This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

Curvature-dimension conditions for symmetric quantum Markov semigroups

Melchior Wirth Institute of Science and Technology Austria (IST Austria), Am Campus 1, 3400 Klosterneuburg, Austria [email protected]  and  Haonan Zhang Institute of Science and Technology Austria (IST Austria), Am Campus 1, 3400 Klosterneuburg, Austria [email protected]
Abstract.

Following up on the recent work on lower Ricci curvature bounds for quantum systems, we introduce two noncommutative versions of curvature-dimension bounds for symmetric quantum Markov semigroups over matrix algebras. Under suitable such curvature-dimension conditions, we prove a family of dimension-dependent functional inequalities, a version of the Bonnet–Myers theorem and concavity of entropy power in the noncommutative setting. We also provide examples satisfying certain curvature-dimension conditions, including Schur multipliers over matrix algebras, Herz–Schur multipliers over group algebras and depolarizing semigroups.

2020 Mathematics Subject Classification:
Primary 81S22; Secondary 46L57, 47C99, 49Q22, 81R05

1. Introduction

Starting with the celebrated work by Lott–Villani [LV09] and Sturm [Stu06a, Stu06b], recent years have seen a lot of research interest in extending the notion of Ricci curvature, or more precisely lower Ricci curvature bounds, beyond the realm of classical differential geometry to spaces with singularities [AGS14a, AGS14b, AGS15, EKS15], discrete spaces [Maa11, EM12, Mie13] or even settings where there is no underlying space at all as for example in noncommutative geometry [CM17, MM17, Wir18, CM20, LJL20, WZ20, DR20].

Most of these approaches take as their starting point either the characterization of lower Ricci curvature bound in terms of convexity properties of the entropy on Wasserstein space [vRS05] or in terms of Bakry–Émery’s Γ2\Gamma_{2}-criterion [BÉ85], which derives from Bochner’s formula, and in many settings, these two approaches yield equivalent or at least closely related notions of lower Ricci curvature bounds.

One of the reasons to seek to extend the notion of Ricci curvature beyond Riemannian manifolds is that lower Ricci curvature bounds have strong geometric consequences and are a powerful tool in proving functional inequalities. This motivated the investigation of lower Ricci curvature bounds in the noncommutative setting, or for quantum Markov semigroups.

From a positive noncommutative lower Ricci curvature bound in terms of the Γ2\Gamma_{2}-condition, Junge and Zeng [JZ15a, JZ15b] derived a LpL_{p}-Poincaré-type inequality and transportation inequalities, and under such non-negative lower Ricci curvature bounds Junge and Mei proved LpL_{p}-boundedness of Riesz transform [JM10]. Following Lott–Sturm–Villani, Carlen and Maas [CM14, CM17, CM20] studied the noncommutative lower Ricci curvature bound via the geodesic semi-convexity of entropy by introducing a noncommutative analog of the 22-Wasserstein metric. The similar approach was carried out by the first-named author in the infinite-dimensional setting in [Wir18]. These two notions of lower Ricci curvature bounds are in general different, but they can both be characterized in terms of a gradient estimate [Wir18, CM20, WZ20]. A stronger notion of lower Ricci curvature bound, which implies the bound in terms of Γ2\Gamma_{2}-condition and in terms of transportation, was introduced by Li, Junge and LaRacuente [LJL20]. See also the further work of Li [Li20], and Brannan, Gao and Junge [BGJ20a, BGJ20b].

However, for many applications in geometric consequences such as the Bonnet–Myers theorem, and functional inequalities such as the concavity of entropy power, a lower bound on the Ricci curvature is not sufficient, but one needs an upper bound on the dimension as well. This leads to the curvature-dimension condition, whose noncommutative analog will be the main object of this article. As a finite-dimensional analog of lower Ricci curvature bounds, the curvature-dimension condition also admits various characterizations. Similar to the “infinite-dimensional” setting, two main approaches describing curvature-dimension conditions are Γ2\Gamma_{2}-criterion following Bakry–Émery and convexity properties of entropy on the 22-Wasserstein space in the spirit of Lott–Sturm–Villani. For metric measure spaces, the equivalence of various characterizations on curvature-dimension conditions and their applications have been extensively studied beginning with [EKS15].

While the notion of dimension is built into the definition of manifolds, it is not obvious in the extended settings and requires new definitions. The goal of this article is to provide such a definition of dimension (upper bounds) in the context of quantum Markov semigroups in a way that it fits well with the previously developed notions of lower Ricci curvature bounds in this framework. This definition allows us to prove interesting consequences on the geometry of the state space as well as some functional inequalities.

Furthermore, for quantum Markov semigroups satisfying an intertwining condition, which already appeared in [CM17] and subsequent work, we provide an easily verifiable upper bound on the dimension, namely the number of partial derivatives in the Lindblad form of the generator. This sufficient condition enables us to prove the curvature-dimension condition in various concrete examples such as quantum Markov semigroups of Schur multipliers and semigroups generated by conditionally negative definite length functions on group algebras.

It should be mentioned that a notion of dimension for a quantum diffusion semigroup already appeared implicitly in the work of König and Smith on the quantum entropy power inequality [KS14]. In particular, from their entropy power inequality one may also derive the concavity of entropy power for the associated quantum diffusion semigroup. See [DPT18, HKV17, AB20] for more related work. This example fits conceptually well with our framework as it satisfies the intertwining condition and the dimension in the entropy power considered there is the number of partial derivatives in the Lindblad form of the generator, although the semigroup acts on an infinite-dimensional algebra and is therefore not covered by our finite-dimensional setting. Here we consider the concavity of the entropy power for arbitrary symmetric quantum Markov semigroups over matrix algebras.

In this paper we will focus on two noncommutative analogues of curvature-dimension conditions: the Bakry–Émery curvature dimension condition BE(K,NK,N), formulated via the Γ2\Gamma_{2}-condition, and the gradient estimate GE(K,NK,N), which is in the spirit of Lott–Sturm–Villani when the reference operator mean is chosen to be the logarithmic mean. They are generalizations of “infinite-dimensional” notions BE(K,K,\infty) and GE(K,K,\infty) in previous work, but let us address one difference in the “finite-dimensional” setting, i.e. N<N<\infty. As we mentioned above, in the “infinite-dimensional” case, i.e. N=N=\infty, GE(K,K,\infty) recovers BE(K,K,\infty) if the operator mean is the left/right trivial mean. However, this is not the case when N<N<\infty; BE(K,NK,N) is stronger than GE(K,NK,N) for the left/right trivial mean.

This article is organized as follows. Section 2 collects preliminaries about quantum Markov semigroups and noncommutative differential calculus that are needed for this paper. In Section 3 we study the noncommutative Bakry–Émery curvature-dimension condition BE(K,NK,N), its applications and the complete version. In Section 4 we investigate the noncommutative gradient estimate GE(K,NK,N) for arbitrary operator means, give an equivalent formulation in the spirit of the Γ2\Gamma_{2}-criterion, and also introduce their complete form. Section 5 is devoted to the gradient estimate GE(K,NK,N), its connection to the geodesic (K,N)(K,N)-convexity of the (relative) entropy and applications to dimension-dependent functional inequalities. In Section 6 we give some examples of quantum Markov semigroups for which our main results apply. In Section 7 we discuss how to extend the theory from this article to quantum Markov semigroups that are not necessarily tracially symmetric and explain the main challenge in this case.

Acknowledgments

H.Z. is supported by the European Union’s Horizon 2020 research and innovation programme under the Marie Skłodowska-Curie grant agreement No. 754411. M.W. acknowledges support from the European Research Council (ERC) under the European Union’s Horizon 2020 research and innovation programme (grant agreement No 716117) and from the Austrian Science Fund (FWF) through grant number F65. Both authors would like to thank Jan Maas for fruitful discussions and helpful comments.

2. Quantum Markov semigroups and noncommutative differential calculus

In this section we give some background material on quantum Markov semigroups, their generators, first-order differential calculus and operator means.

2.1. Quantum Markov semigroups

Throughout we fix a finite-dimensional von Neumann algebra \mathcal{M} with a faithful tracial state τ\tau. By the representation theory of finite-dimensional CC^{\ast}-algebras, \mathcal{M} is of the form j=1nMkj()\bigoplus_{j=1}^{n}M_{k_{j}}(\mathbb{C}) and τ=j=1nαjtrMkj()\tau=\bigoplus_{j=1}^{n}\alpha_{j}\mathrm{tr}_{M_{k_{j}}(\mathbb{C})} with αj0\alpha_{j}\geq 0, j=1nαjkj=1\sum_{j=1}^{n}\alpha_{j}k_{j}=1. Here Mn()M_{n}(\mathbb{C}) denotes the full nn-by-nn matrix algebra and trMn()\mathrm{tr}_{M_{n}(\mathbb{C})} is the usual trace over Mn()M_{n}(\mathbb{C}).

Denote by +\mathcal{M}_{+} the set of positive semi-definite matrices in \mathcal{M}. A density matrix is a positive element ρ\rho\in\mathcal{M} with τ(ρ)=1\tau(\rho)=1. The set of all density matrices is denoted by 𝒮()\mathcal{S(M)} and the set of all invertible density matrices by 𝒮+()\mathcal{S}_{+}(\mathcal{M}). We write L2(,τ)L_{2}(\mathcal{M},\tau) for the Hilbert space obtained by equipping \mathcal{M} with the inner product

,:×,(x,y)τ(xy).\langle\cdot,\cdot\rangle\colon\mathcal{M}\times\mathcal{M}\to\mathbb{C},\,(x,y)\mapsto\tau(x^{\ast}y).

The adjoint of a linear operator T:T\colon\mathcal{M}\to\mathcal{M} with respect to this inner product is denoted by TT^{\dagger}.

A family (Pt)t0(P_{t})_{t\geq 0} of linear operators on \mathcal{M} is called a quantum Markov semigroup if

  1. (a)

    P0=idP_{0}=\mathrm{id}_{\mathcal{M}}, Ps+t=PsPtP_{s+t}=P_{s}P_{t} for s,t0s,t\geq 0,

  2. (b)

    PtP_{t} is completely positive for every t0t\geq 0,

  3. (c)

    Pt𝟏=𝟏P_{t}\mathbf{1}=\mathbf{1} for every t0t\geq 0,

  4. (d)

    tPtt\mapsto P_{t} is continuous.

The generator of (Pt)(P_{t}) is

:,x=limt01t(xPt(x)).\mathcal{L}\colon\mathcal{M}\to\mathcal{M},\,\mathcal{L}x=\lim_{t\searrow 0}\frac{1}{t}(x-P_{t}(x)).

It is the unique linear operator on \mathcal{M} such that Pt=etP_{t}=e^{-t\mathcal{L}}. Let us remark that sign conventions differ and sometimes -\mathcal{L} is called the generator of (Pt)(P_{t}).

Let σ𝒮+()\sigma\in\mathcal{S}_{+}(\mathcal{M}). The quantum Markov semigroup (Pt)(P_{t}) is said to satisfy the σ\sigma-detailed balance condition (σ\sigma-DBC) if

τ(Pt(x)yσ)=τ(xPt(y)σ)\tau(P_{t}(x)y\sigma)=\tau(xP_{t}(y)\sigma)

for x,yx,y\in\mathcal{M} and t0t\geq 0. In the special case σ=𝟏\sigma=\mathbf{1} we say that (Pt)(P_{t}) is tracially symmetric or symmetric, and denote

(a,b):=a,b.\mathcal{E}(a,b):=\langle a,\mathcal{L}b\rangle.

A tracially symmetric quantum Markov semigroup (Pt)(P_{t}) is ergodic if 𝟏\mathbf{1} is the unique invariant state of (Pt)(P_{t}).

Although it is not necessary to formulate the curvature-dimension conditions, we will deal exclusively with tracially symmetric quantum Markov semigroups since all examples where we can verify the conditions fall into that class. As a special case of Alicki’s theorem [Ali76, Theorem 3] (see also [CM17, Theorem 3.1]) the generator \mathcal{L} of a tracially symmetric quantum Markov semigroup on =Mn()\mathcal{M}=M_{n}(\mathbb{C}) is of the form

=j𝒥jj,\mathcal{L}=\sum_{j\in\mathcal{J}}\partial_{j}^{\dagger}\partial_{j},

where 𝒥\mathcal{J} is a finite index set, j=[vj,]\partial_{j}=[v_{j},\cdot\,] for some vjv_{j}\in\mathcal{M}, and for every j𝒥j\in\mathcal{J} there exists a unique j𝒥j^{\ast}\in\mathcal{J} such that vj=vjv_{j}^{\ast}=v_{j^{\ast}}. We call the operators j\partial_{j} partial derivatives. Using the derivation operator :=(j)j𝒥:^:=j𝒥\partial:=(\partial_{j})_{j\in\mathcal{J}}:\mathcal{M}\to\hat{\mathcal{M}}:=\oplus_{j\in\mathcal{J}}\mathcal{M}, we may also write =\mathcal{L}=\partial^{\dagger}\partial.

