Hostname: page-component-745bb68f8f-d8cs5 Total loading time: 0 Render date: 2025-02-06T09:34:58.420Z Has data issue: false hasContentIssue false

Complete positivity order and relative entropy decay

Published online by Cambridge University Press:  06 February 2025

Li Gao*
Affiliation:
School of Mathematics and Statistics Wuhan University, Wuhan, Hubei 430072, P.R.China
Marius Junge
Affiliation:
Department of Mathematics University of Illinois at Urbana-Champaign, Urbana, IL 61801, USA; E-mail: mjunge@illinois.edu
Nicholas LaRacuente
Affiliation:
Department of Computer Science Indiana University, Bloomington, IN 47408, USA; E-mail: nick.laracuente@gmail.com
Haojian Li
Affiliation:
Zentrum Mathematik Technische Universität München, Garching, 85748, Germany; E-mail: lihaojianmath@gmail.com
*
E-mail: gao.li@whu.edu.cn (corresponding author)

Abstract

We prove that for a GNS-symmetric quantum Markov semigroup, the complete modified logarithmic Sobolev constant is bounded by the inverse of its complete positivity mixing time. For classical Markov semigroups, this gives a short proof that every sub-Laplacian of a Hörmander system on a compact manifold satisfies a modified log-Sobolev inequality uniformly for scalar and matrix-valued functions. For quantum Markov semigroups, we show that the complete modified logarithmic Sobolev constant is comparable to the spectral gap up to the logarithm of the dimension. Such estimates are asymptotically tight for a quantum birth-death process. Our results, along with the consequence of concentration inequalities, are applicable to GNS-symmetric semigroups on general von Neumann algebras.

Type
Analysis
Creative Commons
Creative Common License - CCCreative Common License - BY
This is an Open Access article, distributed under the terms of the Creative Commons Attribution licence (https://creativecommons.org/licenses/by/4.0/), which permits unrestricted re-use, distribution, and reproduction in any medium, provided the original work is properly cited.
Copyright
© The Author(s), 2025. Published by Cambridge University Press

1 Introduction

The time evolution of dynamical systems is a central topic in ergodic theory, probability theory, geometry and analysis. Similarly, decay properties of dissipative quantum systems also naturally arise in quantum many-body systems, quantum information theory and high energy physics. The aim of this article is to provide a new framework of decay estimates that applies for both classical and quantum systems in the non-ergodic setting. Here, ergodicity means the system admits a unique equilibrium state, also termed primitive in mathematical physics literature, whereas non-ergodic systems admit multiple equilibrium states.

Logarithmic Sobolev inequality (LSI) is a powerful functional inequality in deriving the mixing time of Markovian evolution. LSI was first introduced in the seminal works of Gross [Reference Gross35, Reference Gross34] as an equivalent reformulation of Nelson’s Hypercontractivity (HC) [Reference Nelson58, Reference Nelson59]. It has been widely studied on manifolds and graphs for the deep connections to geometry and concentration phenomenon. However, attempts to translate the notion of hypercontractivity to the matrix-valued setting or the non-ergodic setting failed miserably [Reference Bardet and Rouzé7], due to the lack of uniform convexity of certain noncommutative spaces [Reference Junge and Parcet42]. This results in a roadblock for the standard argument connecting hypercontractivity, entropy decay and mixing time, as well as the lack of tensorization property used in many-body systems.

We propose a new, direct approach to entropy decay that also applies to fully noncommutative, non-ergodic setting. Let $T_t=e^{-Lt}$ be a quantum Markov semigroup on a finite von Neumann algebra ${\mathcal M}$ with generator L (i.e., a semigroup of completely positive trace-preserving maps). We aim to establish the exponential entropy decay,

(1.1) $$ \begin{align} D(T_t(\rho)\|E(\rho)) \hspace{.1cm} \le \hspace{.1cm} e^{-2\alpha_1 t} D(\rho\|E(\rho)) \end{align} $$
$$ \begin{align*} \text{ or equivalently } \hspace{.1cm}\hspace{.1cm}\hspace{.1cm}\hspace{.1cm}\hspace{.1cm}\hspace{.1cm}\hspace{.1cm}\hspace{.1cm}\hspace{.1cm} 2\alpha_1 & D(\rho)\|E(\rho))\le \tau\Big(L(\rho)(\ln\rho-\ln E(\rho))\Big), \end{align*} $$

where $D(\rho \|\sigma ) \hspace {.1cm} = \hspace {.1cm} \tau (\rho \ln \rho -\rho \ln \sigma )$ is the quantum relative entropy for two density operators $\rho ,\sigma $ and $\tau $ can be any normal faithful trace on ${\mathcal M}$ . The equilibrium state $E(\rho )$ associated to any initial density $\rho $ is given by the ergodic mean $\displaystyle E(\rho )=\lim _{t\to \infty }\frac {1}{t}\int _0^t T_s(\rho ) ds$ . It turns out that the simple properties of relative entropy enable us to prove a direct link between positivity order and entropy decay. Indeed, let us for simplicity assume that the semigroup is trace symmetric

$$\begin{align*}\tau(T_t(x)y)=\tau(xT_t(y))\hspace{.1cm} \text{ for }\hspace{.1cm} x,y\in{\mathcal M}, t\ge 0.\end{align*}$$

Under this assumption, we discover the following entropy difference lemma:

(1.2) $$ \begin{align} D(\rho\|T_{2t}(\sigma))\hspace{.1cm} \le \hspace{.1cm} D_{T_t}(\rho) + D(\rho\|\sigma) , \quad \text{ where }\quad D_{\Phi}(\rho) :\hspace{.1cm} = \hspace{.1cm} \tau(\rho\ln \rho)-\tau(\Phi(\rho)\ln \Phi(\rho)). \end{align} $$

The new quantity $D_{\Phi }(\rho )$ is the loss of von Neumann entropy under a channel map $\Phi $ . Our second ingredient is a stability estimate inspired by the positivity order condition by Gao and Rouzé [Reference Gao and Rouzé31] (see also [Reference LaRacuente46]) that

(1.3) $$ \begin{align} (1-\varepsilon)E(x) \hspace{.1cm} \le \hspace{.1cm} T_t(x) \hspace{.1cm} \le \hspace{.1cm} (1+\varepsilon) E(x) , \hspace{.1cm} \forall\hspace{.1cm} x\ge 0\Longrightarrow D(\rho\|E(\rho)) \hspace{.1cm} \le \hspace{.1cm} C_\varepsilon D(\rho\|T_t(\rho)) \end{align} $$

for some constant $C_{\varepsilon }$ only depending on $\varepsilon $ and the index of the ergodic mean projection E. Now, suppose the condition (1.3) holds for time $t(\varepsilon )$ and find

$$ \begin{align*} D(\rho\|E(\rho))\le C_\varepsilon D(\rho\|T_{t(\varepsilon)}(\rho)) \le C_\varepsilon \bigg( D_{\frac{T_{t(\varepsilon)}}{2n}}(\rho)+ D(\rho\|T_{\frac{(n-1)t(\varepsilon)}{n}}(\rho))\bigg)\hspace{.1cm} \le \hspace{.1cm} n C_{\varepsilon} D_{T_{t(\varepsilon)/2n}}(\rho) , \end{align*} $$

where we apply (1.2) iteratively to the term $D(\rho \|T_{\frac {(n-1)t(\varepsilon )}{n}}(\rho ))$ . Taking the limit $n\to \infty $ , we derive the inequality

$$ \begin{align*}D\big(\rho\|E(\rho)\big) \hspace{.1cm} \le \hspace{.1cm} \frac{t(\varepsilon)}{2} C_{\varepsilon} \tau(L(\rho)\ln \rho),\end{align*} $$

which is the differential version of (1.1) with $\alpha _1= \frac {1}{C_\varepsilon t(\varepsilon )}$ , called the modified logarithmic Sobolev inequality (in short, MLSI). The largest possible constant $\alpha _1$ in (1.1) is called the MLSI constant.

1.1 MLSI for GNS-symmetric semigroups

Many dynamics in quantum information processing are not trace symmetric. One major application of open systems is state preparation by simulating time evolution governed by a Lindbladian

(1.4) $$ \begin{align} \textstyle L(x) \hspace{.1cm} = \hspace{.1cm} i[h,x]+\sum_j 2a_j^*xa_j-a_j^*a_jx-xa_j^*a_j. \end{align} $$

A natural one is the Davies semigroup, which converges to the thermal Gibbs state $\phi =\frac {e^{-\beta H}}{{\text {tr}}(e^{-\beta H} )}$ . For any finite inverse temperature $\beta>0$ , the Davies semigroup is never trace symmetric but satisfies the following detailed balance condition

$$\begin{align*}\phi(T_t(x)y)\hspace{.1cm} = \hspace{.1cm} \phi(xT_t(y)), \hspace{.1cm} \forall\hspace{.1cm} x,y\end{align*}$$

with respect to the Gibbs state $\phi $ , which we call GNS-symmetry. In this context, a breakthrough result of MLSI constant $\alpha _1$ was made by Gao and Rouzé [Reference Gao and Rouzé31] that

(1.5) $$ \begin{align} \alpha_1\ge\frac{\lambda(L)}{C(E)} \end{align} $$

for every GNS symmetric semigroups in finite dimensions. Here, $\lambda (L)$ is the spectral gap of the semigroup generator L, and $C(E)=\inf \{\mu \hspace {.1cm} |\hspace {.1cm} x\le \mu E (x)\hspace {.1cm}, \hspace {.1cm} \text { for all } x\ge 0 \}$ is the Pimsner-Popa index for the condition expectation $\displaystyle E=\lim _{t\to \infty }T_t$ . An important consequence of Gao and Rouzé’s estimate (1.5) is the positivity of the complete MLSI constant $\alpha _c(L)=\inf _{n}\alpha _1(L\otimes \operatorname {\mathrm {\operatorname {id}}}_{\mathbb {M}_n})$ (in short, CMLSI constant),

(1.6) $$ \begin{align} \alpha_c\ge\frac{\lambda}{C_{cb}(E)}>0 ,\end{align} $$

because the complete Pimsner-Popa index $C_{cb}(E)=\sup _{n}C(E\otimes \operatorname {\mathrm {\operatorname {id}}}_{\mathbb {M}_n})$ is finite in finite dimensions. The CMLSI constant is of particular interest because it satisfies the tensorization property $\alpha _c(T_t\otimes S_t)=\min \{\alpha _c(T_t), \alpha _c(S_t)\}$ , while the MLSI constant $\alpha _1$ does not.

Our ‘positivity order implies entropy decay’ argument above gives an exponential improvement to (1.6) in terms of the dimension constant $C_{cb}(E)$ .

Theorem 1.1 (cf. Theorem 3.2 and 4.10).

Let $T_t:{\mathcal M}\to {\mathcal M}$ be a quantum Markov semigroup GNS-symmetric to a faithful normal state $\phi $ . Then the optimal CMLSI constant satisfies

$$\begin{align*}\alpha_{c}\hspace{.1cm} \ge \hspace{.1cm} \frac{1}{2t_{cb}(0.1)}, \text{ where } \hspace{.1cm} t_{cb}(\varepsilon):=\inf\{t>0 \hspace{.1cm} |\hspace{.1cm} (1-\varepsilon) E\le_{cp} T_t\le_{cp} (1+\varepsilon)E \}. \end{align*}$$

Here, $\Psi \le _{cp}\Phi $ means $\Psi -\Phi $ is a completely positive map. Moreover, in finite dimensions,

(1.7) $$ \begin{align}\alpha_1\ge \alpha_{c}\ge \frac{\lambda}{2\ln (10 C_{cb}(E))}.\end{align} $$

The quantity $t_{cb}$ , called CB return time, is the mixing time in terms of complete positivity order. Similar terms of complete positivity have been also considered in the quantum setting for the study of approximate unitary t-design ([Reference Brandao, Harrow and Horodecki12]). In the fully non-ergodic noncommutative setting, $t_{cb}$ was originally introduced in [Reference Gao, Junge and LaRacuente29] via completely bounded (CB) $L_1\to L_\infty $ norm, whose connection to complete positivity order and spectral gap relies heavily on operator space theory (see Section 3.2).

The proof to GNS-symmetric cases uses the ideas of Haagerup reduction [Reference Haagerup, Junge and Xu36], a method to derive results for type III von Neumann algebras by reducing them to cases of tracial von Neumann algebras. Thanks to this machinery, our estimate in trace-symmetric settings can be salvaged to a GNS-symmetric semigroup on general $\sigma $ -finite von Neumann algebras, including both classical systems and quantum systems. A particular interesting example is a matrix version of the classical n-level death-birth process which admits an invariant state $\rho _{n}\varpropto (e^{-\beta k})_{k=0,..,n}$ and a Lindbladian given by nearest neighbor interactions. In this example, we show that both the spectral gap is are uniformly controlled, and

$$\begin{align*}\lambda\sim \Theta(1)\hspace{.1cm} , \hspace{.1cm}\alpha_1\sim \alpha_{c}\sim \frac{1}{n}, \hspace{.1cm} t_{cb}\sim \ln(C_{cb}(E))\sim n.\end{align*}$$

Hence, both estimates in our Theorem 1.1 are asymptotically tight for this GNS-symmetric example.

1.2 MLSI for matrix-valued functions

Besides the quantum setting, our results also provide interesting MLSI and concentration inequalities for random matrices of arbitrary size. For a classical Markov semigroup $P_t:L_\infty (\Omega ,\mu )\to L_\infty (\Omega ,\mu )$ on some probability space $(\Omega ,\mu )$ , the notion of CMLSI is basically a uniform MLSI for positive matrix-valued random variables $g:\Omega \to {\mathbb M}_n$ of all dimensions $n\ge 1$ ,

(1.8) $$ \begin{align} \mu\circ {\text{tr}} (g\ln g - E_\mu(g) \ln E_\mu(g)) \leq \frac{1}{2\alpha} \mu \circ {\text{tr}}( (L g) \ln g ). \end{align} $$

Here, $\mu (f)=\int fd\mu $ is the scalar valued mean, $E_\mu (g)=\int gd\mu \in {\mathbb M}_n$ is the matrix valued mean, and ${\text {tr}}$ is the standard matrix trace. In this setting, the CB return time $t_{cb}$ is simply the $L_\infty $ -mixing time

$$\begin{align*}t_{b}(\varepsilon)=\{t>0| \parallel \! T_t-E_\mu:L_1(\Omega)\to L_\infty(\Omega) \! \parallel_{}\le \varepsilon\},\end{align*}$$

which is accessible by kernel estimates derived from harmonic analysis. As a consequence of Theorem 1.1, we obtain CMLSI for all sub-Laplacians of Hörmander system.

Theorem 1.2. Let $(M,g)$ be a compact connected Riemannian manifold without boundary, and ${\omega } d\operatorname {vol}$ be a probability measure with a smooth density $\omega $ with respect to the volume form $d\operatorname {vol}$ . Suppose $H=\{X_i\}_{i=1}^k\subset TM$ is a family of vectors fields satisfying the Hörmander’s condition that at every point $x\in M$ ,

$$\begin{align*}T_xM=\operatorname{span}\{[X_{i_1},[X_{i_2},\cdots, [X_{i_{n-1}}, X_{i_n}]]] \hspace{.1cm} | \hspace{.1cm} 1\leqslant i_1,i_2\cdots i_n\leqslant k , n\ge 1\}. \end{align*}$$

Then the horizontal heat semigroup $P_t=e^{-\Delta _H t}$ generated by the sub-Laplacian

$$\begin{align*}\Delta_H=\sum_{i} X_i^*X_i=-\sum_{i}X_i^2+(\text{div}_{\mu}(X_i)+X_i(\ln \omega))X_i \end{align*}$$

has CMLSI constant $\alpha _c(\Delta _H)>0$ . Here, $X_i^*$ is the adjoint operator with respect to $L_2(M, \omega d\operatorname {vol})$ .

For scalar-valued functions, the positivity of $\alpha _1(\Delta _H)$ was proved by Ługiewicz and Zegarliński [Reference Lugiewicz and Zegarliński53], using a hypercontractive argument similar from [Reference Diaconis and Saloff-Coste23]. Nevertheless, both [Reference Diaconis and Saloff-Coste23] and [Reference Lugiewicz and Zegarliński53] rely on the Rothaus Lemma [Reference Rothaus68, Reference Bakry3], a crucial step which does not apply for matrix-valued functions (see Section 3.6).

In this setting, our Theorem 1.1 gives a short proof of

(1.9) $$ \begin{align} \boxed{\text{Heat kernel estimate}}+ \boxed{\text{Spectral gap}} \Longrightarrow \boxed{\text{LSI/MLSI}} \end{align} $$

for scalar-valued function, and also extends to matrix-valued setting by replacing LSI with CMLSI. A particular interesting example, also covered in [Reference Gao and Gordina28], is the Lie group $M=SU(2)$ with the canonical sub-Laplacian $\Delta _H=-X^2-Y^2$ , where the Lie algebra $\mathfrak {su}(2)$ is spanned by the Pauli matrices $X,Y$ and $Z=\frac {1}{2}[X,Y]$ . The CMLSI of heat semigroups (standard Laplacians) was obtained in [Reference Li, Junge and LaRacuente49, Reference Brannan, Gao and Junge14] using the Ricci curvature lower bound as a crucial tool. Nevertheless, in the sub-elliptic case the Ricci curvature in the degenerate direction of the vector field $H=\{X_i\}_{i=1}^k$ can be interpreted as $-\infty $ . In [Reference Gao and Gordina28], the curvature condition were substituted by a gradient estimate that was first introduced by Driver and Melcher [Reference Driver and Melcher24] for Heisenberg group, later obtained for nilpotent Lie groups [Reference Melcher54] and $SU(2)$ [Reference Baudoin and Bonnefont8]. Our Theorem 1.2 obtains CMLSI for all sub-Laplacian of Hörmander systems, without using any curvature condition. It implies the following uniform CMLSI constant for trace symmetric Lindbladians as ‘representation’ of Hörmander system on Lie groups.

Corollary 1.3. Let G be a compact Lie group and $H=\{X_1,\cdots ,X_k\}$ be a generating set of its Lie algebra $\mathfrak {g}$ . There exists a constant $\alpha _{c}(\Delta _H)>0$ such that for all unitary representation u, the induced quantum Markov semigroup generated by

$$\begin{align*}L_H(\rho)=-\sum_{i=1}^k [d_u(X_i),[d_u(X_i),\rho]]\end{align*}$$

satisfies $\alpha _{c}(L_H)\ge \alpha _{c}(\Delta _H)>0$ . Here, $d_u$ is the Lie algebra homomorphism induced by u.

1.3 Concentration inequalities

An important application of MLSI is to derive concentration inequalities. This was first discovered by Otto and Villani [Reference Otto and Villani61], later extended to the discrete case by Erbar and Maas [Reference Erbar and Maas26], and more recently to the noncommutative setting in [Reference Rouzé and Datta69, Reference Gao, Junge and LaRacuente29, Reference Carlen and Maas16]. As an application of our MLSI estimate for GNS-symmetric semigroups, we derive concentration inequalities for a general faithful invariant state $\phi $ . Recall that the Lipschitz semi-norm

$$\begin{align*}\|x\|_{\text{Lip}} \hspace{.1cm} = \hspace{.1cm}: \max\{ \parallel \! \Gamma_L(x,x) \! \parallel_{}^{\frac{1}{2}}, \hspace{.1cm} \parallel \! \Gamma_{L}(x^*,x^*) \! \parallel_{}^{\frac{1}{2}}\}.\end{align*}$$

The Lipschitz semi-norm is defined through the gradient form (or Carré du Champ operator)

$$\begin{align*}\Gamma_L(x,y) \hspace{.1cm} = \hspace{.1cm} \frac{1}{2}\Big(L(x^*)y+x^*L(y)-L(x^*y)\Big), \hspace{.1cm} \forall x,y\in \text{dom}(L). \end{align*}$$

Theorem 1.4. Let ${\mathcal M}$ be a $\sigma $ -finite von Neumann algebra and let $T_t=e^{-tL}$ be a GNS- $\phi $ -symmetric quantum Markov semigroup with positive MLSI constant $\alpha _{1}(L)>0$ . Then there exists a universal constant c such that for $2\le p <\infty $ ,

$$\begin{align*}\alpha\|x-E(x)\|_{L_p({\mathcal M},\phi)}\hspace{.1cm} \le \hspace{.1cm} c\sqrt{p}\parallel \! x \! \parallel_{\textrm{Lip}} .\end{align*}$$

Moreover, for any $t>0$ , there exists a projection e such that

$$\begin{align*}\|e(x-E(x))e\|_{\infty} \le t \quad \mbox{and} \quad \phi(1-e)\le 2 \exp\left( -\frac{\alpha^2t^2}{16ec^2\parallel \! x \! \parallel_{\textrm{Lip}}^2}\right) .\end{align*}$$

As a special case, we obtain the following matrix concentration inequalities which can be compared to the work of Tropp [Reference Tropp75].

Corollary 1.5. Let $S_1,\cdots ,S_n$ be an independent sequence of random $d\times d$ -matrices such that $\parallel \! S_i-\mathbb {E}S_i \! \parallel _{\infty }\le M \hspace {.1cm} , \hspace {.1cm} a.e.$ Then, we have the matrix Bernstein inequality that for the sum $Z=\sum _{k=1}^nS_k$ ,

(1.10) $$ \begin{align} {\mathbb E}\parallel \! Z-{\mathbb E} Z \! \parallel_{\infty}\le 2ce^{-1/2} \sqrt{ (v(Z)+M^2)\log d}\end{align} $$

and the matrix Chernoff bound

$$\begin{align*}P(|Z-\mathbb{E}Z|>t)\le 2 d\exp\Big( -\frac{t^2}{64ec^2(v(Z)+M^2)}\Big),\end{align*}$$

where

$$\begin{align*}v(Z)=\max \{ \parallel \! {\mathbb E}((Z-\mathbb{E}Z)^*(Z-\mathbb{E}Z)) \! \parallel_{\hspace{.1cm}}, \parallel \! {\mathbb E}((Z-\mathbb{E}Z)^*(Z-\mathbb{E}Z)) \! \parallel_{}\}.\end{align*}$$

In particular, the inequality (1.10) improves the term $M\log d$ in Tropp’s result [Reference Tropp75] to $M\sqrt {\log d}$ . For more details, see Example 5.18.

After the acceptance of this paper, we get to know the sub-Gaussian type estimate (1.10) of matrix concentration was obtained by Huang and Tropp [Reference Huang and Tropp38, Reference Huang and Tropp39] via matrix-valued Poincare inequality and matrix-valued Bakry-Émery curvature condition. Actually, in the introduction of [Reference Huang and Tropp39] they raise the question whether the sub-gaussian estimate can be obtained by matrix-valued Log-Sobolev inequality. Our result answers this question.

1.4 Outline of the paper

We organize our paper as follows to make it accessible for readers from different backgrounds. In Section 2, we provide a brief review of quantum information basics in the setting of tracial von Neumann algebras. We prove our key entropy difference lemma (Lemma 2.1) and an improved data processing inequality (Theorem 2.5). Building upon these results, we discuss the functional inequalities of symmetric quantum Markov semigroups in Section 3. We prove our main Theorem 1.1 in the trace symmetric case and its consequence Theorem 1.2 for classical Markov semigroups. We also illustrate the failure of the matrix-valued logarithmic Sobolev ineuqality in Proposition 3.15. The discussion up to this point does not involve much technicality beyond the basic concepts of finite von Neumann algebras. Readers from quantum information and classical analysis are welcome to consider examples such as the matrix algebra ${\mathbb M}_n$ and matrix-valued functions $L_\infty (\Omega , {\mathbb M}_n)$ .

In Section 4, we dive into the GNS-symmetric cases. Here, we discuss the Haagerup reduction for channels and entropic quantities, deriving Theorem 1.1 (Theorem 4.10 and Corollary 4.13) in its full generality. Section 5 collects applications of our general results Theorem 1.4 and Corollary 1.5. We conclude the paper in Section 6 with some discussions on remaining open questions.

Notations. We use calligraphic letters ${\mathcal M},{\mathcal N}$ for von Neumann algebras and denote ${\mathbb M}_n$ as the algebra of $n\times n$ as complex matrices. We use $\tau $ as the trace on von Neumann algebra, and ${\text {tr}}$ as the standard matrix trace. The identity operator is denoted by $1$ , and the identity map between spaces is denoted as $\operatorname {\mathrm {\operatorname {id}}}$ , sometimes specified with subscript like $1_{\mathcal M}$ and $\operatorname {\mathrm {\operatorname {id}}}_{{\mathbb M}_n}$ . We write $ a^*$ as the adjoint element of a and $\Phi _*$ for a pre-adjoint map of $\Phi $ .

2 Entropy contraction of symmetric Markov maps

2.1 States, channels and entropies

We briefly review some basic information-theoretic concepts in the noncommutative setting. Recall that a von Neumann algebra ${\mathcal M}$ is a unital $*$ -subalgebra of $B(H)$ closed under weak $^*$ -topology. A linear functional $\phi :{\mathcal M}\to \mathbb {C}$ is called a state if it is positive, meaning $\phi (x^*x)\ge 0$ for any $x\in {\mathcal M}$ , and additionally, $\phi (1)=1$ . We say $\phi $ is normal if $\phi $ is weak $^*$ -continuous. Throughout the paper, we will only consider normal states and denote $S({\mathcal M})$ as the normal state space of ${\mathcal M}$ . We write $s(\phi )$ as the support projection of a state $\phi $ , which is the minimal projection e such that $\phi (x)=\phi (exe)\hspace {.1cm}, \forall \hspace {.1cm} x\in {\mathcal M}$ . A normal state $\phi $ is faithful if $s(\phi )=1$ . For two normal states $\rho $ and $\sigma $ , the relative entropy is defined as

(2.1) $$ \begin{align}D(\rho||\sigma)=\begin{cases} \langle \xi_\rho| \log\Delta(\rho/\sigma) |\xi_\rho \rangle, & \mbox{if } s(\rho)\le s(\sigma)\\ +\infty, & \mbox{otherwise}. \end{cases},\end{align} $$

where $\xi _\rho $ is a vector implementing the state $\rho $ , and $\Delta (\rho /\sigma )$ is the relative modular operator. This form of definition (2.1) was introduced by Araki [Reference Araki2] for general von Neumann algebras.

In this section, we will focus on the case that ${\mathcal M}$ is a finite von Neumann algebra. Namely, ${\mathcal M}$ is equipped with a normal faithful tracial state $\tau $ . The tracial noncommutative $L_p$ -space $L_p({\mathcal M},\tau )$ is defined as the completion of ${\mathcal M}$ with respect to the p-norm $\parallel \! a \! \parallel _{p}=\tau (|a|^{p})^{1/p}$ . We identify $L_\infty ({\mathcal M})\cong {\mathcal M}$ , and also $ L_1({\mathcal M})\cong {\mathcal M}_*$ via the trace duality

$$ \begin{align*}d_\phi\in L_1({\mathcal M})\longleftrightarrow \phi\in {\mathcal M}_*,\hspace{.1cm} \phi(x)=\tau(d_\phi x).\end{align*} $$

We say $\rho \in L_1({\mathcal M})$ is a density operator if $\rho \ge 0$ and $\tau (\rho )=1$ , which corresponds to a normal state in the above identification. We will often identify normal states with their density operators if no ambiguity. Via this identification, relative entropy reduces to the original definition of Umegaki [Reference Umegaki76],

$$\begin{align*}D(\rho||\sigma)=\tau(\rho\log \rho-\rho\log \sigma ),\end{align*}$$

provided this trace is well defined. For example, for $\rho $ and $\sigma $ in the bounded state space

$$ \begin{align*} S_{B}({\mathcal M})=\{\rho\in S({\mathcal M})\hspace{.1cm} |\hspace{.1cm} \mu_1 1\le \rho \le \mu_2 1\text{ for some } \mu_1,\mu_2>0 \}, \end{align*} $$

the Umegaki’s formula is always well defined and finite. For this reason, we will mostly work with bounded states from $S_{B}({\mathcal M})$ and derive results for general case $S({\mathcal M})$ by approximation. When the second state $\sigma =1$ , this gives the entropy functional

$$\begin{align*}H(\rho):=D(\rho|| 1)=\tau(\rho\log\rho).\end{align*}$$

Note that the standard convention of von Neumann entropy in quantum information literature is often with an additional negative sign .

We say a linear map $T:{\mathcal M}\to {\mathcal M}$ is a quantum Markov map if T is normal, unital and completely positive. Recall that T is unital if $T( 1)=1$ . The pre-adjoint map $T_*:{\mathcal M}_*\to {\mathcal M}_*$ is called a quantum channel, which sends normal states to normal states. In the tracial setting, $T_*:L_1({\mathcal M})\to L_1({\mathcal M})$ given by

$$\begin{align*}\tau(T_*(\rho)y)=\tau(\rho T(y)), \hspace{.1cm} \forall\hspace{.1cm} y\in {\mathcal M}, \rho\in L_1({\mathcal M}),\end{align*}$$

is completely positive and trace-preserving (in short, CPTP). A fundamental inequality about quantum channel is the data processing inequality (also called monotonicity of relative entropy)

(2.2) $$ \begin{align} D(\rho||\sigma)\ge D(T_*(\rho)||T_*(\sigma)), \hspace{.1cm} \forall \rho,\sigma\in S({\mathcal M}). \end{align} $$

The data processing inequality states that two quantum states cannot become more distinguishable under a quantum channel. The data processing inequality remains valid for T being positive but not necessarily completely positive; see [Reference Müller-Hermes and Reeb56, Reference Frenkel27]. The main technical result of this work is an improved data processing inequality for quantum channels under symmetric conditions (Theorem 2.5).

2.2 Entropy contraction for unital quantum channels

We start our discussion on entropy contraction of unital quantum channels. The restriction of $\Phi $ on ${\mathcal M}$ is bounded and normal; thus, $\Phi $ can be viewed as the $L_1$ -norm extension of its restriction $\Phi :{\mathcal M}\to {\mathcal M}$ . By duality, its adjoint $\Phi ^*:{\mathcal M}\to {\mathcal M}$ is a trace-preserving quantum Markov map and hence also extends to a unital quantum channel.

For a state $\rho $ with $H(\rho )<\infty $ , we define the entropy difference of $\Phi $ ,

$$\begin{align*}D_\Phi(\rho):=H(\rho)-H(\Phi(\rho)).\end{align*}$$

Non-negativity of the entropy difference $D_\Phi (\rho )\ge 0$ follows from data processing inequality (2.2) and $\Phi (1)=1$ ,

$$\begin{align*}H(\rho)=D(\rho||1)\geq D(\Phi(\rho)||\Phi(1))= H(\Phi(\rho)).\end{align*}$$

We start with the key lemma in our argument.

Lemma 2.1 (Entropy difference lemma).

