |^2. \end{aligned}$$

Observe that, fixing the eigenvalues \(s_i\)’s and \(r_k\)’s, the entropy \(\text {Ent}_{2, \sigma }(X)\) is a linear function of A and \(\mathcal {E}_{2, \mathcal {L}}(X)\) is concave function of A. On the other hand, since both \(\{|1\rangle , \dots , |d\rangle \}\) and \(\{|{{\tilde{1}}}\rangle , \dots , |{{\tilde{d}}}\rangle \}\) form orthonormal bases, A is a doubly stochastic matrix. Then by Birkhoff’s theorem, A can be written as a convex combination of permutations matrices. We conclude that if an inequality of the form

$$\begin{aligned} \beta \Big (\sum _{k=1}^d r_k \log r_k - \sum _{i, k=1}^d a_{ik}r_k \log s_i\Big )\le 1- \Big ( \sum _{i, k=1}^d a_{ik} \sqrt{s_i r_k} \Big )^2, \end{aligned}$$

holds for all permutation matrices A, then it holds for all doubly stochastic A, and then for all \(\sigma , \rho \) with the given eigenvalues. We note that A is a permutation matrix when \(\{|1\rangle , \dots , |d\rangle \}\) and \(\{|{{\tilde{1}}}\rangle , \dots , |{{\tilde{d}}}\rangle \}\) are the same bases (up to some permutation) which means that \(\sigma \) and \(\rho \) commute. Therefore, a log-Sobolev inequality of the form

$$\begin{aligned} \beta \text {Ent}_{2, \sigma } \big ( \Gamma _{\sigma }^{-1/2}(\rho ) \big ) \le \mathcal {E}_{2, \mathcal {L}}\big (\Gamma _\sigma ^{-1/2}(\rho )\big ), \end{aligned}$$

holds for all \(\rho \) if and only if it holds for all \(\rho \) that commute with \(\sigma \). That is, to find the log-Sobolev constant

$$\begin{aligned} \alpha _2(\mathcal {L}) = \inf _{\rho } \frac{\mathcal {E}_{2, \mathcal {L}}\big (\Gamma _\sigma ^{-1/2}(\rho )\big )}{\text {Ent}_{2, \sigma } \big ( \Gamma _{\sigma }^{-1/2}(\rho ) \big ) }, \end{aligned}$$

we may restrict to those \(\rho \) that commute with \(\sigma \). This optimization problem over such \(\rho \) is equivalent to computing the 2-log-Sobolev constant of the classical simple Lindblad generator, and has been solved in Theorem A.1 of [18]. \(\quad \square \)

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Beigi, S., Datta, N. & Rouzé, C. Quantum Reverse Hypercontractivity: Its Tensorization and Application to Strong Converses. Commun. Math. Phys. 376, 753–794 (2020). https://doi.org/10.1007/s00220-020-03750-z

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s00220-020-03750-z

Navigation