2.2. Noncommutative differential calculus and operator means

Let us shortly recall the definition and some basic properties of operator means. Let \mathcal{H} be an infinite-dimensional Hilbert space. A map Λ:B()+×B()+B()+\Lambda\colon B(\mathcal{H})_{+}\times B(\mathcal{H})_{+}\to B(\mathcal{H})_{+} is called an operator connection if it satisfies the following properties.

  1. (a)

    monotonicity: if ACA\leq C and BDB\leq D, then Λ(A,B)Λ(C,D)\Lambda(A,B)\leq\Lambda(C,D),

  2. (b)

    transformer inequality: CΛ(A,B)CΛ(CAC,CBC)C\Lambda(A,B)C\leq\Lambda(CAC,CBC) for any A,B,CB()+A,B,C\in B(\mathcal{H})_{+},

  3. (c)

    continuity: AnAA_{n}\searrow A and BnBB_{n}\searrow B imply Λ(An,Bn)Λ(A,B)\Lambda(A_{n},B_{n})\searrow\Lambda(A,B).

An operator connection Λ\Lambda is called an operator mean if it additionally satisfies

  1. (d)

    Λ(id,id)=id\Lambda(\mathrm{id}_{\mathcal{H}},\mathrm{id}_{\mathcal{H}})=\mathrm{id}_{\mathcal{H}}.

Here by AnAA_{n}\searrow A we mean A1A2A_{1}\geq A_{2}\geq\cdots and AnA_{n} converges strongly to AA. The operator connection Λ\Lambda is symmetric if Λ(A,B)=Λ(B,A)\Lambda(A,B)=\Lambda(B,A) for all A,BB()+A,B\in B(\mathcal{H})_{+}.

Lemma 2.1.

Let Λ\Lambda be an operator connection. Then for λ0\lambda\geq 0, A,B,C,DB()+A,B,C,D\in B(\mathcal{H})_{+} and unitary UB()U\in B(\mathcal{H}), we have

  1. (a)

    positive homogeneity: Λ(λA,λB)=λΛ(A,B)\Lambda(\lambda A,\lambda B)=\lambda\Lambda(A,B),

  2. (b)

    concavity: Λ(A,C)+Λ(B,D)Λ(A+B,C+D)\Lambda(A,C)+\Lambda(B,D)\leq\Lambda(A+B,C+D),

  3. (c)

    unitary invariance: Λ(UAU,UBU)=UΛ(A,B)U\Lambda(U^{\ast}AU,U^{\ast}BU)=U^{\ast}\Lambda(A,B)U.

If Λ\Lambda is an operator mean, then additionally

  1. (d)

    Λ(A,A)=A\Lambda(A,A)=A.

Proof.

See equations (II0), (2.1), Theorem 3.3 and Theorem 3.5 in [KA80]. ∎

While operator connections are initially only defined for bounded operators on an infinite-dimensional Hilbert space, one can easily extend this definition to operators on finite-dimensional Hilbert spaces as follows. If Λ\Lambda is an operator connection, \mathcal{H} is a finite-dimensional Hilbert space and A,BB()+A,B\in B(\mathcal{H})_{+}, then one can define Λ(A,B)\Lambda(A,B) as VΛ(VAV,VBV)VV^{\ast}\Lambda(VAV^{\ast},VBV^{\ast})V, where VV is an isometric embedding of \mathcal{H} into an infinite-dimensional Hilbert space. The unitary invariance from the previous lemma ensures that this definition does not depend on the choice of the embedding VV.

Let L(ρ)L(\rho) and R(ρ)R(\rho) be the left and right multiplication operators, respectively, and fix an operator mean Λ\Lambda. For ρ+\rho\in\mathcal{M}_{+} we define

ρ^=Λ(L(ρ),R(ρ)).\hat{\rho}=\Lambda(L(\rho),R(\rho)).

Of particular interest for us are the cases when Λ\Lambda is the logarithmic mean

Λlog(L(ρ),R(ρ))=01L(ρ)sR(ρ)1s𝑑s,\Lambda_{\text{log}}(L(\rho),R(\rho))=\int_{0}^{1}L(\rho)^{s}R(\rho)^{1-s}\,ds,

or the left/right trivial mean

Λleft(L(ρ),R(ρ))=L(ρ),Λright(L(ρ),R(ρ))=R(ρ).\Lambda_{\text{left}}(L(\rho),R(\rho))=L(\rho),\leavevmode\nobreak\ \leavevmode\nobreak\ \Lambda_{\text{right}}(L(\rho),R(\rho))=R(\rho).

With Λ=Λlog\Lambda=\Lambda_{\text{log}} being the logarithmic mean, we have the chain rule identity for log\log (see [CM17, Lemma 5.5] for a proof):

ρ=ρ^logρ=01ρs(logρ)ρ1s𝑑s.\partial\rho=\hat{\rho}\partial\log\rho=\int_{0}^{1}\rho^{s}(\partial\log\rho)\rho^{1-s}ds.

Here and in what follows, we use the notation

ρ^(x1,,xn):=(ρ^x1,,ρ^xn).\hat{\rho}(x_{1},\dots,x_{n}):=(\hat{\rho}x_{1},\dots,\hat{\rho}x_{n}).

3. Bakry–Émery curvature-dimension condition BE(K,NK,N)

This section is devoted to the noncommutative analog of the Bakry–Émery curvature-dimension condition BE(K,NK,N) defined by the Γ2\Gamma_{2}-criterion. After giving the definition, we will show that it is satisfied for certain generators in Lindblad form, where the dimension parameter NN is given by the number of partial derivatives. We will then prove that BE(K,N)\mathrm{BE}(K,N) implies an improved Poincaré inequality. In the final part of this section we study a complete version of BE(K,N)\mathrm{BE}(K,N), called CBE(K,N)\mathrm{CBE}(K,N), and show that it has the expected tensorization properties.

3.1. Bakry–Émery curvature-dimension condition BE(K,NK,N)

Let (Pt)(P_{t}) be a quantum Markov semigroup on \mathcal{M} with generator \mathcal{L}. The associated carré du champ operator Γ\Gamma is defined as

Γ(a,b):=12(ab+(a)b(ab)),\Gamma(a,b):=\frac{1}{2}\left(a^{\ast}\mathcal{L}b+(\mathcal{L}a)^{\ast}b-\mathcal{L}(a^{\ast}b)\right),

and the iterated carré du champ operator Γ2\Gamma_{2} is defined as

Γ2(a,b):=12(Γ(a,b)+Γ(a,b)Γ(a,b)).\Gamma_{2}(a,b):=\frac{1}{2}\left(\Gamma(a,\mathcal{L}b)+\Gamma(\mathcal{L}a,b)-\mathcal{L}\Gamma(a,b)\right).

As usual, we write Γ(a)\Gamma(a) for Γ(a,a)\Gamma(a,a) and Γ2(a)\Gamma_{2}(a) for Γ2(a,a)\Gamma_{2}(a,a).

Proposition 3.1.

Let KK\in\mathbb{R} and N(0,]N\in(0,\infty]. For a quantum Markov semigroup (Pt)(P_{t}) over \mathcal{M} with generator \mathcal{L}, the following are equivalent:

  1. (a)

    for any t0t\geq 0 and any aa\in\mathcal{M}:

    Γ(Pta)e2KtPtΓ(a)1e2KtKN|Pta|2,\Gamma(P_{t}a)\leq e^{-2Kt}P_{t}\Gamma(a)-\frac{1-e^{-2Kt}}{KN}|\mathcal{L}P_{t}a|^{2},
  2. (b)

    for any aa\in\mathcal{M}:

    Γ2(a)KΓ(a)+1N|a|2.\Gamma_{2}(a)\geq K\Gamma(a)+\frac{1}{N}|\mathcal{L}a|^{2}.

If this is the case, we say the semigroup (Pt)(P_{t}) satisfies Bakry–Émery curvature-dimension condition BE(K,N)\mathrm{BE}(K,N).

Proof.

The proof is essentially based on the following identities: For s[0,t]s\in[0,t],

ddsPs((Ptsa)(Ptsa))=2PsΓ(Ptsa),\frac{d}{ds}P_{s}((P_{t-s}a)^{\ast}(P_{t-s}a))=2P_{s}\Gamma(P_{t-s}a),

and

ddsPsΓ(Ptsa)=2PsΓ2(Ptsa),\frac{d}{ds}P_{s}\Gamma(P_{t-s}a)=2P_{s}\Gamma_{2}(P_{t-s}a),

which follow by direct computations. To prove (a)(b)(a)\implies(b), we set

ϕ(t):=e2KtPtΓ(a)Γ(Pta)1e2KtKN|Pta|2.\phi(t):=e^{-2Kt}P_{t}\Gamma(a)-\Gamma(P_{t}a)-\frac{1-e^{-2Kt}}{KN}|\mathcal{L}P_{t}a|^{2}.

Since ϕ(t)0\phi(t)\geq 0 for all t0t\geq 0 and ϕ(0)=0\phi(0)=0, we have ϕ(0)0\phi^{\prime}(0)\geq 0, which is nothing but (b).

To show (b)(a)(b)\implies(a), we put for any t>0t>0:

φ(s):=e2KsPsΓ(Ptsa),s[0,t].\varphi(s):=e^{-2Ks}P_{s}\Gamma(P_{t-s}a),\leavevmode\nobreak\ \leavevmode\nobreak\ s\in[0,t].

Then by assumption and Kadison-Schwarz inequality,

φ(s)=2e2KsPs(Γ2(Ptsa)KΓ(Ptsa))2e2KsNPs(|Ptsa|2)2e2KsN|Pta|2.\varphi^{\prime}(s)=2e^{-2Ks}P_{s}\left(\Gamma_{2}(P_{t-s}a)-K\Gamma(P_{t-s}a)\right)\geq\frac{2e^{-2Ks}}{N}P_{s}\left(|\mathcal{L}P_{t-s}a|^{2}\right)\geq\frac{2e^{-2Ks}}{N}|\mathcal{L}P_{t}a|^{2}.

So

φ(t)φ(0)=0tφ(s)𝑑s2N0te2Ks𝑑s|Pta|2=1e2KtKN|Pta|2,\varphi(t)-\varphi(0)=\int_{0}^{t}\varphi^{\prime}(s)ds\geq\frac{2}{N}\int_{0}^{t}e^{-2Ks}ds|\mathcal{L}P_{t}a|^{2}=\frac{1-e^{-2Kt}}{KN}|\mathcal{L}P_{t}a|^{2},

which proves (a). ∎

Remark 3.2.

From the proof one can see that the function

t1e2KtKN,t\mapsto\frac{1-e^{-2Kt}}{KN},

in (a) can be replaced by any ff such that f(0)=0f(0)=0 and f(0)=2/Nf^{\prime}(0)=2/N.

Remark 3.3.

The notion BE(K,N)\mathrm{BE}(K,N) is clearly consistent: If (Pt)(P_{t}) satisfies BE(K,N)\mathrm{BE}(K,N), then it also satisfies BE(K,N)\mathrm{BE}(K^{\prime},N^{\prime}) for all KKK^{\prime}\leq K and NNN^{\prime}\geq N.

Remark 3.4.

While all our examples of quantum Markov semigroups satisfying BE\mathrm{BE} are tracially symmetric, let us point out that this is not necessary for the definition nor for the results in the rest of this section with the exception of Proposition 3.7. See also the discussion in Section 7.

We shall give a sufficient condition for Bakry–Émery curvature-dimension condition BE(K,NK,N). Before that we need a simple inequality.

Lemma 3.5.

For any aj,1jda_{j},1\leq j\leq d, in a C*-algebra, we have

j=1d|aj|21d|j=1daj|2.\sum_{j=1}^{d}|a_{j}|^{2}\geq\frac{1}{d}\left|\sum_{j=1}^{d}a_{j}\right|^{2}.
Proof.

In fact,

dj=1d|aj|2|j=1daj|2=12j,k=1d(ajaj+akakajakakaj)=12j,k=1d|ajak|20.d\sum_{j=1}^{d}|a_{j}|^{2}-\left|\sum_{j=1}^{d}a_{j}\right|^{2}=\frac{1}{2}\sum_{j,k=1}^{d}\left(a_{j}^{\ast}a_{j}+a_{k}^{\ast}a_{k}-a_{j}^{\ast}a_{k}-a_{k}^{\ast}a_{j}\right)=\frac{1}{2}\sum_{j,k=1}^{d}|a_{j}-a_{k}|^{2}\geq 0.\qed
Definition 3.6.