Let $\Phi :L_1({\mathcal M})\to L_1({\mathcal M})$ be a unital quantum channel and $\Phi ^*$ be its adjoint. Then for two bounded states $\rho ,\omega \in {S_B}({\mathcal M})$ ,

$$\begin{align*}D(\rho\|\Phi^*\Phi(\omega)) \le D_\Phi(\rho) + D(\rho\|\omega) \le \tau((\operatorname{\mathrm{\operatorname{id}}}-\Phi^*\Phi)(\rho)\ln \rho ) + D(\rho\|\omega) .\end{align*}$$

Proof. By duality, $\Phi ^*$ is also completely positive unital. Then,

$$ \begin{align*} D(\rho\|\Phi^*\Phi(\omega))=\hspace{.1cm}&\tau(\rho\ln \rho -\rho \ln \Phi^*\Phi(\omega)) \\ =\hspace{.1cm}&\tau(\rho\ln \rho-\Phi(\rho)\log \Phi(\rho))+\tau( \Phi(\rho)\log \Phi(\rho)-\rho \ln \Phi^*\Phi(\omega)) \\ =\hspace{.1cm}&D_\Phi(\rho)+\tau\big( \Phi(\rho)\log \Phi(\rho)-\rho \ln \Phi^*\Phi(\omega)\big) \\ \overset{(1)}{\le}\hspace{.1cm} &D_\Phi(\rho)+\tau\Big( \Phi(\rho)\log \Phi(\rho)-\rho \Phi^* \big(\ln \Phi(\omega)\big)\Big) \\ = \hspace{.1cm}&D_\Phi(\rho)+\tau\Big( \Phi(\rho)\log \Phi(\rho)-\Phi(\rho)\ln \Phi(\omega)\Big) \\ =\hspace{.1cm} &D_\Phi(\rho)+D(\Phi(\rho)\|\Phi(\omega)) \\ \overset{(2)}{\le}\hspace{.1cm} &D_\Phi(\rho)+D(\rho\|\omega), \end{align*} $$

where (2) follows from the monotonicity of relative entropy. The inequality (1) uses the operator concavity [Reference Choi18] of logarithm function $t\mapsto \ln t$ that for any positive operator $x\ge 0$ ,

$$\begin{align*}\Phi^* (\ln x)\le \ln \Phi^*(x).\end{align*}$$

This proves the first inequality. For the second part, it suffices to notice that

$$ \begin{align*} D_\Phi(\rho)=&\tau(\rho\log \rho-\Phi(\rho)\log \Phi(\rho)) \le \tau(\rho\log \rho-\Phi(\rho)\Phi(\log \rho)) =\tau(\rho\log \rho-\Phi^*\Phi(\rho)\log \rho), \end{align*} $$

where we use the operator concavity $\Phi (\ln x)\le \ln \Phi (x)$ again.

We iterate the above lemma as follows:

$$ \begin{align*}D(\rho|| (\Phi^* \Phi)^n(\rho))\le\hspace{.1cm} D_\Phi(\rho)+D(\rho|| (\Phi^* \Phi)^{n-1}(\rho)) \le\hspace{.1cm} n D_\Phi(\rho)+D(\rho|| \rho)=n D_\Phi(\rho). \end{align*} $$

Then a relevant question is what would be the limit of $(\Phi ^* \Phi )^n(\rho )$ as $n \to \infty $ . This leads to the multiplicative domain of $\Phi $ . Recall that the multiplicative domain of a unital completely positive map $\Phi $ is

$$\begin{align*}{\mathcal N}_\Phi=\{ x\in{\mathcal M} \ |\ \Phi(y) \Phi(x)=\Phi(yx),\hspace{.1cm} \Phi(x) \Phi(y)= \Phi(xy),\forall y\in\mathcal{M}\}.\end{align*}$$

When $\Phi $ is normal, ${\mathcal N}_\Phi \subset \mathcal {M}$ is a von Neumann subalgebra ([Reference Luczak52, Theorem 1]). A linear map ${E:{\mathcal M}\to {\mathcal M}}$ is called a conditional expectation if E is a unital completely positive map and idempotent $E\circ E=E$ . When ${\mathcal M}$ is a finite von Neumann algebra, for any subalgebra ${\mathcal N}\subset {\mathcal M}$ , there always exists a (unique) trace-preserving conditional expectation E onto ${\mathcal N}$ such that

(2.3) $$ \begin{align}\tau(xy)=\tau(xE(y)), \hspace{.1cm} x\in {\mathcal N}, y\in {\mathcal M}. \end{align} $$

Such E is a unital quantum channel.

Proposition 2.2. Let $\Phi :L_1({\mathcal M})\to L_1({\mathcal M})$ be a unital quantum channel, and let $E:{\mathcal M}\to {\mathcal N}$ be the trace-preserving conditional expectation onto the multiplicative domain ${\mathcal N}:={\mathcal N}_\Phi $ . Then

  1. i) $\Phi :{\mathcal N}\to \Phi ({\mathcal N})$ is a $*$ -isomorphism with inverse $\Phi ^*:\Phi ({\mathcal N})\to {\mathcal N}$ . Moreover, $\Phi ({\mathcal N})$ is the multiplicative domain for $\Phi ^*$ , and

    (2.4) $$ \begin{align} (\Phi^* \Phi)\circ E=E\circ (\Phi^* \Phi)=E, \hspace{.1cm} E_0\circ \Phi=\Phi\circ E, \end{align} $$
    where $E_0:{\mathcal M}\to \Phi ({\mathcal N})$ is the trace-preserving conditional expectation onto $\Phi ({\mathcal N})$ .
  2. ii) $\Phi $ is an isometry on $L_2({\mathcal N})$ . If, in addition, $\parallel \! \Phi (\operatorname {\mathrm {\operatorname {id}}}-E):L_2({\mathcal M})\to L_2({\mathcal M}) \! \parallel _{2}<1$ , then ${E=\lim _{n}(\Phi ^* \Phi )^n}$ as a map from $L_2({\mathcal M})$ to $L_2({\mathcal M})$ .

Proof. It is clear that $\Phi $ is a $*$ -homomorphism on ${\mathcal N}$ . For any $x,y\in L_2({\mathcal N})\subset L_2({\mathcal M})$ ,

$$\begin{align*}\tau( y(\Phi^*\circ \Phi)(x))=\tau( \Phi(y) \Phi(x))=\tau(\Phi(xy))=\tau(xy).\end{align*}$$

Thus, $\Phi ^*\circ \Phi |_{{\mathcal N}}=\operatorname {\mathrm {\operatorname {id}}}_{\mathcal N}$ is the identity map. This verifies $(\Phi ^* \Phi )\circ E=E$ . Since $E^*=E$ , $E\circ (\Phi ^* \Phi )=E$ follows from taking the adjoint. Thus, $\Phi :{\mathcal N}\to \Phi ({\mathcal N})$ is a $*$ -isomorphism with inverse $\Phi ^*$ . Denoting ${\mathcal N}_0$ as the multiplicative domain for $\Phi ^*$ , we have $\Phi ({\mathcal N})\subset {\mathcal N}_0$ . Conversely, we also have $\Phi ^*(N_0)\subset {\mathcal N}$ by switching the role of $\Phi =(\Phi ^*)_*$ . Then $\Phi ({\mathcal N})={\mathcal N}_0$ since $\Phi $ is bijective on ${\mathcal N}$ . For ii), we note that by (2.4),

$$ \begin{align*}(\operatorname{\mathrm{\operatorname{id}}}-E)\Phi^* \Phi(\operatorname{\mathrm{\operatorname{id}}}-E)=(\operatorname{\mathrm{\operatorname{id}}}-E)(\Phi^* \Phi-E)=\Phi^* \Phi-E, \hspace{.1cm} &\hspace{.1cm} (\Phi^* \Phi-E)^n= (\Phi^* \Phi)^n-E. \end{align*} $$

Therefore,

$$ \begin{align*}&\parallel \! \Phi^* \Phi-E:L_2({\mathcal M})\to L_2({\mathcal M}) \! \parallel_{}=\parallel \! \Phi(\operatorname{\mathrm{\operatorname{id}}}-E) \! \parallel_{2}^2<1, \\ &\parallel \! (\Phi^* \Phi)^n-E:L_2({\mathcal M})\to L_2({\mathcal M}) \! \parallel_{}=\parallel \! (\Phi^* \Phi-E)^n:L_2({\mathcal M})\to L_2({\mathcal M}) \! \parallel_{}=\parallel \! \Phi(\operatorname{\mathrm{\operatorname{id}}}-E) \! \parallel_{2}^{2n}, \end{align*} $$

which goes to $0$ as $n\to \infty $ .

In order to estimate entropic quantities, we will use the approximation in terms of complete positivity. Recall that for a density operator $\sigma \in S({\mathcal M})$ with full support, the Bogoliubov-Kubo-Mori (BKM) metric for $X\in {\mathcal M}$ is defined by

$$ \begin{align*}\gamma_{\sigma}(X):=\int_{0}^\infty \tau(X^*(\sigma+s)^{-1}X(\sigma+s)^{-1})ds.\end{align*} $$

The BKM metric is a Riemannian metric on the space of states with full support that is monotone under any quantum channel $\Psi $ ,

$$\begin{align*}\gamma_{\Psi(\sigma)}(\Psi(X))\le \gamma_{\sigma}(X), \forall X\in {\mathcal M}. \end{align*}$$

It connects to the relative entropy as follows ([Reference Gao and Rouzé31, Lemma 2.2]):

(2.5) $$ \begin{align} D(\rho||\sigma)=\int_0^1 \int_{0}^s \gamma_{\rho_t}(\rho-\sigma)dtds=\int_0^1 (1-t) \gamma_{\rho_t}(\rho-\sigma)dt, \end{align} $$

where $\rho _t=t\rho +(1-t) \sigma $ for $t\in [0,1]$ . It is proved in [Reference Gao and Rouzé31, Lemma 2.1 & 2.2] that if $\rho \le c\sigma $ ,

(2.6) $$ \begin{align} &c\gamma_{\rho}(X)\le \gamma_{\sigma}(X), \hspace{.1cm} \forall X\in {\mathcal M} \nonumber\\ &k(c) \gamma_{\sigma}(\rho-\sigma) \le D(\rho||\sigma) \le \gamma_{\sigma}(\rho-\sigma), \end{align} $$

where $k(c)=\frac {c\ln c-c+1}{(c-1)^2}$ . The above discussion remains valid if $s(\rho )\le s(\sigma )$ and $X\in s(\sigma ){\mathcal M} s(\sigma )$ . For two positive maps $\Psi $ and $\Phi $ , we write $\Phi \le \Psi $ if $\Psi -\Phi $ is positive.

Lemma 2.3. Let E be a conditional expectation (not necessarily trace-preserving) and $\Psi $ be a quantum Markov map such that

$$\begin{align*}(1-\varepsilon)E\le \Psi \le (1+\varepsilon)E.\end{align*}$$

Assume that $E\circ \Psi =E$ . Then for any $\rho \in S({\mathcal M})$ ,

$$\begin{align*}D(\rho||\Psi_*(\rho))\ge \Big(\frac{1-\varepsilon}{1+\varepsilon}-\frac{\varepsilon}{(1-\varepsilon)k(2)}\Big) D(\rho||E_*(\rho)).\end{align*}$$

In particular, for $\varepsilon =\frac {1}{10}$ ,

$$\begin{align*}D(\rho||\Psi_*(\rho))\ge \frac{1}{2} D(\rho||E_*(\rho)).\end{align*}$$

Proof. By assumption, $\Psi _*=(1-\varepsilon )E_*+\varepsilon \Psi _0$ for some unital positive map $\Psi _0\le 2E_*$ . We denote $\sigma =E_*(\rho ), \tilde {\sigma }=\Phi _*(\rho )$ and $\omega =\Psi _0(\rho )$ . Then $\tilde {\sigma }=(1-\varepsilon )\sigma +\varepsilon \omega $ . Note that for any bounded state $\sigma \in S_B({\mathcal M})$ , $X\mapsto \sqrt {\gamma _\sigma (X)}$ is a Hilbert space norm. Then by the triangle inequality,

$$ \begin{align*}\sqrt{\gamma(\rho-\tilde{\sigma})}=\ &\sqrt{\gamma(\rho-(1-\varepsilon)\sigma-\varepsilon \omega)}\\ =\ &\sqrt{\gamma((\rho-\sigma)+\varepsilon (\sigma-\omega))} \\ \ge\ &\sqrt{\gamma(\rho-\sigma)}-\varepsilon\sqrt{\gamma(\sigma-\omega),} \end{align*} $$

where $\gamma $ can be $\gamma _\phi $ for any bounded state $\phi \in S_B({\mathcal M})$ . Then

$$ \begin{align*}\gamma(\rho-\tilde{\sigma}) \ge\ &\gamma(\rho-\sigma)-2\varepsilon\sqrt{\gamma(\rho-\sigma)}\sqrt{\gamma(\sigma-\omega)}+\varepsilon^2\gamma(\sigma-\omega)\\ \ge\ &\gamma(\rho-\sigma)-2\varepsilon\sqrt{\gamma(\rho-\sigma)}\sqrt{\gamma(\sigma-\omega)}\\ \ge\ &\gamma(\rho-\sigma)-\varepsilon\gamma(\rho-\sigma)-\varepsilon\gamma(\sigma-\omega)\\ =\ &(1-\varepsilon)\gamma(\rho-\sigma)-\varepsilon\gamma(\sigma-\omega). \end{align*} $$

Now take $\rho _t=t\rho +(1-t)\sigma $ and $\tilde {\rho }_t=t\rho +(1-t)\tilde {\sigma }$ ,

$$ \begin{align*} D(\rho||\tilde{\sigma})&=\int_0^1 (1-t)\gamma_{\tilde{\rho}_t}(\rho-\tilde{\sigma})dt \\ &\ge (1-\varepsilon)\int_0^1 (1-t) \gamma_{\tilde{\rho}_t}(\rho-\sigma)dt-\varepsilon\int_0^1(1-t) \gamma_{\tilde{\rho}_t}(\sigma-\omega)dt. \end{align*} $$

For the first term, because $\tilde {\rho }_t\le (1+\varepsilon )\rho _t$ ,

$$ \begin{align*} \int_0^1 (1-t) \gamma_{\tilde{\rho}_t}(\rho-\sigma)dt \ge (1+\varepsilon)^{-1}\int_0^1 (1-t) \gamma_{\rho_t}(\rho-\sigma)dtds=(1+\varepsilon)^{-1}D(\rho||\sigma). \end{align*} $$

For the second term, consider that $\tilde {\rho }_t\ge (1-\varepsilon )(1-t)\sigma $ ,

$$ \begin{align*} \int_0^1 (1-t)\gamma_{\tilde{\rho}_t}(\sigma-\omega)dt \le\ & \frac{1}{(1-\varepsilon)}\int_0^1 \gamma_{\sigma}(\omega-\sigma)dt\\ =\ &\frac{1}{(1-\varepsilon)}\gamma_{\sigma}(\omega-\sigma)\\ \overset{(1)}{\le}\ & \frac{1}{(1-\varepsilon)k(2)}D(\omega||\sigma) \\ =\ & \frac{1}{(1-\varepsilon)k(2)}D(\Psi_*(\rho)||\sigma)\overset{(2)}{\le} \frac{1}{(1-\varepsilon)k(2)}D(\rho||\sigma). \end{align*} $$

Here, the inequality (1) above uses $\omega \le 2\sigma $ and (2.6). The inequality (2) above follows from the monotonicity of relative entropy and the fact $\Psi _*(\sigma )=\sigma $ . Combining the estimated above, we obtained

$$\begin{align*}D(\rho||\tilde{\sigma})\ge \frac{1-\varepsilon}{1+\varepsilon} D(\rho||\sigma)-\varepsilon k(2)^{-1}D(\rho||\sigma)=\Big(\frac{1-\varepsilon}{1+\varepsilon}-\frac{\varepsilon}{(1-\varepsilon)k(2)}\Big)D(\rho||\sigma) ,\end{align*}$$

where $k(2)=2\ln 2-1$ . The above inequality is nontrivial for $\varepsilon $ such that

$$\begin{align*}\frac{1-\varepsilon}{1+\varepsilon}-\frac{\varepsilon}{(1-\varepsilon)k(2)}>0. \end{align*}$$

Taking $\varepsilon =0.1$ , the above expression is approximately $0.53>\frac {1}{2}$ .

Remark 2.4. This lemma is related to [Reference LaRacuente47, Corollary 2.15] and is a variant of [Reference Gao and Rouzé31, Theorem 5.3], which proves for GNS symmetric $\Phi $ ,

(2.7) $$ \begin{align}D(\rho||(\Phi_*)^2(\rho))\ge (1-\varepsilon^2 k(2)^{-1}) D(\rho||E_*(\rho)).\end{align} $$

Compared with [Reference LaRacuente47, Corollary 2.15], the above Lemma assumes a simpler condition and may achieve a stronger constant in certain regimes of interest. The above Lemma improves (2.7) from two points: 1) does not need any symmetric assumption; 2) remove the square in $\Phi _*^2$ . When $\Psi _*=\Phi _*^2$ is a square, (2.7) could yield better bound that for $\varepsilon =0.4$ ,

$$\begin{align*}(1-(0.4)^2 k(2)^{-1})>\frac{1}{2}>\frac{1}{4}> \frac{1-0.4^2}{1+0.4^2}-\frac{0.4^2}{(1-0.4^2)k(2)}.\end{align*}$$

Putting the above lemma together, we obtain the following entropy contraction of unital quantum channels.

Theorem 2.5. Let $\Phi $ be a unital quantum channel and let $E:{\mathcal M}\to {\mathcal N}$ be the trace-preserving conditional expectation onto the multiplicative domain $\mathcal {N}$ of $\Phi $ . Define the CB return time

(2.8) $$ \begin{align} k_{cb}(\Phi):=\inf \{k\in \mathbb{N}^+\hspace{.1cm} |\hspace{.1cm} 0.9 E \le_{cp} (\Phi^*\Phi)^{k}\le_{cp} 1.1 E \} \ . \end{align} $$

Then for any state $\rho \in S({\mathcal M})$ ,

(2.9) $$ \begin{align} D(\Phi(\rho)||\Phi\circ E(\rho))\le \Big (1-\frac{1}{2k_{cb}(\Phi)} \Big ) D(\rho||E(\rho))\ .\end{align} $$

Furthermore, for any finite von Neumann algebra $\mathcal {Q}$ and state $\rho \in S({\mathcal M} \overline {\otimes }\mathcal {Q}),$

(2.10) $$ \begin{align} D( \Phi\otimes \operatorname{\mathrm{\operatorname{id}}} (\rho)||(\Phi\circ E)\otimes \operatorname{\mathrm{\operatorname{id}}} (\rho))\le \Big (1-\frac{1}{2k_{cb}(\Phi)} \Big ) D(\rho|| E\otimes \operatorname{\mathrm{\operatorname{id}}} (\rho)).\end{align} $$

Proof. It suffices to consider a bounded state $\rho \in S_B({\mathcal M})$ . Note that by the conditional expectation property (2.3),

$$ \begin{align*}&D(\rho||E(\rho))=\tau(\rho\log \rho-\rho\log E(\rho))=\tau(\rho\log \rho)-\tau(E(\rho)\log E(\rho))=H(\rho)-H(E(\rho)),\\ &D(\Phi(\rho)||\Phi\circ E(\rho))\kern1.5pt{=}\kern1.5pt D(\Phi(\rho)||E_0\circ \Phi(\rho))\kern1.5pt{=}\kern1.5pt H(\Phi(\rho))\kern1.5pt{-}\kern1.5pt H(E_0 \circ \Phi(\rho)) \kern1.5pt{=}\kern1.5pt H(\Phi(\rho))\kern1.5pt{-}\kern1.5pt H(\Phi\circ E(\rho)), \end{align*} $$

where we used the property $\Phi \circ E=E_0 \circ \Phi $ from Proposition 2.2. Moreover, $H(E(\rho ))=H(\Phi \circ E(\rho ))$ as $\Phi $ is a trace-preserving $*$ -isomorphism on ${\mathcal N}$ . Thus, we have

$$\begin{align*}D_\Phi(\rho)=H(\rho)-H(\Phi(\rho))=D(\rho||E(\rho))-D(\Phi(\rho)||\Phi \circ E(\rho)).\end{align*}$$

Iterating the entropy difference Lemma 2.1, we have

$$ \begin{align*}D(\rho\|(\Phi^*\Phi)^{k}(\rho)) \hspace{.1cm} \le\hspace{.1cm} & D_\Phi(\rho) + D(\rho\|(\Phi^*\Phi)^{k-1}(\rho))\\ \le\hspace{.1cm} & k D_\Phi(\rho) + D(\rho\|\rho)\\ = \hspace{.1cm}& k (D(\rho||E(\rho))-D(\Phi(\rho)||\Phi \circ E(\rho)). \end{align*} $$

Now using Lemma 2.3, for $k=k_{cb}(\Phi )$ ,

$$\begin{align*}D(\rho||E(\rho))\le 2 D(\rho\|(\Phi^*\Phi)^{k}\rho))\le 2k\big (D(\rho||E(\rho))-D(\Phi(\rho)||\Phi \circ E(\rho))\big). \end{align*}$$

Rearranging the terms gives the assertion. The general case $\rho \in S({\mathcal M})$ can be obtained via approximation $\rho _\varepsilon =(1-\varepsilon )\rho +\varepsilon 1$ as [Reference Brannan, Gao and Junge14, Lemma A.2]. The same argument applies to $\operatorname {\mathrm {\operatorname {id}}}_{\mathcal Q}\otimes \Phi $ , because the CB return time $k_{cb}(\operatorname {\mathrm {\operatorname {id}}}_{\mathcal Q}\otimes \Phi )=k_{cb}(\Phi )$ is same as of $\Phi $ by the definition.

The above theorem is an improved data processing inequality for the relative entropy between a state $\rho $ and its conditional expectation $E(\rho )$ . Here, ${\mathcal N}$ is the ‘decoherence free’ subalgebra. Indeed, for any two states $\sigma _1,\sigma _2\in {\mathcal N},$

$$\begin{align*}D(\sigma_1||\sigma_2)\ge D(\Phi(\sigma_1)||\Phi(\sigma_2))\ge D(\Phi^*\Phi(\sigma_1)||\Phi^*\Phi(\sigma_2))=D(\sigma_1||\sigma_2)\end{align*}$$

does not decay. Outside the ‘decoherence free’ subalgebra ${\mathcal N}$ , the relative entropy from a state $\rho $ to its projection $E(\rho )$ on ${\mathcal N}$ is strictly contractive under every use of the channel $\Phi $ .

For $\Phi $ being a symmetric quantum Markov map, we have $\Phi =\Phi _*$ . Moreover, Proposition 2.2 reduces to

$$\begin{align*}\Phi\circ E=E\circ \Phi\hspace{.1cm} , \hspace{.1cm} \Phi^2\circ E= E\circ \Phi^2=E.\end{align*}$$

Then

$$ \begin{align*} D(\Phi^{2}(\rho)\|E(\rho))&=D(\Phi^{2}(\rho)\|\Phi^{2}\circ E(\rho))=D(\Phi^{2}(\rho)\|\Phi\circ E\circ \Phi (\rho))\\ &\leq (1-\frac{1}{2k_{cb}(\Phi)})D(\Phi(\rho)\|E\circ\Phi(\rho))=(1-\frac{1}{2k_{cb}(\Phi)})D(\Phi(\rho)\|\Phi\circ E(\rho))\\ &\leq (1-\frac{1}{2k_{cb}(\Phi)})^{2}D(\rho\|E(\rho)). \end{align*} $$

We can iterate the entropy contraction above and obtain the discrete time entropy decay,

$$ \begin{align*} D(\Phi^{2n}(\rho)|| E(\rho))\le (1-\frac{1}{2k_{cb}(\Phi)})^{2n} D(\rho||E(\rho)). \end{align*} $$

3 Complete modified log-Sobolev inequality for symmetric Markov semigroups

3.1 Functional inequalities

In this section, we discuss a continuous time relative entropy decay for symmetric quantum Markov semigroups. We first review some basics of quantum Markov semigroups. A quantum Markov semigroup $(T_t)_{t\ge 0}:{\mathcal M}\to {\mathcal M}$ is a family of maps satisfying

  1. i) for each $t\ge 0$ , $T_t$ is a quantum Markov map (i.e., normal, completely positive and unital)

  2. ii) $T_0=\operatorname {\mathrm {\operatorname {id}}}_{\mathcal M}$ and $T_s\circ T_t=T_{s+t}$ for any $s,t \ge 0$ .

  3. iii) for $x\in {\mathcal M}$ , $t\mapsto T_t(x)$ is weak $^*$ -continuous.

The generator of the semigroup is defined as

$$\begin{align*}\hspace{.1cm} Lx=w^*\text{-}\lim_{t\to 0} \frac{1}{t}(x-T_t(x))\hspace{.1cm} \end{align*}$$

on the domain of L that the limit exists. In this section, we still consider ${\mathcal M}$ as a finite von Neumann algebra equipped with a normal faithful tracial state $\tau $ . Given $(T_t)_{t\ge 0}$ is symmetric (or more specifically, trace-symmetric), that is,

$$\begin{align*}\tau(x^*T_t(y))=\tau(T_t(x)^*y)\hspace{.1cm} , \hspace{.1cm} \forall x,y\in {\mathcal M}, t\ge 0,\end{align*}$$

the generator L is a positive, symmetric operator, densely defined on $L_2({\mathcal M})$ . Its kernel is the fixed-point subspace ${\mathcal N}:=\ker (L)=\{x\in {\mathcal M}\hspace {.1cm} | \hspace {.1cm} T_t(x)=x, \forall t\ge 0\}$ , which coincides with the common multiplicative domain of all $T_t$ – hence a von Neumann subalgebra. Moreover, each $T_t$ is an ${\mathcal N}$ -bimodule map

$$\begin{align*}T_t(axb)=aT_t(x)b, \hspace{.1cm} \forall\hspace{.1cm} a,b\in {\mathcal N} ,x\in {\mathcal M}.\end{align*}$$

In particular, we have

$$\begin{align*}T_t\circ E= E\circ T_t=E,\end{align*}$$

where $E:{\mathcal M}\to {\mathcal N}$ is the trace-preserving conditional expectation onto the fixpoint algebra ${\mathcal N}$ . We say $(T_t)$ is ergodic if ${\mathcal N}={\mathbb C} 1$ is trivial. Note that in the mathematical physics literature, it is common to use primitive instead of ergodic. In this case, the semigroup admits a unique invariant state – namely, the trace $\tau $ up to normalization. We will consider symmetric quantum Markov semigroups that are not necessarily ergodic.

Recall that a semigroup is equivalently determined by its Dirichlet form

$$\begin{align*}{\mathcal E}:L_2({\mathcal M})\to [0,\infty]\hspace{.1cm} ,\hspace{.1cm} {\mathcal E}(x,x)=\tau(x^*L x). \end{align*}$$

We write $\operatorname {\mathrm {\operatorname {dom}}} (L)$ for the domain of L and $\operatorname {\mathrm {\operatorname {dom}}} ({\mathcal E})$ for the domain of ${\mathcal E}$ . The Dirichlet subalgebra is defined as ${\mathcal A}_{\mathcal E}:=\operatorname {\mathrm {\operatorname {dom}}} ({\mathcal E})\cap {\mathcal M}$ . It was proved [Reference Davies and Lindsay21] that ${\mathcal A}_{\mathcal E}$ is a dense $*$ -subalgebra of ${\mathcal M}$ and a core of $L^{1/2}$ . We denote by

$$\begin{align*}S({\mathcal A}_{\mathcal E})=S({\mathcal M})\cap {\mathcal A}_{\mathcal E} , \hspace{.1cm} S_B({\mathcal A}_{\mathcal E})=S_B({\mathcal M})\cap {\mathcal A}_{\mathcal E}\end{align*}$$

the set of bounded density operators from ${\mathcal A}_{\mathcal E}$ . We now introduce the formal definitions of functional inequalities for quantum Markov semigroups.

Definition 3.1. Let $T_t=e^{-Lt}:{\mathcal M}\to {\mathcal M}$ be a symmetric quantum Markov semigroup and $E:{\mathcal M}\to {\mathcal N}$ be the trace-preserving conditional expectation onto its fixed point space. We say $T_t$ satisfies

  1. i) the Poincaré inequality (PI) for $\lambda>0$ if

    (3.1) $$ \begin{align} \lambda\parallel \! x-E(x) \! \parallel_{2}^2\,{\le}\, {\mathcal E}(x,x), \hspace{.1cm} \forall x\in {\mathcal A}_{\mathcal E}, \end{align} $$
  2. ii) the log-Sobolev inequality (LSI) for $\alpha>0$ if

    (3.2) $$ \begin{align} \alpha \tau\big(|x|^2\ln |x^2|-E(|x^2|)\ln E(|x^2|)\big)\le 2{\mathcal E}(x,x), \hspace{.1cm} \forall x\in {\mathcal A}_{\mathcal E}, \end{align} $$
  3. iii) the modified log-Sobolev inequality (MLSI) for $\alpha>0$ if

    (3.3) $$ \begin{align} 2\alpha D(\rho||E(\rho))\le {\mathcal E}(\rho,\ln\rho), \hspace{.1cm} \forall \rho\in S_B({\mathcal A}_{\mathcal E}), \end{align} $$
  4. iv) the complete modified log-Sobolev inequality (CMLSI) for $\alpha>0$ if $\operatorname {\mathrm {\operatorname {id}}}_{\mathcal Q}\otimes T_t$ satisfies $\alpha $ -MLSI inequality for any finite von Neumann algebra ${\mathcal Q}$ .

The optimal (largest possible) constants for PI, LSI, MLSI and CMLSI will be denoted respectively as $\lambda (L), \alpha _2(L),\alpha _{1}(L)$ and $\alpha _{c}(L)$ (or $\lambda , \alpha _2,\alpha _1$ and $\alpha _{c}$ in short if the generator is clear).

The Poincaré inequality (3.1) is equivalent to the spectral gap of L as a positive operator. LSI (3.2) is equivalent to hypercontractivity [Reference Olkiewicz and Zegarlinski60]

(3.4) $$ \begin{align} \parallel \! T_t:L_2({\mathcal M})\to L_p({\mathcal M}) \! \parallel_{}\,{\le}\, 1\hspace{.1cm} \text{ if } \hspace{.1cm} p\le 1+e^{2\alpha t} .\end{align} $$

MLSI (3.3) is known to be equivalent to the exponential decay of relative entropy ([Reference Bardet5, Theorem 3.2] and [Reference Brannan, Gao and Junge14, Proposition A.3]) that

(3.5) $$ \begin{align} D(T_t(\rho)||E(\rho))\le e^{-2\alpha t}D(\rho||E(\rho)), \hspace{.1cm} \forall \rho\in S({\mathcal M}).\end{align} $$

The equivalence of (3.3) and (3.5) is obtained by differentiating the relative entropy for $T_t$ at $0$ , which leads to the entropy production on the R.H.S of MLSI

$$\begin{align*}I_L(\rho):={\mathcal E}(\rho,\ln\rho)=-\frac{d}{dt}\vert_{t=0} D(T_t(\rho)||E(\rho))=\tau(L(\rho) \ln \rho) .\end{align*}$$

It is well known that

$$\begin{align*}\alpha_{2}\leq\alpha_{1}\leq \lambda . \end{align*}$$

The main motivation to consider CMLSI over MLSI and LSI is the tensorization property

(3.6) $$ \begin{align} \alpha_{c}(L_1\otimes \operatorname{\mathrm{\operatorname{id}}}+\operatorname{\mathrm{\operatorname{id}}}\otimes L_2)=\min\{\alpha_{c}(L_1), \hspace{.1cm} \alpha_{c}(L_2)\},\end{align} $$

which in the quantum cases fails for $\alpha _1$ [Reference Brannan, Gao and Junge14, Section 4.4], and is only known to hold for $\alpha _2$ for limited examples in small dimensions. The main result of this section is Theorem 1.1, which asserts a lower bound

$$\begin{align*}\alpha_c(L)\ge \frac{1}{2t_{cb}(L)}\end{align*}$$

by the inverse of CB return time

(3.7) $$ \begin{align} t_{cb}(L)=\inf\{t>0 \hspace{.1cm} |\hspace{.1cm} (1-0.1) E\le_{cp} T_t\le_{cp} (1+0.1)E \}. \end{align} $$

Here, we set $\varepsilon =0.1$ for the notation $t_{cb}(\varepsilon )$ in Theorem 1.1 because of Lemma 2.3.