Suppose that \mathcal{L} is the generator of the tracially symmetric quantum Markov semigroup (Pt)(P_{t}) with the Lindblad form:

=j=1djj,\mathcal{L}=\sum_{j=1}^{d}\partial_{j}^{\dagger}\partial_{j}, (LB)

where j()=[vj,]\partial_{j}(\cdot)=[v_{j},\cdot] with the adjoint being j()=[vj,]\partial_{j}^{\dagger}(\cdot)=[v_{j}^{\ast},\cdot], and {vj}={vj}\{v_{j}\}=\{v_{j}^{\ast}\}. Then we say (Pt)(P_{t}) satisfies the KK-intertwining condition for some KK\in\mathbb{R} if

jPt=eKtPtj, 1jd,\partial_{j}P_{t}=e^{-Kt}P_{t}\partial_{j},\leavevmode\nobreak\ \leavevmode\nobreak\ 1\leq j\leq d,

or equivalently

j=j+Kj, 1jd.\partial_{j}\mathcal{L}=\mathcal{L}\partial_{j}+K\partial_{j},\leavevmode\nobreak\ \leavevmode\nobreak\ 1\leq j\leq d.
Proposition 3.7.

Suppose that the generator \mathcal{L} of the tracially symmetric quantum Markov semigroup (Pt)(P_{t}) admits the Lindblad form (LB). Then for any aa,

Γ2(a)=Rej=1d(jaja)ja+j,k=1d|kja|2.\Gamma_{2}(a)=\operatorname{Re}\sum_{j=1}^{d}(\partial_{j}\mathcal{L}a-\mathcal{L}\partial_{j}a)^{\ast}\partial_{j}a+\sum_{j,k=1}^{d}|\partial_{k}^{\dagger}\partial_{j}a|^{2}. (3.1)

If (Pt)(P_{t}) satisfies the KK-intertwining condition for KK\in\mathbb{R}, then (Pt)(P_{t}) satisfies BE(K,d)\mathrm{BE}(K,d).

Proof.

Note that

(ja)=avjvja=j(a).(\partial_{j}a)^{\ast}=a^{\ast}v_{j}^{\ast}-v_{j}^{\ast}a^{\ast}=-\partial_{j}^{\dagger}(a^{\ast}).

This, together with the Leibniz rule for j\partial_{j}’s (so also j\partial_{j}^{\dagger}’s), and the fact that {j}={j}\{\partial_{j}\}=\{\partial_{j}^{\dagger}\}, yields

(ab)\displaystyle\mathcal{L}(a^{\ast}b) =j=1d(jja)b+ajjb+(ja)(jb)+(ja)(jb)\displaystyle=\sum_{j=1}^{d}(\partial_{j}^{\dagger}\partial_{j}a^{\ast})b+a\partial_{j}^{\dagger}\partial_{j}b+(\partial_{j}^{\dagger}a)(\partial_{j}b)+(\partial_{j}a)(\partial_{j}^{\dagger}b)
=(a)b+ab+j=1d(ja)(jb)+(ja)(jb)\displaystyle=(\mathcal{L}a)^{\ast}b+a^{\ast}\mathcal{L}b+\sum_{j=1}^{d}(\partial_{j}^{\dagger}a^{\ast})(\partial_{j}b)+(\partial_{j}a^{\ast})(\partial_{j}^{\dagger}b)
=(a)b+abj=1d((ja)(jb)+(ja)(jb))\displaystyle=(\mathcal{L}a)^{\ast}b+a^{\ast}\mathcal{L}b-\sum_{j=1}^{d}\left((\partial_{j}a)^{\ast}(\partial_{j}b)+(\partial_{j}^{\dagger}a)^{\ast}(\partial_{j}^{\dagger}b)\right)
=(a)b+ab2j=1d(ja)(jb).\displaystyle=(\mathcal{L}a)^{\ast}b+a^{\ast}\mathcal{L}b-2\sum_{j=1}^{d}(\partial_{j}a)^{\ast}(\partial_{j}b).

So by definition, the carré du champ operator is given by:

Γ(a,b)=12(ab+(a)b(ab))=j=1d(ja)jb.\Gamma(a,b)=\frac{1}{2}\left(a^{\ast}\mathcal{L}b+(\mathcal{L}a)^{\ast}b-\mathcal{L}(a^{\ast}b)\right)=\sum_{j=1}^{d}(\partial_{j}a)^{\ast}\partial_{j}b. (3.2)

The above computations yield

Γ(a,(a))+Γ((a),a)=j=1d(ja)ja+(ja)ja=2Rej=1d(ja)ja,\displaystyle\Gamma(a,\mathcal{L}(a))+\Gamma(\mathcal{L}(a),a)=\sum_{j=1}^{d}(\partial_{j}\mathcal{L}a)^{\ast}\partial_{j}a+(\partial_{j}a)^{\ast}\partial_{j}\mathcal{L}a=2\operatorname{Re}\sum_{j=1}^{d}(\partial_{j}\mathcal{L}a)^{\ast}\partial_{j}a,

and

Γ(a)\displaystyle\mathcal{L}\Gamma(a) =j=1d((ja)ja)\displaystyle=\sum_{j=1}^{d}\mathcal{L}\left((\partial_{j}a)^{\ast}\partial_{j}a\right)
=j=1d((ja)ja+(ja)ja2k=1d(kja)kja)\displaystyle=\sum_{j=1}^{d}\left((\mathcal{L}\partial_{j}a)^{\ast}\partial_{j}a+(\partial_{j}a)^{\ast}\mathcal{L}\partial_{j}a-2\sum_{k=1}^{d}(\partial_{k}\partial_{j}a)^{\ast}\partial_{k}\partial_{j}a\right)
=2Rej=1d(ja)ja2j,k=1d|kja|2.\displaystyle=2\operatorname{Re}\sum_{j=1}^{d}(\mathcal{L}\partial_{j}a)^{\ast}\partial_{j}a-2\sum_{j,k=1}^{d}|\partial_{k}\partial_{j}a|^{2}.

Thus

Γ2(a)=\displaystyle\Gamma_{2}(a)= 12(Γ(a,(a))+Γ((a),a)Γ(a))\displaystyle\frac{1}{2}\left(\Gamma(a,\mathcal{L}(a))+\Gamma(\mathcal{L}(a),a)-\mathcal{L}\Gamma(a)\right)
=\displaystyle= Rej=1d(ja)jaRej=1d(ja)ja+j,k=1d|kja|2\displaystyle\operatorname{Re}\sum_{j=1}^{d}(\partial_{j}\mathcal{L}a)^{\ast}\partial_{j}a-\operatorname{Re}\sum_{j=1}^{d}(\mathcal{L}\partial_{j}a)^{\ast}\partial_{j}a+\sum_{j,k=1}^{d}|\partial_{k}\partial_{j}a|^{2}
=\displaystyle= Rej=1d(jaja)ja+j,k=1d|kja|2,\displaystyle\operatorname{Re}\sum_{j=1}^{d}(\partial_{j}\mathcal{L}a-\mathcal{L}\partial_{j}a)^{\ast}\partial_{j}a+\sum_{j,k=1}^{d}|\partial_{k}^{\dagger}\partial_{j}a|^{2},

where in the last equality we used again the fact that {j}={j}\{\partial_{j}\}=\{\partial_{j}^{\dagger}\}. This proves (3.1). If (Pt)(P_{t}) satisfies the KK-intertwining condition, then

Rej=1d(jaja)(ja)=Kj=1d(ja)(ja)=KΓ(a).\operatorname{Re}\sum_{j=1}^{d}(\partial_{j}\mathcal{L}a-\mathcal{L}\partial_{j}a)^{\ast}(\partial_{j}a)=K\sum_{j=1}^{d}(\partial_{j}a)^{\ast}(\partial_{j}a)=K\Gamma(a).

Moreover, by Lemma 3.5 we get

j,k=1d|kja|2j=1d|jja|21d|j=1djja|2=1d|a|2.\sum_{j,k=1}^{d}\lvert\partial_{k}^{\dagger}\partial_{j}a\rvert^{2}\geq\sum_{j=1}^{d}\lvert\partial_{j}^{\dagger}\partial_{j}a\rvert^{2}\geq\frac{1}{d}\left\lvert\sum_{j=1}^{d}\partial_{j}^{\dagger}\partial_{j}a\right\rvert^{2}=\frac{1}{d}\lvert\mathcal{L}a\rvert^{2}.

Therefore (Pt)(P_{t}) satisfies BE(K,d)\mathrm{BE}(K,d):

Γ2(a)=Rej=1d(jaja)ja+j,k=1d|kja|2KΓ(a)+1d|a|2.\Gamma_{2}(a)=\operatorname{Re}\sum_{j=1}^{d}(\partial_{j}\mathcal{L}a-\mathcal{L}\partial_{j}a)^{\ast}\partial_{j}a+\sum_{j,k=1}^{d}|\partial_{k}^{\dagger}\partial_{j}a|^{2}\geq K\Gamma(a)+\frac{1}{d}\lvert\mathcal{L}a\rvert^{2}.\qed

3.2. Applications

In this subsection we present two applications of the Bakry–Émery curvature-dimension condition, namely a Poincaré inequality and a Bonnet–Myers theorem.

It is well known that when K>0K>0, the dimensionless bound BE(K,)\mathrm{BE}(K,\infty) implies that the smallest non-zero eigenvalue of the generator is at least KK. As a simple application of the dimensional variant we show that this bound can be improved.

Proposition 3.8 (Poincaré inequality).

Let K>0K>0 and N>1N>1. If (Pt)(P_{t}) satisfies BE(K,N)\mathrm{BE}(K,N) and λ1\lambda_{1} is the smallest non-zero eigenvalue of \mathcal{L}, then

λ1KNN1.\lambda_{1}\geq\frac{KN}{N-1}.
Proof.

By BE(K,N)\mathrm{BE}(K,N) we have

a22=τ(Γ2(a))Kτ(Γ(a))+1Nτ(|a|2)=Ka,a2+1Na22.\displaystyle\lVert\mathcal{L}a\rVert_{2}^{2}=\tau(\Gamma_{2}(a))\geq K\tau(\Gamma(a))+\frac{1}{N}\tau(\lvert\mathcal{L}a\rvert^{2})=K\langle\mathcal{L}a,a\rangle_{2}+\frac{1}{N}\lVert\mathcal{L}a\rVert_{2}^{2}.

In particular, if a=λ1a\mathcal{L}a=\lambda_{1}a and a2=1\lVert a\rVert_{2}=1, then

λ12Kλ1+1Nλ12,\displaystyle\lambda_{1}^{2}\geq K\lambda_{1}+\frac{1}{N}\lambda_{1}^{2},

from which the desired inequality follows. ∎

To state the Bonnet–Myers theorem, we recall the definition of the metric dΓd_{\Gamma} on the space of density matrices that is variously known as quantum L1L^{1}-Wasserstein distance, Connes distance or spectral distance. It is given by

dΓ(ρ0,ρ1)=sup{τ(a(ρ1ρ0))a=a,Γ(a)𝟏}d_{\Gamma}(\rho_{0},\rho_{1})=\sup\{\tau(a(\rho_{1}-\rho_{0}))\mid a=a^{\ast}\in\mathcal{M},\,\Gamma(a)\leq\mathbf{1}\}

for ρ0,ρ1𝒮()\rho_{0},\rho_{1}\in\mathcal{S(M)}.

Proposition 3.9.

Let K,N(0,)K,N\in(0,\infty). If a symmetric quantum Markov semigroup (Pt)(P_{t}) is ergodic and satisfies Bakry–Émery curvature-dimension condition BE(K,N)\mathrm{BE}(K,N), then

dΓ(ρ,𝟏)π2NKd_{\Gamma}(\rho,\mathbf{1})\leq\frac{\pi}{2}\sqrt{\frac{N}{K}}

for all ρ𝒮()\rho\in\mathcal{S}(\mathcal{M}).

In particular,

supρ0,ρ1𝒮()dΓ(ρ0,ρ1)πKN.\sup_{\rho_{0},\rho_{1}\in\mathcal{S(M)}}d_{\Gamma}(\rho_{0},\rho_{1})\leq\pi\sqrt{\frac{K}{N}}.
Proof.

The proof follows the same line as that of [LMP18, Theorem 2.4]. The condition BE(K,N)\mathrm{BE}(K,N) implies

1e2KtKN(Pta)2e2KtPtΓ(a),\displaystyle\frac{1-e^{-2Kt}}{KN}(\mathcal{L}P_{t}a)^{2}\leq e^{-2Kt}P_{t}\Gamma(a),

for any a=aa=a^{\ast}\in\mathcal{M}. If Γ(a)𝟏\Gamma(a)\leq\mathbf{1}, we have

PtaKN1e2Kt1.\displaystyle\lVert\mathcal{L}P_{t}a\rVert_{\infty}\leq\sqrt{KN}\sqrt{\frac{1}{e^{2Kt}-1}}.

Thus for any ρ𝒮()\rho\in\mathcal{S}(\mathcal{M}),

|τ((Ptaa)ρ)|0t|ddsτ((Psa)ρ)|𝑑sKN01e2Ks1𝑑s=π2NK.\displaystyle\lvert\tau((P_{t}a-a)\rho)\rvert\leq\int_{0}^{t}\left\lvert\frac{d}{ds}\tau((P_{s}a)\rho)\right\rvert\,ds\leq\sqrt{KN}\int_{0}^{\infty}\frac{1}{\sqrt{e^{2Ks}-1}}\,ds=\frac{\pi}{2}\sqrt{\frac{N}{K}}.

Therefore

τ(a(ρ𝟏))=τ((aPta)ρ)+τ(Pta(ρ𝟏))π2NK+τ(a(Ptρ𝟏)).\displaystyle\tau(a(\rho-\mathbf{1}))=\tau((a-P_{t}a)\rho)+\tau(P_{t}a(\rho-\mathbf{1}))\leq\frac{\pi}{2}\sqrt{\frac{N}{K}}+\tau(a(P_{t}\rho-\mathbf{1})).

Since (Pt)(P_{t}) is assumed to be ergodic, we have Ptρ𝟏P_{t}\rho\to\mathbf{1} as tt\to\infty, and we end up with

dΓ(ρ,𝟏)=supΓ(a)1τ(a(ρ𝟏))π2NK.d_{\Gamma}(\rho,\mathbf{1})=\sup_{\Gamma(a)\leq 1}\tau(a(\rho-\mathbf{1}))\leq\frac{\pi}{2}\sqrt{\frac{N}{K}}.\qed

3.3. Complete BE(K,NK,N)

In many applications it is desirable to have estimates that are tensor-stable in the sense that they hold not only for (Pt)(P_{t}), but also for (PtidMn())(P_{t}\otimes\mathrm{id}_{M_{n}(\mathbb{C})}) with a constant independent of nn\in\mathbb{N}. Even in the case K=K=\infty, it seems to be unknown if this is true for the Bakry–Émery estimate. For that reason we introduce the complete Bakry–Émery estimate CBE(K,N)\mathrm{CBE}(K,N), which has this tensor stability by definition. We will show that this stronger estimate also holds for quantum Markov semigroup satisfying the KK-intertwining condition, and moreover, this estimate behaves as expected under arbitrary tensor products.

Definition 3.10.

Let KK\in\mathbb{R} and N>0N>0. We say that the quantum Markov semigroup (Pt)(P_{t}) satisfies CBE(K,N)\mathrm{CBE}(K,N) if

[Γ(Ptxj,Ptxk)]j,ke2Kt[PtΓ(xj,xk)]j,k1e2KtKN[(Ptxj)(Ptxk)]j,k,\displaystyle[\Gamma(P_{t}x_{j},P_{t}x_{k})]_{j,k}\leq e^{-2Kt}[P_{t}\Gamma(x_{j},x_{k})]_{j,k}-\frac{1-e^{-2Kt}}{KN}[(\mathcal{L}P_{t}x_{j})^{\ast}(\mathcal{L}P_{t}x_{k})]_{j,k},

for all x1,,xnx_{1},\dots,x_{n}\in\mathcal{M} and t>0t>0.

Just as in Proposition 3.1 one can show that CBE(K,N)\mathrm{CBE}(K,N) is equivalent to

[Γ2(xj,xk)]j,kK[Γ(xj,xk)]j,k+1N[(xj)(xk)]j,k[\Gamma_{2}(x_{j},x_{k})]_{j,k}\geq K[\Gamma(x_{j},x_{k})]_{j,k}+\frac{1}{N}[(\mathcal{L}x_{j})^{\ast}(\mathcal{L}x_{k})]_{j,k}

for all x1,,xnx_{1},\dots,x_{n}\in\mathcal{M} and t0t\geq 0.

For N=N=\infty, this criterion was introduced in [JZ15a] for group von Neumann algebras under the name algebraic Γ2\Gamma_{2}-condition.

To show that CBE(K,N)\mathrm{CBE}(K,N) for (Pt)(P_{t}) is equivalent to BE(K,N)\mathrm{BE}(K,N) for (PtidMn())(P_{t}\otimes\mathrm{id}_{M_{n}(\mathbb{C})}) with constants independent of nn, we need the following elementary lemma.

Lemma 3.11.

Let 𝒜,\mathcal{A},\mathcal{B} be two C*-algebras. If x=[xjk]Mn(𝒜)x=[x_{jk}]\in M_{n}(\mathcal{A}), y=[yjk]Mn()y=[y_{jk}]\in M_{n}(\mathcal{B}) are positive, then

j,kxjkyjk0.\sum_{j,k}x_{jk}\otimes y_{jk}\geq 0.
Proof.

By assumption there are a=[ajk]Mn(𝒜)a=[a_{jk}]\in M_{n}(\mathcal{A}), b=[bjk]Mn()b=[b_{jk}]\in M_{n}(\mathcal{B}) such that

xjk\displaystyle x_{jk} =laljalk,\displaystyle=\sum_{l}a_{lj}^{\ast}a_{lk},
yjk\displaystyle y_{jk} =mbmjbmk.\displaystyle=\sum_{m}b_{mj}^{\ast}b_{mk}.

Thus

j,kxjkyjk\displaystyle\sum_{j,k}x_{jk}\otimes y_{jk} =j,k,l,maljalkbmjbmk\displaystyle=\sum_{j,k,l,m}a_{lj}^{\ast}a_{lk}\otimes b_{mj}^{\ast}b_{mk}
=l,m(jaljbmj)(kalkbmk)\displaystyle=\sum_{l,m}\left(\sum_{j}a_{lj}^{\ast}\otimes b_{mj}^{\ast}\right)\left(\sum_{k}a_{lk}\otimes b_{mk}\right)
=l,m|jaljbmj|2\displaystyle=\sum_{l,m}\left\lvert\sum_{j}a_{lj}\otimes b_{mj}\right\rvert^{2}
0.\displaystyle\geq 0.\qed
Proposition 3.12.

Let (Pt)(P_{t}) be a quantum Markov semigroup on \mathcal{M}. For KK\in\mathbb{R} and N(0,]N\in(0,\infty], the following assertions are equivalent:

  1. (a)

    (Pt)(P_{t}) satisfies CBE(K,N)\mathrm{CBE}(K,N).

  2. (b)

    (PtidMn())(P_{t}\otimes\mathrm{id}_{M_{n}(\mathbb{C})}) satisfies BE(K,N)\mathrm{BE}(K,N) for all nn\in\mathbb{N}.

Proof.

(a)\implies(b): Write Γ,Γ2\Gamma,\Gamma_{2} for the (iterated) carré du champ associated with (Pt)(P_{t}) and Γ,Γ2\Gamma^{\otimes},\Gamma_{2}^{\otimes} for the same forms associated with (PtidMn())(P_{t}\otimes\mathrm{id}_{M_{n}(\mathbb{C})}).

A direct computation shows

Γ2(jxjyj)\displaystyle\Gamma_{2}^{\otimes}\left(\sum_{j}x_{j}\otimes y_{j}\right) =j,kΓ2(xj,xk)yjyk,\displaystyle=\sum_{j,k}\Gamma_{2}(x_{j},x_{k})\otimes y_{j}^{\ast}y_{k},
Γ(jxjyj)\displaystyle\Gamma^{\otimes}\left(\sum_{j}x_{j}\otimes y_{j}\right) =j,kΓ(xj,xk)yjyk,\displaystyle=\sum_{j,k}\Gamma(x_{j},x_{k})\otimes y_{j}^{\ast}y_{k},
|(id𝒩)(jxjyj)|2\displaystyle\left\lvert(\mathcal{L}\otimes\mathrm{id}_{\mathcal{N}})\left(\sum_{j}x_{j}\otimes y_{j}\right)\right\rvert^{2} =j,k(xj)(xk)yjyk.\displaystyle=\sum_{j,k}(\mathcal{L}x_{j})^{\ast}(\mathcal{L}x_{k})\otimes y_{j}^{\ast}y_{k}.

Hence

Γ2(jxjyj)KΓ(jxjyj)1N|(id𝒩)(jxjyj)|2\displaystyle\quad\;\Gamma_{2}^{\otimes}\left(\sum_{j}x_{j}\otimes y_{j}\right)-K\Gamma^{\otimes}(\sum_{j}x_{j}\otimes y_{j})-\frac{1}{N}\left\lvert(\mathcal{L}\otimes\mathrm{id}_{\mathcal{N}})\left(\sum_{j}x_{j}\otimes y_{j}\right)\right\rvert^{2}
=j,k(Γ2(xj,xk)KΓ(xj,xk)1N(xj)(xk))yjyk,\displaystyle=\sum_{j,k}(\Gamma_{2}(x_{j},x_{k})-K\Gamma(x_{j},x_{k})-\frac{1}{N}(\mathcal{L}x_{j})^{\ast}(\mathcal{L}x_{k}))\otimes y_{j}^{\ast}y_{k},

and the result follows from Lemma 3.11 and (a).

(b)\implies(a): Let x=jxj|1j|x=\sum_{j}x_{j}\otimes\ket{1}\bra{j}. The computations from (a)\implies(b) show

Γ2(x)=j,kΓ2(xj,xk)|jk|\Gamma_{2}^{\otimes}(x)=\sum_{j,k}\Gamma_{2}(x_{j},x_{k})\otimes\ket{j}\bra{k}

and similar formulas for Γ\Gamma^{\otimes} and idMn()\mathcal{L}\otimes\mathrm{id}_{M_{n}(\mathbb{C})}. Using the \ast-isomorphism Mn()Mn(),j,kxjk|jk|[xjk]j,k\mathcal{M}\otimes M_{n}(\mathbb{C})\to M_{n}(\mathcal{M}),\,\sum_{j,k}x_{jk}\otimes\ket{j}\bra{k}\mapsto[x_{jk}]_{j,k}, assertion (a) follows. ∎

In the following two results we will give two classes of examples for which the condition CBE\mathrm{CBE} is satisfied.

Proposition 3.13.

Suppose that the generator \mathcal{L} of the quantum Markov semigroup (Pt)(P_{t}) admits the Lindblad form (LB) with dd partial derivatives 1,,d\partial_{1},\dots,\partial_{d}. If (Pt)(P_{t}) satisfies the KK-intertwining condition for KK\in\mathbb{R}, then (Pt)(P_{t}) satisfies CBE(K,d)\mathrm{CBE}(K,d).

Proof.

A direct computation shows that idMn()\mathcal{L}\otimes\mathrm{id}_{M_{n}(\mathbb{C})} admits a Lindblad form with derivations 1idMn(),,didMn()\partial_{1}\otimes\mathrm{id}_{M_{n}(\mathbb{C})},\dots,\partial_{d}\otimes\mathrm{id}_{M_{n}(\mathbb{C})}. Now the claim is a direct consequence of Propositions 3.7 and 3.12. ∎

Proposition 3.14.

If \mathcal{M} is commutative and (Pt)(P_{t}) satisfies BE(K,N)\mathrm{BE}(K,N), then it also satisfies CBE(K,N)\mathrm{CBE}(K,N).

Proof.