Theorem 3.2. Let $T_t=e^{-Lt}:{\mathcal M}\to {\mathcal M}$ be a symmetric quantum Markov semigroup and $E:{\mathcal M} \to {\mathcal N}$ be the trace-preserving conditional expectation onto the fixed point subalgebra ${\mathcal N}$ . Define the CB return time as

$$\begin{align*}t_{cb}(L)=\inf \Big \{ t>0\hspace{.1cm} |\hspace{.1cm} 0.9E\le_{cp} T_t\le_{cp} 1.1E \Big \}.\end{align*}$$

Then

$$\begin{align*}\frac{1}{2t_{cb}(L)}\hspace{.1cm} \le \hspace{.1cm} \alpha_{c}(L)\hspace{.1cm} \le \hspace{.1cm}\alpha_{1}(L).\end{align*}$$

Proof. Let $t_m=t_{cb}(L)/2m$ for some $m\in \mathbb {N}_+$ . Since $T_t$ is symmetric, $T_{t_m}^*T_{t_m}=T_{t_m}T_{t_m}=T_{2t_m}$ . Hence, $T_{t_m}$ has discrete return time $k_{cb}(T_{t_m})= m$ . By the Lemma 2.1, for any $\rho \in S_B({\mathcal M})$ ,

$$ \begin{align*}D(T_{t_m}(\rho)||E(\rho))\le & (1-\frac{1}{2m}) D(\rho||E(\rho)). \end{align*} $$

Write $t_{cb}=t_{cb}(L)$ . Now assume further $\rho \in \cup _{t>0} T_t({\mathcal M})\subset \operatorname {\mathrm {\operatorname {dom}}} (L)$ . We have by Theorem 2.5,

$$ \begin{align*}I(\rho)&=\lim_{t\to 0}\frac{ D(\rho||E(\rho))-D(T_{t}(\rho)||E(\rho))}{t}\\ &=\lim_{m\to \infty}\frac{ D(\rho||E(\rho))-D(T_{\frac{t_{cb}}{2m}}(\rho)||E(\rho))}{\frac{t_{cb}}{2m}}\\ &\ge \lim_{m\to \infty}\frac{\frac{1}{2m}D(\rho||E(\rho))}{\frac{t_{cb}}{2m}} =\frac{1}{t_{cb}} D(\rho||E(\rho)). \end{align*} $$

The entropy decay for general $\rho \in S({\mathcal M})$ can be obtained by approximation as in the Appendix [Reference Brannan, Gao and Junge14, Appendix]). This proves $\alpha _{1}(L)\ge \frac {1}{2t_{cb}(L)}$ . The same argument applies to $\operatorname {\mathrm {\operatorname {id}}}_{\mathcal Q}\otimes L$ yields the assertion $\alpha _{c}(L)\ge \frac {1}{2t_{cb}(L)}$ .

Remark 3.3. a) For LSI constant $\alpha _2$ , the $\Omega (\frac {1}{t_{cb}})$ lower bounds were obtained for ergodic semigroups in both classical [Reference Diaconis and Saloff-Coste23] and quantum setting [Reference Temme, Pastawski and Kastoryano74]. These bounds as well as our bound for $\alpha _c$ are asymptotic tight (See Example 5.6 and Section 5.3).

b) In [Reference Brannan, Gao and Junge14], a similar estimate $\alpha _c\ge \Omega (\frac {1}{t_{cb}})$ was obtained for semigroups that admits non-negative entropic Ricci curvature lower bound. The entropy Ricci curvature lower bound for quantum Markov semigroup was introduced by Carlen and Mass [Reference Carlen and Maas15] using $\lambda $ -displacement convexity of entropy functionals H w.r.t to certain noncommutative Wasserstein distance, inspired from Lott and Villani [Reference Lott and Villani51], and Sturm’s [Reference Sturm71] work on metric measure spaces. For heat semigroups on Riemmannian manifold, the entropy Ricci curvature lower bound follows from a lower bound of the Ricci curvature tensor. Nevertheless, in the noncommutative case, these entropy Ricci curvature lower bounds for quantum Markov semigroup are in general hard to verify. So far, most examples rely on certain interwining relation $\nabla T_t= e^{-\lambda t}\tilde {T}_t \nabla $ between the semigroup $T_t$ and a gradient operator $\nabla $ (see [Reference Carlen and Maas15, Reference Brannan, Gao and Junge13, Reference Wirth and Zhang81]).

c) Our Theorem 1.1 here does not rely on any curvature conditions, which uses only information theoretic tools such as entropic quantities and inequalities. To the best of our knowledge, this direct proof is even novel in the classical setting. It is worth pointing out that the definition of relative entropy as well as its exponential decay of relative entropy is independent of the choice of the trace, which also shows the naturalness of our approach and the extension to non-tracial von Neumann algebras in Section 4.

3.2 CB return time

We now consider a common scenario where the CB return time $t_{cb}$ is finite. The original motivation for the notion, despite defining using CP (completely positive) order (3.7), is the following characterization using CB (completely bounded) norm. Recall that a linear map $\Psi :{\mathcal M}\to {\mathcal M}$ is called a ${\mathcal N}$ -bimodule map if

$$\begin{align*}\Psi(axb)=a\Psi(x)b, \hspace{.1cm} \forall\hspace{.1cm} a,b\in {\mathcal N}, x\in {\mathcal M}.\end{align*}$$

Proposition 3.4. Let ${\mathcal N}\subset {\mathcal M}$ be a subalgebra and $E:{\mathcal M}\to {\mathcal N}$ be the trace-preserving conditional expectation. Let $\Psi :{\mathcal M}\to {\mathcal M}$ be an ${\mathcal N}$ -bimodule $*$ -preserving map. For any $\varepsilon>0$ , the following two conditions are equivalent:

  1. i) $(1-\varepsilon ) E\le _{cp} \Psi \le _{cp} (1+\varepsilon ) E\hspace {.1cm} $ ;

  2. ii) $\parallel \! \Psi -E:L_\infty ^1({\mathcal N}\subset {\mathcal M})\to L_\infty ({\mathcal M}) \! \parallel _{cb}\le \varepsilon $ .

The condition ii) above is the completely bounded norm from the space $L_\infty ^1({\mathcal N}\subset {\mathcal M})$ to ${\mathcal M}$ . $L_\infty ^1({\mathcal N}\subset {\mathcal M})$ is called a conditional $L_\infty $ space, defined as the completion of ${\mathcal M}$ with respect to the norm

$$\begin{align*}\parallel \! x \! \parallel_{L_\infty^1({\mathcal N}\subset{\mathcal M})}=\sup_{a,b\in{\mathcal N} \hspace{.1cm} ,\hspace{.1cm} \|a\|_2=\|b\|_2=1}\parallel \! axb \! \parallel_{1}, \end{align*}$$

where the supremum takes over all $a,b\in L_2({\mathcal N})$ with $\|a\|_2=\|b\|_2=1$ . The operator space structure of $L_\infty ^1({\mathcal N}\subset {\mathcal M})$ is given by the identification

$$\begin{align*}{\mathbb M}_n(L_\infty^1({\mathcal N}\subset{\mathcal M}))=L_\infty^1({\mathbb M}_n({\mathcal N})\subset {\mathbb M}_n({\mathcal M}))\hspace{.1cm}\end{align*}$$

(see [Reference Junge and Parcet42] and [Reference Gao, Junge and LaRacuente30, Appendix]). Proposition 3.4 is relatively self-evident in the ergodic case ${\mathcal N}=\mathbb {C}1$ , $L_\infty ^1({\mathcal N}\subset {\mathcal M})\cong L_1({\mathcal M})$ , which we illustrate below.

Example 3.5 (Classical case).

Let $(\Omega ,\mu )$ be a probability space. Let $P:L_\infty (\Omega )\to L_\infty (\Omega )$ be a linear map with kernel $P(f)(x)=\int _{\Omega }k(x,y)f(y)d\mu (y)$ . It is clear that P is $*$ -preserving (i.e., $P(\bar {f})=\overline {P(f)}$ if k is real); P is a positive map if and only if the kernel function $k\ge 0$ . Recall the expectation map

$$ \begin{align*}E_\mu:L_\infty(\Omega)\to \mathbb{C}\textbf{1}\hspace{.1cm} , \hspace{.1cm} E(f)=(\int_{\Omega}f\mu)\textbf{1},\end{align*} $$

where $\textbf {1}$ is the unit constant function. The kernel of $E_\mu $ is the constant function $ \textbf {1}$ on the product space $\Omega \times \Omega $ . The following equivalence is self-evident:

(3.8) $$ \begin{align} (1-\varepsilon)E\le P \le (1+\varepsilon)E \nonumber \Longleftrightarrow\hspace{.1cm} &\varepsilon E\le P-E \le \varepsilon E \nonumber \\ \Longleftrightarrow\hspace{.1cm} &\varepsilon \textbf{1}\le k-\textbf{1} \le \varepsilon \textbf{1}\nonumber \\ \Longleftrightarrow\hspace{.1cm} &\parallel \! k-\textbf{1} \! \parallel_{L_\infty(\Omega\otimes \Omega)}\le \varepsilon\nonumber \\ \Longleftrightarrow\hspace{.1cm} &\parallel \! P-E:L_1(\Omega)\to L_\infty(\Omega) \! \parallel_{}\le \varepsilon. \end{align} $$

To see the equivalence in terms of complete positivity and completely bounded norm in Proposition 3.4, it suffices to notice that every positive (resp. bounded) map to $L_\infty (\Omega )$ is automatically completely positive (resp. completely bounded with same norm [Reference Smith70]).

Example 3.6 (Quantum case).

The above argument also applies to the noncommutative ergodic case ${\mathcal N}=\mathbb {C}1\subset {\mathcal M}$ . The correspondence between the map P and its kernel k generalizes to the isomorphism between the map T and its Choi operator $C_T\in {\mathcal M}^{op}\overline {\otimes }{\mathcal M}$

$$\begin{align*}T(x)=\tau \otimes \operatorname{\mathrm{\operatorname{id}}} (C_T(x\otimes 1)),\hspace{.1cm} x\in L_1({\mathcal M})\cong ({\mathcal M}^{op})_*,\end{align*}$$

where ${\mathcal M}^{op}$ is the opposite algebra of ${\mathcal M}$ . The isomorphism $T\mapsto C_T$ is not only positivity-preserving by Choi’s Theorem ( T is CP iff $C_T\ge 0$ ), but also isometric by Effros-Ruan Theorem (see [Reference Effros and Ruan25, Reference Blecher and Paulsen10]),

$$\begin{align*}\parallel \! T:L_1({\mathcal M})\to L_\infty({\mathcal M}) \! \parallel_{cb}=\parallel \! C_T \! \parallel_{{\mathcal M}^{op}\overline{\otimes}{\mathcal M}}.\end{align*}$$

Then the equivalence in Proposition 3.4 follows as (3.8).

For the general case of a ${\mathcal N}$ -bimodule map T with a nontrivial ${\mathcal N}$ , the above isomorphism holds with more involved module Choi operator, which we refer to the discussion in Section 4.3 and also [Reference Bardet, Capel, Gao, Lucia, Pérez-García and Rouzé6, Lemma 5.1] and [Reference Gao, Junge and LaRacuente29, Lemma 3.15] for the complete proof of Proposition 3.4.

With Proposition 3.4, the CB-return time can be equivalently defined as

(3.9) $$ \begin{align} t_{cb}(L):=\inf \{\hspace{.1cm} t>0\hspace{.1cm} |\hspace{.1cm} \parallel \! T_{t}-E:L_\infty^1({\mathcal N}\subset{\mathcal M})\to L_\infty({\mathcal M}) \! \parallel_{cb}\le 0.1 \}. \end{align} $$

It is known that $t_{cb}$ is finite whenever $T_t$ satisfies the Poincaré inequality and one-point ultra-contractive estimate.

Proposition 3.7. Let $T_t:{\mathcal M}\to {\mathcal M}$ be a symmetric quantum Markov semigroup and $E:{\mathcal M}\to {\mathcal N}$ be the trace-preserving conditional expectation onto the fixed point space. Suppose

  1. i) $T_t$ satisfies the Poincaré inequality: $\lambda>0$ such that $\parallel \! T_{t}-E: L_2({\mathcal M})\to L_2({\mathcal M}) \! \parallel _{}\le e^{-\lambda t}\hspace {.1cm} ,\hspace {.1cm} \forall t\ge 0$ ;

  2. ii) There exists $t_0\ge 0$ such that $\parallel \! T_{t_0}: L_\infty ^1({\mathcal N}\subset {\mathcal M})\to L_\infty ({\mathcal M}) \! \parallel _{cb}\le C_0$ .

Then $ t_{cb}\hspace {.1cm} \le \hspace {.1cm} \frac {1}{\lambda }\ln (10 C_0)+t_0$ .

Proof. This is now a standard argument similar to [Reference Brannan, Gao and Junge14, Proposition 3.8] and [Reference Gao and Rouzé31, Lemma B.1].

Remark 3.8. For the special case of $t_0=0$ and $T_0=\operatorname {\mathrm {\operatorname {id}}}:{\mathcal M}\to {\mathcal M}$ , we consider

$$\begin{align*}\parallel \! \operatorname{\mathrm{\operatorname{id}}}: L_\infty^1({\mathcal N}\subset {\mathcal M})\to L_\infty({\mathcal M}) \! \parallel_{cb}=\inf \{\hspace{.1cm} \mu>0\hspace{.1cm} |\hspace{.1cm} \operatorname{\mathrm{\operatorname{id}}} \le_{cp} \mu E\}:=C_{cb}(E).\end{align*}$$

$C_{cb}(E)$ was introduced in [Reference Gao, Junge and LaRacuente30] as the complete bounded version of Popa and Pimsner’s subalgebra index [Reference Pimsner and Popa67],

$$\begin{align*}C(E):=\inf\{\mu>0 \hspace{.1cm} |\hspace{.1cm} \rho\le \mu E\rho, \hspace{.1cm} \forall \rho\in {\mathcal M}_{+}\}, \hspace{.1cm} C_{cb}(E):=\sup_{n}C(E\otimes \operatorname{\mathrm{\operatorname{id}}}_{\mathbb{M}_n} ).\end{align*}$$

When ${\mathcal M}$ is finite dimensional, both the index $C(E)$ and $C_{cb}(E)$ are finite and admit the explicit formula [Reference Pimsner and Popa67, Theorem 6.1]. In this case, one can take $t_0=0$ in above Proposition 3.7 and yields

$$\begin{align*}t_{cb}\hspace{.1cm} \le \hspace{.1cm} \frac{\ln (10 C_{cb}(E))}{\lambda}.\end{align*}$$

3.3 Classical Markov semigroups

In the remainder of this section, we focus on applications toward classical Markov map. We postpone the discussion of truly noncommutative semigroups to Section 4. Let $T_t=e^{-Lt}:L_\infty (\Omega ,\mu )\to L_\infty (\Omega ,\mu )$ be an ergodic Markov semigroup symmetric to the probability measure $\mu $ . Note that in the ergodic case $L_\infty ^1(\mathbb {C}1\subset L_\infty (\Omega ) )=L_1(\Omega ,\mu )$ , and by Smith’s lemma [Reference Smith70], any bounded map $T:L_1(\Omega ,\mu )\to L_\infty (\Omega ,\mu )$ is automatic completely bounded

$$\begin{align*}\parallel \! T: L_1(\Omega,\mu)\to L_\infty(\Omega,\mu) \! \parallel_{}=\parallel \! T: L_1(\Omega,\mu)\to L_\infty(\Omega,\mu) \! \parallel_{cb}. \end{align*}$$

Then the CB return time $t_{cb}$ reduces to the standard $L_\infty $ -mixing time

$$ \begin{align*} t_{b}(\varepsilon)=&\inf \{ \hspace{.1cm} t>0\hspace{.1cm} |\hspace{.1cm} \parallel \! T_t-E:L_1(\Omega,\mu)\to L_\infty(\Omega,\mu) \! \parallel_{}\le \varepsilon \}. \end{align*} $$

Then by a combination of Theorem 1.1 and Proposition 3.7, we obtain the following theorem.

Theorem 3.9. Let $T_t=e^{-tL}:L_\infty (\Omega ,\mu )\to L_\infty (\Omega ,\mu )$ be an ergodic Markov semigroup symmetric to the probability measure $\mu $ . Suppose

  1. i) $T_t$ satisfies $\lambda $ -Poincaré inequality for some $\lambda>0$ : for $f\in \operatorname {\mathrm {\operatorname {dom}}}(L^{1/2})$ ,

    (3.10) $$ \begin{align} \lambda\mu(|f-E_\mu(f)|^2)\le \int f(Lf)d\mu .\end{align} $$
  2. ii) There exists $t_0>0$ such that

    (3.11) $$ \begin{align}\parallel \! T_{t_0}: L_1(\Omega,\mu)\to L_\infty(\Omega,\mu) \! \parallel_{}\le C_0.\end{align} $$

Then

(3.12) $$ \begin{align}\alpha_1\ge \alpha_{c}\ge \frac{\lambda}{2(\lambda t_0+\ln C_0+2 )}. \end{align} $$

This result can be compared to the bound of Diaconis and Saloff-Coste [Reference Diaconis and Saloff-Coste23, Theoem 3.10], which statesFootnote 1

(3.13) $$ \begin{align} \alpha_1\ge \alpha_{2}\ge \frac{\lambda}{\lambda t_0+\ln (C_0)+1}. \end{align} $$

In particular, $\alpha _1\ge \alpha _{2}\ge \frac {2}{t_b(e^{-2})}$ for the alternative $L_\infty $ -mixing time

$$ \begin{align*} t_b(e^{-2})=\inf \{ t>0\hspace{.1cm} |\hspace{.1cm} \parallel \! T_t-E_\mu:L_1(\Omega,\mu)\to L_\infty(\Omega,\mu) \! \parallel_{}\le \frac{1}{e^2}\}. \end{align*} $$

By the comparison $e^{-3}<0.1<e^{-2}$ , we have $t_b(e^{-2})\le t_{b}(0.1)\le \frac {3}{2}t_b(e^{-2})$ . Hence, in terms of lower bound for $\alpha _1$ , (3.13) and (3.12) are equivalent up to absolute constants. The difference is that (3.13) lower bounds the LSI constant $\alpha _2$ and our estimate (3.12) bounds the CMLSI constant $\alpha _{c}$ .

For finite Markov chains with $|\Omega |<\infty $ , we have finite index

$$\begin{align*}C_{cb}(E_\mu)=C(E_\mu)=\inf\{C>0\hspace{.1cm} |\hspace{.1cm} f\le C\mu(f) \hspace{.1cm} \forall f\ge 0\}= \parallel \! \mu^{-1} \! \parallel_{\infty},\end{align*}$$

where $\mu $ is a strictly positive probability density function. It was proved in [Reference Diaconis and Saloff-Coste23] that

(3.14) $$ \begin{align} &\frac{1}{t_b(e^{-2})}\le \lambda \le \frac{2+\log\parallel \! \mu^{-1} \! \parallel_{\infty}}{2t_b(e^{-2})},\qquad \end{align} $$
(3.15) $$ \begin{align} &\frac{1}{t_b(e^{-2})}\le \alpha_2 \le \frac{4+\log\log \parallel \! \mu^{-1} \! \parallel_{\infty}}{2t_b(e^{-2})}. \end{align} $$

Combined with our Theorem 1.1, we obtain the following:

Corollary 3.10. For a finite Markov chain $T_t:l_\infty (\Omega ,\mu )\to l_\infty (\Omega ,\mu )$ symmetric to $\mu $ ,

$$\begin{align*}\min \Big \{ \frac{4\alpha_2}{3(4+\log\log \parallel \! \mu^{-1} \! \parallel_{\infty})}, \frac{\lambda}{2\log (10\parallel \! \mu^{-1} \! \parallel_{\infty})} \Big \} \le \alpha_c\le \alpha_1\le \lambda.\end{align*}$$

Proof. Note that $t_b(e^{-2})\le t_{cb}(0.1)\le \frac {3}{2}t_b(e^{-2})$ . Then by Theorem 1.1 and (3.15),

$$\begin{align*}\alpha_c\ge \frac{1}{2t_{cb}(0.1)}\ge \frac{1}{3t_{b}(e^{-2})} \ge \frac{2\alpha_2}{3(4+\log\log \parallel \! \mu^{-1} \! \parallel_{\infty})}.\end{align*}$$

The other lower bound $\alpha _c\ge \frac {\lambda }{2\log (10\parallel \! \mu ^{-1} \! \parallel _{\infty })}$ follows from Theorem 3.9 by choosing $t_0=0$ .

Example 3.11. When $\Omega $ is not finite, here is a simple example with $\alpha _c(L)>0$ , but the ultra-contractivity (3.11) is never satisfied for finite $t_0$ . Take $L=I-E_\mu $ . It generates the so-called depolarizing semigroup

$$\begin{align*}T_t=e^{-t}\operatorname{\mathrm{\operatorname{id}}}+(1-e^{-t})E_\mu, \hspace{.1cm} T_t(f)= e^{-t}f+(1-e^{-t})\mu(f)\textbf{1},\end{align*}$$

where $\textbf {1}$ is the unit constant function. Then for any $t<\infty $ ,

$$\begin{align*}\parallel \! T_{t}-E_\mu: L_1(\Omega,\mu)\to L_\infty(\Omega,\mu) \! \parallel_{}=\parallel \! e^{-t}\operatorname{\mathrm{\operatorname{id}}}: L_1(\Omega,\mu)\to L_\infty(\Omega,\mu) \! \parallel_{}=e^{-t}C(E_\mu),\end{align*}$$

which is infinite whenever $L_{\infty }(\Omega ,\mu )$ is infinite dimensional. However, it follows from direct calculation that $\alpha _{c}(I-E_\mu )\geq \frac {1}{2}$ .

3.4 Hörmander system

We now discuss the application to Markov semigroups on smooth manifolds generated by sub-Laplacians. Let $(M,g)$ be a d-dimensional compact connected Riemannian manifold without boundary and let $d\mu =\omega d\operatorname {vol}$ be a probability measure with smooth density $\omega $ w.r.t the volume form $d\operatorname {vol}$ . A family of vector fields $H=\{X_i\}_{i=1}^k\subset TM$ with $k\leqslant d$ is called a Hörmander system if at every point $x\in M$ , the tangent space at x can be spanned by the iterated Lie brackets of $X_i$ ’s

(Hörmander condition) $$ \begin{align} T_xM=\operatorname{span}\{[X_{i_1},[X_{i_2},\cdots, [X_{i_{n-1}}, X_{i_n}]]] \hspace{.1cm} | \hspace{.1cm} 1\leqslant i_1,i_2\cdots i_n\leqslant k \}. \end{align} $$

By compactness, we can assume there is a global constant $l_H$ such that for every point $x\in M$ , we need at most $l_H$ th iterated Lie bracket in above expression (also called strong Hörmander condition). Denote $\nabla =(X_1, \cdots , X_k)$ and by $X_i^*$ the adjoint of $X_i$ on $L^{2}(M,d\mu )$ . Under the Hörmander condition, the sub-Laplacian

$$\begin{align*}\Delta_H=\nabla^*\nabla= \sum_{i} X_i^*X_i=-\sum_{i}X_i^2+(\text{div}_{\mu}(X_i)+X_i(\ln \omega))X_i \end{align*}$$

is a symmetric operator on $L^{2}(M,d\mu )$ which generates an ergodic Markov semigroup $P_t=e^{-\Delta _H t}$ , often called the horizontal heat semigroup. Here, $\operatorname {div}_{\mu }(X)$ is the divergence of X w.r.t to $\mu $ . When $H=\{X_i\}_{i=1}^d$ forms an orthonormal frame to the Riemannian metric, $\Delta _H=\Delta $ recovers the (weighted) Laplace-Beltrami operator and $P_t=e^{-\Delta t}$ is the (weighted) heat semigroup on M.

The gradient form (Carré du Champ operator) of $\Delta _H$ is given by

$$\begin{align*}\Gamma(f,g):=\frac{1}{2}(f\Delta_H(g)+\Delta_H(f)g-\Delta_H(fg))= \sum_{i}\langle X_i f,X_ig\rangle. \end{align*}$$

It follows from the product rule of derivatives that $\Gamma $ is diffusive (i.e., $\Gamma (fg,h)=f\Gamma (g,h)+g\Gamma (f,h)$ ). For diffusion semigroups, it is known [Reference Bakry, Gentil and Ledoux4, Theorem 5.2.1] that the MLSI constant $\alpha _1$ and the LSI constant $\alpha _2$ coincide (i.e., $\alpha :=\alpha _1=\alpha _2$ ). The positivity

$$\begin{align*}\alpha(\Delta_H)>0 \hspace{.1cm} \end{align*}$$

for any Hörmander’s system $H=\{X_i\}_{i=1}^k$ on a compact connected Riemannian manifold without boundary was proved in [Reference Lugiewicz and Zegarliński53, Theorem 3.1]. Our Theorem 1.2 improves this to $\alpha _c(\Delta _H)>0$ .

Proof of Theorem 1.2.

Recall the following Sobolev-type inequality (see, for example, [Reference Lugiewicz and Zegarliński53, Lemma 2.1]):

(3.16) $$ \begin{align} \|f\|_{q} \hspace{.1cm} \le \hspace{.1cm} C \big( \langle \Delta_H f,f\rangle + \parallel \! f \! \parallel_{2}^2\big)^{1/2} , \end{align} $$

where $q=\frac {2dl_H}{dl_H-2}>2$ and $l_H$ is globoal Lie bracket length needed in the strong Hörmander condition. By Varopoulos’ Theorem (see [Reference Varopoulos, Saloff-Coste and Coulhon77, Chapter 2]) on the dimension of semigroups, this implies the following ultra-contractive estimate:

(3.17) $$ \begin{align} \parallel \! e^{-\Delta_Ht}:L_{1}(M,\mu)\to L_{\infty}(M,\mu) \! \parallel_{}\leqslant C^{\prime} t^{-m/2} \text{ for } 0 \le t \le 1 \text{ and some }C'>0, \end{align} $$

where $m=dl_{H}$ . Also, it was proved in [Reference Lugiewicz and Zegarliński53, Theorem 2.3] that $\Delta _H$ satisfies the Poincaré inequality: $\lambda (\Delta _H)>0$ . Combining these with our Theorem 3.9 yields the assertion.

The Sobolev-type inequality (3.16) is also used in [Reference Lugiewicz and Zegarliński53] by Lugiewicz and Zegarlínski to prove that $\alpha _2(\Delta _H)>0$ . Their proof relies on the Rothaus lemma, and so does the discrete case by Diaconis and Saloff-Coste [Reference Diaconis and Saloff-Coste23]. However, we will see in Section 3.6 that this approach is out of scope for showing the CMLSI constant $\alpha _c(\Delta _H)>0$ .

Example 3.12. The special unitary group $\operatorname {SU}\left ( 2 \right )$ is

$$\begin{align*}\operatorname{SU}\,\left( 2 \right)=\{ cI+xX+yY+zZ: c^2+x^2+y^2+z^2=1, x,y,x, c \in \mathbb{R} \}, \end{align*}$$

where $X,Y,Z$ are the skew-Hermitian Pauli unitary

$$\begin{align*}X=\left[\begin{array}{cc} 0& 1\\ -1& 0 \end{array}\right],Y=\left[\begin{array}{cc} 0& i\\ i& 0 \end{array}\right], Z=\left[\begin{array}{cc} i& 0\\ 0& -i \end{array}\right]. \end{align*}$$

The Lie algebra is $\mathfrak {su}(2)=\operatorname {span}_{\mathbb {R}}\{X,Y,Z\}$ with Lie bracket rules as

(3.18) $$ \begin{align}[X,Y]=2Z\ , [Y,Z]=2X\ , [Z,X]=2Y. \end{align} $$

The canonical sub-Riemannian structure is given by $H=\{X,Y\}$ , which is a generating set of $\mathfrak {g}$ because $[X,Y]=2Z$ . The associated sub-Laplacian is

(3.19) $$ \begin{align} \Delta_H=-(X^2+Y^2). \end{align} $$

The semigroup $P_t=e^{-\Delta _H t}$ on $\operatorname {SU}\left ( 2 \right )$ has been studied as a prototype of horizontal heat semigroups. In particular, Baudoin and Bonnefont in [Reference Baudoin and Bonnefont8] proved that

(3.20) $$ \begin{align} \Gamma(P_t f,P_t f)\leqslant C e^{-4 t}P_t(\Gamma(f, f)), \end{align} $$

for some constant $C>0$ . In [Reference Gao and Rouzé31], Gao and Gordina based on (3.20) proved the CMLSI constant that

$$\begin{align*}\alpha_c(\Delta_H)\ge (2\int_0^\infty C e^{-4 t}dt )^{-1}=\frac{2}{C}.\end{align*}$$

The gradient estimate (3.20), as a weaker variant of Bakry-Emery curvature dimension condition, has been found useful to derive CMLSI in [Reference Gao and Rouzé31]. Nevertheless, this weaker gradient estimate is only known for only a limited number of examples in the sub-Riemannian setting [Reference Driver and Melcher24, Reference Melcher54]. Our result avoids this condition and obtains CMLSI for general Hörmander systems.

Example 3.13. Let $n\ge 3$ . The special unitary group $\operatorname {SU}\left ( n \right )$ is

$$\begin{align*}\operatorname{SU}\left( n \right)=\{ u\in {\mathbb M}_n\hspace{.1cm}|\hspace{.1cm} u^*u=1, \hspace{.1cm} \det(u)=1\}. \end{align*}$$

The Lie algebra $\mathfrak {su}(n)$ is the space of all the skew-Hermitian matrices, and a natural basis $\mathfrak {su}(n)$ is given by $\{X_{j,k},Y_{j,k},Z_{k} \hspace {.1cm} | \hspace {.1cm} 1\le j< k\le n\}$ where

$$\begin{align*}X_{j,k}=e_{jk}-e_{kj}\hspace{.1cm} ,\hspace{.1cm} Y_{j,k}=i(e_{jk}+e_{kj})\hspace{.1cm} ,\hspace{.1cm} Z_k=i(e_{11}-e_{kk}), \end{align*}$$

which is $n^2-1$ dimensional. Let $V=\{1,\cdots , n\}$ be a vertex set and $E\subset V\times V$ as an edge set. The set

$$\begin{align*}H_E=\{X_{j,k},Y_{j,k} \hspace{.1cm} |\hspace{.1cm} (j,k)\in E\}\end{align*}$$

is a generating set if and only if $(V,E)$ is a connected graph. The associated sub-Laplacian

$$\begin{align*}\Delta_E=-\sum_{(j,k)\in E} X_{j,k}^2+Y_{j,k}^2\hspace{.1cm} \end{align*}$$

is a generalization of (3.19). Theorem (1.2) implies that $\alpha _c(\Delta _E)>0$ for all connected $(V,E)$ , despite the gradient estimate (3.20) is not known for this type of generator.