By assumption, C(X)\mathcal{M}\cong C(X) for a compact space XX. We have to show

[Γ2(fj,fk)(x)]j,kK[Γ(fj,fk)(x)]j,k+1N[(fj)(x)¯(fk)(x)]j,k\displaystyle[\Gamma_{2}(f_{j},f_{k})(x)]_{j,k}\geq K[\Gamma(f_{j},f_{k})(x)]_{j,k}+\frac{1}{N}[\overline{(\mathcal{L}f_{j})(x)}(\mathcal{L}f_{k})(x)]_{j,k}

for xXx\in X, which follows from

j,kαj¯αkΓ2(fj,fk)(x)\displaystyle\sum_{j,k}\overline{\alpha_{j}}\alpha_{k}\Gamma_{2}(f_{j},f_{k})(x) =Γ2(jαjfj)(x)\displaystyle=\Gamma_{2}\left(\sum_{j}\alpha_{j}f_{j}\right)(x)
KΓ(jαjfj)(x)+1N|(jαjfj)(x)|2\displaystyle\geq K\Gamma\left(\sum_{j}\alpha_{j}f_{j}\right)(x)+\frac{1}{N}\left\lvert\mathcal{L}\left(\sum_{j}\alpha_{j}f_{j}\right)(x)\right\rvert^{2}
=Kj,kαj¯αkΓ(fj,fk)(x)+1Nj,kαj¯αk(fj)(x)¯(fk)(x)\displaystyle=K\sum_{j,k}\overline{\alpha_{j}}\alpha_{k}\Gamma(f_{j},f_{k})(x)+\frac{1}{N}\sum_{j,k}\overline{\alpha_{j}}\alpha_{k}\overline{(\mathcal{L}f_{j})(x)}(\mathcal{L}f_{k})(x)

for any αj\alpha_{j}\in\mathbb{C}. ∎

Before we state the tensorization property of CBE\mathrm{CBE}, we need another elementary inequality.

Lemma 3.15.

Let 𝒜\mathcal{A} be a C*-algebra. If a,b𝒜a,b\in\mathcal{A} and λ>0\lambda>0, then

|a+b|2(1+λ)|a|2+(1+λ1)|b|2.\lvert a+b\rvert^{2}\leq(1+\lambda)\lvert a\rvert^{2}+(1+\lambda^{-1})\lvert b\rvert^{2}.
Proof.

In fact,

(1+λ)|a|2+(1+λ1)|b|2\displaystyle(1+\lambda)\lvert a\rvert^{2}+(1+\lambda^{-1})\lvert b\rvert^{2} =|a+b|2+λ|a|2+λ1|b|2abba\displaystyle=|a+b|^{2}+\lambda\lvert a\rvert^{2}+\lambda^{-1}\lvert b\rvert^{2}-a^{\ast}b-b^{\ast}a
=|a+b|2+|λ1/2aλ1/2b|2\displaystyle=|a+b|^{2}+\lvert\lambda^{1/2}a-\lambda^{-1/2}b\rvert^{2}
|a+b|2.\displaystyle\geq|a+b|^{2}.\qed
Proposition 3.16.

Let \mathcal{M}, 𝒩\mathcal{N} be finite-dimensional von Neumann algebras and let (Pt)(P_{t}), (Qt)(Q_{t}) be tracially symmetric quantum Markov semigroups on \mathcal{M} and 𝒩\mathcal{N}, respectively. If (Pt)(P_{t}) satisfies CBE(K,N)\mathrm{CBE}(K,N) and (Qt)(Q_{t}) satisfies CBE(K,N)\mathrm{CBE}(K^{\prime},N^{\prime}), then (PtQt)(P_{t}\otimes Q_{t}) satisfies CBE(min{K,K},N+N)\mathrm{CBE}(\min\{K,K^{\prime}\},N+N^{\prime}).

Proof.

We use superscripts for the (iterated) carré du champ to indicate the associated quantum Markov semigroup. Let κ=min{K,K}\kappa=\min\{K,K^{\prime}\}. We have

Γ2PQκΓPQ=(Γ2Pid𝒩κΓPid𝒩)+(Γ2idQκΓidQ)+2ΓPΓQ,\displaystyle\Gamma_{2}^{P\otimes Q}-\kappa\Gamma^{P\otimes Q}=(\Gamma_{2}^{P\otimes\mathrm{id}_{\mathcal{N}}}-\kappa\Gamma^{P\otimes\mathrm{id}_{\mathcal{N}}})+(\Gamma_{2}^{\mathrm{id}_{\mathcal{M}}\otimes Q}-\kappa\Gamma^{\mathrm{id}_{\mathcal{M}}\otimes Q})+2\Gamma^{P}\otimes\Gamma^{Q},

where

(ΓPΓQ)(jxjyj):=j,kΓP(xj,xk)ΓQ(yj,yk).(\Gamma^{P}\otimes\Gamma^{Q})\left(\sum_{j}x_{j}\otimes y_{j}\right):=\sum_{j,k}\Gamma^{P}(x_{j},x_{k})\otimes\Gamma^{Q}(y_{j},y_{k}).

By CBE(κ,N)\mathrm{CBE}(\kappa,N) for (Pt)(P_{t}) and CBE(κ,N)\mathrm{CBE}(\kappa,N^{\prime}) for (Qt)(Q_{t}) we have

(Γ2Pid𝒩κΓPid𝒩)(jxjyj)1N|jPxjyj|2,\displaystyle(\Gamma_{2}^{P\otimes\mathrm{id}_{\mathcal{N}}}-\kappa\Gamma^{P\otimes\mathrm{id}_{\mathcal{N}}})\left(\sum_{j}x_{j}\otimes y_{j}\right)\geq\frac{1}{N}\left\lvert\sum_{j}\mathcal{L}_{P}x_{j}\otimes y_{j}\right\rvert^{2},
(Γ2idQκΓidQ)(jxjyj)1N|jxjQyj|2.\displaystyle(\Gamma_{2}^{\mathrm{id}_{\mathcal{M}}\otimes Q}-\kappa\Gamma^{\mathrm{id}_{\mathcal{M}}\otimes Q})\left(\sum_{j}x_{j}\otimes y_{j}\right)\geq\frac{1}{N^{\prime}}\left\lvert\sum_{j}x_{j}\otimes\mathcal{L}_{Q}y_{j}\right\rvert^{2}.

Moreover,

(ΓPΓQ)(jxjyj)0(\Gamma^{P}\otimes\Gamma^{Q})\left(\sum_{j}x_{j}\otimes y_{j}\right)\geq 0

by Lemma 3.11.

Finally,

1N|jPxjyj|2+1N|jxjQyj|21N+N|jPxjyj+xjQyj|2\frac{1}{N}\left\lvert\sum_{j}\mathcal{L}_{P}x_{j}\otimes y_{j}\right\rvert^{2}+\frac{1}{N^{\prime}}\left\lvert\sum_{j}x_{j}\otimes\mathcal{L}_{Q}y_{j}\right\rvert^{2}\geq\frac{1}{N+N^{\prime}}\left\lvert\sum_{j}\mathcal{L}_{P}x_{j}\otimes y_{j}+x_{j}\otimes\mathcal{L}_{Q}y_{j}\right\rvert^{2}

by Lemma 3.15, which shows BE(κ,N+N)\mathrm{BE}(\kappa,N+N^{\prime}) for (PtQt)(P_{t}\otimes Q_{t}). To prove CBE(κ,N+N)\mathrm{CBE}(\kappa,N+N^{\prime}), we can simply apply the same argument to (PtidMn())(P_{t}\otimes\mathrm{id}_{M_{n}(\mathbb{C})}) and (QtidMn())(Q_{t}\otimes\mathrm{id}_{M_{n}(\mathbb{C})}) for arbitrary nn\in\mathbb{N}. ∎

4. The gradient estimate GE(K,N)\mathrm{GE}(K,N)

4.1. Gradient estimate GE(K,N)\mathrm{GE}(K,N) and a sufficient condition

In [CM14, CM17, Wir18, CM20], a noncommutative analog of the 22-Wasserstein metric was constructed on the set of quantum states. Among other things, it gives rise to a notion of (entropic) lower Ricci curvature bound via geodesic semi-convexity of the entropy. This allows to prove a number of functional inequalities under strictly positive lower Ricci curvature bound, including the modified log-Sobolev inequality that (seemingly) cannot be produced under the Bakry–Émery curvature-dimension condition BE(K,K,\infty).

This entropic lower Ricci curvature bound is captured in the following gradient estimate

Ptaρ2e2KtaPtρ2,\lVert\partial P_{t}a\rVert_{\rho}^{2}\leq e^{-2Kt}\lVert\partial a\rVert_{P_{t}\rho}^{2}, (GE(K,)(K,\infty))

or equivalently

Rea,ρ^a+12ddt|t=0Ptρ^a,aKaρ2,\operatorname{Re}\langle\partial\mathcal{L}a,\hat{\rho}\partial a\rangle+\frac{1}{2}\left\langle\frac{d}{dt}\big{|}_{t=0}\widehat{P_{t}\rho}\partial a,\partial a\right\rangle\geq K\lVert\partial a\rVert_{\rho}^{2}, (4.1)

where the notations ρ^\hat{\rho} and ρ\|\cdot\|_{\rho} correspond to the logarithmic mean Λlog\Lambda_{\text{log}}. Recall Section 2 for more details. The fact that logarithmic mean comes into play lies in the use of chain rule

ρ^jlogρ=jρ, 1jd.\hat{\rho}\partial_{j}\log\rho=\partial_{j}\rho,\leavevmode\nobreak\ \leavevmode\nobreak\ 1\leq j\leq d.

In fact, for the gradient estimate (GE(K,)(K,\infty)) and its equivalent form (4.1) one can work with any operator mean. This not only makes the theory more flexible, but also includes the Bakry–Émery curvature-dimension condition BE(K,K,\infty) as a special case. Indeed, one recovers BE(K,K,\infty) by replacing the logarithmic mean in (4.1) with the left/right trivial mean. In the next section we discuss the connection of GE(K,NK,N) and (K,N)(K,N)-convexity of the (relative) entropy.

The study of (GE(K,)(K,\infty)) for arbitrary operator means was started in [Wir18, WZ20]. Here we continue to work within this framework and focus on the “finite-dimensional” version of (GE(K,)(K,\infty)) or (4.1), which we call gradient estimate GE(K,N)\mathrm{GE}(K,N).

Definition 4.1.

Let Λ\Lambda be an operator mean and (Pt)(P_{t}) be a symmetric quantum Markov semigroup whose generator takes the Lindblad form (LB). We say that (Pt)(P_{t}) satisfies the gradient estimate GE(K,N)\mathrm{GE}(K,N) for K,N(0,]K\in\mathbb{R},N\in(0,\infty] if

Ptaρ2e2KtaPtρ21e2KtKN|(a,Ptρ)|2,\lVert\partial P_{t}a\rVert_{\rho}^{2}\leq e^{-2Kt}\lVert\partial a\rVert_{P_{t}\rho}^{2}-\frac{1-e^{-2Kt}}{KN}\lvert\mathcal{E}(a,P_{t}\rho)\rvert^{2}, (GE(K,N)(K,N))

for any t0t\geq 0, aa\in\mathcal{M} and ρ𝒮+()\rho\in\mathcal{S}_{+}(\mathcal{M}).

Remark 4.2.

It is obvious that when N=N=\infty, (GE(K,N)(K,N)) becomes the gradient estimate GE(K,)\mathrm{GE}(K,\infty). From the definition it is not immediately clear that if (Pt)(P_{t}) satisfies the gradient estimate GE(K,N)\mathrm{GE}(K,N), then it also satisfies the gradient estimate GE(K,N)\mathrm{GE}(K^{\prime},N^{\prime}) whenever KKK^{\prime}\leq K and NNN^{\prime}\geq N. But this can be seen from the following equivalent formulation in the flavor of the Γ2\Gamma_{2}-condition.

Remark 4.3.

If ρ𝒮()\rho\in\mathcal{S(M)} is not invertible, one can apply GE(K,N)(K,N) to ρϵ=ρ+ϵ𝟏1+ϵ\rho^{\epsilon}=\frac{\rho+\epsilon\mathbf{1}}{1+\epsilon} and let ϵ0\epsilon\searrow 0 to see that GE(K,N)(K,N) still remains true.

Proposition 4.4.

For any operator mean Λ\Lambda and any symmetric quantum Markov semigroup (Pt)(P_{t}), the gradient estimate GE(K,N)(K,N) holds if and only if

Rea,ρ^a12dG(ρ)(ρ)a,aKaρ2+1N|(a,ρ)|2\operatorname{Re}\langle\partial\mathcal{L}a,\hat{\rho}\partial a\rangle-\frac{1}{2}\langle dG(\rho)(\mathcal{L}\rho)\partial a,\partial a\rangle\geq K\lVert\partial a\rVert_{\rho}^{2}+\frac{1}{N}\lvert\mathcal{E}(a,\rho)\rvert^{2} (4.2)

for any ρ𝒮+()\rho\in\mathcal{S}_{+}(\mathcal{M}) and any aa\in\mathcal{M}. Here dG(ρ)dG(\rho) denotes the Fréchet derivative of G(ρ):=ρ^=Λ(L(ρ),R(ρ))G(\rho):=\hat{\rho}=\Lambda(L(\rho),R(\rho)).

Proof.