3.5 Transference semigroups

Let us discuss an immediate application of $\alpha _c(\Delta _H)>0$ to symmetric Quantum Markov semigroups. Let G be a compact Lie group and $H=\{X_1,\cdots ,X_k\}$ be a generating set of its Lie algebra $\mathfrak {g}$ . Then $\{X_1,\cdots ,X_k\}$ satisfies the Hörmander condition, and its sub-Laplacian $\Delta _H=-\sum _{k}X_i^2$ generates a Markov semigroup $P_t=e^{-\Delta _H t}$ symmetric to the Haar measure. Let $u:G\to {\mathbb M}_n$ be a finite dimensional unitary representation and $d_u:\mathfrak {g}\to i({\mathbb M}_n)_{s.a.}$ be the corresponding Lie algebra morphism. $P_t=e^{-\Delta _H t}$ induces a quantum Markov semigroup $T_t=e^{-L_Ht}:{\mathbb M}_m\to {\mathbb M}_m$ with generator in the Lindbladian form [Reference Lindblad50],

$$\begin{align*}L_H(\rho)=-\sum_{i=1}^k [d_u(X_i),[d_u(X_i),\rho]].\end{align*}$$

$T_t$ is called a transference semigroup of $P_t$ by the following commuting diagram:

(3.21)

where the transference map $\pi _u$ is a $*$ -endomorphism given by

$$ \begin{align*} \pi_u: {\mathbb M}_m\to L_{\infty}(G, {\mathbb M}_m),\hspace{.1cm} \pi_u(\rho)(g)=u(g)^*\rho u(g), \end{align*} $$

which embeds ${\mathbb M}_m$ into $L^{\infty }(G, {\mathbb M}_m)$ . Then the quantum semigroup $T_t$ is the restriction of the matrix-valued extension of classical semigroup $P_t\otimes \operatorname {\mathrm {\operatorname {id}}}_{{\mathbb M}_m}$ on the image of $\pi ({\mathbb M}_m)$ . Such a transference relation holds fro any unitary representation. We obtain the following dimension-free estimates both spectral gap and CMLSI constant (see [Reference Gao, Junge and LaRacuente29, Section 4]):

$$\begin{align*}\alpha_{c}(\Delta_H)\le \alpha_{c}(L_H), \lambda(\Delta_H)\le \lambda(L_H),\end{align*}$$

which are independent of the choice of the unitary representation. Then Corollary 1.3 follows immediately from Theorem 1.2.

3.6 Failure of matrix valued log-Sobolev inequality

As mentioned above, a standard analysis approach to MLSI through hypercontractivity or LSI relies on the Rothaus Lemma (see, for example, [Reference Rothaus68, Reference Bakry3])

$$\begin{align*}H(|f|^2)\le H(|f-E_\mu(f)|^2)+\parallel \! f-E_\mu(f) \! \parallel_{2}^2.\end{align*}$$

Here, we show that the Rothaus Lemma, LSI and hypercontractivity all fail for matrix-valued functions for any classical Markov semigroups. This is a strong indication that the approach by Diaconis and Saloff-Coste’s hypercontractive [Reference Diaconis and Saloff-Coste23] estimate (also used in [Reference Lugiewicz and Zegarliński53]) cannot be used in proving lower bounds for the CMLSI constants.

The following lemma calculates the derivatives of the entropy functional $H(\rho )=\tau (\rho \log \rho )$ . Recall the BKM metric of a operator $X\in {\mathcal M}$ at a base state $\rho $ is

$$\begin{align*}\gamma_\rho(X)=\int_{0}^\infty\tau(X^*(\rho+s)^{-1}X (\rho+s)^{-1}).\end{align*}$$

Lemma 3.14. Let $t\mapsto \rho _t \in S_B({\mathcal M}), t\in (a,b)$ be a smooth family of bounded density operator. Define the function $F(t)=H(\rho _t)=\tau (\rho _t\log \rho _t)$ . Then

$$ \begin{align*}F'(t)=\tau(\rho_t'(\log \rho_t+1)),\hspace{.1cm} F"(t)=\tau(\rho_t"(\log \rho_t+1))+\gamma_{\rho_t}(\rho^{\prime}_t), \end{align*} $$

where $\rho _t'$ and $\rho _t"$ are the first and second order derivative of $\rho _t$ .

Proof. The formula for $F'$ follows from [Reference Wirth79, Lemma 5.8]. For the second derivative, recall the noncommutative chain rule

$$\begin{align*}\frac{d}{dt}(\log \rho_t)=\int_{0}^\infty(\rho_t+s)^{-1}\rho_t'(\rho_t+s)^{-1}ds.\end{align*}$$

By calculating the second derivative, we obtain the second assertion

$$ \begin{align*} F"(t)=&\tau(\rho_t"(\log \rho_t+1))+\int_{0}^\infty \tau(\rho_t'(\rho_t+s)\rho_t'(\rho_t+s))ds \hspace{.1cm} = \hspace{.1cm} \tau(\rho_t"(\log \rho_t+1))+\gamma_{\rho_t}(\rho^{\prime}_t).\\[-42pt] \end{align*} $$

Proposition 3.15. Let $T_t=e^{-tL}:L_\infty (\Omega ,\mu )\to L_\infty (\Omega ,\mu )$ be an ergodic symmetric Markov semigroup. Let $\alpha _R,\alpha _2,\alpha _h$ be the optimal (largest) constant such that the following inequalities hold for any $f\in L_\infty (\Omega ,{\mathbb M}_2)\cap {\mathcal A}_{\mathcal E}$ ,

(Rothaus) $$ \begin{align} &\alpha_R\hspace{.1cm} D\Big(|f|^2 \left| \right|E_\mu(|f|^2)\Big)\le D\Big(|\hat{f}|^2 \left| \right| E_\mu(|\hat{f}|^2)\Big)+\parallel \! \hat{f} \! \parallel_{2}^2,\qquad\qquad\ \ \qquad \end{align} $$
(LSI) $$ \begin{align} &\alpha_2\hspace{.1cm} D(f^2||E_\mu(f^2))\le 2{\mathcal E}(f,f)\qquad\quad\qquad\qquad\qquad\qquad\qquad\qquad\qquad\qquad \end{align} $$
(Hypercontractivity) $$ \begin{align} &\parallel \! T_tf \! \parallel_{L_2({\mathbb M}_2,L_{p(t)}(\Omega))}\le \parallel \! f \! \parallel_{L_2({\mathbb M}_2,L_2(\Omega))} \text{ for } p(t)=1+e^{2\alpha_h t} \end{align} $$

where $E_\mu (f)=(\int f d\mu )1_\Omega $ is the expectation map and $\hat {f}=f-E_\mu (f)$ is the mean zero part of f. Then $\alpha _R=\alpha _2=\alpha _h=0$ .

Proof. We write $\tau (f)=\frac {1}{2}\int {\text {tr}}(f)d\mu $ for the normalized trace on $L_\infty (\Omega ,{\mathbb M}_2)$ . We start with constant $\alpha _R$ in the Rothaus lemma. Without loss of generality, we may assume there is a measurable set $X\subset \Omega $ such that $\mu (X)=r$ for some $0<r<1$ . Let $\eta \in (0,1)$ . Then $h_0=(1-r)1_X-r 1_{X^c}$ is a real mean zero function. Consider the matrix valued function $f_\varepsilon =f+\varepsilon h$ where

$$\begin{align*}f=\left[\begin{array}{cc}1+\eta &0 \\ 0& 1-\eta \end{array}\right]\textbf{1} , \hspace{.1cm} h=\left[\begin{array}{cc}0 & h_0 \\ h_0& 0 \end{array}\right], \end{align*}$$

where f is a constant matrix valued function. Then $E_\mu f_\varepsilon =f, \hat {f}_\varepsilon =\varepsilon h$ and

$$ \begin{align*}&f_\varepsilon^2=(f+\varepsilon h)^2=f^2+\varepsilon(fh+hf)+\varepsilon^2 h^2=f^2+2\varepsilon h+\varepsilon^2 h^2. \end{align*} $$

Then $E_\mu (|f_\varepsilon |^2)=f^2+\varepsilon ^2 h^2$ and $E_\mu (|\hat {f}_\varepsilon |^2)=E_\mu ( h^2)\varepsilon ^2$ . Using Lemma 3.14, the Taylor expansion of the left-hand side of (LSI) is

$$ \begin{align*} D\Big(|f|^2 \left| \right|E_\mu(|f|^2)\Big)&=D(f^2+2\varepsilon h+\varepsilon^2 h^2\left| \right|f^2+\varepsilon^2 h^2)\\ &=H(f^2+2\varepsilon h+\varepsilon^2 h^2)-H(f^2+\varepsilon^2 h^2)\\ &=2\tau(h\log f)\varepsilon+\big(2\tau(h^2(\log f+1))+\gamma_{f}(2h)\big)\varepsilon^2+O(\varepsilon^3)\\&\quad-\tau(2E(h^2)(\log f+1))\varepsilon^2+O(\varepsilon^3)\\ &=\gamma_{f}(2h)\varepsilon^2+O(\varepsilon^3) , \end{align*} $$

where we used the fact $\tau (h\log f)=0$ and $\tau (h^2\log f-E_\mu (h^2)\log f)=0$ . For the right-hand side of the Rothaus lemma, we find

$$ \begin{align*} D(|\hat{f}_\varepsilon|^2||E_\mu(|\hat{f}_\varepsilon|^2))=D(h^2||E_\mu( h^2))\varepsilon^2, \hspace{.1cm} \parallel \! \hat{f}_\varepsilon \! \parallel_{2}^2=\parallel \! h \! \parallel_{2}^2\varepsilon^2. \end{align*} $$

While both $D(h^2||E_\mu ( h^2))$ and $\parallel \! h \! \parallel _{2}^2$ are finite, we have

$$ \begin{align*}\gamma_{f}(2h)&=4\int_{0}^\infty\tau(h(f+s)^{-1}h (f+s)^{-1})ds\\ &=4\int_{0}^\infty \int_\Omega {\text{tr}}(\left[\begin{array}{cc}\frac{1}{(1-\eta+s)(1+\eta+s)}h^2 &0 \\ 0& \frac{1}{(1-\eta+s)(1+\eta+s)}h^2 \end{array}\right]\textbf{1}_\Omega )d\mu ds\\ &=4\Big(\int_{0}^\infty \frac{1}{(1-\eta+s)(1+\eta+s)} ds\Big)\parallel \! h \! \parallel_{2}^2 \\ &= \frac{2}{\eta}\ln\frac{1+\eta}{1-\eta}\parallel \! h \! \parallel_{2}^2. \end{align*} $$

Note that we can choose $\eta \to 1$ and $\frac {1}{2\eta }\ln (\frac {1+\eta }{1-\eta })\to +\infty $ , which implies $\alpha _R=0$ . The same example applies to LSI by choosing a mean zero function $h_0$ such that ${\mathcal E}(h_0,h_0)<\infty $ . For the hypercontractivity, for $p\ge 2$ we recall the norms

$$ \begin{align*} \parallel \! f \! \parallel_{L_2({\mathbb M}_2, L_2(\Omega))}&=\parallel \! f \! \parallel_{L_2({\mathbb M}_2)\otimes L_2(\Omega)}=(\int {\text{tr}}(f^*f)d\mu)^{1/2}. \\ \parallel \! f \! \parallel_{L_2({\mathbb M}_2, L_p(\Omega))}&=\inf_{x,y\in ({\mathbb M}_2)_+,\hspace{.1cm} \parallel \! \hspace{.1cm} x\hspace{.1cm} \! \parallel_{2r}=\parallel \! \hspace{.1cm} y\hspace{.1cm} \! \parallel_{2r}=1}\parallel \! x^{-1}fy^{-1} \! \parallel_{L_2({\mathbb M}_2, L_2(\Omega))},\end{align*} $$

where the infimum takes over all positive invertible $x,y\in {\mathbb M}_2$ with unit $2r$ -norm for $\frac {1}{r}=\frac {1}{2}-\frac {1}{p}$ . Since $T_t$ is a bimodule map for $\mathbb {C}1\otimes {\mathbb M}_2\subset L_\infty (\Omega ,{\mathbb M}_2)$ , we can equivalently consider the norm

$$\begin{align*}\parallel \! T_t: L_2({\mathbb M}_2, L_2(\Omega))\to L_2({\mathbb M}_2, L_p(\Omega)) \! \parallel_{}=\parallel \! T_t: L_2({\mathbb M}_2, L_2(\Omega))\to L_2^a({\mathbb M}_2, L_p(\Omega)) \! \parallel_{},\end{align*}$$

where the asymmetric amalgamated $L_2^a({\mathbb M}_2, L_p(\Omega ))$ space is equipped with norm

$$ \begin{align*}\parallel \! f \! \parallel_{L_2^a({\mathbb M}_2, L_p(\Omega))}=\inf_{a\in ({\mathbb M}_2)_+, \parallel \! \hspace{.1cm} a\hspace{.1cm} \! \parallel_{r}=1}\parallel \! fa^{-1} \! \parallel_{L_2({\mathbb M}_2, L_2(\Omega))}. \end{align*} $$

In particular,

$$\begin{align*}\parallel \! f \! \parallel_{L_2^a({\mathbb M}_2, L_p(\Omega))}^2=\parallel \! f^*f \! \parallel_{L_1({\mathbb M}_2, L_{\frac{p}{2}}(\Omega))},\end{align*}$$

and we have

$$\begin{align*}D(|f|^2||E(|f|^2))=\lim_{q\to 1^+}\frac{\parallel \! |f|^2 \! \parallel_{L_1({\mathbb M}_2, L_q(\Omega))} -\parallel \! |f|^2 \! \parallel_{1}}{q-1 }.\end{align*}$$

Now define $p(t)=2q(t)=1+e^{2\alpha _h t}$

$$\begin{align*}G(t)=\parallel \! T_t f \! \parallel_{L_2^a({\mathbb M}_2, L_p(t)(\Omega))}^2=\parallel \! |T_tf|^2 \! \parallel_{L_1({\mathbb M}_2, L_{q(t)}(\Omega))}. \end{align*}$$

By assumption $G(t)\le 1$ , we have

$$\begin{align*}G'(0)=-2{\mathcal E}(f,f)+\alpha_hD(|f|^2||E(|f|^2))\le 0,\end{align*}$$

which implies $\alpha _h\le \alpha _2=0$ . Note, however, that $\alpha _h\ge 0$ because $T_t$ is always contractive on $L_2({\mathbb M}_2, L_2(\Omega ))$ . Hence, $\alpha _h=0$ , and the proof is complete.

Remark 3.16. Similar to [Reference Bardet and Rouzé7, Corollary 5.2], the above proposition implies that for $p\neq 2$ , neither $L_2^a({\mathbb M}_2, L_p(\Omega ))$ or $L_2({\mathbb M}_2, L_p(\Omega ))$ are uniformly convex.

4 Entropy contraction for GNS symmetric quantum channels

4.1 State symmetric quantum channels

Let ${\mathcal M}$ be a von Neumann algebra and $\phi $ a normal faithful state. We have the GNS cyclic representation $\{\pi _\phi , H_\phi , \eta _\phi \}$ , which is a $*$ -isomorphism $\pi _\phi :{\mathcal M}\to H_\phi $ with a cyclic and separating vector $\eta _\phi $ such that

$$\begin{align*}\phi(x)=\langle \eta_\phi, \pi_\phi(x)\eta_\phi \rangle , \hspace{.1cm} x\in {\mathcal M}.\end{align*}$$

By identifying ${\mathcal M}\cong \pi _\phi ({\mathcal M})$ , the modular automorphism group $\alpha ^\phi _t$ for $t\in \mathbb {R}$ is defined as

$$\begin{align*}\alpha^\phi_t:{\mathcal M}\to {\mathcal M}\hspace{.1cm} ,\hspace{.1cm} \alpha^\phi_t(x)=\Delta^{it}x \Delta^{-it} , \hspace{.1cm} x\in {\mathcal M},\end{align*}$$

where $\Delta $ is the modular operator of $\phi $ , defined as follows:

$$\begin{align*}\Delta=S^*\bar{S}, \hspace{.1cm} S(\pi_\phi(x)\eta_\phi)=\pi_\phi(x^*)\eta_\phi .\end{align*}$$

We consider the following two symmetric conditions with respect to a state $\phi $ .

Definition 4.1. We say a quantum Markov map $\Phi :{\mathcal M} \to {\mathcal M}$ is GNS-symmetric with respect to $\phi $ (in short, GNS- $\phi $ -symmetric) if

$$\begin{align*}\phi(\Phi(x)y)=\phi(x\Phi(y)), \hspace{.1cm} \forall \hspace{.1cm} x,y\in {\mathcal M}\hspace{.1cm}; \end{align*}$$

ii) We say $\Phi $ is KMS-symmetric with respect to $\phi $ (in short, KMS- $\phi $ -symmetric) if

$$\begin{align*}\langle \Delta^{\frac{1}{4}} x\eta_\phi, \Delta^{\frac{1}{4}}\Phi(y)\eta_\phi \rangle=\langle \Delta^{\frac{1}{4}} \Phi(x)\eta_\phi, \Delta^{\frac{1}{4}}y\eta_\phi \rangle, \hspace{.1cm} \forall \hspace{.1cm} x,y\in {\mathcal M}. \end{align*}$$

Correspondingly, we call the pre-adjoint $\Phi _*:{\mathcal M}_*\to {\mathcal M}_*$ a GNS- or KMS- $\phi $ -symmetric quantum channel.

Both definitions are generalizations of the detailed balance condition for classical Markov chains and imply that $\phi =\phi \circ \Phi =\Phi _*(\phi )$ is an invariant state of $\Phi $ . It is proven in [Reference Giorgetti, Parzygnat, Ranallo and Russo32, Reference Wirth80] that the GNS- $\phi $ -symmetric quantum Markov map is equivalent to KMS- $\phi $ -symmetric plus that $\Phi $ commutes with the modular group

$$\begin{align*}\alpha_t^{\phi}\circ\Phi\hspace{.1cm} = \hspace{.1cm} \Phi\circ\alpha_t^{\phi}, \hspace{.1cm} t\in {\mathbb R}.\end{align*}$$

The commutation to modular group is also called Accardi-Cecchini condition in [Reference Giorgetti, Parzygnat, Ranallo and Russo32] for a study of quantum Bayes rule [Reference Parzygnat62, Reference Parzygnat and Russo65, Reference Parzygnat and Buscemi63, Reference Parzygnat and Fullwood64].

For simplicity, we will consider a semifinite von Neumann algebra ${\mathcal M}$ equipped with a normal faithful semi-finite trace $\tau $ , but our discussion applies to general von Neumann algebras with proper interpretation of notations. In the tracial setting, we can write $\phi (x)=\tau (d_\phi x)$ using the density operator $d_\phi $ of $\phi $ . Then the modular automorphism group is given by

$$\begin{align*}\alpha_t^{\phi}(x) \hspace{.1cm} = \hspace{.1cm} d_{\phi}^{-it}xd_{\phi}^{it}, \hspace{.1cm} x\in {\mathcal M} , t\in \mathbb{R}.\end{align*}$$

Let $\Phi _*:L_1({\mathcal M})\to L_1({\mathcal M})$ be the pre-adjoint quantum channel via trace duality. The KMS- $\phi $ -symmetry is equivalent to

(4.1) $$ \begin{align} \Phi_*(d_\phi^{1/2}xd_\phi^{1/2}) \hspace{.1cm} = \hspace{.1cm} d_\phi^{1/2}\Phi(x)d_\phi^{1/2}, \hspace{.1cm} \forall x\in {\mathcal M}. \end{align} $$

For $1\le p\le \infty $ , the weighted $L_p$ -space $L_p({\mathcal M},\phi )$ is the completion of ${\mathcal M}$ under the norm

$$\begin{align*}\parallel \! x \! \parallel_{p,\phi}=\parallel \! d_\phi^{1/2p} x d_\phi^{1/2p} \! \parallel_{p},\end{align*}$$

where $\parallel \! y \! \parallel _{p}=\tau (|y|^p)^{1/p}$ is the tracial p-norm. For $p=2$ , $L_p({\mathcal M},\phi )$ is a Hilbert space with KMS-inner product $\parallel \! x \! \parallel _{2,\phi }^2=\langle \Delta ^{\frac {1}{4}}x\eta _\phi ,\Delta ^{\frac {1}{4}}x\eta _\phi \rangle $ . By equation (4.1), $\Phi $ is also a contraction on $L_1({\mathcal M},\phi )$ , and hence a contraction on $L_p({\mathcal M},\phi )$ for all $1\le p\le \infty $ by complex interpolation.

The lemma below is an analog of Proposition 2.2.

Proposition 4.2. Let $\Phi :{\mathcal M}\to {\mathcal M}$ be a GNS- $\phi $ -symmetric quantum Markov map for a normal faithful state $\phi $ . Denote ${\mathcal N}$ as the multiplicative domain of $\Phi $ . Then

  1. i) ${\mathcal N}$ is invariant under $\alpha _t^\phi $ . Hence, there exists a $\phi $ -preserving normal conditional expectation ${E:{\mathcal M}\to {\mathcal N}}$ .

  2. ii) $\Phi |_{\mathcal N}$ is an involutive $*$ -automorphism satisfying

    (4.2) $$ \begin{align} \Phi^2\circ E=E\circ \Phi^2=E, \hspace{.1cm} E\circ \Phi=\Phi\circ E. \end{align} $$

    Moreover, $\Phi ^2$ is a ${\mathcal N}$ -bimodule map satisfying $\Phi ^2(axb)=a\Phi ^2(x)b$ for any $a,b\in {\mathcal N}$ and $x\in {\mathcal M}$ .

  3. iii) $\Phi $ is an isometry on $L_2({\mathcal N},\phi )$ . If, in addition,

    $$\begin{align*}{\parallel \! \Phi(\operatorname{\mathrm{\operatorname{id}}}-E):L_2({\mathcal M},\phi)\to L_2({\mathcal M},\phi) \! \parallel_{2}<1},\end{align*}$$
    then $E=\lim _{n}\Phi ^{2n}$ as a map from $L_2({\mathcal M},\phi )$ to $L_2({\mathcal M},\phi )$ .

Proof. It suffices to explain i). The rest follows similar as Proposition 2.2 (see also [Reference Gao and Rouzé31, Lemma 2.5] for the finite dimensional case). Indeed, since $\Phi $ commutes with $\alpha _t^\phi $ , for $x\in {\mathcal N}$ ,

$$ \begin{align*} \Phi\big(\alpha_t^\phi(x)y\big)&=\Phi\big(\alpha_t^\phi(x \alpha_{t}^{\phi}\circ\alpha_{-t}^\phi(y) )\big)=\alpha_t^\phi\circ\Phi\big(x\alpha_{t}^{\phi}\circ \alpha_{-t}^\phi(y) \big)\\ &=\alpha_t^\phi\Big(\Phi(x)\Phi(\alpha_{t}^{\phi}\circ\alpha_{-t}^\phi(y)) \Big)=\alpha_t^\phi\circ\Phi(x) \alpha_t^\phi\circ \Phi\circ \alpha_{-t}^\phi(y)=\Phi(\alpha_t^\phi(x))\Phi(y). \end{align*} $$

Multiplicativity on the other side is similar, implying $ \alpha _t^\phi (x)\in {\mathcal N}$ . By Takesaki’s theorem [Reference Takesaki72], there exists $\phi $ -preserving conditional expectation satisfying the defining property

$$\begin{align*}\phi(xy)=\phi(xE(y)) \hspace{.1cm} \forall \hspace{.1cm} x\in {\mathcal N}, y\in {\mathcal M},\end{align*}$$

from which the GNS- $\phi $ -symmetry follows.

4.2 Haagerup’s reduction

A von Neumann algebra ${\mathcal M}$ is called type III if it does not admit a nontrivial semifinite trace. We briefly review the basics of Haagerup’s construction and refer to [Reference Haagerup, Junge and Xu36] for more details. The key idea is to consider the additive subgroup $G=\bigcup _{n\in {\mathbb N}} 2^{-n}\mathbb {Z} \subset \mathbb {R}$ of the automorphism group. Let ${\mathcal M}\subset B(H)$ be a von Neumann algebra and $\phi $ be a normal faithful state. One can define the crossed product by the action $\alpha ^\phi :G\curvearrowright {\mathcal M}$

$$\begin{align*}\hat{{\mathcal M}}\hspace{.1cm} = \hspace{.1cm} {\mathcal M}\rtimes_{\alpha^{\phi}}G . \end{align*}$$

$\hat {{\mathcal M}}$ can be considered as the von Neumann subalgebra $ \hat {{\mathcal M}}=\{\pi ({\mathcal M}),\lambda (G)\}" \subset {\mathcal M}\overline \otimes B(\ell _2(G))$ generated by the embeddings

(4.3) $$ \begin{align} \pi:{\mathcal M}\to {\mathcal M}\rtimes_{\alpha^{\phi}}G, \hspace{.1cm} &\pi(a)\hspace{.1cm} = \hspace{.1cm} \sum_{g} \alpha_{g^{-1}}(a) \otimes |{g}\rangle\langle{g}|\nonumber\\ \lambda:G\to {\mathcal M}\rtimes_{\alpha^{\phi}}G,\hspace{.1cm} &\lambda(g)(|{x}\rangle\otimes |{h}\rangle)=|{x}\rangle\otimes |{gh}\rangle\hspace{.1cm} , \hspace{.1cm} \forall \hspace{.1cm} |{x}\rangle\in H, |{h}\rangle\in \ell_2(G) . \end{align} $$

Basically, $\pi $ is the transference homomorphism ${\mathcal M}\to \ell _\infty (G,{\mathcal M})$ , and $\lambda $ is the left regular representation on $\ell _2(G)$ . The set of finite sums $\{\sum _g a_g \lambda (g)\hspace {.1cm} | \hspace {.1cm} a_g \in {\mathcal M} \}\subset \hat {{\mathcal M}}$ forms a dense $w^*$ -subalgebra of $\hat {{\mathcal M}}$ . In the following, we identify ${\mathcal M}$ with $\pi ({\mathcal M})\subset \hat {M}$ (resp. a with $\pi (a)$ ) and view ${\mathcal M}\subset \hat {{\mathcal M}}$ as a subalgebra. The state $\phi $ admits a natural extension as a normal faithful state on $\hat {{\mathcal M}}$

$$\begin{align*}\hat{\phi}(\sum_g a_g \lambda(g)) \hspace{.1cm} = \hspace{.1cm} \phi(a_0) .\end{align*}$$

Moreover,

$$\begin{align*}E_{\mathcal M}:\hat{{\mathcal M}}\to {\mathcal M}\hspace{.1cm}, \hspace{.1cm} E_{{\mathcal M}}(\sum_g a_g\lambda(g))=a_0\end{align*}$$

is the canonical normal conditional expectation such that $\phi \circ E_{\mathcal M}\hspace {.1cm} = \hspace {.1cm} \hat {\phi }$ .

The main object in Haagerup’s construction is an increasing family of subalgebras

$$\begin{align*}{\mathcal M}_n \hspace{.1cm} = \hspace{.1cm} \hat{{\mathcal M}}_{\psi_n}:=\{x\in \hat{{\mathcal M}} \hspace{.1cm} |\hspace{.1cm} \alpha_t^{\psi_n}(x)=x \hspace{.1cm}, \hspace{.1cm}\forall\hspace{.1cm} t\in {\mathbb R}\},\end{align*}$$

given by the centralizer algebra $\hat {{\mathcal M}}_{\psi _n}$ for a suitable family of states $\psi _n$ so that $\bigcup _n {\mathcal M}_n$ is $w^*$ -dense in $\hat {{\mathcal M}}$ . The state $\psi _n$ is defined via a Radon-Nikodym density w.r.t to $\hat {\phi }$

$$\begin{align*}\psi_n(x) \hspace{.1cm} = \hspace{.1cm} \hat{\phi}(e^{-a_n}x) \hspace{.1cm} ,\hspace{.1cm} a_n \hspace{.1cm} = \hspace{.1cm} -i2^n\text{Log} (\lambda(2^{-n})) .\end{align*}$$

Here, $\text {Log}$ is the principal branch of the logarithmic function with $0\le \text {Log}(z)<2\pi $ . Each subalgebra ${\mathcal M}_n$ contains $\lambda (G)$ , and there exists normal conditional expectation $E_{{\mathcal M}_n}:\hat {{\mathcal M}}\to {\mathcal M}_n$ . Indeed, by the definition of $\psi _n$ , the modular group $\alpha _t^{\psi _n}$ is $2^{-n}$ periodic. The explicit form (see [Reference Haagerup, Junge and Xu36, Lemma 2.3]) is given by

$$\begin{align*}E_{{\mathcal M}_n}=2^{n}\int_0^{2^{-n}}\alpha_t^{\psi_n} dt.\end{align*}$$

The normalized state $\tau _n=\frac {\psi _n}{\psi _n(1)}$ is a normalized trace on ${\mathcal M}_n$ . The key properties of ${\mathcal M}_n$ are summarized in [Reference Haagerup, Junge and Xu36, Theorem 2.1 & Lemma 2.7], which we state below.

Theorem 4.3. With above notations, ${\mathcal M}_n$ is an increasing family of von Neumann subalgebras satisfying the following properties

  1. (1) Each $({\mathcal M}_n,\tau _n)$ is a finite von Neumann algebra.

  2. (2) $\bigcup _{n\ge 1} {\mathcal M}_n$ is weak $^*$ -dense in $\hat {{\mathcal M}}$ .

  3. (3) There exists a $\hat {\phi }$ -preserving normal faithful conditional expectation $E_{{\mathcal M}_n}:\hat {{\mathcal M}}\to {\mathcal M}_n$ such that

    $$\begin{align*}\hat{\phi}\circ E_{{\mathcal M}_n}=\hat{\phi}\hspace{.1cm}, \hspace{.1cm} \alpha_t^{\hat{\phi}}\circ E_{{\mathcal M}_n}= E_{{\mathcal M}_n}\circ \alpha_t^{\hat{\phi}}. \end{align*}$$

    Moreover, $E_{{\mathcal M}_n}(x)\to x$ in $\sigma $ -strong topology for any $x\in \hat {{\mathcal M}}$ .

We now look at the Haagerup reduction on the states. For a state $\rho \in S({\mathcal M})$ , $\hat {\rho }=\rho \circ E_{\mathcal M}$ is the canonical extension on $\hat {{\mathcal M}}$ . We denote $\rho _n:=\hat {\rho }|_{{\mathcal M}_n}\in {\mathcal M}_{n,*}$ as the restriction state of $\hat {\rho }$ on the subalgebra ${\mathcal M}_n \subset \hat {{\mathcal M}}$ . Note that the predual ${\mathcal M}_{n,*}$ can be viewed as a subspace of $\hat {{\mathcal M}}_*$ via the embedding

$$\begin{align*}\iota_{n,*}: \hat{{\mathcal M}}_{n, *}\to \hat{{\mathcal M}}_*\hspace{.1cm}, \iota_{n,*}(\omega)=\omega\circ E_{{\mathcal M}_n}. \end{align*}$$

Via this identification, $\rho _n=\hat {\rho }|_{{\mathcal M}_n}\circ E_{{\mathcal M}_n}=\hat {\rho }\circ E_{{\mathcal M}_n}=E_{{\mathcal M}_n,*}(\hat {\rho })\in \hat {{\mathcal M}}_*$ . Moreover, by the weak $^*$ -density of the family ${\mathcal M}_n$ , $\rho _n\to \hat {\rho }$ converges in the weak topology. An immediate consequence is the following approximation of relative entropy.