Assume that (Pt)(P_{t}) satisfies GE(K,N)(K,N). Set

ϕ(t):=e2KtaPtρ2Ptaρ21e2KtKN|(a,Ptρ)|2.\phi(t):=e^{-2Kt}\lVert\partial a\rVert_{P_{t}\rho}^{2}-\lVert\partial P_{t}a\rVert_{\rho}^{2}-\frac{1-e^{-2Kt}}{KN}\lvert\mathcal{E}(a,P_{t}\rho)\rvert^{2}.

Then ϕ(t)0\phi(t)\geq 0 and ϕ(0)=0\phi(0)=0. Therefore ϕ(0)0\phi^{\prime}(0)\geq 0, that is,

ddt|t=0Ptρ^a,a2Kaρ2+ρ^a,a+ρ^a,a2N|(a,ρ)|20.\left\langle\frac{d}{dt}\big{|}_{t=0}\widehat{P_{t}\rho}\partial a,\partial a\right\rangle-2K\lVert\partial a\rVert_{\rho}^{2}+\langle\hat{\rho}\partial\mathcal{L}a,\partial a\rangle+\langle\hat{\rho}\partial a,\partial\mathcal{L}a\rangle-\frac{2}{N}\lvert\mathcal{E}(a,\rho)\rvert^{2}\geq 0.

This is nothing but (4.2), since dG(ρ)(ρ)=ddt|t=0Ptρ^dG(\rho)(\mathcal{L}\rho)=-\frac{d}{dt}\big{|}_{t=0}\widehat{P_{t}\rho}.

Now suppose that (Pt)(P_{t}) satisfies (4.2). Fix t>0t>0 and put

φ(s):=e2KsPtsaPsρ2, 0st.\varphi(s):=e^{-2Ks}\lVert\partial P_{t-s}a\rVert_{P_{s}\rho}^{2},\leavevmode\nobreak\ \leavevmode\nobreak\ 0\leq s\leq t.

Then applying (4.2) to (ρ,a)=(Psρ,Ptsa)(\rho,a)=(P_{s}\rho,P_{t-s}a), we get

φ(s)=\displaystyle\varphi^{\prime}(s)= e2Ks(Psρ^Ptsa,Ptsa+Psρ^Ptsa,Ptsa\displaystyle e^{-2Ks}\left(\langle\widehat{P_{s}\rho}\partial\mathcal{L}P_{t-s}a,\partial P_{t-s}a\rangle+\langle\widehat{P_{s}\rho}\partial P_{t-s}a,\partial\mathcal{L}P_{t-s}a\rangle\right.
dG(Psρ)(ρ)Ptsa,Ptsa2KPtsaPsρ2)\displaystyle\left.-\left\langle dG(P_{s}\rho)(\mathcal{L}\rho)\partial P_{t-s}a,\partial P_{t-s}a\right\rangle-2K\lVert\partial P_{t-s}a\rVert_{P_{s}\rho}^{2}\right)
\displaystyle\geq 2Ne2Ks|(Ptsa,Psρ)|2\displaystyle\frac{2}{N}e^{-2Ks}\lvert\mathcal{E}(P_{t-s}a,P_{s}\rho)\rvert^{2}
=\displaystyle= 2Ne2Ks|(a,Ptρ)|2.\displaystyle\frac{2}{N}e^{-2Ks}\lvert\mathcal{E}(a,P_{t}\rho)\rvert^{2}.

This, together with the fundamental theorem of calculus, yields

e2KtaPtρ2Ptaρ2=φ(t)φ(0)=0tφ(s)𝑑s1e2KtKN|(a,Ptρ)|2.\displaystyle e^{-2Kt}\lVert\partial a\rVert_{P_{t}\rho}^{2}-\lVert\partial P_{t}a\rVert_{\rho}^{2}=\varphi(t)-\varphi(0)=\int_{0}^{t}\varphi^{\prime}(s)ds\geq\frac{1-e^{-2Kt}}{KN}\lvert\mathcal{E}(a,P_{t}\rho)\rvert^{2}.

Therefore (Pt)(P_{t}) satisfies GE(K,N)(K,N). ∎

Remark 4.5.

Similar to Remark 3.2, the function

t1e2KtKN,t\mapsto\frac{1-e^{-2Kt}}{KN},

in GE(K,N)(K,N) can be replaced by any ff such that f(0)=0f(0)=0 and f(0)=2/Nf^{\prime}(0)=2/N.

Remark 4.6.

In the case N=N=\infty, the gradient estimate GE(K,)\mathrm{GE}(K,\infty) for the left trivial mean is equivalent to the exponential form of BE(K,)\mathrm{BE}(K,\infty). For N<N<\infty this seems to be no longer the case, but one still has one implication: the Bakry–Émery curvature-dimension condition BE(K,NK,N) is stronger than GE(K,N)\mathrm{GE}(K,N) for the left trivial mean. This is a consequence of Cauchy-Schwarz inequality for the state τ(ρ)\tau(\rho\,\cdot\,):

|(a,ρ)|2=|a,ρ|2|a|2,ρ.|\mathcal{E}(a,\rho)|^{2}=|\langle\mathcal{L}a,\rho\rangle|^{2}\leq\langle|\mathcal{L}a|^{2},\rho\rangle.

Similar to BE(K,NK,N), the intertwining condition is also sufficient to prove GE(K,N)\mathrm{GE}(K,N) with the same dimension (upper bound).

Theorem 4.7.

Let (Pt)(P_{t}) be a symmetric quantum Markov semigroup over \mathcal{M} with the Lindblad form (LB). Suppose that (Pt)(P_{t}) satisfies KK-intertwining condition for some KK\in\mathbb{R}. Then for any operator mean Λ\Lambda the quantum Markov semigroup (Pt)(P_{t}) satisfies GE(K,d)\mathrm{GE}(K,d).

Proof.

For aa\in\mathcal{M}, recall that

Pt(aa)(Pta)Pta=0tddsPs((Ptsa)Ptsa)𝑑s=20tPsΓ(Ptsa)𝑑s.\displaystyle P_{t}(a^{\ast}a)-(P_{t}a)^{\ast}P_{t}a=\int_{0}^{t}\frac{d}{ds}P_{s}\left((P_{t-s}a)^{\ast}P_{t-s}a\right)ds=2\int_{0}^{t}P_{s}\Gamma(P_{t-s}a)ds.

Under the KK-intertwining condition, we have (either by Kadison–Schwarz or BE(K,K,\infty))

PsΓ(Ptsa)e2KsΓ(Pta).P_{s}\Gamma(P_{t-s}a)\geq e^{2Ks}\Gamma(P_{t}a).

So

Pt(aa)(Pta)Pta20te2Ks𝑑sΓ(Pta)=e2Kt1KΓ(Pta).P_{t}(a^{\ast}a)-(P_{t}a)^{\ast}P_{t}a\geq 2\int_{0}^{t}e^{2Ks}ds\Gamma(P_{t}a)=\frac{e^{2Kt}-1}{K}\Gamma(P_{t}a). (4.3)

By (3.2) and Lemma 3.5, we get for any (xj)1jd(x_{j})_{1\leq j\leq d}\subset\mathcal{M}

j=1dΓ(Ptxj)=j,k=1d|kPtxj|2=j,k=1d|kPtxj|2j=1d|jPtxj|21d|j=1djPtxj|2.\sum_{j=1}^{d}\Gamma(P_{t}x_{j})=\sum_{j,k=1}^{d}|\partial_{k}P_{t}x_{j}|^{2}=\sum_{j,k=1}^{d}|\partial_{k}^{\dagger}P_{t}x_{j}|^{2}\geq\sum_{j=1}^{d}|\partial_{j}^{\dagger}P_{t}x_{j}|^{2}\geq\frac{1}{d}\left|\sum_{j=1}^{d}\partial_{j}^{\dagger}P_{t}x_{j}\right|^{2}. (4.4)

Let ^=j=1d\hat{\mathcal{M}}=\oplus_{j=1}^{d}\mathcal{M} be equipped with the inner product

(xj),(yj):=j=1dxj,yj,\langle(x_{j}),(y_{j})\rangle:=\sum_{j=1}^{d}\langle x_{j},y_{j}\rangle,

and P^t\hat{P}_{t} be the operator acting on ^\hat{\mathcal{M}} such that P^t(x1,,xd)=(Ptx1,,Ptxd)\hat{P}_{t}(x_{1},\dots,x_{d})=(P_{t}x_{1},\dots,P_{t}x_{d}). Fix ρ𝒮+()\rho\in\mathcal{S}_{+}(\mathcal{M}). For simplicity, let us identify ρ\rho with the element (ρ,,ρ)(\rho,\dots,\rho) in ^\hat{\mathcal{M}}. Then for x=(x1,,xd)^x=(x_{1},\dots,x_{d})\in\hat{\mathcal{M}}, we have by (4.3) and (4.4) that

P^t(xx),ρ(P^tx)P^tx,ρ=\displaystyle\langle\hat{P}_{t}(x^{\ast}x),\rho\rangle-\langle(\hat{P}_{t}x)^{\ast}\hat{P}_{t}x,\rho\rangle= j=1dPt(xjxj)(Ptxj)Ptxj,ρ\displaystyle\sum_{j=1}^{d}\langle P_{t}(x_{j}^{\ast}x_{j})-(P_{t}x_{j})^{\ast}P_{t}x_{j},\rho\rangle
\displaystyle\geq e2Kt1dKj=1dΓ(Ptxj),ρ\displaystyle\frac{e^{2Kt}-1}{dK}\sum_{j=1}^{d}\langle\Gamma(P_{t}x_{j}),\rho\rangle
\displaystyle\geq e2Kt1dK|j=1djPtxj|2,ρ.\displaystyle\frac{e^{2Kt}-1}{dK}\left\langle\left|\sum_{j=1}^{d}\partial_{j}^{\dagger}P_{t}x_{j}\right|^{2},\rho\right\rangle.

From KK-intertwining condition and Cauchy-Schwarz inequality for the state τ(ρ)\tau(\rho\cdot) on \mathcal{M}, this is bounded from below by

1e2KtdK|j=1dPtjxj|2,ρ1e2KtdK|j=1dPtjxj,ρ|2=1e2KtdK|x,Ptρ|2.\frac{1-e^{-2Kt}}{dK}\left\langle\left|\sum_{j=1}^{d}P_{t}\partial_{j}^{\dagger}x_{j}\right|^{2},\rho\right\rangle\geq\frac{1-e^{-2Kt}}{dK}\left|\sum_{j=1}^{d}\left\langle P_{t}\partial_{j}^{\dagger}x_{j},\rho\right\rangle\right|^{2}\\ =\frac{1-e^{-2Kt}}{dK}\left|\left\langle x,\partial P_{t}\rho\right\rangle\right|^{2}.

So we have proved that for any x^x\in\hat{\mathcal{M}}:

x(P^tρ),xP^tx,(P^tx)ρ+1e2KtdKx,|PtρPtρ|(x),\displaystyle\langle x(\hat{P}_{t}\rho),x\rangle\geq\langle\hat{P}_{t}x,(\hat{P}_{t}x)\rho\rangle+\frac{1-e^{-2Kt}}{dK}\langle x,\ket{\partial P_{t}\rho}\bra{\partial P_{t}\rho}(x)\rangle,

or equivalently

R(Ptρ)P^tR(ρ)P^t+1e2KtdK|PtρPtρ|.R(P_{t}\rho)\geq\hat{P}_{t}R(\rho)\hat{P}_{t}+\frac{1-e^{-2Kt}}{dK}\ket{\partial P_{t}\rho}\bra{\partial P_{t}\rho}.

Replacing xx by xx^{\ast}, we obtain

L(Ptρ)P^tL(ρ)P^t+1e2KtdK|PtρPtρ|.L(P_{t}\rho)\geq\hat{P}_{t}L(\rho)\hat{P}_{t}+\frac{1-e^{-2Kt}}{dK}\ket{\partial P_{t}\rho}\bra{\partial P_{t}\rho}.

Note that the second summand is the same in both cases.

Now since Λ\Lambda is an operator mean, we have

Λ(L(Ptρ),R(Ptρ))\displaystyle\Lambda(L(P_{t}\rho),R(P_{t}\rho))\geq Λ(P^tL(ρ)P^t,P^tR(ρ)P^t)+1e2KtdKΛ(|PtρPtρ|,|PtρPtρ|)\displaystyle\Lambda(\hat{P}_{t}L(\rho)\hat{P}_{t},\hat{P}_{t}R(\rho)\hat{P}_{t})+\frac{1-e^{-2Kt}}{dK}\Lambda\left(\ket{\partial P_{t}\rho}\bra{\partial P_{t}\rho},\ket{\partial P_{t}\rho}\bra{\partial P_{t}\rho}\right)
\displaystyle\geq P^tΛ(L(ρ),R(ρ))P^t+1e2KtdK|PtρPtρ|,\displaystyle\hat{P}_{t}\Lambda(L(\rho),R(\rho))\hat{P}_{t}+\frac{1-e^{-2Kt}}{dK}\ket{\partial P_{t}\rho}\bra{\partial P_{t}\rho},

where in the first inequality we used the monotonicity, concavity (Lemma 2.1 (b)) and positive homogeneity (Lemma 2.1 (a)) of Λ\Lambda, and in the second inequality we used the transformer inequality and Lemma 2.1(d). This, together with the KK-intertwining condition, yields

Ptaρ2\displaystyle\lVert\partial P_{t}a\rVert_{\rho}^{2} =Λ(L(ρ),R(ρ))Pta,Pta\displaystyle=\langle\Lambda(L(\rho),R(\rho))\partial P_{t}a,\partial P_{t}a\rangle
=e2KtP^tΛ(L(ρ),R(ρ))P^ta,a\displaystyle=e^{-2Kt}\langle\hat{P}_{t}\Lambda(L(\rho),R(\rho))\hat{P}_{t}\partial a,\partial a\rangle
e2KtΛ(L(Ptρ),R(Ptρ))a,ae2Kte4KtdK|Ptρ,a|2\displaystyle\leq e^{-2Kt}\langle\Lambda(L(P_{t}\rho),R(P_{t}\rho))\partial a,\partial a\rangle-\frac{e^{-2Kt}-e^{-4Kt}}{dK}\lvert\langle\partial P_{t}\rho,\partial a\rangle\rvert^{2}
=e2KtaPtρ2e2Kte4KtdK|(a,Ptρ)|2.\displaystyle=e^{-2Kt}\lVert\partial a\rVert_{P_{t}\rho}^{2}-\frac{e^{-2Kt}-e^{-4Kt}}{dK}\lvert\mathcal{E}(a,P_{t}\rho)\rvert^{2}.

This completes the proof, by Remark 4.5. ∎

4.2. Bonnet–Myers theorem

As a first application of the dimensional gradient estimate GE(K,N)\mathrm{GE}(K,N), we present here a Bonnet–Myers theorem for the noncommutative analog of the Wasserstein distance introduced in [CM17, CM20]. The proof is quite similar (or, in fact, similar to the dual) to the proof of Proposition 3.9.

Let us first recall the definition of the metric. The space 𝒮+()\mathcal{S}_{+}(\mathcal{M}) of invertible density matrices is a smooth manifold and the tangent space at ρ𝒮+()\rho\in\mathcal{S}_{+}(\mathcal{M}) can be canonically identified with the traceless self-adjoint elements of \mathcal{M}. Assume that (Pt)(P_{t}) is a tracially symmetric quantum Markov semigroup with generator \mathcal{L} with Lindblad form (LB).

Fix an operator mean Λ\Lambda. For ρ𝒮+()\rho\in\mathcal{S}_{+}(\mathcal{M}) we define

𝒦ρΛ:,xρ^x=j=1dj(Λ(L(ρ),R(ρ))j(x)).\mathcal{K}_{\rho}^{\Lambda}\colon\mathcal{M}\to\mathcal{M},\,x\mapsto\partial^{\dagger}\hat{\rho}\partial x=\sum_{j=1}^{d}\partial_{j}^{\dagger}(\Lambda(L(\rho),R(\rho))\partial_{j}(x)). (4.5)

The Riemannian metric gΛg^{\Lambda} on 𝒮+()\mathcal{S}_{+}(\mathcal{M}) is defined by

gρΛ(ρ˙1,ρ˙2)=ρ˙1,(𝒦ρΛ)1ρ˙2.g^{\Lambda}_{\rho}(\dot{\rho}_{1},\dot{\rho}_{2})=\langle\dot{\rho}_{1},(\mathcal{K}^{\Lambda}_{\rho})^{-1}\dot{\rho}_{2}\rangle.

The associated distance function on 𝒮+()×𝒮+()\mathcal{S}_{+}(\mathcal{M})\times\mathcal{S}_{+}(\mathcal{M}) is denoted by 𝒲\mathcal{W}. By [CM20, Proposition 9.2], 𝒲\mathcal{W} can be extended to 𝒮()×𝒮()\mathcal{S}(\mathcal{M})\times\mathcal{S}(\mathcal{M}) since

Λ(a𝟏,b𝟏)Λ(min{a,b}𝟏,min{a,b}𝟏)=min{a,b}𝟏\Lambda(a\mathbf{1},b\mathbf{1})\geq\Lambda(\min\{a,b\}\mathbf{1},\min\{a,b\}\mathbf{1})=\min\{a,b\}\mathbf{1}

for all a,b>0a,b>0.

Proposition 4.8.

Fix an operator mean Λ\Lambda. Let K,N(0,)K,N\in(0,\infty). If (Pt)(P_{t}) is ergodic and satisfies gradient estimate GE(K,N)\mathrm{GE}(K,N), then

𝒲(ρ,𝟏)π2NK\mathcal{W}(\rho,\mathbf{1})\leq\frac{\pi}{2}\sqrt{\frac{N}{K}}

for all ρ𝒮+()\rho\in\mathcal{S}_{+}(\mathcal{M}).

In particular,

supρ0,ρ1𝒮+()𝒲(ρ0,ρ1)πKN.\sup_{\rho_{0},\rho_{1}\in\mathcal{S}_{+}(\mathcal{M})}\mathcal{W}(\rho_{0},\rho_{1})\leq\pi\sqrt{\frac{K}{N}}.
Proof.

Since (Pt)(P_{t}) is ergodic, we have Ptρ𝟏P_{t}\rho\to\mathbf{1} as tt\to\infty. Let ρt=Ptρ\rho_{t}=P_{t}\rho for t0t\geq 0. The gradient estimate GE(K,N)\mathrm{GE}(K,N) implies

|a,ρ˙t|=|a,ρt|KNe2Kt1e2Ktaρt=KNe2Kt1aρt\lvert\langle a,\dot{\rho}_{t}\rangle\rvert=\lvert\langle a,\mathcal{L}\rho_{t}\rangle\rvert\leq\sqrt{KN}\sqrt{\frac{e^{-2Kt}}{1-e^{-2Kt}}}\lVert\partial a\rVert_{\rho_{t}}=\frac{\sqrt{KN}}{\sqrt{e^{2Kt}-1}}\lVert\partial a\rVert_{\rho_{t}}

for all aa\in\mathcal{M}. Choosing a=(𝒦ρtΛ)1ρ˙ta=(\mathcal{K}^{\Lambda}_{\rho_{t}})^{-1}\dot{\rho}_{t}, we get

gρtΛ(ρ˙t,ρ˙t)KNe2Kt1𝒦ρtΛa,a=KNe2Kt1gρtΛ(ρ˙t,ρ˙t).g^{\Lambda}_{\rho_{t}}(\dot{\rho}_{t},\dot{\rho}_{t})\leq\frac{\sqrt{KN}}{\sqrt{e^{2Kt}-1}}\sqrt{\langle\mathcal{K}^{\Lambda}_{\rho_{t}}a,a\rangle}=\frac{\sqrt{KN}}{\sqrt{e^{2Kt}-1}}\sqrt{g^{\Lambda}_{\rho_{t}}(\dot{\rho}_{t},\dot{\rho}_{t})}.

Hence

gρtΛ(ρ˙t,ρ˙t)KNe2Kt1,\sqrt{g^{\Lambda}_{\rho_{t}}(\dot{\rho}_{t},\dot{\rho}_{t})}\leq\frac{\sqrt{KN}}{\sqrt{e^{2Kt}-1}},

and we conclude

𝒲(ρ,𝟏)0gρtΛ(ρ˙t,ρ˙t)𝑑t0KNe2Kt1𝑑t=π2KN.\mathcal{W}(\rho,\mathbf{1})\leq\int_{0}^{\infty}\sqrt{g^{\Lambda}_{\rho_{t}}(\dot{\rho}_{t},\dot{\rho}_{t})}\,dt\leq\int_{0}^{\infty}\frac{\sqrt{KN}}{\sqrt{e^{2Kt}-1}}\,dt=\frac{\pi}{2}\sqrt{\frac{K}{N}}.\qed

4.3. Complete GE(K,N)\mathrm{GE}(K,N)

Now we turn to the complete version of GE(K,N)\mathrm{GE}(K,N).

Definition 4.9.

We say that a quantum Markov semigroup (Pt)(P_{t}) satisfies complete gradient estimate CGE(K,N)\mathrm{CGE}(K,N) for KK\in\mathbb{R} and N(0,]N\in(0,\infty] if (PtidMn())(P_{t}\otimes\mathrm{id}_{M_{n}(\mathbb{C})}) satisfies GE(K,N)\mathrm{GE}(K,N) for all nn\in\mathbb{N}.

Similar to Proposition 3.13, the KK-intertwining condition is sufficient for CGE:\mathrm{CGE}\colon

Proposition 4.10.

Suppose that the generator \mathcal{L} of the quantum Markov semigroup (Pt)(P_{t}) admits the Lindblad form (LB) with dd partial derivatives 1,,d\partial_{1},\dots,\partial_{d}. If (Pt)(P_{t}) satisfies the KK-intertwining condition for KK\in\mathbb{R}, then (Pt)(P_{t}) satisfies CGE(K,d)\mathrm{CGE}(K,d).

Also, the complete gradient estimate CGE\mathrm{CGE} is tensor stable.

Proposition 4.11.

Consider two quantum Markov semigroups (Ptj)(P_{t}^{j}) acting on j\mathcal{M}_{j}, j=1,2j=1,2. If (Ptj)(P_{t}^{j}) satisfies CGE(Kj,Nj),j=1,2\mathrm{CGE}(K_{j},N_{j}),j=1,2, then the tensor product (Pt1Pt2)(P_{t}^{1}\otimes P_{t}^{2}) over =12\mathcal{M}=\mathcal{M}_{1}\otimes\mathcal{M}_{2} satisfies CGE(K,N)\mathrm{CGE}(K,N) with K=min{K1,K2}K=\min\{K_{1},K_{2}\} and N=N1+N2N=N_{1}+N_{2}.

Proof.

For each j=1,2j=1,2, we denote by j\mathcal{L}_{j} the generator of (Ptj)(P_{t}^{j}) and j:j^j\partial^{j}:\mathcal{M}_{j}\to\hat{\mathcal{M}}_{j} (to distinguish from partial derivatives j\partial_{j}’s) the corresponding derivation operator so that j=(j)j\mathcal{L}_{j}=(\partial^{j})^{\dagger}\partial^{j}. Denote Pt=Pt1Pt2P_{t}=P_{t}^{1}\otimes P_{t}^{2}. Then its generator is =\mathcal{L}=\partial^{\dagger}\partial, where the derivation operator \partial is given by

=(1id,id2).\partial=(\partial^{1}\otimes\mathrm{id},\mathrm{id}\otimes\partial^{2}).

Since (Ptj)(P_{t}^{j}) satisfies CGE(K,Nj),j=1,2\mathrm{CGE}(K,N_{j}),j=1,2, we have for any a^:=j^ja\in\hat{\mathcal{M}}:=\otimes_{j}\hat{\mathcal{M}}_{j} and ρ𝒮+()\rho\in\mathcal{S}_{+}(\mathcal{M}) that

Ptaρ2=\displaystyle\lVert\partial P_{t}a\rVert_{\rho}^{2}= (1id)(Pt1id)(idPt2)aρ2+(id2)(idPt2)(Pt1id)aρ2\displaystyle\|(\partial^{1}\otimes\mathrm{id})(P_{t}^{1}\otimes\mathrm{id})(\mathrm{id}\otimes P_{t}^{2})a\|_{\rho}^{2}+\|(\mathrm{id}\otimes\partial^{2})(\mathrm{id}\otimes P_{t}^{2})(P_{t}^{1}\otimes\mathrm{id})a\|_{\rho}^{2}
\displaystyle\leq e2Kt((1id)(idPt2)a(Pt1id)ρ2+(id2)(Pt1id)a(idPt2)ρ2)\displaystyle e^{-2Kt}\left(\|(\partial^{1}\otimes\mathrm{id})(\mathrm{id}\otimes P_{t}^{2})a\|_{(P_{t}^{1}\otimes\mathrm{id})\rho}^{2}+\|(\mathrm{id}\otimes\partial^{2})(P_{t}^{1}\otimes\mathrm{id})a\|_{(\mathrm{id}\otimes P_{t}^{2})\rho}^{2}\right)
1e2KtK(1N1|(1id)Pta,ρ|2+1N2|(id2)Pta,ρ|2).\displaystyle-\frac{1-e^{-2Kt}}{K}\left(\frac{1}{N_{1}}\lvert\langle(\mathcal{L}_{1}\otimes\mathrm{id})P_{t}a,\rho\rangle\rvert^{2}+\frac{1}{N_{2}}\lvert\langle(\mathrm{id}\otimes\mathcal{L}_{2})P_{t}a,\rho\rangle\rvert^{2}\right).