Lemma 4.4. Let $\rho $ and $\sigma $ be two normal states of ${\mathcal M}$ . Then

$$\begin{align*}D(\rho||\sigma) \hspace{.1cm} = \hspace{.1cm} D(\hat{\rho}||\hat{\sigma})=\lim_{n\to \infty} D(\rho_n||\sigma_n) .\end{align*}$$

Proof. Let $\iota :{\mathcal M}\subset \hat {{\mathcal M}}$ be the inclusion map. Because $\hat {\rho }=\rho \circ E_{\mathcal M}$ is an extension of $\rho $ , $\iota _*(\hat {\rho })=\hat {\rho }|_{\mathcal M}=\rho $ , and similarly for $\sigma $ . Both $\iota : {\mathcal M} \to \hat {{\mathcal M}}$ and $E_{\mathcal M}:\hat {{\mathcal M}}\to {\mathcal M}$ are quantum Markov maps. Then by the data processing inequality,

$$\begin{align*}D(\rho||\sigma)=D(\iota_*(\hat{\rho})||\iota_*(\hat{\sigma}))\le D(\hat{\rho}||\hat{\sigma})=D(E_{{\mathcal M},*}(\rho)||E_{{\mathcal M},*}(\sigma))\le D(\rho||\sigma). \end{align*}$$

Thus, $D(\rho ||\sigma )=D(\hat {\rho }||\hat {\sigma })$ . As for the limit, we have

$$ \begin{align*} D(\hat{\rho}\|\hat{\sigma})&\leq \liminf_{n}D(\rho_{n}\|\sigma_{n})\\ &=\liminf_{n} D(E_{{\mathcal M}_n,*}(\rho_n)||E_{{\mathcal M}_n,*}(\sigma_{n,*}))\\ &\le D(\hat{\rho}||\hat{\sigma}) , \end{align*} $$

where the equality follows from the lower semi-continuity of relative entropy (see, for example, [Reference Hiai37, Theorem 2.7]). The second inequality is another use the data processing inequality.

We shall also apply the Haagerup’s reduction on GNS-symmetric maps. Let $\Phi :{\mathcal M}\to {\mathcal M}$ be a GNS- $\phi $ -symmetric quantum Markov map. Its canonical extension map

$$\begin{align*}\hat{\Phi}:\hat{{\mathcal M}}\to \hat{{\mathcal M}}\hspace{.1cm}, \hspace{.1cm} \hat{\Phi}(\sum_g a_g\lambda(g)) \hspace{.1cm} = \hspace{.1cm} \sum_g \Phi(a_g)\lambda(g) \hspace{.1cm} \end{align*}$$

is also a GNS- $\hat {\phi }$ -symmetric quantum Markov map. Indeed, $\hat {\Phi }=\Phi \otimes \operatorname {\mathrm {\operatorname {id}}}_{B(\ell _2(G))}|_{\hat {{\mathcal M}}}$ is the restriction of $\Phi \otimes \operatorname {\mathrm {\operatorname {id}}}_{B(\ell _2(G))}$ on $\hat {{\mathcal M}}\subset {\mathcal M}\overline {\otimes } B(\ell _2(G)$ . It is clear that $\hat {\Phi }$ has the multiplicative domain

(4.4) $$ \begin{align}\hat{{\mathcal N}}:={\mathcal N}\rtimes_{\alpha^{\phi}} G, \end{align} $$

where ${\mathcal N}$ is the multiplicative domain of $\Phi $ . In particular, this crossed product is well defined because $\alpha ^\phi _t({\mathcal N})={\mathcal N}$ . Moreover, the $\hat {\phi }$ -preserving conditional expectation $\hat {E}:\hat {{\mathcal M}}\to \hat {{\mathcal N}}$ is nothing but the canonical extension of $E:{\mathcal M}\to {\mathcal N}$ .

Recall that we write $E_{\mathcal M}$ and $E_n$ as the normal conditional expectations from $\hat {{\mathcal M}}$ onto ${\mathcal M}$ and ${\mathcal M}_n$ , respectively. The next lemma shows that the extension $\hat {\Phi }$ is well compatible with the approximation family ${\mathcal M}_n$ .

Lemma 4.5. Let $\Phi :{\mathcal M}\to {\mathcal M}$ be a GNS- $\phi $ -symmetric quantum Markov map. With the notations above,

  1. i) $\hat {\Phi }$ commutes with $E_{\mathcal M}, \hat {E}$ and $E_{{\mathcal M}_n}$ . In particular, $\hat {\Phi }({\mathcal M}_n)\subset {\mathcal M}_n$ .

  2. ii) The restriction $\Phi _n=\hat {\Phi }|_{{\mathcal M}_n}$ is a normal unital completely positive map symmetric with respect to the tracial state $\tau _n$ .

  3. iii) Let ${\mathcal N}_n\subset {\mathcal M}_n$ be the multiplicative domain for $\Phi _n$ . Then the restriction map $E_n:=\hat {E}|_{{\mathcal M}_n}:{\mathcal M}_n\to {\mathcal N}_n$ is the $\tau _n$ -preserving conditional expectation.

Proof. The relation $\hat {\Phi }\circ E_{\mathcal M}=E_{\mathcal M}\circ \hat {\Phi }$ is clear from the definition of $\hat {\Phi }$ , and $\hat {\Phi }\circ \hat {E}=\hat {E}\circ \hat {\Phi }$ follows from Lemma 4.2. Recall that $\psi _n(x)=\hat {\phi }(e^{-a_n}x)$ with density operator $e^{-a_n}\in \lambda (G)"$ and $\lambda (G)$ is in the centralizer of $\hat {\phi }$ [Reference Haagerup, Junge and Xu36, Lemma 2.3]. Then

$$\begin{align*}\alpha_t^{\psi_n} \hspace{.1cm} = \hspace{.1cm} u(t)^*\alpha_t^{\hat{\phi}}u(t)=\text{ad}_{u(t)}\alpha_t^{\hat{\phi}} \end{align*}$$

for the unitary $u(t)=e^{-ita_n}$ . Note that $\hat {\Phi }$ commutes with $\alpha _t^{\hat {\phi }}$ by GNS- $\hat {\phi }$ -symmetry, and also commutes with $\text {ad}_{u(t)}$ because $u(t)\in \lambda (G)"$ is in $\hat {\Phi }$ ’s multiplicative domain. Thus, $\hat {\Phi }$ commutes with $\alpha _t^{\psi _n}$ and hence the conditional expectation $E_{{\mathcal M}_n}=2^{-n}\int _0^{2^{-n}}\alpha _t^{\psi _n}$ . This proves i).

For ii), we note that for $x,y\in {\mathcal M}_n$ ,

$$\begin{align*}\psi_n(x\Phi_n(y))=\hat{\phi}(e^{-a_n}x\hat{\Phi}(y))=\hat{\phi}(\hat{\Phi}(e^{-a_n}x)y)=\hat{\phi}(e^{-a_n}\hat{\Phi}(x)y)=\psi_n(\Phi_n(x)y),\end{align*}$$

where we use the fact that $\hat {\Phi }$ is GNS- $\hat {\phi }$ -symmetric and $e^{-a_n}\in \lambda (G)"$ is in the fixed point subspace of $\hat {\Phi }$ . Finally, iii) follows from applying i) and ii) to $\hat {E}$ .

To summarize the lemma above, we have the following commuting diagrams:

Figure 1 Haagerup reduction of quantum Markov map and conditional expectation.

Basically, $\Phi _n$ is a family of trace symmetric channels approximating $\hat {\Phi }$ , which is in turn a natural extension of $\Phi $ . The same picture holds for the conditional expectations $E_n,\hat {E}$ and E.

4.3 Entropy contraction

We shall now discuss the entropy contraction of GNS- $\phi $ -symmetric channels. The first step is to extend the entropy difference Lemma 2.1. Define the state space that is bounded with respect to $\phi $ ,

$$\begin{align*}S_B({\mathcal M},\phi)=\{\rho \in S({\mathcal M})\hspace{.1cm} |\hspace{.1cm} c^{-1}\phi\le \rho\le c\phi\hspace{.1cm}, \hspace{.1cm} \text{for some } c>0 \}.\end{align*}$$

For all $\rho \in S_B({\mathcal M},\phi )$ , $D(\rho ||\phi ) <\infty $ is finite. Such $S_B({\mathcal M},\phi )$ is a dense subset of $S({\mathcal M})$ because for any $\rho $ and $0<\varepsilon <1$ , $\rho _\varepsilon =(1-\varepsilon )\rho +\varepsilon \phi \in S_B({\mathcal M})$ . For $\rho \in S_B({\mathcal M})$ , we define the entropy difference for a GNS- $\phi $ -symmetric quantum channel $\Phi _*$ as

$$\begin{align*}D_{\Phi_*}(\rho):=D(\rho||\phi)-D(\Phi_*(\rho)||\phi).\end{align*}$$

By data processing inequality and $\Phi _*(\phi )=\phi $ , $D_{\Phi _*}(\rho )\ge 0$ . In the trace symmetric case, $D_{\Phi _*}(\rho )=D(\rho ||1)-D(\Phi _*(\rho )||1)=H(\rho )-H(\Phi _*(\rho ))$ as in Section 2. Let E be the conditional expectation onto the multiplicative domain of $\Phi $ . By the chain rule [Reference Petz66, Theorem 2] that for any E invariant state $\psi \circ E=\psi $ ,

$$\begin{align*}D(\rho||\psi)=D(\rho||E_*(\rho))+D(E_*(\rho)||\psi), \end{align*}$$

we have the alternative expressions $D_{\Phi _*}(\rho )=D(\rho ||E_*(\rho ))-D(\Phi _*(\rho )||\Phi _*E_*(\rho )), $ where we used the property $\Phi _*E_*=E_*\Phi _*$ in Proposition 4.2.

Lemma 4.6. Let $\Phi _*$ be a GNS- $\phi $ -symmetric quantum channel. For any state $\rho ,\omega \in S_{b}({\mathcal M},\phi )$ ,

$$\begin{align*}D(\rho||\Phi_*^2(\omega))\hspace{.1cm} \le \hspace{.1cm} D_{\Phi_*}(\rho)+D(\rho||\omega) .\end{align*}$$

Proof. Recall that we use $\rho _n=\hat {\rho }|_{{\mathcal M}_n}=E_{{\mathcal M}_n,*}(\hat {\rho })$ and $\omega _n=\hat {\omega }|_{{\mathcal M}_n}=E_{{\mathcal M}_n,*}(\hat {\omega })$ as the restriction states on finite von Neumann algebra ${\mathcal M}_n\subset \hat {{\mathcal M}}$ obtained from the Haagerup reduction. By Lemma 4.5, we know that $\Phi _n=\hat {\Phi }|_{{\mathcal M}_n}$ is a quantum Markov map symmetric with respect to the tracial state $\tau _n$ . Thus, by Lemma 2.1 in the tracial case,

$$\begin{align*}D(\rho_n||\Phi_{n}^2(\omega_n)) \hspace{.1cm} \le \hspace{.1cm} D_{\Phi_n}(\rho_n) +D(\rho_n||\omega_n) ,\end{align*}$$

where we identify $\Phi _{n}=\Phi _{n,*}$ by trace symmetry. Here, since $\Phi _n=\Phi |_{{\mathcal M}_n}$ is GNS-symmetric to $\phi _n=\phi |_{{\mathcal M}_n}$ ,

$$\begin{align*}D_{\Phi_n}(\rho_n)=D(\rho_n|| \tau_n)-D(\Phi_n(\rho_n)|| \tau_n)=D(\rho_n||\phi_n)-D(\Phi_n(\rho_n)|| \phi_n) .\end{align*}$$

By the definitions of $\Phi _n$ and $\rho _n$ , and the hat “ $\:\hat {}\:$ ” notation for states on $\hat {{\mathcal M}}$ ,

$$ \begin{align*} &\Phi_{n}(\rho_n)=\hat{\rho}|_{{\mathcal M}_n}\circ \Phi_{n} =\hat{\rho}\circ \hat{\Phi}|_{{\mathcal M}_n}= \widehat{\Phi_*(\rho)}|_{{\mathcal M}_n}= \Phi_*(\rho)_n\hspace{.1cm} ,\\ &\Phi_{n}^2(\rho_n)=\Phi_{n}(\Phi_*(\rho)_n)=\Phi_*^2(\rho)_n . \end{align*} $$

Then by Lemma 4.4, we can approximate every entropic term

$$ \begin{align*}\lim_{n}D(\rho_n||(\Phi_{n})^2(\omega_n))=&\lim_{n}D(\rho_n||\Phi_*^2(\omega)_n)=D(\rho||\Phi_*^2(\omega))\hspace{.1cm} ,\\ \lim_{n} D_{\Phi_n}(\rho_n)=&\lim_{n}D(\rho_n|| \phi_n)-D(\Phi_n(\rho_n)|| \phi_n)\\=&\lim_{n}D(\rho_n||\phi_n)-D(\Phi_*(\rho)_n|| \phi_n)=D_{\Phi_*}(\rho)\hspace{.1cm} ,\\ \lim_{n}D(\rho_n||\omega_n)=&D(\rho||\omega).\\[-42pt] \end{align*} $$

The next lemma shows the CB-return time is also compatible with Haargerup reduction.

Lemma 4.7. Let $\Psi :{\mathcal M}\to {\mathcal M}$ be a GNS- $\phi $ -symmetric quantum Markov map and E be the conditional expectation on its multiplicative domain. Suppose

(4.5) $$ \begin{align} (1-\varepsilon) E \le_{cp} \Psi \le_{cp} (1+\varepsilon)E . \end{align} $$

Then for all $n\in {\mathbb N}$ ,

$$\begin{align*}(1-\varepsilon)E_{n} \le_{cp} \Psi_n \le_{cp} (1+\varepsilon)E_{n} . \end{align*}$$

Moreover, if $0.9 E \le _{cp} \Psi \le _{cp} 1.1E$ and $\Psi \circ E=E$ , then for any $\rho \in S_{B}({\mathcal M},\phi )$ ,

$$\begin{align*}\frac{1}{2}D(\rho||E_*(\rho))\le D(\rho||\Psi_*(\rho)).\end{align*}$$

Proof. The CP order inequality follows from the fact that both maps $E_n$ and $\Psi _n$ are the restriction of $E\otimes \operatorname {\mathrm {\operatorname {id}}}$ and $\Psi \otimes \operatorname {\mathrm {\operatorname {id}}}$ on the subalgebra ${\mathcal M}_n\subset \hat {{\mathcal M}}\subset {\mathcal M}\overline {\otimes } B(\ell _2(G))$ . Then the entropy inequality can be obtained by the tracial case Lemma 2.3 and approximation as in Lemma 4.6.

We then extend the entropy contraction to the GNS-symmetric case.

Theorem 4.8. Let $\Phi :{\mathcal M}\to {\mathcal M}$ be a GNS- $\phi $ -symmetric quantum Markov map and E be the $\phi $ -preserving conditional expectation onto its multiplicative domain ${\mathcal N}$ . Define

$$\begin{align*}k_{cb}(\Phi) \hspace{.1cm} = \hspace{.1cm} \inf\{k\in {\mathbb N}^+ \hspace{.1cm} |\hspace{.1cm} 0.9 E\le_{cp}\Phi^{2k} \le_{cp} 1.1 E\hspace{.1cm} \}. \end{align*}$$

Then, for any $\sigma $ -finite von Neumann algebra ${\mathcal Q}$ , state $\rho \in S({\mathcal M}\overline {\otimes } {\mathcal Q})$ ,

$$\begin{align*}D(\Phi_*\otimes \operatorname{\mathrm{\operatorname{id}}}_{{\mathcal Q}}(\rho)||(\Phi_*\circ E_*)\otimes \operatorname{\mathrm{\operatorname{id}}}_{{\mathcal Q}}(\rho)) \hspace{.1cm} \le \hspace{.1cm} \Big (1-\frac{1}{2k_{cb}(\Phi)}\Big )D(\rho||E_*\otimes \operatorname{\mathrm{\operatorname{id}}}_{{\mathcal Q}}(\rho) ).\end{align*}$$

Proof. For $\rho \in S_{B}({\mathcal M},\phi )$ , the proof is same as the tracial case Theorem 2.5 by using Lemma 4.6 and Lemma 4.7 above. The general case $\rho \in S({\mathcal M})$ can be approximated by $\rho _\varepsilon = (1-\varepsilon )\rho + \varepsilon \phi $ .

Recall that in finite dimensions, the MLSI is defined as the supremum of $\alpha $ such that

$$\begin{align*}2\alpha D(\rho||E_*(\rho))\le I_L(\rho):=\tau(L_*(\rho)(\ln \rho-\ln\phi) ).\end{align*}$$

The right-hand side $I_L(\rho )$ is the entropy production, and the equivalence to entropy decay relies on the de Bruijn identity

(4.6) $$ \begin{align} I_L(\rho)=-\frac{d}{dt}D(T_*(\rho)||E_*(\rho))|_{t=0}.\end{align} $$

In infinite dimensions, the de Bruijn identity (4.6) is less justified even in $B(H)$ with $\dim (H)=+\infty $ (see discussions in [Reference Huber, K¨onig and Vershynina40, Reference König and Smith44]). To avoid this issue, we define the MLSI on Type III von Neumann algebra as follows.

Definition 4.9. For a GNS- $\phi $ -symmetric quantum Markov semigroup $T_t=e^{-tL}:{\mathcal M}\to {\mathcal M}$ , we define the modified log-Sobolev (MLSI) constant $\alpha _1(L)$ as the largest constant $\alpha $ such that

(4.7) $$ \begin{align}D(T_{t,*}(\rho)||E_*(\rho)) \hspace{.1cm} \le \hspace{.1cm} e^{-2\alpha t}D(\rho||E_*(\rho))\hspace{.1cm}, \hspace{.1cm} \forall \rho\in S({\mathcal M}),\end{align} $$

where E is the $\phi $ -preserving conditional expectation onto the fixed point subalgebra ${\mathcal N}$ . The complete MLSI constant is then defined as $\alpha _c(L):=\sup _{\mathcal {{\mathcal Q}}}\alpha (L\otimes \operatorname {\mathrm {\operatorname {id}}}_{{\mathcal Q}})$ , where the supremum is over all $\sigma $ -finite von Neumann algebra ${\mathcal Q}$ .

This definition of MLSI also does not depend on any choice of reference state $\phi $ (see Lemma 4.16). With this definition, we obtain the first half of Theorem 1.1, which is restated below.

Theorem 4.10. Let $T_t=e^{-tL}:{\mathcal M}\to {\mathcal M}$ be a GNS- $\phi $ -symmetric quantum Markov semigroup. Denote $t_{cb} \hspace {.1cm} = \hspace {.1cm} \inf \{t>0 \hspace {.1cm}|\hspace {.1cm} 0.9 E\le _{cp}T_t \le _{cp} 1.1 E\} $ . Then

$$\begin{align*}\alpha_1\ge \alpha_c\ge \frac{1}{2t_{cb}}.\end{align*}$$

Namely, for any $\sigma $ -finite von Neumann algebra ${\mathcal Q}$ and state $\rho \in S({\mathcal M}\overline {\otimes } {\mathcal Q})$ , we have the exponential decay of relative entropy

$$\begin{align*}D(T_{t,*}\otimes \operatorname{\mathrm{\operatorname{id}}}_{{\mathcal Q}}(\rho)||E_*\otimes \operatorname{\mathrm{\operatorname{id}}}_{{\mathcal Q}}(\rho)) \hspace{.1cm} \le \hspace{.1cm} e^{-\frac{t}{t_{cb}}}D(\rho||E_*\otimes \operatorname{\mathrm{\operatorname{id}}}_{{\mathcal Q}}(\rho)) \hspace{.1cm},\hspace{.1cm} t\ge 0.\end{align*}$$

Proof. This can be approximated using the tracial case Theorem 2.5 as Lemma 4.6 above.

Remark 4.11. In the above Haagerup’s reduction, both $\hat {\Phi }$ and $\Phi _n$ are always non-ergodic even given $\Phi $ is ergodic. From this point of view, our consideration for non-ergodic cases is essential even for ergodic $\Phi $ . It also indicates that Haagerup’s reduction does not work for LSI/hypercontractivity.

As we have seen in Proposition 3.7 for the tracial case, a combination of heat kernel estimates and spectral gap allows us to bound CB return time. The same analysis remains valid in the GNS- $\phi $ -symmetric case. For $1\le p\le \infty $ , we define the $\phi $ -weighted conditional $L_\infty ^p({\mathcal N}\subset {\mathcal M}, \phi )$ space as the completion of ${\mathcal M}$ under the norm

$$\begin{align*}\parallel \! x \! \parallel_{L_\infty^p({\mathcal N}\subset {\mathcal M}, \phi)}=\sup\{\parallel \! axb \! \parallel_{p,\phi}\hspace{.1cm} |\hspace{.1cm} a,b\in {\mathcal N},\hspace{.1cm} \parallel \! aa^* \! \parallel_{p,\phi}=\parallel \! b^*b \! \parallel_{p,\phi}=1\hspace{.1cm}\}.\end{align*}$$

For a GNS-symmetric ${\mathcal N}$ -bimodule map $\Psi :{\mathcal M}\to {\mathcal M}$ , the equivalence in Proposition 3.4 also holds,

(4.8) $$ \begin{align}(1-\varepsilon) E\le_{cp} \Psi \le_{cp} (1+\varepsilon) E\hspace{.1cm} \Longleftrightarrow \hspace{.1cm} \parallel \! \Psi-E:L_\infty^1({\mathcal N}\subset{\mathcal M}, \phi)\to L_\infty({\mathcal M}) \! \parallel_{cb}\le \varepsilon. \end{align} $$

Based on that, we have an analog of Proposition 3.7.

Proposition 4.12. Let $T_t:{\mathcal M}\to {\mathcal M}$ be a GNS- $\phi $ -symmetric quantum Markov semigroup and $E:{\mathcal M}\to {\mathcal N}$ be the $\phi $ -preserving conditional expectation onto the fixed point space. Suppose

  1. i) the $\lambda $ -Poincaré inequality that $\parallel \! T_{t}-E: L_2({\mathcal M},\phi )\to L_2({\mathcal M},\phi ) \! \parallel _{}\le e^{-\lambda t}\hspace {.1cm} ,\hspace {.1cm} \forall t\ge 0$ ;

  2. ii) there exists $t_0$ such that $\parallel \! T_{t_0}: L_\infty ^1({\mathcal N}\subset {\mathcal M},\phi )\to L_\infty ({\mathcal M}) \! \parallel _{cb}\le C_0$ .

Then $ t_{cb}\hspace {.1cm} \le \hspace {.1cm} \frac {1}{\lambda }\ln (10 C_0)+t_0$ . In particular, if $C_{cb}(E)<\infty $ , $ t_{cb}\hspace {.1cm} \le \hspace {.1cm} \frac {1}{\lambda }\ln (10 C_{cb}(E))$ .

Proof. The argument is similar to the tracial cases by using the property of $L_\infty ^p({\mathcal N}\subset {\mathcal M},\phi )$ for general von Neumann algebra established in [Reference Junge and Parcet42]. See also [Reference Bardet, Capel, Gao, Lucia, Pérez-García and Rouzé6, Section 5] for the argument in finite dimensional GNS-symmetric cases.

4.4 Applications to finite quantum Markov chains

Let $T_t=e^{-Lt}:{\mathbb M}_d\to {\mathbb M}_d$ be a quantum Markov semigroup on matrix algebra ${\mathbb M}_d$ . Its generator L admits the following Lindbladian form ([Reference Gorini, Kossakowski and Sudarshan33, Reference Lindblad50]):

$$\begin{align*}L(x)= i[h,x]+\sum_{j} \gamma_j(V_j^*[x,V_j]+[ V_j^*,x]V_j), \end{align*}$$

where $h,V_j\in {\mathbb M}_d$ and $h=h^*$ is Hermitian. When $T_t$ is GNS-symmetric, one has the following simplified form [Reference Alicki1, Reference Kossakowski, Frigerio, Gorini and Verri45] that

$$\begin{align*}L(x)=\sum_{j}e^{-w_j/2}\Big(V_j^*[x,V_j]+[ V_j^*,x]V_j\Big), \end{align*}$$

where $\{V_j\}=\{V_j\}^*$ is an orthogonal set with respect to trace inner product and the eigenvector of modular group $\alpha _t^{\phi }(V_j)=e^{-iw_j t}V_j\hspace {.1cm}.$ In finite dimensions, the completely Pimsner-Popa index $C_{cb}(E)$ is always finite. Combining Theorem 4.10 and Proposition 4.12, we obtain the second half of Theorem 1.1 restated as below.

Corollary 4.13. For finite dimensional GNS-symmetric quantum Markov semigroups,

(4.9) $$ \begin{align}\alpha_1\ge \alpha_{c}\ge \frac{\lambda}{2\ln (10 C_{cb}(E))}.\end{align} $$

Corollary 4.13 improves the bound $\alpha _c \ge \frac {\lambda }{2C_{cb}(E)}$ in the previous work of Gao and Rouzé [Reference Gao and Rouzé31].

Remark 4.14. In the ergodic case ${\mathcal N}=\mathbb {C}1$ , the conditional expectation $E_\phi (x)=\phi (x)1$ has index

$$\begin{align*}C(E_\phi)=\parallel \! \phi^{-1} \! \parallel_{\infty}\hspace{.1cm}, C_{cb}(E_\phi)\le \parallel \! \phi^{-1} \! \parallel_{\infty}^2.\end{align*}$$

The above bound (4.9) gives

$$\begin{align*}\alpha_1\ge \alpha_{c}\ge \frac{\lambda}{2\ln10+4 \ln\parallel \! \phi^{-1} \! \parallel_{\infty}}. \end{align*}$$

This can be compared to the bound

(4.10) $$ \begin{align} \alpha_1\ge \alpha_2\ge \frac{2(1-\frac{2}{\parallel \! \phi^{-1} \! \parallel_{\infty}})\lambda}{\ln(\parallel \! \phi^{-1} \! \parallel_{\infty}-1)} \hspace{.1cm}\end{align} $$

proved by Diaconis and Saloff-Coste [Reference Diaconis and Saloff-Coste23] for symmetric classical Markov semigroups. In the quantum case, it is only obtained for unital semigroups [Reference Kastoryano and Temme43] and $d=2$ [Reference Beigi, Datta and Rouzé9]. For both classical and quantum depolarizing semigroups $L(x)=x-\phi (x)1$ , this bound is known to be optimal for $\alpha _2$ , which lower bounds $\alpha _1$ . Our results gives a general $\mathcal {O}(\frac {\lambda }{\parallel \! \phi ^{-1} \! \parallel _{\infty }})$ lower bound for $\alpha _1$ for non-ergodic cases and also the complete constant $\alpha _c$ .

Remark 4.15. The Corollary 3.10 shows that the CMLSI constant $\alpha _{c}$ for a classical Markov semigroup is lower bounded by LSI constant $\alpha _2$ up to a $O(\log \log \parallel \! \mu ^{-1} \! \parallel _{\infty })$ term. This argument does not work for Quantum Markov semigroup $T_t:\mathbb {M}_d\to {\mathbb M}_d$ on matrix algebras, although (3.15) remains valid for ergodic quantum Markov semigroups. The difference is that for matrix algebra, the bounded return time

$$\begin{align*}t_b(e^{-2}):=\frac{1}{2}\inf\{t>0\hspace{.1cm} |\hspace{.1cm} \parallel \! T_t-E:L_1({\mathbb M}_d,\phi)\to L_\infty({\mathbb M}_d) \! \parallel_{}<1/e^2\}\end{align*}$$

and the CB return time of completely bounded norm

$$ \begin{align*} t_{cb}=\inf\{t>0 \hspace{.1cm} | \hspace{.1cm} \parallel \! T_t-E_\mu:L_1({\mathbb M}_d,\phi)\to L_\infty({\mathbb M}_d) \! \parallel_{cb}<1/10\hspace{.1cm}\} \end{align*} $$

are quite different. In the classical setting, we used the fact

$$\begin{align*}\parallel \! T:L_1(\Omega)\to L_\infty(\Omega) \! \parallel_{}=\parallel \! T:L_1(\Omega)\to L_\infty(\Omega) \! \parallel_{cb}.\end{align*}$$

So the $t_b(e^{-2})$ and $t_{cb}(0.1)$ are comparable by absolute constants. In the noncommutative setting, we only have

$$\begin{align*}\parallel \! T_t-E_\mu:L_1({\mathbb M}_d,\phi)\to L_\infty({\mathbb M}_d) \! \parallel_{cb}\le d\parallel \! T_t-E_\mu:L_1({\mathbb M}_d)\to L_\infty({\mathbb M}_d) \! \parallel_{}.\end{align*}$$

In the trace symmetric case, $\parallel \! \mu ^{-1} \! \parallel _{\infty }=d$ and $t_{cb}(0.1)\le \frac {3}{2}t_b(e^{-2}) +\ln d$ ,

$$\begin{align*}\alpha_{c}\ge \frac{1}{2t_{cb}(0.1)}\ge \frac{1}{3t_b(e^{-2})+ 2\ln d}\sim O(\frac{\alpha_2}{\ln d}),\end{align*}$$

which is worse than the lower bound in the previous remark as $\alpha _2\le \lambda $ .

4.5 Independence of invariant state

The next lemma shows that the GNS-symmetry is also independent of the choice of invariant state $\phi $ .

Lemma 4.16. Let $T:{\mathcal M}\to {\mathcal M}$ be a GNS- $\phi $ -symmetric quantum Markov map for a normal faithful state $\phi $ . Denote $E:{\mathcal M}\to {\mathcal N}$ as the $\phi $ -preserving conditional expectation onto the multiplicative domain. Suppose $\psi $ is an another normal faithful state invariant under E (i.e., $\psi \circ E=\psi $ ). Then $T:{\mathcal M}\to {\mathcal M}$ is also GNS- $\psi $ -symmetric.

Proof. Without loss of generality, we assume $\psi \le C\phi $ for some $C>0$ . We first view them as the states on the subalgebra ${\mathcal N}$ by restriction. By [Reference Takesaki73, Theorem 3.17], there exists $h\in {\mathcal N}$ such that

$$\begin{align*}\psi( x )=\phi(h^*xh ) \hspace{.1cm}, \hspace{.1cm} \forall x\in {\mathcal N}. \end{align*}$$

This identity actually also holds for $y\in {\mathcal M}$ . Indeed, because of $\phi \circ E=\phi $ and $\psi \circ E=\psi $ ,

$$\begin{align*}\psi( y )=\psi( E(y) )=\phi(h^*E(y)h )=\phi(E(h^*yh) )=\phi(h^*yh ) \hspace{.1cm}, \hspace{.1cm} \forall y\in {\mathcal M}. \end{align*}$$

Moreover, one can replace h by $T(h)$ , because

$$\begin{align*}\psi( x )=\psi( T(x) )=\phi(h^*T(x)h )=\phi\circ T(T(h^*)xT(h) )=\phi(T(h^*)xT(h) ),\end{align*}$$

where we use the fact that $T^2(h)=h$ . Thus, the GNS-symmetry with respect to $\psi $ follows that for $x,y\in {\mathcal M}$ ,

$$ \begin{align*} \psi( xT(y) )&=\phi(h^*xT(y)h )=\phi(h^*xT(yT(h))) =\phi(T(h^*x)yT(h))=\phi(T(h^*)T(x)yT(h))\\ &=\psi( T(x)y ) ,\end{align*} $$

where we used the multiplicative property of $T(axb)=T(a)T(x)T(b)$ for $a,b\in {\mathcal N}$ . The general case can be obtained via $\psi _{\varepsilon }=(1-\varepsilon )\psi +\varepsilon \phi $ .

We remark that if one has convergence $\lim _{n}\Phi ^{2n}=E$ in $L_2$ -norm, the above E-invariant condition $\phi \circ E=\phi $ can be replaced by $\phi =\Phi ^{2}\circ \phi $ .

Note that the left-hand side of (4.8) only relies on complete positivity. Indeed, the $L_\infty ^1({\mathcal N}\subset {\mathcal M}, \phi )$ norm at the right hand-side is also independent of the choice of the invariant state $\phi =\phi \circ E$ .

Lemma 4.17. Let $\phi $ be a normal faithful state and $E:{\mathcal M}\to {\mathcal N}$ be a $\phi $ -preserving conditional expectation. Suppose $\psi =\psi \circ E$ is another normal faithful state preserved by E. Then,

$$\begin{align*}\parallel \! x \! \parallel_{L_\infty^p({\mathcal N}\subset{\mathcal M}, \phi)}=\parallel \! x \! \parallel_{L_\infty^p({\mathcal N}\subset{\mathcal M}, \psi)}\hspace{.1cm}, \hspace{.1cm} \forall x\in {\mathcal M}.\end{align*}$$

The identity extends to all $x\in L_\infty ^p({\mathcal N}\subset {\mathcal M}, \phi ) $ .

Proof. Note that if both $\phi $ and $\psi $ are E invariant, then $d_{\psi }^{-\frac {1}{2p}}d_\phi ^{\frac {1}{2p}}$ is affiliated to ${\mathcal N}$ . Indeed, as argued in Lemma 4.16, if $\psi \le C\phi $ , then $d_\psi = hd_\phi h^*$ for some $h\in {\mathcal N}$ , and the general case follows from approximation $\psi \le \frac {1}{\varepsilon }((1-\varepsilon )\phi +\varepsilon \psi )$ . Then we have

$$\begin{align*}\parallel \! aa^* \! \parallel_{\phi, p}=\parallel \! d_{\phi}^{\frac{1}{2p}}aa^*d_{\phi}^{\frac{1}{2p}} \! \parallel_{p}=\parallel \! d_{\psi}^{-\frac{1}{2p}}d_{\phi}^{\frac{1}{2p}}aa^*d_{\phi}^{\frac{1}{2p}}d_{\psi}^{-\frac{1}{2p}} \! \parallel_{\psi,2p} .\end{align*}$$

Denote $a_1=d_{\psi }^{-\frac {1}{2p}}d_\phi ^{\frac {1}{2p}}a$ and $b_1=bd_\phi ^{\frac {1}{2p}}d_{\psi }^{-\frac {1}{2p}}$ . For $x\in {\mathcal M}$ ,

$$ \begin{align*} \parallel \! x \! \parallel_{L_\infty^p({\mathcal N}\subset{\mathcal M}, \phi)}=&\sup_{\parallel \! \hspace{.1cm} aa^*\hspace{.1cm} \! \parallel_{\phi,2p}=\parallel \! \hspace{.1cm} b^*b\hspace{.1cm} \! \parallel_{\phi,2p}=1}\parallel \! axb \! \parallel_{\phi,p} =\sup_{\parallel \! \hspace{.1cm} aa^*\hspace{.1cm} \! \parallel_{\phi,p}=\parallel \! \hspace{.1cm} b^*b\hspace{.1cm} \! \parallel_{\phi,p}=1}\parallel \! d_{\psi}^{-\frac{1}{2p}}d_\phi^{\frac{1}{2p}}axbd_{\phi}^{\frac{1}{2p}}d_{\psi}^{-\frac{1}{2p}} \! \parallel_{\psi,p} \\=& \sup_{\parallel \! \hspace{.1cm} a_1a_1^*\hspace{.1cm} \! \parallel_{2p,\psi}=\parallel \! \hspace{.1cm} b_1b_1^*\hspace{.1cm} \! \parallel_{2p,\psi}=1}\parallel \! a_1xb_1 \! \parallel_{\psi, p}=\parallel \! x \! \parallel_{L_\infty^p({\mathcal N}\subset{\mathcal M},\psi)}, \end{align*} $$

where the supremum are for $a,b\in {\mathcal N}$ .

Remark 4.18. For finite ${\mathcal M}$ , one particular invariant state of E used in [Reference Bardet and Rouzé7, Reference Bardet, Capel, Gao, Lucia, Pérez-García and Rouzé6] is $\phi _{{\text {tr}}}=E_*(1)$ . This state is convenient because $\phi _{{\text {tr}}}|_{\mathcal N}$ is a trace. Then by Lemma 4.17, we have

$$\begin{align*}\parallel \! x \! \parallel_{L_\infty^p({\mathcal N}\subset {\mathcal M}, \phi)}=\parallel \! x \! \parallel_{L_\infty^p({\mathcal N}\subset {\mathcal M}, \phi_{\text{tr}})}=\sup\{\parallel \! axb \! \parallel_{p,\phi_{\text{tr}}}\hspace{.1cm} |\hspace{.1cm} a,b\in {\mathcal N},\hspace{.1cm} \parallel \! a \! \parallel_{p,\phi_{\text{tr}}}=\parallel \! b \! \parallel_{p,\phi_{\text{tr}}}=1\hspace{.1cm}\},\end{align*}$$

where we used the fact $L_p({\mathcal N},\phi _{\text {tr}})$ is a tracial $L_p$ -space. We will use this point to simplify the discussion in Section 5.4.

5 Applications and examples

5.1 Entropy contraction coefficients

In this section, we discuss the implications of our results on contraction coefficients studied in [Reference Diaconis and Saloff-Coste23, Reference Del Moral, Ledoux and Miclo22, Reference Müller-Hermes, Stilck França and Wolf57, Reference Gao and Rouzé31]. These are analogs of functional inequalities for a single quantum channel.

Definition 5.1. Let $\Phi :{\mathcal M}\to {\mathcal M}$ be a quantum Markov map GNS- $\phi $ -symmetric to a normal faithful state $\phi $ and $E:{\mathcal M}\to {\mathcal N}$ be the $\phi $ -preserving conditional expectation onto the multiplicative domain of $\Phi $ . We define

  1. i) the $L_2$ -contraction coefficient:

    (5.1) $$ \begin{align}\lambda(\Phi):=\parallel \! \Phi(\operatorname{\mathrm{\operatorname{id}}}-E):L_2({\mathcal M},\phi)\to L_2({\mathcal M},\phi) \! \parallel_{}. \end{align} $$
  2. ii) the entropy contraction coefficient:

    $$\begin{align*}\alpha(\Phi):=\sup_{\rho} \frac{D(\Phi_*(\rho)||\Phi_*\circ E_*(\rho))}{D(\rho||E_*(\rho))}.\end{align*}$$
  3. iii) the complete entropy contraction coefficient $\alpha _c(\Phi ):=\sup _{{\mathcal Q}}\alpha (\operatorname {\mathrm {\operatorname {id}}}_{\mathcal Q}\otimes \Phi )$ where the supremum is over all $\sigma $ -finite von Neumann algebras ${\mathcal Q}$ .

The condition $\lambda (\Phi )<1$ can be viewed as a Poincaré inequality for a quantum channel $\Phi $ , which implies the exponential convergence in $L_2$ ,

$$\begin{align*}\parallel \! \Phi^n(X)-E(X) \! \parallel_{L_2({\mathcal M},\phi)}\le \lambda(\Phi)^n\parallel \! X-E(X) \! \parallel_{L_2({\mathcal M},\phi)}\to 0.\end{align*}$$

Similarly, the entropy contraction coefficient gives the convergence in relative entropy

$$\begin{align*}D(\Phi^n(\rho)||\Phi^n\circ E(\rho))\le \alpha(\Phi)^n D(\rho||E(\rho)).\end{align*}$$

The complete constant $\alpha _c(\Phi )$ controls not only the entropy contraction of $\Phi $ but also $\operatorname {\mathrm {\operatorname {id}}}_{\mathcal Q}\otimes \Phi $ with any environment system $\mathcal {Q}$ . This leads to the tensorization property of $\alpha _c$ that for two GNS-symmetric quantum channels [Reference Gao and Rouzé31],

(5.2) $$ \begin{align}\alpha_c(\Phi_1\otimes \Phi_2)=\max\{\alpha_c(\Phi_1),\alpha_c(\Phi_2)\}.\end{align} $$

For classical Markov maps, the tensorization property (5.2) is known to also hold for the non-complete constant $\alpha $ . Nevertheless, for the quantum Markov map (channel), this is not the case, and $\alpha (\Phi )$ in general can be strictly less than $\alpha _c(\Phi )$ (see [Reference Brannan, Gao and Junge14, Section 4.4]).

In finite dimensions, the existence of strictly contractive constant $\alpha _c(\Phi )<1$ was obtained in [Reference Gao and Rouzé31, Theorem 4.1]. Our results give an explicit estimate for $\alpha _c(\Phi )$ .

Corollary 5.2. Let $\Phi $ be a GNS-symmetric quantum Markov map,

$$\begin{align*}\lambda(\Phi)\le \alpha(\Phi)\le \alpha_c(\Phi)\le (1-\frac{1}{2k_{cb}(\Phi)})\le \left(1-\frac{-\ln \lambda(\Phi)}{\ln (10 C_{cb}(E))}\right).\end{align*}$$

Proof. The estimate follows from Theorem 4.8 and a discrete time analog of Proposition 4.12.

Remark 5.3. In the ergodic trace symmetric case ${\mathcal N}=\mathbb {C}1$ and ${\mathcal M}={\mathbb M}_d$ , we have the trace map $E(x)={\text {tr}}(x)\frac {1}{d}$ and the CB-index $C_{cb}(E)= d^2$ . The above estimate implies

(5.3) $$ \begin{align} \lambda(\Phi)\le \alpha(\Phi)\le \alpha_c(\Phi)\le (1-\frac{-\ln \lambda(\Phi)}{\ln (10 d^2)}). \end{align} $$

This can be compared to [Reference Müller-Hermes, Stilck França and Wolf57, Theorem 4.2] and [Reference Kastoryano and Temme43, Corollary 27],

(5.4) $$ \begin{align} \alpha(\Phi)\le 1-\frac{1}{2}\alpha_{2}(\operatorname{\mathrm{\operatorname{id}}}-\Phi^*\Phi)\le 1-\frac{(1-\lambda(\Phi)^2)^2(1-\frac{2}{d})}{\ln(d-1)},\end{align} $$

where $\alpha _{2}(\operatorname {\mathrm {\operatorname {id}}}-\Phi ^2)$ is the LSI constant of $\operatorname {\mathrm {\operatorname {id}}}-\Phi ^2$ as a generator of quantum Markov semigroup. The two upper bounds in (5.3) and (5.4) are comparable, as both are asymptotically $\Theta (\frac {-\ln \lambda (\Phi )}{\ln d})$ . The strength of our results is that (5.3) also bounds the complete constant $\alpha _c(\Phi )$ which has the tensorization property.

Remark 5.4. Our Lemma 2.1 implies

$$\begin{align*}1-\alpha_1(\operatorname{\mathrm{\operatorname{id}}}-\Phi^2)\le \alpha(\Phi),\end{align*}$$

where $\alpha _1$ is MLSI constant of the semigroup generator $(\operatorname {\mathrm {\operatorname {id}}}-\Phi ^*\Phi )$ . For a classical Markov map, it was proved by Del Moral, Ledoux and Miclo [Reference Del Moral, Ledoux and Miclo22] that there exists a universal constant $0<c<1$ such that

(5.5) $$ \begin{align} 1-\alpha_1(\operatorname{\mathrm{\operatorname{id}}}-\Phi^*\Phi)\le \alpha(\Phi)\le 1-c\alpha_1(\operatorname{\mathrm{\operatorname{id}}}-\Phi^*\Phi).\end{align} $$

To the best of our knowledge, the above upper bound in (5.5) is open in the quantum case.

5.2 Graph random walks

Let $G=(V,E)$ be a finite undirected graph with $|V|=d$ and the edge set $E\subset V\times V$ . The discrete time random walk on G is a finite Markov chain given by the stochastic matrix

$$\begin{align*}K_G(u,v)=\begin{cases} \frac{1}{d(u)}, & \mbox{if } (u,v)\in E \\ 0, & \mbox{otherwise}. \end{cases}\end{align*}$$

Here, $d(u)$ is the degree of vertex $u\in V$ . Then $K_G:l_\infty (V)\to l_\infty (V)$ is a Markov map. The $K_G$ admits a unique station distribution $\pi (u)=\frac {d(u)}{2m}$ , where $|E|=m$ . It is clear that $K_G$ is symmetric to the measure $\pi $ , also called reversible. Hence, $K_G$ is an ergodic unital channel on $L_\infty (V,\pi )$ as $\pi (K_G(f))=\pi (f)$ . The expectation map is $E_\pi (f)=\pi (f)1$ whose index is

$$\begin{align*}C_{cb}(E_\pi)=\parallel \! \pi^{-1} \! \parallel_{\infty}.\end{align*}$$

$K_G$ is connected and $E_\pi $ are symmetric operators on $L_2(V,\pi )$ and

$$\begin{align*}\lambda(K_G)=\parallel \! K_G-E_\pi:L_2(V,\pi)\to L_2(V,\pi) \! \parallel_{}<1\end{align*}$$

if $K_G$ not bipartite (in the bipartite case $K_G$ has eigenvalue $-1$ ). Then our results imply

(5.6) $$ \begin{align}\alpha(K_G)\le \alpha_c(K_G)\le (1-\frac{1}{2k_{cb}(K_G)})\le (1-\frac{-\ln \lambda(K_G)}{\ln (10 \parallel \! \pi^{-1} \! \parallel_{\infty})}). \end{align} $$

Example 5.5 (Cyclic graphs).

Let us consider the cyclic graph $C_d=(V,E)$ with $d\ge 4$ where $V=\{1,\cdots ,d\}$ and $E=\{(j,j+1)| j=1,\cdots , d \}$ . Here, the addition is understood in the sense of ‘mod d’. Then

$$\begin{align*}K_{C_d}(i,j)=\begin{cases} \frac{1}{2}, & \mbox{if } |i-j|=1 \\ 0, & \mbox{otherwise}. \end{cases}\end{align*}$$

As $C_d$ is 2-regular, $K_{C_d}$ is symmetric to the uniform distribution $\pi (i)=1/d$ . It is known that $K_{C_d}$ has spectrum

$$\begin{align*}\lambda_j=\cos(\frac{2\pi j}{d})\hspace{.1cm} ,\hspace{.1cm} j=0,\cdots, d-1.\end{align*}$$

The associated eigenvector is $e_j=\frac {1}{\sqrt {d}}(1,\omega ^j,\omega ^{2j},\cdots , \omega ^{(d-1)j})$ where $\omega =\exp (\frac {2\pi i}{d})$ . When $d=2m+1$ is odd, $\pi $ is the unique stationary measure, and $E_\pi $ is the projection onto the vector $e_0$ . We have

$$ \begin{align*} K_G^k-E_\pi= (K_G-E_\pi)^k=\sum_{j=1}^{2m}\lambda_j^k |{e_j}\rangle\langle{e_j}|. \end{align*} $$

By triangle inequality, we have

$$ \begin{align*} \parallel \! K_G^k-E_\pi:L_1(V,\pi)\to L_\infty(V,\pi) \! \parallel_{}&\le \sum_{j=1}^{2m}|\lambda_j|^k= 2\sum_{j=1}^{m}\cos(\frac{\pi j}{d})^k\\ &\le 2\frac{d}{\pi}\int_{0}^{\pi/2}\cos^k(x)dx =2\frac{d}{\pi}W_k\le 2Cd\sqrt{\frac{1}{2k\pi }}, \end{align*} $$

where $C>0$ is some absolute constant by fact that the Wallis integrals $W_k=\int _{0}^{\pi /2}\cos ^k(x)dx\sim \sqrt {\frac {\pi }{2k}}$ . Thus,

$$\begin{align*}k_{cb}(K_{C_d})\le \frac{(10Cd)^2}{\pi}\sim \mathcal{O}(d^2), \end{align*}$$

and (5.6) implies

$$\begin{align*}\alpha(K_{C_d})\ge \alpha_c(K_{C_d})\ge 1-\mathcal{O}(d^{-2}).\end{align*}$$

By Miclo’s result (5.5), this is asymptotically tight because the MLSI constant $\alpha _1(I-K_G^2)\sim \mathcal {O}(d^{-2})$ (see Example 5.6 below for detials). The similar asymptotic estimate also holds for even circle $d=2m$ .

For the continuous time random walk, we consider $w:E\to (0,\infty )$ to be a positive weighted function on the edge set E. The (weighted) graph Laplacian is given by the matrix

$$ \begin{align*} L_G(u,v)=\begin{cases} \sum_{e=(u,u')\in E}w_e, & \mbox{if } u=v \\ -w_e, & \mbox{if } (u,v)\in E \\ 0, & \mbox{otherwise}. \end{cases} \end{align*} $$

$L_G$ generates the continuous time random walk $T_t=e^{-L_G t}$ as a Markov semigroup and is symmetric with to the uniform distribution $\pi $ on V. $T_t$ is ergodic if and only if $G=(V,E)$ is connected. The expectation map $E_\pi (f)=\pi (f)\textbf {1}$ has index $C_{cb}(E_\pi )=d$ . Then Corollary 4.13,

(5.7) $$ \begin{align}\frac{ \lambda(L_G)}{2(\ln d+\ln 10)}\le \alpha_c(L_G)\le \alpha(L_G)\le \lambda(L_G). \end{align} $$

This lower bound of $\alpha _c(L_G)$ has better dependence on the dimension d than [Reference Li, Junge and LaRacuente49, Lemma 5.2].

Example 5.6 (Cyclic graphs).

Let us again consider the cyclic graph $C_d$ with d vertices. For the uniformly weighted case $w_e\equiv 1$ , $L_{C_d}$ is a circulant matrix

$$\begin{align*}L_{C_d}(i,j)=\begin{cases} 2, & \mbox{if } i=j \\ -1, & \mbox{if } |i-j|=1 \\ 0, & \mbox{otherwise}. \end{cases}\end{align*}$$

Thus, $L_{C_d}=2(I-K_{C_d})$ where $K_{C_d}$ is the random walk kernel in Example 5.5, and $ L_{C_d}$ has spectrum $\lambda _j=2(1-\cos \frac {2\pi j}{d})$ . As discussed in [Reference Diaconis and Saloff-Coste23, Example 3.6],

$$\begin{align*}\parallel \! T_t-E:L_1(V,\pi)\to L_\infty(V,\pi) \! \parallel_{}\le 2\exp(-\frac{4t}{d^2})(\sqrt{1+d^2/4t}).\hspace{.1cm}\end{align*}$$

Choosing $t_0=d^2$ , we have

$$\begin{align*}\parallel \! T_t-E:L_1(V,\pi)\to L_\infty(V,\pi) \! \parallel_{}\le 2e^{-4}\sqrt{5/4}<\frac{1}{10}.\end{align*}$$

Thus, by Theorem 1.1,

$$\begin{align*}\frac{1}{2d^2}\le \alpha_c(L_{C_d})\le \alpha_1(L_{C_d})\le 2(1-\cos\frac{2\pi}{d})=\frac{8\pi^2}{d^2}+\mathcal{O}(\frac{1}{d^4}). \end{align*}$$

This shows that for this example, our inverse of $t_{cb}$ bound for $\alpha _c$ is tight up to absolute constant. Note that the LSI constant $\alpha _2(L_{C_d})$ is also of $\Theta (\frac {1}{d^2})$ .

We refer to [Reference Diaconis and Saloff-Coste23, Reference Bobkov and Tetali11] more examples on spectral gap $\lambda $ , Log-Sobolev constants $\alpha _2$ , $\alpha _1$ , and $L_\infty $ mixing time $t_b$ of finite Markov chains.

5.3 A noncommutative Birth-Death process

Let us illustrate our estimate with a noncommutative birth-death process. This example is a generalization of graph Laplacians on matrix algebras (see [Reference Li, Junge and LaRacuente49, Reference Junge, LaRacuente and Rouzé41] for similar constructions). To fix the notation, let $G=(V, E)$ be an undirected graph with $n=|V|$ vertices and edge set E. For each edge $(r,s)\in E$ , we introduce the edge Lindbladian on ${\mathbb M}_n$ ,

$$ \begin{align*} L_{rs}(x)&=e^{\beta_{rs}/2}L_{e_{rs}}(x) + e^{-\beta_{rs}}L_{e_{sr}}(x) \\ &= e^{\beta_{rs}/2}(e_{ss}x+xe_{ss}-2e_{sr}xe_{rs}) +e^{-\beta_{rs}/2}(e_{rr}x+xe_{rr}-2e_{rs}xe_{sr})\hspace{.1cm}, \end{align*} $$

where $e_{rs}\in {\mathbb M}_n$ is the matrix unit with $1$ at the $(r,s)$ position. The total Lindbladian is a weighted sum over the edge set E,

$$ \begin{align*} L&\hspace{.1cm} = \hspace{.1cm} \sum_{(r,s)\in E} w(r,s) L_{rs}\\ &\hspace{.1cm} = \hspace{.1cm} 2\sum_{s\in V}\left(\sum_{(r,s)\in E}w(r,s)e^{\beta_{rs}/2}\right)(e_{ss}x+xe_{ss})-4\sum_{(r,s)\in E}w(r,s) e^{\beta_{rs}/2} e_{sr}xe_{rs}, \end{align*} $$

where we assume $\beta _{rs}=-\beta _{sr}$ and $w(r,s)=w(s,r)>0$ for the GNS-symmetry condition. Note that for $j\neq k$ ,

$$ \begin{align*} &L(e_{jk})=2(\sum_{(r,k)\in E}w(r,k)e^{\beta_{rk}/2} + \sum_{(r,j)\in E} w(r,j)e^{\beta_{r,j}/2})e_{jk} \hspace{.1cm},\\ &L(e_{jj}) = 4 \sum_{(r,j)\in E} w(r,j)(e^{\beta_{r,j}/2} e_{jj}-e^{-\beta_{r,j}/2}e_{rr}) . \end{align*} $$

Let us collect some relevant facts of such a Lindbladian L as noncommutative extension of graph Laplacian.

  1. i) Denote $\ell _{\infty }(V)\subset {\mathbb M}_n$ as the diagonal subalgebra. $L(\ell _{\infty }(V))\subset \ell _{\infty }(V)$ , and $L|_{\ell _{\infty }(V)}$ is a weighted graph Laplacian;

  2. ii) For $r\neq s$ , the matrix unit $e_{rs}$ is an eigenvector of L

    $$\begin{align*}L(e_{rs}) \hspace{.1cm} = \hspace{.1cm} \gamma_{rs} e_{rs} ,\end{align*}$$
    where $\gamma _{rs}=2(\sum _{(r,j)\in E}w(r,j)e^{\beta _{rj}/2}+\sum _{(k,s)\in E}w(k,s)e^{-\beta _{ks}/2})$ .
  3. iii) $\ker (L)\subset \ell _{\infty }(V)$ , and $\ker (L)=\mathbb {C}1$ if $\mathcal {G}=(V, E)$ is connected.

  4. iv) Let $\mu =(\mu _k)\in \ell _{\infty }(V)$ be a density operator in the diagonal subalgebra. Then L is GNS- $\mu $ -symmetric if $e^{\beta _{rs}}=\mu _s/\mu _r$ for any $s\neq r$ .

Assume $L=\sum _{(s,r)\in E}L_{sr}$ is an ergodic graph Lindbladian satisfying GNS- $\mu $ -symmetric condition for a diagonal density operator $\mu $ . Denote $E_d$ as the projection onto diagonal subalgebra. We can decompose the semigroup $T_t=e^{-tL}$ on the diagonal part and off diagonal part.

(5.8) $$ \begin{align} T_t=T_tE_d +T_t(\operatorname{\mathrm{\operatorname{id}}}-E_d):=T_t^{diag}+T_t^{off}. \end{align} $$

It is clear from i) and ii) that $T_tE_d$ is a classical graph random walk and $T_t(\operatorname {\mathrm {\operatorname {id}}}-E_d)$ is a Schur multiplier on ${\mathbb M}_n$ . Using this decomposition, we consider the CB-return time of the semigroup

$$\begin{align*}t_{cb}(\varepsilon):=\inf\{t>0\hspace{.1cm} | \hspace{.1cm} \parallel \! T_t-E_\mu: L_1({\mathbb M}_n,\mu)\to {\mathbb M}_n \! \parallel_{cb}\leq \epsilon\}\end{align*}$$

satisfying

$$\begin{align*}t_{cb}(2\varepsilon)\le t^{diag}_{cb}(\varepsilon)+t^{off}_{cb}(\varepsilon),\end{align*}$$

where $t^{diag}_{cb}$ and $t^{off}_{cb}$ are the CB-return time for the diagonal part $T_{t}E_{d}$ and off diagonal part $T_t(\operatorname {\mathrm {\operatorname {id}}}-E_d)$ , respectively, where

$$ \begin{align*}t_{cb}^{diag}(\epsilon)&=:\inf\{t>0\hspace{.1cm}|\hspace{.1cm} \|T_{t}E_{d}-E_{\mu}:L_{1}(v,\mu)\to L_{\infty}(V)\|_{cb}\leq \epsilon \}\\ t_{cb}^{off}(\epsilon)&=:\inf\{t>0\hspace{.1cm}|\hspace{.1cm} \|T_{t}(\operatorname{\mathrm{\operatorname{id}}}-E_{d}):L_{1}(\mathbb{M}_{n},\mu)\to \mathbb{M}_{n}\|_{cb}\leq \varepsilon\}.\end{align*} $$

For the diagonal part, $t_{cb}^{diag}(\epsilon )$ is a classical $L_\infty $ mixing time, i.e. the smallest t such that

$$\begin{align*}\parallel \! T_tE_d-E_\mu: L_1(V,\mu)\to L_\infty(V) \! \parallel_{}\le \varepsilon .\end{align*}$$

For the off-diagonal term, we deduce from the Effros-Ruan isomorphism that a Schur multiplier map

$$ \begin{align*}\parallel \! T_t(\operatorname{\mathrm{\operatorname{id}}}-E_d): L_1({\mathbb M}_n,\mu)\to {\mathbb M}_n \! \parallel_{cb}=&\parallel \! \sum_{r\neq s} \mu_{r}^{-1/2}e^{-\gamma_{rs}t}\mu_{s}^{-1/2}e_{rs}\otimes e_{rs} \! \parallel_{\infty}\\ =&\parallel \! \sum_{r\neq s} \mu_{r}^{-1/2}e^{-\gamma_{rs}t}\mu_{s}^{-1/2}e_{rs} \! \parallel_{\infty}.\end{align*} $$

Note that for each t,

$$\begin{align*}A_t=\sum_{r\neq s} \mu_{r}^{-1/2}e^{-\gamma_{rs}t}\mu_{s}^{-1/2}e_{rs}\end{align*}$$

is a symmetric matrix with positive entry. A standard application of Schur’s lemma for matrices with positive entries implies

$$ \begin{align*} \parallel \! A_t \! \parallel_{\infty}\le \sup_{r}\Big(\sum_{s}\mu_{r}^{-1/2}e^{-\gamma_{rs}t}\mu_{s}^{-1/2}\Big), \end{align*} $$

which gives us an estimate for the off diagonal term $t_{cb}^{off}(\epsilon )$ .

Now we consider the birth-death process on a finite state space $V=\{1,\cdots ,n\}$ , which we denote as $L^{BD}_n$ . The corresponding edge E set consists of only successive vertices $E=\{(j,j+1)| 1\le j\le n-1\}$ . The simplest case chooses the uniform weight $w(r,s)=1$ for $(r,s)\in E$ and allows only one Bohr frequency $e^{-\beta }=\frac {\mu _{j}}{\mu _{j+1}}$ , and the resulting stationary measure is the well-studied thermal state

$$\begin{align*}\mu=Z_{\beta}^{-1}(e^{-\beta j})_{j=1}^n,\end{align*}$$

where $Z_{\beta }=\sum _{j=1}^{n} e^{-\beta j}$ is the normalization constant. In this case, $\gamma _{rs}=8(\cosh \beta ) t$ , and the off diagonal CB norm can be estimated by

$$ \begin{align*} \parallel \! A_t \! \parallel_{\infty}\le & \sup_{r}\Big(\sum_{s=1}^n e^{\beta r/2}e^{\beta s/2}\Big) Z_\beta e^{-8(\cosh \beta) t} \\ \le & e^{\beta \frac{n-2}{2}}\frac{1-e^{n\beta /2}}{1-e^{\beta /2}} \frac{ 1-e^{-n\beta }}{ 1-e^{-\beta}} e^{-8(\cosh \beta) t}. \end{align*} $$

Thus, $t_{cb}^{off}(\varepsilon )\le C_1(\beta )n$ for some constant $C_1(\beta )$ depending on $\beta $ . For the classic part, we refer to [Reference Miclo55] and [Reference Chen17] for the fact that the spectral gap is of order $O(1)$ ; that is,

$$\begin{align*}c(\beta) \le \lambda(L^{diag}_n)\le C_{2}(\beta) \end{align*}$$

for all $n\in {\mathbb N}$ . For the commutative system on the diagonal part, this implies (see also [Reference Diaconis and Saloff-Coste23])

$$\begin{align*}t^{diag}_{cb}(\varepsilon) \hspace{.1cm} \le \hspace{.1cm} 2c(\beta)^{-1}(2+ |\log \mu_n|)\hspace{.1cm} \le \hspace{.1cm} C_2(\beta)n ,\end{align*}$$

(for $\varepsilon =e^{-2}$ , but here, the actual value of $\varepsilon $ does not change the asymptotic estimate). However, we have based on [Reference Miclo55] that

$$\begin{align*}t^{diag}_{cb}(0.1)\ge \alpha_1(L^{diag}_n)^{-1}\hspace{.1cm} \ge \hspace{.1cm} c(\beta)n.\end{align*}$$

Combining the diagonal and off diagonal part, we know $t_{cb}(L^{BD}_n)\sim n$ . It turns out CMLSI constant has asymptotic $\alpha _c(L^{BD}_n)\sim \frac {1}{n}$ , which indicates our estimate $\alpha _{c}\ge \frac {1}{2t_{cb}}$ is asymptotically tight for this example.

Theorem 5.7. For $\beta> 0$ , there exist constants $c(\beta ), C(\beta )>0$ such that the CMLSI constant of noncommutative birth-death process $L_n^{BD}$ satisfies

$$\begin{align*}\frac{c(\beta)}{n} \hspace{.1cm} \le \hspace{.1cm} \alpha_c(L_n^{BD})\le \alpha_1(L^{BD}_n) \hspace{.1cm} \le \hspace{.1cm} \frac{C(\beta)}{n} .\end{align*}$$

The same $\Theta (\frac {1}{n})$ asymptotic holds for $t_{cb}(L^{BD}_n)^{-1}$ .

Proof. It suffices to show that

$$\begin{align*}\alpha_c(L^{BD}_{n})\le \alpha_1(L^{BD}_n)\le \frac{C(\beta)}{n} .\end{align*}$$

For this, we consider the function in the commutative system on the diagonal

$$\begin{align*}f(k)=\frac{Z(\beta)}{n}e^{\beta k} \text{ and } \sum_{k=1}^nf(k)\mu(k)=\sum_{k=1}^{n} \frac{Z(\beta)}{n}e^{\beta k}\frac{1}{Z(\beta)}e^{-\beta k}=1\end{align*}$$

so that $\rho :=f\mu $ represents a probability density. The relative entropy term satisfies

$$\begin{align*}D(\rho||\mu)= D(f\mu||\mu)\hspace{.1cm} = \hspace{.1cm} \sum_k \frac{e^{-\beta k}}{Z(\beta)} f(k)(\beta k+\ln Z(\beta)-\ln n)\ \hspace{.1cm} = \hspace{.1cm} \ln Z(\beta)-\ln n+ \beta \frac{n+1}{2} .\end{align*}$$

Our density is $\rho \equiv (\frac {1}{n})$ , and the reference density is $\mu (k)=\frac {e^{-\beta k}}{Z(\beta )}$ .

Denote $a_k=|k\rangle \langle k+1|$ . On the diagonal, we have

$$ \begin{align*} \frac{1}{2}L_*(f)&= \sum_k e^{\beta/2}(a_ka_k^*f-a_k^*fa_k)+ e^{-\beta/2}(a_k^*a_kf-a_kfa_k^*) \\ &= \sum_k e^{\beta/2}(e_{k}f(k)-f(k)e_{k+1}) + e^{-\beta/2}(f(k+1)e_{k+1}-f(k+1)e_k) \\ &= \frac{1}{Z(\beta)n} (e^{\beta/2}(e_0-e_{n}) + e^{-\beta/2}(e_n-e_0)) . \end{align*} $$

We have

$$ \begin{align*} L_{n,*}^{BD}(f)(k)=\begin{cases} 4 (e^{\beta/2}-e^{-\beta/2}), &\quad \text{if} \quad k=1;\\ 0,&\quad \text{if} \quad k=2,n-1;\\ 4 (e^{-\beta/2}-e^{\beta/2}), &\quad \text{if} \quad k=n.\\ \end{cases} \end{align*} $$

Note that

$$\begin{align*}\ln \rho-\ln \mu \hspace{.1cm} = \hspace{.1cm} \ln f \hspace{.1cm} = \hspace{.1cm} \big(\beta k-\ln (Z(\beta)n)\big)_{k=1}^n .\end{align*}$$

Then we have the entropy production

$$\begin{align*}I_{L_{n}^{BD}}(\rho)=\tau( L_{n,*}^{BD}(f) \ln f) \sim c(\beta) \hspace{.1cm} \end{align*}$$

for some constant $c(\beta )$ only depending on $\beta $ . This holds for $n\hspace {.1cm} \ge \hspace {.1cm} n_0$ large enough.

Remark 5.8. When $\beta> 0$ , $\sum _{k=1}^n e^{-\beta k} = O(1)$ is a geometric series. In the case that $\beta = 0$ , the above birth-death process reduces to a ‘broken’ version of the cyclic graph (linear graph) as in Example 5.6 with $\alpha _c(L_n) \sim 1/n^2$ .

5.4 Noncommutative concentration inequality

In this section, we show that $\text {CMLSI}$ of a GNS- $\phi $ -symmetric semigroup implies concentration inequalities for the state $\phi $ . The key quantity in the discussion is the Lipschitz semi-norm

$$\begin{align*}\|x\|_{\text{Lip}}^2 \hspace{.1cm} = \hspace{.1cm} \max\{ \parallel \! \Gamma_L(x,x) \! \parallel_{}\hspace{.1cm}, \hspace{.1cm} \parallel \! \Gamma_{L}(x^*,x^*) \! \parallel_{} \}, \end{align*}$$

where the gradient form (or Carré du Champ operator) is

$$\begin{align*}\Gamma_L(x,y) \hspace{.1cm} = \hspace{.1cm} \frac{1}{2}\Big(L(x^*)y+x^*L(y)-L(x^*y)\Big)\hspace{.1cm}, \hspace{.1cm} \forall x,y\in \text{dom}(L). \end{align*}$$

Note that $\|\cdot \|_{\text {Lip}}$ is a semi-norm (satisfying triangle inequality) because $\Gamma _L$ is completely positive bilinear form. Our first lemma is to show that $\|x\|_{\text {Lip}}$ can be approximated by Haagerup reduction.

Lemma 5.9. Let $x\in {\mathcal M}$ . Then for all $n\in {\mathbb N}$ ,

$$\begin{align*}\|E_{{\mathcal M}_n}(x)\|_{\text{Lip}}\hspace{.1cm} \le \hspace{.1cm} \|x\|_{\text{Lip}} .\end{align*}$$

Proof. Recall the conditional expectation $E_{{\mathcal M}_n}:\hat {{\mathcal M}}\to {\mathcal M}_n$ is given by

$$\begin{align*}E_{{\mathcal M}_n}(x)=2^n\int_{0}^{2^{-n}} \alpha_t^{\psi_n}(x)d t. \end{align*}$$

Note that $\alpha _t^{\psi _n}$ is an inner automorphism on ${\mathcal M}\rtimes _\alpha 2^{-n}\mathbb {Z}\cong L_\infty (\mathbb {T}, {\mathcal M}) $ . We note that for a modular automorphism $\alpha _t$ such that $L\alpha _t=\alpha _tL$ ,

$$\begin{align*}\Gamma_L(\alpha_t(x),\alpha_t(y)) \hspace{.1cm} = \hspace{.1cm} \alpha_t(\Gamma_L(x,y)) , \end{align*}$$

which implies $\parallel \! x \! \parallel _{\text {Lip}}=\parallel \! \alpha _t(x) \! \parallel _{\text {Lip}}$ . Here, both $\alpha _t^{\hat {\phi }}$ and $\alpha _t^{\psi _n}$ commute with $\hat {L}=\operatorname {\mathrm {\operatorname {id}}}_{\mathbb {T}}\otimes L$ by the GNS-symmetricness of $\hat {L}$ . Then by triangle inequality,

$$ \begin{align*} \parallel \! E_{{\mathcal M}_n}(x) \! \parallel_{\text{Lip}} &= \Big \| 2^n\int_0^{2^{-n}} \alpha_{t}^{\psi_n}(x) dt \Big \|_{\text{Lip}}\le 2^n\int_0^{2^{-n}} \parallel \! x \! \parallel_{\text{Lip}}dt=\parallel \! x \! \parallel_{\text{Lip}}. \\[-42pt] \end{align*} $$

Lemma 5.10. Let ${\mathcal M}_0,{\mathcal N}\subset {\mathcal M}$ be two subalgebras and $\phi $ be a normal faithful state. Suppose $E_{0}:{\mathcal M}\to {\mathcal M}_0$ and $E:{\mathcal M}\to {\mathcal N}$ are $\phi $ -preserving conditional expectations onto ${\mathcal M}_0$ and ${\mathcal N}$ , respectively. Suppose $E\circ E_0=E_0\circ E$ satisfy the commuting square condition

where ${\mathcal N}_{0}\subset {\mathcal N}$ is a subalgebra. Then for any $p\in [1,\infty ]$ and any $x\in {\mathcal M}$ ,

$$\begin{align*}\parallel \! E_0(x) \! \parallel_{L_\infty^p({\mathcal N}_0\subset {\mathcal M}_0,\phi)}=\parallel \! E_0(x) \! \parallel_{L_\infty^p({\mathcal N}\subset {\mathcal M},\phi)}\le \parallel \! x \! \parallel_{L_\infty^p({\mathcal N}\subset {\mathcal M},\phi)}.\end{align*}$$

In other words, $L_\infty ^p({\mathcal N}_0\subset {\mathcal M}_0,\phi )\subset L_\infty ^p({\mathcal N}\subset {\mathcal M},\phi )$ as a $1$ -complemented subspace with projection  $E_0$ .

Proof. We can assume $\phi =\phi _{\text {tr}}$ in the Remark 4.18. Using commuting square assumption, we know $E_0(a)\in {\mathcal N}_0$ for $a\in {\mathcal N}$ . By definition,

$$ \begin{align*} \parallel \! E_{0}(x) \! \parallel_{L_\infty^p({\mathcal N}_0\subset {\mathcal M}_0, \phi)}= &\sup_{a,b\in\hspace{.1cm} {\mathcal N}_0}\parallel \! aE_{0}(x)b \! \parallel_{\phi,p} \le \sup_{a,b\in\hspace{.1cm} {\mathcal N}_0}\parallel \! E_0(axb) \! \parallel_{\phi,p}\\ \le &\sup_{a,b\in\hspace{.1cm} {\mathcal N}_0}\parallel \! axb \! \parallel_{\phi,p} \le \parallel \! x \! \parallel_{L_\infty^p({\mathcal N}\subset {\mathcal M}, \phi)}, \end{align*} $$

where the supremum is for all $a,b$ in the corresponding subalgebra with $\parallel \! a \! \parallel _{\phi ,p}=\parallel \! b \! \parallel _{\phi ,p}=1$ . Now it suffices to show the other direction

$$ \begin{align*}\parallel \! x \! \parallel_{L_\infty^p({\mathcal N}_0\subset {\mathcal M}_0,\phi)}\ge \parallel \! x \! \parallel_{L_\infty^p({\mathcal N}\subset {\mathcal M},\phi)},\end{align*} $$

for $x\in {\mathcal M}_0$ . For that, we revoke that for $\frac {1}{p}+\frac {1}{q}=1$ , $L_1^{p'}({\mathcal N}\subset {\mathcal M})\subset L_\infty ^p({\mathcal N}\subset {\mathcal M}, \phi )^*$ is as a weak $^*$ -dense subspace [Reference Junge and Parcet42, Proposition 4.5]. Here, for $x\in {\mathcal M}$ ,

$$\begin{align*}\parallel \! y \! \parallel_{L_1^{q}({\mathcal N}\subset {\mathcal M})}=\inf_{y=azb}\parallel \! a \! \parallel_{2p,\phi}\parallel \! y \! \parallel_{q,\phi}\parallel \! b \! \parallel_{2p,\phi},\end{align*}$$

where the infimum is over all factorization $y=azb$ with $a,b\in {\mathcal N}, z\in {\mathcal M}$ . The duality pairing is given by the KMS inner product,

$$\begin{align*}\langle x,y \rangle=\tau(x^*d_{\phi}^{1/2} y d_{\phi}^{1/2})=\langle x,y \rangle_{\phi}. \end{align*}$$

Indeed, it was proved in [Reference Junge and Parcet42, Corollary 3.13] that

$$\begin{align*}E_0: L_1^{q}({\mathcal N}\subset {\mathcal M})\to L_1^{q}({\mathcal N}_0\subset {\mathcal M}_0)\end{align*}$$

is a contraction by the commuting square condition. Therefore, for $x\in {\mathcal N}_0$ , by the KMS- $\phi $ -symmetry of $E_0$ ,

$$ \begin{align*}\parallel \! x \! \parallel_{L_\infty^p({\mathcal N}\subset {\mathcal M}, \phi)}=&\sup_{\parallel \! \hspace{.1cm} y\hspace{.1cm} \! \parallel_{L_1^{q}({\mathcal N}\subset {\mathcal M})}=1}\langle x,y \rangle_\phi \\ =& \sup_{\parallel \! \hspace{.1cm} y\hspace{.1cm} \! \parallel_{L_1^{q}({\mathcal N}\subset {\mathcal M})}=1}\langle x,E_0(y) \rangle_\phi \\ \le& \sup_{\parallel \! \hspace{.1cm} z\hspace{.1cm} \! \parallel_{L_1^{q}({\mathcal N}_0\subset {\mathcal M}_0)}=1}\langle x,z \rangle_\phi=\parallel \! x \! \parallel_{L_\infty^p({\mathcal N}_0\subset {\mathcal M}_0, \phi)} .\\[-47pt] \end{align*} $$

Lemma 5.11. For $x\in {\mathcal M}$ , $\lim _{n}\|E_{{\mathcal M}_n}(x)\|_{L_\infty ^p({\mathcal N}_n\subset {\mathcal M}_n, \psi _n)}= \|x\|_{L_\infty ^p({\mathcal N}\subset {\mathcal M}, \phi )}$ .

Proof. Recall the commuting square condition $E_{{\mathcal M}_n}\circ \hat {E}=\hat {E}\circ E_{{\mathcal M}_n}$ . By Lemma 4.17 & 5.10,

$$ \begin{align*}\parallel \! E_{{\mathcal M}_n}(x) \! \parallel_{L_\infty^p({\mathcal N}_n\subset {\mathcal M}_n, \psi_n)}=&\parallel \! E_{{\mathcal M}_n}(x) \! \parallel_{L_\infty^p({\mathcal N}_n\subset {\mathcal M}_n,\hat{\phi})} =\parallel \! E_{{\mathcal M}_n}(x) \! \parallel_{L_\infty^p({\mathcal N}\subset {\mathcal M},\hat{\phi})}\le \parallel \! x \! \parallel_{L_\infty^p({\mathcal N}\subset {\mathcal M},\hat{\phi})}. \end{align*} $$

The other direction follows from the weak $^*$ -convergence $E_{{\mathcal M}_n}(x) \to x$ . Fix $\frac {1}{q}+\frac {1}{p}=1$ . For any $\varepsilon>0$ , there exists $a_0,b_0\in \hat {{\mathcal N}}$ and $y_0\in \hat {{\mathcal M}}$ such that

$$ \begin{align*} &\parallel \! aa^* \! \parallel_{p,\hat{\phi}}=\parallel \! b^*b \! \parallel_{p,\hat{\phi}}=\parallel \! y \! \parallel_{\hat{\phi},q}=1\hspace{.1cm}, \hspace{.1cm} &\hat{\tau}(d_{\hat{\phi}}^{1/2}axb d_{\hat{\phi}}^{1/2}y)\ge \parallel \! x \! \parallel_{L_\infty^p(\hat{{\mathcal N}}\subset \hat{{\mathcal M}}, \hat{\phi})}-\varepsilon. \end{align*} $$

By the weak $^*$ -density, we can choose $n_1,n_2,n_3$ and $n_4\ge \max \{n_1,n_2,n_3\}$ inductively such that

$$ \begin{align*} \tau(d_\phi^{1/2}E_{{\mathcal M}_{n_1}}(a)E_{{\mathcal M}_{n_4}}(x)E_{{\mathcal M}_{n_2}}(b)d_\phi^{1/2}E_{{\mathcal M}_{n_3}}(y))>\tau(d_\phi^{1/2}axbd_\phi^{1/2}y)-\varepsilon >\parallel \! x \! \parallel_{L_\infty^p(\hat{{\mathcal N}}\subset \hat{{\mathcal M}}, \hat{\phi})}-2\varepsilon. \end{align*} $$

Since $E_{{\mathcal M}_{n}}(\hat {{\mathcal N}})={\mathcal N}_n$ (see the commuting diagram after Lemma 4.5), we have

$$ \begin{align*}&\parallel \! E_{{\mathcal M}_{n_1}}(a)E_{{\mathcal M}_{n_1}}(a)^* \! \parallel_{\hat{\phi},p}\le \parallel \! E_{{\mathcal M}_{n_1}}(aa^*) \! \parallel_{\hat{\phi},p}\le \parallel \! aa^* \! \parallel_{\hat{\phi},p}= 1\\ &\parallel \! E_{{\mathcal M}_{n_3}}(b^*)E_{{\mathcal M}_{n_3}}(b) \! \parallel_{\hat{\phi},p}\le \parallel \! b^*b \! \parallel_{\hat{\phi},p}= 1\hspace{.1cm}, \\ & \parallel \! E_{{\mathcal M}_{n_4}}(y) \! \parallel_{\hat{\phi},q}\le \parallel \! y \! \parallel_{\hat{\phi},q}=1 \end{align*} $$

by the KMS- $\hat {\phi }$ -symmetry of $E_{{\mathcal M}_n}$ . Then, for $n\ge n_4=\max \{n_1,n_2,n_3,n_4\}$ ,

$$ \begin{align*} \parallel \! E_{{\mathcal M}_{n}}(x) \! \parallel_{L_\infty^p(\hat{{\mathcal N}}_n\subset \hat{{\mathcal M}}_n, \hat{\phi})}&\ge\, \parallel \! E_{{\mathcal M}_{n_4}}(x) \! \parallel_{L_\infty^p(\hat{{\mathcal N}}_n\subset \hat{{\mathcal M}}_n, \hat{\phi})}\\ &\ge \tau(d_{\hat{\phi}}^{1/2}E_{{\mathcal M}_{n_1}}(a)E_{{\mathcal M}_{n_4}}(x)E_{{\mathcal M}_{n_2}}(b) d_{\hat{\phi}}^{1/2}E_{{\mathcal M}_{n_3}}(y))\\ &\ge \,\parallel \! x \! \parallel_{L_\infty^p(\hat{{\mathcal N}}\subset \hat{{\mathcal M}}, \hat{\phi})}-2\varepsilon. \end{align*} $$

This proves

$$\begin{align*}\lim_{n}\|E_{{\mathcal M}_n}(x)\|_{L_\infty^p({\mathcal N}_n\subset {\mathcal M}_n, \psi_n)}= \|x\|_{L_\infty^p(\hat{{\mathcal N}}\subset \hat{{\mathcal M}}, \hat{\phi})}.\end{align*}$$

Finally, the assertion follows from

$$\begin{align*}\|x\|_{L_\infty^p(\hat{{\mathcal N}}\subset \hat{{\mathcal M}}, \hat{\phi})}=\|x\|_{L_\infty^p({\mathcal N}\subset {\mathcal M}, \phi)},\end{align*}$$

as a consequence of $E_0\circ \hat {E}=\hat {E}\circ E_0$ by Lemma 5.10.

Now we restate and prove Theorem 1.4.

Theorem 5.12. Let ${\mathcal M}$ be a $\sigma $ -finite von Neumann algebra and $T_t=e^{-tL}$ be a GNS- $\phi $ -symmetric quantum Markov semigroup. Suppose $T_t$ satisfies $\text {MLSI}$ with parameter $\alpha>0$ . There exists an universal constant c such that for $2\le p <\infty $ ,

$$\begin{align*}\alpha\|x-E(x)\|_{L_p({\mathcal M},\phi)}\le \alpha\|x-E(x)\|_{L_{\infty}^p({\mathcal N}\subset {\mathcal M},\phi)}\hspace{.1cm} \le \hspace{.1cm} c\sqrt{p}\parallel \! x \! \parallel_{\textrm{Lip}} .\end{align*}$$

Proof. We first show that if $T_t$ satisfies $\alpha $ -MLSI, so does the approximation semigroup.

$$\begin{align*}T_{n,t}=\hat{T}_t|_{{\mathcal M}_n}:{\mathcal M}_n\to {\mathcal M}_n.\end{align*}$$

Indeed, as we see in the discussion above, ${\mathcal M}_n\subset {\mathcal M}\rtimes _{\alpha ^{\phi }_t} 2^{-n}\mathbb {Z}\cong L_\infty (\mathbb {T}, {\mathcal M})$ , and the extension $\hat {T}_t=T_t\otimes \operatorname {\mathrm {\operatorname {id}}}_{\mathbb {T}}$ has $\alpha $ -MLSI (because $L_\infty (\mathbb {T})$ is a commutative space). Note that since ${\mathcal M}_n\subset {\mathcal M}\rtimes _{\alpha ^{\phi }_t} 2^{-n}\mathbb {Z}\subset {\mathcal M}\rtimes _{\alpha ^{\phi }_t} G$ , the restriction $E_{{\mathcal M}_n}:{\mathcal M}\rtimes _{\alpha ^{\phi }_t} 2^{-n}\mathbb {Z}\to {\mathcal M}_n$ is also a conditional expectation. Then for any $\rho ,\sigma \in S({\mathcal M}_n)$ , we have

$$\begin{align*}D(E_{{\mathcal M}_n,*}\rho|| E_{{\mathcal M}_n,*}\sigma)\le D(\rho|| \sigma)=D(\rho|_{{\mathcal M}_n}|| \sigma|_{{\mathcal M}_n})\le D(E_{{\mathcal M}_n,*}\rho|| E_{{\mathcal M}_n,*}\sigma).\end{align*}$$

Using the commutation relation $T_{n,t}\circ E_{{\mathcal M}_n}=E_{{\mathcal M}_n}\circ \hat {T}_t$ and $E_{{\mathcal M}_n}\circ \hat {E}=E_n\circ E_{{\mathcal M}_n} $ , we have for $\rho \in S({\mathcal M}_n)$

$$ \begin{align*}D(T_{t,n,*}\rho|| E_{n,*}\rho)&=D(E_{{\mathcal M}_n,*}T_{t,n,*}\rho|| E_{{\mathcal M}_n,*}E_{n,*}\rho) =D(\hat{T}_{t,*}E_{{\mathcal M}_n,*}\rho|| \hat{E}_{*}E_{{\mathcal M}_n,*}\rho)\\ &\le e^{-2\alpha t} D(E_{{\mathcal M}_n,*}\rho|| \hat{E}_{*} E_{{\mathcal M}_n,*}\rho)\\&= e^{-2\alpha t} D(E_{{\mathcal M}_n,*}\rho|| E_{{\mathcal M}_n,*}E_{n,*}\rho)= e^{-2\alpha t} D(\rho|| E_{n,*}\rho) . \end{align*} $$

Thus, $T_{n,t} $ has $\alpha $ - $\text {MLSI}$ on ${\mathcal M}_n$ . Note that $T_{n,t}$ is both GNS- $\hat {\phi }$ -symmetric for the extension state $\hat {\phi }$ and also symmetric for the trace $\psi _n$ . Now, we may use the tracial version of the concentration inequality [Reference Gao, Junge and LaRacuente29, Theorem 6.10] that for $x\in {\mathcal M}_n$ ,

$$\begin{align*}\alpha\parallel \! E_{{\mathcal M}_n}(x)-E_nE_{{\mathcal M}_n}(x) \! \parallel_{L_\infty^p({\mathcal N}_n\subset {\mathcal M}_n)}\le C\sqrt{p}\parallel \! E_{{\mathcal M}_n}(x) \! \parallel_{\text{Lip}}.\end{align*}$$

Now by the approximation of Lemma 5.11 and independence of $L_\infty ^p({\mathcal N}_n\subset {\mathcal M}_n)$ on the reference state, for $x\in {\mathcal M}$ ,

$$ \begin{align*} \parallel \! {x-E(x)} \! \parallel_{L_\infty^p({\mathcal N}\subset {\mathcal M},\phi)}&=\lim_{n}\parallel \! E_{{\mathcal M}_n}(x-E(x)) \! \parallel_{L_\infty^p({\mathcal N}_n\subset {\mathcal M}_n,\psi_n)}\\&=\lim_{n}\parallel \! E_{{\mathcal M}_n}(x)-E_{n}E_{{\mathcal M}_n}(x) \! \parallel_{L_\infty^p({\mathcal N}_n\subset {\mathcal M}_n,\psi_n)} \\ &\le C\sqrt{p}\parallel \! E_{{\mathcal M}_n}(x) \! \parallel_{\text{Lip}}\le C\sqrt{p}\parallel \! x \! \parallel_{\text{Lip}}. \end{align*} $$

The other inequality

$$\begin{align*}\parallel \! y \! \parallel_{L_\infty^p({\mathcal N}\subset {\mathcal M},\phi)}\ge \parallel \! y \! \parallel_{L_p({\mathcal M},\phi)}\end{align*}$$

is clear from definition of $L_\infty ^p({\mathcal N}\subset {\mathcal M},\phi )$ .

For Gaussian type concentration property, we introduce the following definition.

Definition 5.13. For an operator O, we say that

$$\begin{align*}\text{Prob}_{\phi}(|O|>t) \hspace{.1cm} \le \hspace{.1cm} \varepsilon \end{align*}$$

if there exists a projection e such that

$$\begin{align*}\|eOe\|_{\infty} \le t \quad \mbox{and} \quad \phi(1-e)\le \varepsilon .\end{align*}$$

The next lemma is a Chebyshev inequality for $\phi $ -weighted $L_p$ norm.

Lemma 5.14. Let $x\in L_p({\mathcal M},\phi )$ and $1<p<\infty $ . Then

$$\begin{align*}\text{Prob}_\phi(|x|>t) \hspace{.1cm} \le \hspace{.1cm} 2\Big ( \frac{t}{4} \Big )^{-p}\|x\|^{p}_{p,\phi} .\end{align*}$$

Proof. We start with a positive element $x=y^2$ and assume $\|x\|_{p,\phi }=M$ . Then we have

$$\begin{align*}M=\|x\|_{p,\phi} \hspace{.1cm} = \hspace{.1cm} \|d_{\phi}^{1/2p}xd_{\phi}^{1/2p}\|_p \hspace{.1cm} = \hspace{.1cm} \|yd_{\phi}^{1/2p}\|_{2p}^2 .\end{align*}$$

Recall the asymmetric Kosaki $L_p$ -space

$$\begin{align*}\parallel \! y \! \parallel_{L_{2p}^c({\mathcal M},\phi)}:=\|yd_{\phi}^{1/2p}\|_{2p} , \end{align*}$$

and the complex interpolation relation [Reference Junge and Parcet42]

$$\begin{align*}L_{2p}^c({\mathcal M},\phi)=[{\mathcal M},L_2^c({\mathcal M},\phi)]_{1/p} , \end{align*}$$

and the relation between real and complex interpolation

$$\begin{align*}L_{2p}^c({\mathcal M},\phi)=[{\mathcal M},L_2^c({\mathcal M},\phi)]_{1/p}\subset [{\mathcal M},L_2^c({\mathcal M},\phi)]_{1/p,\infty} . \end{align*}$$

By the definition of real interpolation space, for every $s>0$ , we have a decomposition $y \hspace {.1cm} = \hspace {.1cm} y_1+ y_2$ such that

$$\begin{align*}\|y_1\|_{\infty}+ s\parallel \! y_2 \! \parallel_{L_2^c({\mathcal M},\phi)} \hspace{.1cm} \le \hspace{.1cm} s^{1/p}M^{1/2} .\end{align*}$$

Then by Chebychev’s inequality for the spectral projection $e=e_{[0,a]}(y_2^*y_2)$ , we have

$$\begin{align*}a \phi(1-e)\hspace{.1cm} \le \hspace{.1cm} \phi(y_2^*y_2) \hspace{.1cm} \le \hspace{.1cm} s^{2/p-2}M \quad \mbox{and} \quad \|y_2e\|_{\infty}^2\hspace{.1cm} = \hspace{.1cm} \|ey_2^*y_2e\|_{\infty}\hspace{.1cm} \le \hspace{.1cm} a .\end{align*}$$

Choose $a=s^{2/p}M$ and deduce that

$$\begin{align*}\|exe\|_{\infty}=\|ye\|_{\infty}^2 \hspace{.1cm} \le \hspace{.1cm} (\|y_1e\|_{\infty}+ \|y_2e\|_{\infty})^2\hspace{.1cm} \le \hspace{.1cm} (s^{1/p}M^{1/2}+s^{1/p}M^{1/2})^{2} = 4a .\end{align*}$$

Then for $t=4a$ and

$$\begin{align*}\phi(1-e) \hspace{.1cm} \le \hspace{.1cm} a^{-1} s^{2/p-2}M \hspace{.1cm} = \hspace{.1cm} s^{-2} \hspace{.1cm} = \hspace{.1cm} (\frac{t}{4M})^{-p}=(\frac{t}{4})^{-p}M^p . \end{align*}$$

For an arbitrary x, we may write $x=x_1x_2$ such that

$$\begin{align*}\|d_\phi^{1/2p}x_1\|_{2p}\hspace{.1cm} = \hspace{.1cm} \|x_2d_\phi^{1/2p}\|_{2p} \hspace{.1cm} = \hspace{.1cm} \parallel \! x \! \parallel_{p,\phi}=M .\end{align*}$$

Then for each $s>0$ , we have decomposition

$$\begin{align*}x_1\hspace{.1cm} = \hspace{.1cm} x_{11}+x_{12} \hspace{.1cm} ,\hspace{.1cm} x_2=x_{21}+x_{22} \hspace{.1cm} \end{align*}$$

with

$$ \begin{align*}& \|x_{11}\|_{\infty}+ s\parallel \! x_{12} \! \parallel_{L_2^c({\mathcal M},\phi)}\hspace{.1cm} \le \hspace{.1cm} s^{1/p}M^{1/2} \hspace{.1cm} ,\hspace{.1cm} \|x_{21}\|_{\infty}+ s\parallel \! x_{22} \! \parallel_{L_2^c({\mathcal M},\phi)} \hspace{.1cm} \le \hspace{.1cm} s^{1/p}M^{1/2} .\end{align*} $$

We then use the Chebychev inequality for $e=e_{[0,a]}(x_{12}^{*}x_{12}+x_{22}^*x_{22})$ ,

$$\begin{align*}a\phi(1-e)\le \phi(x_{12}^{*}x_{12} +x_{22}^*x_{22}) \hspace{.1cm} \le \hspace{.1cm} 2s^{2/p-2}M .\end{align*}$$

Take $a=s^{2/p}M$ ,

$$ \begin{align*} \|exe\|_{\infty} &= \|e(x_1x_2)e\|_{\infty} \hspace{.1cm} = \hspace{.1cm} \|e(x_{11}+x_{12})(x_{21}+x_{22})e\|_{\infty} \\ & \le \hspace{.1cm} \|x_{11}x_{22}\|_{\infty} + \|ex_{12}x_{21}\|_{\infty}+ \|x_{11}x_{22}e\|_{\infty} + \|ex_{12}x_{22}e\|_{\infty} \\ &\le \hspace{.1cm} 4s^{2/p}M . \end{align*} $$

Thus, for $t=4s^{2/p}M$ , by Chebychev’s inequality for e,

$$ \begin{align*} \phi(1-e) &\leq \frac{1}{a}\phi(x_{12}^{*}x_{12}+x_{22}^{*}x_{22}) \le a^{-1} 2s^{2/p-2}M \hspace{.1cm} = \hspace{.1cm} 2s^{-2} \hspace{.1cm} = \hspace{.1cm} 2 (\frac{t}{4M})^{-p}.\\[-42pt] \end{align*} $$

Corollary 5.15. Let $T_t=e^{-tL}$ be a GNS- $\phi $ -symmetric quantum Markov semigroup. Suppose $T_t$ satisfies $\alpha $ -MLSI. Then for any $x\in {\mathcal M}$ and $t>0$ ,

$$\begin{align*}\text{Prob}_{\phi}(|x-E_{fix}(x)|>t) \hspace{.1cm} \le \hspace{.1cm} 2 \exp \Big ( -\frac{2}{e} \Big (\frac{\alpha t}{4c\parallel \! x \! \parallel_{\text{Lip}}} \Big )^2 \Big ), \end{align*}$$

where c is a universal constant as in Theorem 5.12.

Proof. By Lemma 5.14 and Theorem 5.12, we have

$$\begin{align*}\text{Prob}_{\phi}(|x-E(x)|>t)\le 2 (t/4)^{-p}\parallel \! x-E(x) \! \parallel_{L_p({\mathcal M},\phi)}^p\le 2 \big(\frac{4c\parallel \! x \! \parallel_{\text{Lip}}\sqrt{p}}{\alpha t}\big)^{p}.\end{align*}$$

Minimizing over p gives $p=\frac {1}{e}(\frac {\alpha t}{4c\parallel \! x \! \parallel _{\text {Lip}}})^2$ , which implies

$$ \begin{align*} &\text{Prob}_{\phi}(|x-E(x)|>t)\le 2 \exp( -\frac{\alpha^2t^2}{16ec^2\parallel \! x \! \parallel_{\text{Lip}}^2}).\\[-48pt] \end{align*} $$

Remark 5.16. In the ergodic case, the above results can be compared to [Reference Rouzé and Datta69, Theorem 8], which states that for self-adjoint $x=x^*$ ,

$$ \begin{align*} \phi(e_{\{|x-E(x)|>t\}})\le \exp\Big( -\frac{\alpha t^2}{8\parallel \! d_\phi^{-1/2}xd_\phi^{1/2} \! \parallel_{\tilde{\text{Lip}}}^2}\Big) \end{align*} $$

with a different Lipschitz norm $\parallel \! \cdot \! \parallel _{\tilde {\text {Lip}}}^2$ . Our Corollary 5.15 here uses a more natural definition of the Lipschitz norm and applies to non-ergodic cases. Nevertheless, the projection we have for

$$\begin{align*}\text{Prob}_{\phi}(|x-E(x)|>t)\end{align*}$$

is not necessarily a spectral projection $e_{\{|x-E(x)|>t\}}$ and will depend on the state $\phi $ .

Remark 5.17. In the operator valued setting, let ${\mathcal Q}$ be any finite von Neumann algebra and $T_t\otimes \operatorname {\mathrm {\operatorname {id}}}_{{\mathcal Q}}$ be the amplification semigroup on ${\mathcal Q}\overline {\otimes } {\mathcal M}$ . The conditional expectation for $T_t\otimes \operatorname {\mathrm {\operatorname {id}}}_{{\mathcal Q}}$ is $E\otimes \operatorname {\mathrm {\operatorname {id}}}_{{\mathcal Q}}$ . Note that by Lemma 4.17, $T_t\otimes \operatorname {\mathrm {\operatorname {id}}}_{{\mathcal Q}}$ is GNS-symmetric to the product state $ \phi \otimes \sigma $ , for any state $\sigma \in S({\mathcal Q})$ and any invariant state $\phi \in E_*(S({\mathcal M}))$ . This means we obtain

$$\begin{align*}\text{Prob}_{\phi\otimes \sigma}( |x-E_{fix}(x)|>t) \le 2 e^{-\frac{\alpha^2t^2}{C\parallel \! x \! \parallel_{\text{Lip}}^2}} \end{align*}$$

for any product state $\phi \otimes \sigma $ of this specific form. The projection of course depends on both $\phi $ and $\sigma $ .

We illustrate our result with a special case as matrix concentration inequalities.

Example 5.18 (Matrix concentration inequality).

Let $S_1,\cdots ,S_n$ be an independent sequence of random $d\times d$ -matrices $S_1,\cdots ,S_n$ such that

$$\begin{align*}\parallel \! S_i-\mathbb{E}S_i \! \parallel_{\infty}\le M \hspace{.1cm} , \hspace{.1cm} a.e. \end{align*}$$

Tropp in [Reference Tropp75, Corollary 6.1.2] proved the following matrix Bernstein inequality that for the sum ${Z=\sum _{k=1}^nS_k}$ ,

$$\begin{align*}{\mathbb E}\parallel \! Z-\mathbb{E}Z \! \parallel_{\infty}\le \sqrt{2v(Z)\log(2d)}+\frac{1}{3}M\log(2d)\end{align*}$$

and the matrix Chernoff bound

$$\begin{align*}P(|Z-\mathbb{E}Z|>t) \le 2d\exp\big(-\frac{t^2}{v(Z)+\frac{t}{3}M}\big),\end{align*}$$

where

$$\begin{align*}v(Z)=\max \{ \parallel \! {\mathbb E}((Z-\mathbb{E}Z)^*(Z-\mathbb{E}Z)) \! \parallel_{\hspace{.1cm}}, \parallel \! {\mathbb E}((Z-\mathbb{E}Z)^*(Z-\mathbb{E}Z)) \! \parallel_{}\}.\end{align*}$$

Now to apply our results, we recall that the depolarizing semigroup with generator $L(f):=(I-E_\mu )(f)=f-\mu (f)\textbf {1}_\Omega $ on any probability space $(\Omega ,\mu )$ has $\alpha _c\ge \frac {1}{2}$ (a simple fact by convexity of relative entropy). For a random matrix $f:\Omega \to {\mathbb M}_d$ , the Lipschitz norm is

(5.9) $$ \begin{align} \parallel \! f \! \parallel_{\text{Lip}}^2&= \frac{1}{2}\max\{\parallel \! \hat{f}^*\hat{f}+E_\mu(\hat{f}^*\hat{f}) \! \parallel_{\infty}\hspace{.1cm}, \hspace{.1cm} \parallel \! \hat{f}\hat{f}^*+E_\mu(\hat{f}\hat{f}^*) \! \parallel_{\infty}\}\nonumber\\ &\le \frac{1}{2}(\parallel \! f \! \parallel_{\infty}^2+v(f)), \end{align} $$

where $\hat {f}=f-E_\mu (f)$ is the mean zero part.

Now we consider for each $k=1,\cdots , n$ , $S_k:\Omega _k\to M_d$ as a random matrix on $(\Omega _k,\mu _k)$ . Then on the product space $(\Omega ,\mu )=(\Omega _1,\mu _1)\times \cdots \times (\Omega _n,\mu _n)$ , we have by Theorem 5.12 for $Z=\sum _{k}S_k$

$$ \begin{align*} {\mathbb E}\parallel \! Z-{\mathbb E} Z \! \parallel_{\infty}\le \Big ( \frac{1}{d}{\mathbb E}\parallel \! Z-{\mathbb E} Z \! \parallel_{p}^p \Big )^{1/p}\le d^{1/p}\parallel \! Z-{\mathbb E} Z \! \parallel_{L_\infty(M_d, L_p(\Omega) )} \le 2cd^{1/p}\sqrt{p}\parallel \! Z \! \parallel_{\text{Lip}} , \end{align*} $$

where $\parallel \! \cdot \! \parallel _{p}$ is the p-norm for the normalized trace ( ${\text {tr}}(1)=1$ ). Applying (5.9) and optimizing p gives

$$\begin{align*}{\mathbb E}\parallel \! Z-{\mathbb E} Z \! \parallel_{\infty}\le 2ce^{-1/2} \sqrt{ (v(Z)+M^2)\log d}.\end{align*}$$

For the matrix Chernoff bound, we use Corollary 5.15

$$\begin{align*}P(|Z-\mathbb{E}Z|>t)\le d\text{Prob}_{\mu\otimes \frac{{\text{tr}}}{d}}(|Z-\mathbb{E}Z|>t)\le 2 d\exp\Big( -\frac{t^2}{64ec^2(v(Z)+M^2)}\Big).\end{align*}$$

6 Final discussion

1. Positivity and complete positivity. The central quantity in this work is the CB return time $t_{cb}$ and $k_{cb}$ defined via complete positivity. Alternatively, one can consider positive maps and positivite mixing time. Indeed, the entropy difference Lemma 2.1

$$\begin{align*}D(\rho\|\Phi^*\Phi(\omega)) \le D_\Phi(\rho) + D(\rho\|\omega)\end{align*}$$

holds for a positive unital trace-preserving map $\Phi $ . This is because the operator concavity

$$ \begin{align*} \Phi(\ln x)\leq \ln \Phi(x),\quad \forall x\geq 0 \end{align*} $$

of the logarithmic function holds for any unital positive map $\Phi $ [Reference Choi18], and the monotonicity of relative entropy

$$ \begin{align*} D(\rho\|\sigma)\geq D(\Phi(\rho)\|\Phi(\sigma)) \end{align*} $$

was proved for any positive trace-preserving map $\Phi $ in [Reference Müller-Hermes and Reeb56] (see also [Reference Frenkel27]). Thus, both inequalities used in the proof of Lemma 2.1 hold for positive maps. Also, the conditions in Lemma 2.3 also only require positivity order

(6.1) $$ \begin{align} (1-\varepsilon)E\le\Psi \le (1+\varepsilon)E, \end{align} $$

where $\Phi \ge \Psi $ means $\Phi -\Psi $ is a positive map but not necessarily completely positive. Combining these two relaxed lemmas for positive maps, we have an analog of Theorem 1.1.

Theorem 6.1. i) For a positive unital trace-preserving map $\Phi :{\mathcal M}\to {\mathcal M}$ ,

$$\begin{align*}\alpha(\Phi)\le 1-\frac{1}{2k(\Phi)}\hspace{.1cm} \text{ where } \hspace{.1cm} k(\Phi):=\inf\{k\in \mathbb{N}^+ \hspace{.1cm} |\hspace{.1cm} 0.9 E\le \Phi^{2k}\le 1.1E\}.\end{align*}$$

ii) For a trace symmetric positive unital semigroup $T_t=e^{-tL}:{\mathcal M}\to {\mathcal M}$ ,3

$$\begin{align*}\alpha(L)\ge \frac{1}{2t(L)} \hspace{.1cm} \text{ where } \hspace{.1cm} t(L):=\inf\{t\in \mathbb{N}^+ \hspace{.1cm} |\hspace{.1cm} 0.9 E\le T_t\le 1.1E\}.\end{align*}$$

Applying the above theorem to $\Phi \otimes \operatorname {\mathrm {\operatorname {id}}}_{\mathcal {Q}}$ and $T_t\otimes \operatorname {\mathrm {\operatorname {id}}}_{\mathcal {Q}}$ for any finite von Neumann algebra $\mathcal {Q}$ actually yields our main Theorem 1.1 for trace symmetric cases. It remains open whether this observation holds for GNS-symmetric cases.

Problem 6.2. Does Theorem 6.1 with positivity conditions hold for GNS-symmetric cases?

The obstruction is that in the Haagerup reduction, we need the complete positivity and CB return time $k_{cb}(\Phi )$ of $\Phi $ to imply positivity and positivity mixing time $k(\Phi )$ of the extension $\hat {\Phi }$ , similar for the semigroup $T_t$ . One possible approach is to avoid using Haagerup reduction, and prove Lemma 4.6 directly.

The comparison between positivity and complete positivity has a deep root in the entanglement theory of quantum physics (see [Reference Chruściński and Pascazio20]). From the mathematical point of view, although the positivity looks a more flexible condition, it lacks connection to CB norms as Proposition 3.4. Indeed, there is no non-complete analog of Choi’s theorem [Reference Choi19]

$$\begin{align*}C_T\in ({\mathcal M}\otimes {\mathcal M}^{op})_+ \hspace{.1cm} \Longleftrightarrow \hspace{.1cm} T(x)=\tau \otimes \operatorname{\mathrm{\operatorname{id}}} (C_T(x\otimes 1))\hspace{.1cm} \text{ is } \hspace{.1cm} CP.\end{align*}$$

Therefore, despite that the estimate of $\alpha _1(L)$ only requires $t(L)$ , our kernel estimate Proposition 3.7 only applies to $t_{cb}(L)$ .

2. GNS and KMS symmetry. Both GNS-symmetry and KMS-symmetry are noncommutative generalizations for the detailed balance condition of classical Markov chains. As observed in [Reference Carlen and Maas15], GNS-symmetry is the strongest generalization of detail balance condition, and KMS is the weakest, which means the assumption of GNS-symmetry is the most restrictive. It is natural to ask whether our main results (c.f. Theorem 4.10 & 4.8) can be obtained for KMS-symmetric channels or semigroups.

Problem 6.3. Do entropy decay results Theorem 4.10 and 4.8 or the entropy difference Lemma 4.6 hold for KMS-symmetric maps?

The key property of a GNS-symmetric map $\Phi $ is the commutation with modular group $\Phi \circ \alpha _t^\phi =\alpha _t^\phi \circ \Phi $ . This has been used to ensure the compatibility of Haagerup reduction with channel and semigroups (see Lemma 4.5). One can ask whether the same commuting diagram Figure 1 can be obtained for KMS Markov maps. That will allow us to use Haagerup reduction to obtain the entropy difference Lemma (4.6) for KMS-symmetric channels. Another approach is, again, to avoid using Haagerup reduction and prove the KMS-case directly. At the moment of writing, this is not unclear to us even on finite dimensional matrix algebras.

From a mathematical physics perspective, it is also interesting to explore the relative entropy decay beyond GNS symmetry. For instance, one has a Lindbladian of the form $x \mapsto i[h, x] + L(x)$ such that L is GNS-symmetric and the adjoint action $\text {ad}(e^{-iht})$ commutes with L. Then the associated semigroup is $e^{-iht}e^{-tL}(\cdot )e^{iht}$ , which has the same entropy decay as $e^{-tL}$ . Such Lindbladians are considered in [Reference LaRacuente48]. Indeed, there is also numerical and theoretical evidence that adding an nonzero Hamiltonian part can destroy the exponential entropy decay. We refer to [Reference LaRacuente48] for more discussion on entropy decay beyond symmetry conditions.

3. MLSI and CMLSI constant. By the results of this work and also previous works [Reference Li, Junge and LaRacuente49, Reference Brannan, Gao and Junge14, Reference Gao and Gordina28], we now know the positivity of CMLSI constant $\alpha _c>0$ for many cases of classical Markov semigroups with the (non-complete) MLSI constant $\alpha>0$ . That is, $\alpha \ge \alpha _c>0$ for

  1. i) finite Markov chains [Reference Li, Junge and LaRacuente49, Reference Gao and Gordina28];

  2. ii) heat semigroups on manifold with curvature lower bound [Reference Brannan, Gao and Junge14];

  3. iii) sub-Laplacians of Hörmander system on a compact Riemannian manifold.

It remains open whether MLSI constant $\alpha $ and CMLSI constant $\alpha _c$ coincide for classical semigroups. This would be in the similar spirit that the bounded norm (resp. positivity) and the complete bounded norm (resp. complete positivity) coincide for a classical map on $L_\infty (\Omega ,\mu )$ .

Problem 6.4. Does $\alpha =\alpha _c$ for a classical symmetric Markov semigroup $T_t:L_\infty (\Omega ,\mu )\to L_\infty (\Omega ,\mu )$ ?

For a quantum Markov semigroup, a counterexample is the qubit depolarizing semigroup

$$\begin{align*}T_t:{\mathbb M}_2\to {\mathbb M}_2 \hspace{.1cm} , \hspace{.1cm} T_t(\rho)=e^{-t}\rho+(1-e^{-t})\frac{1}{2},\end{align*}$$

which has $\frac {1}{2}\le \alpha _c(T_t)<\alpha (T_t)=1$ because of entangled states [Reference Brannan, Gao and Junge14, Section 4.3]. It is natural to ask whether $\alpha _c<\alpha $ also holds for classical depolarizing channel.

Another interesting example is the heat semigroup on the unit torus $\mathbb {T}=\{z\in \mathbb {C}\hspace {.1cm} | \hspace {.1cm} |z|=1\}$ ,

$$\begin{align*}P_t: L_\infty(\mathbb{T})\to L_\infty(\mathbb{T})\hspace{.1cm}, P_t(z^n)=e^{-n^2t}z^n.\end{align*}$$

It was proved by [Reference Weissler78] that $\alpha (P_t)=\lambda (P_t)=1$ . The best known bound for CMLSI is $\alpha _c(P_t)\ge \frac {1}{6}$ . It is open whether the gap can be closed.

Problem 6.5. Does the heat semigroup $P_t$ on the torus $\mathbb {T}$ have $\alpha _c(P_t)=\alpha (P_t)=1$ ?

Acknowledgements

LG is grateful to the support of NSF grant DMS-2154903 and AMS-Simons Travel Grants. Nicholas LaRacuente was supported as an IBM Postdoc at The University of Chicago. MJ was partially supported by NSF Grant DMS-2247114 and NSF RAISE-TAQS 1839177. HL acknowledges support by the DFG cluster of excellence 2111 (Munich Center for Quantum Science and Technology).

Footnotes

1 Note that the LSI constant in [Reference Diaconis and Saloff-Coste23] is defined as half of our $\alpha _2$ here.

References

Alicki, R., ‘On the detailed balance condition for non-hamiltonian systems’, Rep. Math. Phys. 10(2) (1976), 249258.CrossRefGoogle Scholar
Araki, H., ‘Relative entropy of states of von neumann algebras’, Publ. Res. Inst. Math. Sci. 11(3) (1976), 809833.CrossRefGoogle Scholar
Bakry, D., ‘L’hypercontractivité et son utilisation en théorie des semigroupes’, in Lectures on Probability Theory (1994), 1114.Google Scholar
Bakry, D., Gentil, I., Ledoux, M., et al., Analysis and Geometry of Markov Diffusion Operators vol. 103 (Springer, New York, 2014).CrossRefGoogle Scholar
Bardet, I., ‘Estimating the decoherence time using noncommutative functional inequalities’, Preprint, 2017, arXiv:1710.01039.Google Scholar
Bardet, I., Capel, Á., Gao, L., Lucia, A., Pérez-García, D. and Rouzé, C., ‘Entropy decay for davies semigroups of a one dimensional quantum lattice’, Communications in Mathematical Physics. 405(2) 2024, 42.CrossRefGoogle Scholar
Bardet, I. and Rouzé, C., ‘Hypercontractivity and logarithmic sobolev inequality for non-primitive quantum markov semigroups and estimation of decoherence rates’, in Annnales Henri Poincar´e (Springer, New York, 2022), 165.Google Scholar
Baudoin, F. and Bonnefont, M., ‘The subelliptic heat kernel on su (2): representations, asymptotics and gradient bounds’, Math. Z. 263(3) (2009), 647672.CrossRefGoogle Scholar
Beigi, S., Datta, N. and Rouzé, C., ‘Quantum reverse hypercontractivity: Its tensorization and application to strong converses’, Comm. Math. Phys. 376 (2020), 753794.CrossRefGoogle Scholar
Blecher, D. P. and Paulsen, V. I., ‘Tensor products of operator spaces’, J. Funct. Anal. 99(2) (1991), 262292.CrossRefGoogle Scholar
Bobkov, S. G. and Tetali, P., ‘Modified logarithmic sobolev inequalities in discrete settings’, J. Theoret. Probab. 19(2) (2006), 289336.CrossRefGoogle Scholar
Brandao, F. G. S. L., Harrow, A. W. and Horodecki, M., ‘Local random quantum circuits are approximate polynomial-designs’, Commun. Math. Phys. 346 (2016), 397434.CrossRefGoogle Scholar
Brannan, M., Gao, L. and Junge, M., ‘Complete logarithmic sobolev inequality via ricci curvature bounded below ii’, J. Topol. Anal. (2021), 154.Google Scholar
Brannan, M., Gao, L. and Junge, M.., ‘Complete logarithmic sobolev inequalities via ricci curvature bounded below’, Adv. Math. 394 (2022), 108129.CrossRefGoogle Scholar
Carlen, E. A. and Maas, J.., ‘Gradient flow and entropy inequalities for quantum markov semigroups with detailed balance’, J. Funct. Anal. 273(5) (2017), 18101869.CrossRefGoogle Scholar
Carlen, E. A. and Maas, J., ‘Non-commutative calculus, optimal transport and functional inequalities in dissipative quantum systems’, J. Stat. Phys. 178(2) (2020), 219378.CrossRefGoogle ScholarPubMed
Chen, M.-F., Eigenvalues, Inequalities, and Ergodic Theory (Probability and Its Applications) (Springer-Verlag London, Ltd., London, 2005).Google Scholar
Choi, M.-D., ‘A schwarz inequality for positive linear maps on C*-algebras’, Illinois J. Math. 18(4) (1974), 565574.CrossRefGoogle Scholar
Choi, M.-D., ‘Completely positive linear maps on complex matrices’, Linear Algebra Appl. 10(3) (1975), 285290.CrossRefGoogle Scholar
Chruściński, D. and Pascazio, S., ‘A brief history of the gkls equation’, Open Syst. Inf. Dyn. 24(03) (2017), 1740001.CrossRefGoogle Scholar
Davies, E. B. and Lindsay, J. M., ‘Non-commutative symmetric markov semigroups’, Math. Z. 210(1) (1992), 379411.CrossRefGoogle Scholar
Del Moral, P., Ledoux, M. and Miclo, L., ‘On contraction properties of markov kernels’, Probab. Theory Related Fields 126(3) (2003), 395420.CrossRefGoogle Scholar
Diaconis, P. and Saloff-Coste, L., ‘Logarithmic sobolev inequalities for finite markov chains’, Ann. Appl. Probab. 6(3) (1996), 695750.CrossRefGoogle Scholar
Driver, B. K. and Melcher, T., ‘Hypoelliptic heat kernel inequalities on the heisenberg group’, J. Funct. Anal. 221(2) (2005), 340365.CrossRefGoogle Scholar
Effros, E. G. and Ruan, Z.-J.. Operator Spaces no. 23 (Oxford University Press on Demand, 2000).Google Scholar
Erbar, M. and Maas, J., ‘Ricci curvature of finite markov chains via convexity of the entropy’, Arch. Ration. Mech. Anal. 206 (2012), 9971038.CrossRefGoogle Scholar
Frenkel, P. E., ‘Integral formula for quantum relative entropy implies data processing inequality’, Quantum 7 (2023), p. 1102.Google Scholar
Gao, L. and Gordina, M., ‘Complete modified logarithmic sobolev inequality for sub-laplacian on su(2)’, Journal of Functional Analysis, Accepted 2022, arXiv:2203.12731.Google Scholar
Gao, L., Junge, M. and LaRacuente, N., ‘Fisher information and logarithmic sobolev inequality for matrix-valued functions’, in Annales Henri Poincaré vol. 21 (Springer, New York, 2020), 34093478.Google Scholar
Gao, L., Junge, M. and LaRacuente, N., ‘Relative entropy for von neumann subalgebras’, Int. J. Math. 31(06) (2020), 2050046.CrossRefGoogle Scholar
Gao, L. and Rouzé, C., ‘Complete entropic inequalities for quantum markov chains’, Arch. Ration. Mech. Anal. (2022), 156.Google Scholar
Giorgetti, L., Parzygnat, A. J., Ranallo, A. and Russo, B. P., ‘Bayesian inversion and the tomita–takesaki modular group’, Quart. J. Math. 74(3) (2023), 9751014.CrossRefGoogle Scholar
Gorini, V., Kossakowski, A. and Sudarshan, E. C. G., ‘Completely positive dynamical semigroups of n-level systems’, J. Math. Phys. 17(5) (1976), 821825.CrossRefGoogle Scholar
Gross, L., ‘Hypercontractivity and logarithmic sobolev inequalities for the clifford-dirichlet form’, Duke Math. J. 42(3) (1975), 383396.CrossRefGoogle Scholar
Gross, L., ‘Logarithmic sobolev inequalities’, Amer. J. Math. 97(4) (1975), 10611083.CrossRefGoogle Scholar
Haagerup, U., Junge, M. and Xu, Q., ‘A reduction method for noncommutative lp-spaces and applications’, Trans. Amer. Math. Soc. 362(4) (2010), 1252165.Google Scholar
Hiai, F., Quantum f-Divergences in von Neumann Algebras (Springer, New York, 2021).CrossRefGoogle Scholar
Huang, D. and Tropp, J. A., ‘From Poincaré inequalities to nonlinear matrix concentration’, Bernoulli 27(3) (2021), 17241744.CrossRefGoogle Scholar
Huang, D. and Tropp, J. A., ‘Nonlinear matrix concentration via semigroup methods’, Electronic Journal of Probability 26 (2021).CrossRefGoogle Scholar
Huber, S., K¨onig, R. and Vershynina, A., ‘Geometric inequalities from phase space translations’, J. Math. Phys. 58(1) (2017), 012206.CrossRefGoogle Scholar
Junge, M., LaRacuente, N. and Rouzé, C., ‘Stability of logarithmic sobolev inequalities under a noncommutative change of measure’, Journal of Statistical Physics, 190(2) (2023), p. 30.CrossRefGoogle Scholar
Junge, M. and Parcet, J., Mixed-Norm Inequalities and Operator Space L p Embedding Theory (American Mathematical Soc., 2010).CrossRefGoogle Scholar
Kastoryano, M. J. and Temme, K., ‘Quantum logarithmic sobolev inequalities and rapid mixing’, J. Math. Phys. 54(5) (2013), 052202.CrossRefGoogle Scholar
König, R. and Smith, G., ‘The entropy power inequality for quantum systems’, IEEE Trans. Inform. Theory 60(3) (2014) 15361548.CrossRefGoogle Scholar
Kossakowski, A., Frigerio, A., Gorini, V. and Verri, M., ‘Quantum detailed balance and kms condition’, Comm. Math.Phys. 57(2) (1977), 97110.CrossRefGoogle Scholar
LaRacuente, N., ‘Quasi-factorization and multiplicative comparison of subalgebra-relative entropy’, J. Math. Phys. 63(12) (2022).CrossRefGoogle Scholar
LaRacuente, N., Quasi-factorization and Multiplicative Comparison of Subalgebra-Relative Entropy, Journal of Mathematical Physics, 63(12) 2022.CrossRefGoogle Scholar
LaRacuente, N., ‘Self-restricting noise and exponential decay in quantum dynamics’, Preprint, 2022, arXiv:2203.03745.Google Scholar
Li, H., Junge, M. and LaRacuente, N.. ‘Graph hörmander systems’, In Annales Henri Poincaré (Springer International Publishing, Cham, 2024), 154.Google Scholar
Lindblad, G., ‘On the generators of quantum dynamical semigroups’, Comm. Math. Phys. 48(2) (1976), 119130.CrossRefGoogle Scholar
Lott, J. and Villani, C., ‘Ricci curvature for metric-measure spaces via optimal transport’, Ann. Math. (2009), 903991.CrossRefGoogle Scholar
Luczak, A., ‘Mixing and asymptotic properties of markov semigroups on von neumann algebras’, Math. Z. 235(3) (2000), 615626.Google Scholar
Lugiewicz, P. and Zegarliński, B., ‘Coercive inequalities for h¨ormander type generators in infinite dimensions’, J. Funct. Anal. 247(2) (2007), 438476.CrossRefGoogle Scholar
Melcher, T., ‘Hypoelliptic heat kernel inequalities on lie groups’, Stochastic Process. Appl. 118(3) (2008), 368388.CrossRefGoogle Scholar
Miclo, L., ‘An example of application of discrete Hardy’s inequalities’, Markov Process. Related Fields 5(3) (1999), 319330.Google Scholar
Müller-Hermes, A. and Reeb, D., ‘Monotonicity of the quantum relative entropy under positive maps’, in Annales Henri Poincaré vol. 18 (Springer, New York, 2017), 17771788.Google Scholar
Müller-Hermes, A., Stilck França, D. and Wolf, M. M., ‘Entropy production of doubly stochastic quantum channels’, J. Math. Phys. 57(2) (2016), 022203.CrossRefGoogle Scholar
Nelson, E., ‘A quartic interaction in two dimensions’, in Mathematical Theory of Elementary Particles, Proc. Conf., Dedham, Mass., 1965 (MIT Press, 1966), 6973.Google Scholar
Nelson, E., ‘Construction of quantum fields from markoff fields’, J. Funct. Anal. 12(1) (1973), 97112.CrossRefGoogle Scholar
Olkiewicz, R. and Zegarlinski, B., ‘Hypercontractivity in noncommutative lpspaces, J. Funct. Anal. 161(1) (1999), 246285.CrossRefGoogle Scholar
Otto, F. and Villani, C., ‘Generalization of an inequality by talagrand and links with the logarithmic sobolev inequality’, J. Funct. Anal. 173(2) (2000), 361400.CrossRefGoogle Scholar
Parzygnat, A. J., ‘Inverses, disintegrations, and bayesian inversion in quantum markov categories’, Preprint, 2020, arXiv:2001.08375.Google Scholar
Parzygnat, A. J. and Buscemi, F., ‘Axioms for retrodiction: Achieving time-reversal symmetry with a prior’, Quantum 7 (2023), 1013.CrossRefGoogle Scholar
Parzygnat, A. J. and Fullwood, J., ‘From time-reversal symmetry to quantum bayes’ rules’, PRX Quantum 4(2) (2023), 020334.CrossRefGoogle Scholar
Parzygnat, A. J. and Russo, B. P., ‘A non-commutative bayes’ theorem’, Linear Algebra Appl. 644 (2022), 2894.CrossRefGoogle Scholar
Petz, D., ‘On certain properties of the relative entropy of states of operator algebras’, Math. Z. 206(1) (1991), 351361.CrossRefGoogle Scholar
Pimsner, M. and Popa, S., ‘Entropy and index for subfactors’, in Annales scientifiques de l’Ecole normale supérieure vol. 19 (1986), 57106.CrossRefGoogle Scholar
Rothaus, O. S., ‘Analytic inequalities, isoperimetric inequalities and logarithmic sobolev inequalities’, J. Funct. Anal. 64(2) (1985), 296313.CrossRefGoogle Scholar
Rouzé, C. and Datta, N., ‘Concentration of quantum states from quantum functional and transportation cost inequalities’, J. Math. Phys. 60(1) (2019), 012202.CrossRefGoogle Scholar
Smith, R. R., ‘Completely bounded maps between c*-algebras’, J. Lond. Math. Soc. 2(1) (1983), 157166.CrossRefGoogle Scholar
Sturm, K.-T., ‘On the geometry of metric measure spaces’, Acta Math. 196(1) (2006), 65131.CrossRefGoogle Scholar
Takesaki, M., ‘Conditional expectations in von neumann algebras’, J. Funct. Anal. 9(3) (1972), 306321.CrossRefGoogle Scholar
Takesaki, M. et al, Theory of Operator Algebras II vol. 125 (Springer, New York, 2003).Google Scholar
Temme, K., Pastawski, F. and Kastoryano, M. J., ‘Hypercontractivity of quasi-free quantum semigroups’, J. Phys. A 47(40) (2014), 405303.CrossRefGoogle Scholar
Tropp, J. A. et al, ‘An introduction to matrix concentration inequalities’, Foundations and Trends® in Machine Learning 8(1–2) (2015), 1230.Google Scholar
Umegaki, H., ‘Conditional expectation in an operator algebra, iv (entropy and information)’, in Kodai Mathematical Seminar Reports vol. 14 (Department of Mathematics, Tokyo Institute of Technology, 1962), 5985.Google Scholar
Varopoulos, N. Th., Saloff-Coste, L. and Coulhon, Th., ‘Analysis and geometry on groups’, in Proceedings of the International Congress of Mathematicians vol. 1 (1991), 951957.Google Scholar
Weissler, F. B., ‘Logarithmic sobolev inequalities and hypercontractive estimates on the circle’, J. Funct. Anal. 37(2) (1980), 218234.CrossRefGoogle Scholar
Wirth, M., ‘A noncommutative transport metric and symmetric quantum markov semigroups as gradient flows of the entropy’, Preprint, 2018, arXiv:1808.05419.Google Scholar
Wirth, M., ‘Christensen-evans theorem and extensions of gns-symmetric quantum markov semigroups’, Journal of Functional Analysis, 287(3) (2024), p. 110475.CrossRefGoogle Scholar
Wirth, M. and Zhang, H., ‘Complete gradient estimates of quantum markov semigroups’, Comm. Math. Phys. 387(2) (2021), 761791.CrossRefGoogle ScholarPubMed
Figure 0

Figure 1 Haagerup reduction of quantum Markov map and conditional expectation.