As we have proven in [WZ20, Theorem 4.1], for the first summand one has

(1id)(idPt2)a(Pt1id)ρ2+(id2)(Pt1id)a(idPt2)ρ2aPtρ2.\|(\partial^{1}\otimes\mathrm{id})(\mathrm{id}\otimes P_{t}^{2})a\|_{(P_{t}^{1}\otimes\mathrm{id})\rho}^{2}+\|(\mathrm{id}\otimes\partial^{2})(P_{t}^{1}\otimes\mathrm{id})a\|_{(\mathrm{id}\otimes P_{t}^{2})\rho}^{2}\leq\lVert\partial a\rVert_{P_{t}\rho}^{2}.

As for the second summand, note that =1id+id2\mathcal{L}=\mathcal{L}_{1}\otimes\mathrm{id}+\mathrm{id}\otimes\mathcal{L}_{2}. So by Cauchy-Schwarz inequality,

1N|Pta,ρ|2\displaystyle\frac{1}{N}\lvert\langle\mathcal{L}P_{t}a,\rho\rangle\rvert^{2} =|(1id)Pta,ρ+(id2)Pta,ρ|2N1+N2\displaystyle=\frac{\lvert\langle(\mathcal{L}_{1}\otimes\mathrm{id})P_{t}a,\rho\rangle+\langle(\mathrm{id}\otimes\mathcal{L}_{2})P_{t}a,\rho\rangle\rvert^{2}}{N_{1}+N_{2}}
1N1|(1id)Pta,ρ|2+1N2|(id2)Pta,ρ|2.\displaystyle\leq\frac{1}{N_{1}}\lvert\langle(\mathcal{L}_{1}\otimes\mathrm{id})P_{t}a,\rho\rangle\rvert^{2}+\frac{1}{N_{2}}\lvert\langle(\mathrm{id}\otimes\mathcal{L}_{2})P_{t}a,\rho\rangle\rvert^{2}.

All combined, we obtain

Ptaρ2e2KtaPtρ21e2KtKN|Pta,ρ|2.\lVert\partial P_{t}a\rVert_{\rho}^{2}\leq e^{-2Kt}\lVert\partial a\rVert_{P_{t}\rho}^{2}-\frac{1-e^{-2Kt}}{KN}\lvert\langle\mathcal{L}P_{t}a,\rho\rangle\rvert^{2}.\qed

5. Geodesic (K,N)(K,N)-convexity of the (relative) entropy and relation to the gradient estimate GE(K,N)\mathrm{GE}(K,N)

In the case of the logarithmic mean, the given quantum Markov semigroup is the gradient flow of the (relative) entropy with respect to the transport distance 𝒲\mathcal{W}. In this case, the gradient estimate GE(K,)\mathrm{GE}(K,\infty) is equivalent to geodesic KK-convexity of the (relative) entropy with respect to 𝒲\mathcal{W}, and several functional inequalities can be obtained using gradient flow techniques.

Similarly, the gradient estimate GE(K,N)\mathrm{GE}(K,N) is equivalent to geodesic (K,N)(K,N)-convexity of the (relative) entropy with respect to 𝒲\mathcal{W}, a notion introduced by Erbar, Kuwada and Sturm [EKS15], and again, gradient flow techniques allow to deduce several dimensional functional inequalities from the abstract theory of (K,N)(K,N)-convex functions on Riemannian manifolds.

5.1. (K,N)(K,N)-convexity for the (relative) entropy

Let (M,g)(M,g) be a Riemannian manifold and KK\in\mathbb{R}, N(0,]N\in(0,\infty]. A function SC2(M)S\in C^{2}(M) is called (K,N)(K,N)-convex if

HessS(x)[v,v]1Ng(S(x),v)2Kg(v,v)\operatorname{Hess}S(x)[v,v]-\frac{1}{N}g(\nabla S(x),v)^{2}\geq Kg(v,v)

for all xMx\in M and vTxMv\in T_{x}M.

With the function

UN:M,UN(x)=exp(1NS(x)),U_{N}\colon M\to\mathbb{R},\,U_{N}(x)=\exp\left(-\frac{1}{N}S(x)\right),

the (K,N)(K,N)-convexity of SS can equivalently be characterized by

HessUNKNUN.\operatorname{Hess}U_{N}\leq-\frac{K}{N}U_{N}.

For N=N=\infty, one obtains the usual notion of KK-convexity. Moreover, the notion of (K,N)(K,N)-convexity is obviously monotone in the parameters KK and NN in the sense that if SS is (K,N)(K,N)-convex, then SS is also (K,N)(K^{\prime},N^{\prime})-convex for KKK^{\prime}\leq K and NNN^{\prime}\geq N.

Our focus will be on the case when FF is the (relative) entropy and the Riemannian metric is the one introduced in [CM17, CM20], whose definition was recalled in Subsection 4.2.

If F:𝒮+()F\colon\mathcal{S}_{+}(\mathcal{M})\to\mathbb{R} is smooth, its Frechét derivative can be written as

dF(ρ)=τ(x)dF(\rho)=\tau(x\,\cdot)

for a unique traceless self-adjoint xx\in\mathcal{M}. This element xx shall be denoted by DF(ρ)DF(\rho). In particular, if F(ρ)=τ(ρlogρ)F(\rho)=\tau(\rho\log\rho), then DF(ρ)=logρ+cDF(\rho)=\log\rho+c for some cc\in\mathbb{R}.

By [CM17, Theorem 7.5], the gradient of FF is given by (recall (4.5) for 𝒦ρΛ\mathcal{K}_{\rho}^{\Lambda})

gΛF(ρ)=𝒦ρΛDF(ρ).\nabla_{g^{\Lambda}}F(\rho)=\mathcal{K}_{\rho}^{\Lambda}DF(\rho). (5.1)

Of particular interest to us is the case when FF is the (relative) entropy, that is, the functional

Ent:𝒮+()(0,),Ent(ρ)=τ(ρlogρ).\operatorname{Ent}\colon\mathcal{S}_{+}(\mathcal{M})\to(0,\infty),\,\operatorname{Ent}(\rho)=\tau(\rho\log\rho).

If we choose Λ\Lambda to be the logarithmic mean Λlog\Lambda_{\log}, then ρt=Ptρ\rho_{t}=P_{t}\rho satisfies the gradient flow equation

ρ˙t=gΛEnt(ρt)\dot{\rho}_{t}=-\nabla_{g^{\Lambda}}\operatorname{Ent}(\rho_{t})

for any ρ𝒮+()\rho\in\mathcal{S}_{+}(\mathcal{M}) [CM17, Theorem 7.6]. For this reason, we fix the operator mean Λ\Lambda to be the logarithmic mean in this section.

To formulate the metric formulations of (K,N)(K,N)-convexity, we need the following notation: For θ,κ\theta,\kappa\in\mathbb{R} and t[0,1]t\in[0,1] put

cκ(θ)={cos(κθ),if κ0,cosh(κθ),if κ<0,sκ(θ)={κ1/2sin(κθ),if κ>0,θ,if κ=0,(κ)1/2sinh(κθ),if κ<0,σκ(t)(θ),={sκ(tθ)sκ(θ),κθ20 and κθ2<π2,t,κθ2=0,+,κθ2π2.\displaystyle\begin{split}c_{\kappa}(\theta)&=\begin{cases}\cos(\sqrt{\kappa}\theta),&\text{if }\kappa\geq 0,\\ \cosh(\sqrt{-\kappa}\theta),&\text{if }\kappa<0,\end{cases}\\ s_{\kappa}(\theta)&=\begin{cases}\kappa^{-1/2}\sin(\sqrt{\kappa}\theta),&\text{if }\kappa>0,\\ \theta,&\text{if }\kappa=0,\\ (-\kappa)^{-1/2}\sinh(\sqrt{\kappa}\theta),&\text{if }\kappa<0,\end{cases}\\ \sigma_{\kappa}^{(t)}(\theta),&=\begin{cases}\frac{s_{\kappa}(t\theta)}{s_{\kappa}(\theta)},&\kappa\theta^{2}\neq 0\text{ and }\kappa\theta^{2}<\pi^{2},\\ t,&\kappa\theta^{2}=0,\\ +\infty,&\kappa\theta^{2}\geq\pi^{2}.\end{cases}\end{split} (5.2)

The following theorem is a quite direct consequence of the abstract theory of (K,N)(K,N)-convex functions and the computation of the gradient and Hessian on (𝒮+(),g)(\mathcal{S}_{+}(\mathcal{M}),g) carried out in [CM17, CM20]. Nonetheless, it implies some interesting functional inequalities, as we shall see in the following subsection.

Theorem 5.1.

Fix the logarithmic mean Λ=Λlog\Lambda=\Lambda_{\log}. Let KK\in\mathbb{R} and N(0,]N\in(0,\infty]. Further let

UN(ρ)=exp(1NEnt(ρ)).U_{N}(\rho)=\exp\left(-\frac{1}{N}\operatorname{Ent}(\rho)\right).

The the following assertions are equivalent:

  1. (a)

    The (relative) entropy Ent\operatorname{Ent} is (K,N)(K,N)-convex on (𝒮+(),gΛ)(\mathcal{S}_{+}(\mathcal{M}),g^{\Lambda}).

  2. (b)

    For all ρ,ν𝒮+()\rho,\nu\in\mathcal{S}_{+}(\mathcal{M}), the following Evolution Variational Inequality holds for all t0t\geq 0:

    d+dtsK/N(12𝒲(Ptρ,ν))2+KsK/N(12𝒲(Ptρ,ν))2N2(1UN(ν)UN(Ptρ)).\displaystyle\frac{d^{+}}{dt}s_{K/N}\left(\frac{1}{2}\mathcal{W}(P_{t}\rho,\nu)\right)^{2}+Ks_{K/N}\left(\frac{1}{2}\mathcal{W}(P_{t}\rho,\nu)\right)^{2}\leq\frac{N}{2}\left(1-\frac{U_{N}(\nu)}{U_{N}(P_{t}\rho)}\right). (EVIK,N)
  3. (c)

    For any constant speed geodesic (ρt)t[0,1](\rho_{t})_{t\in[0,1]} in 𝒮+()\mathcal{S}_{+}(\mathcal{M}) one has

    UN(ρt)σK/N(1t)(𝒲(ρ0,ρ1))UN(ρ0)+σK/N(t)(𝒲(ρ0,ρ1))UN(ρ1),t[0,1].U_{N}(\rho_{t})\geq\sigma^{(1-t)}_{K/N}(\mathcal{W}(\rho_{0},\rho_{1}))U_{N}(\rho_{0})+\sigma^{(t)}_{K/N}(\mathcal{W}(\rho_{0},\rho_{1}))U_{N}(\rho_{1}),\leavevmode\nobreak\ \leavevmode\nobreak\ t\in[0,1].
  4. (d)

    The semigroup (Pt)(P_{t}) satisfies GE(K,N)\mathrm{GE}(K,N).

Proof.

(a) \iff (b)\iff(c): These equivalences follow from abstract theory of (K,N)(K,N)-convex functionals on Riemannian manifolds [EKS15, Lemmas 2.2, 2.4].

(a)\iff(d): With the identification of the gradient from (5.1) and the Hessian from [CM20, Proposition 7.16], one sees that the defining inequality of the (K,N)(K,N)-convexity of DD coincides with the equivalent formulation of GE(K,N)\mathrm{GE}(K,N) given in Proposition 4.4. ∎

5.2. Dimension-dependent functional inequalities

Let us first collect some consequences of (K,N)(K,N) convexity that were already observed in [EKS15], adapted to our setting. Recall that Ent(ρ)=τ(ρlogρ)\operatorname{Ent}(\rho)=\tau(\rho\log\rho). We use the notation

(ρ)=τ((ρ)logρ)\mathcal{I}(\rho)=\tau((\mathcal{L}\rho)\log\rho)

for the Fisher information.

It satisfies the de Bruijn identity

ddtEnt(Ptρ)=(Ptρ).\frac{d}{dt}\operatorname{Ent}(P_{t}\rho)=-\mathcal{I}(P_{t}\rho).

The following inequalities (b) (c) and (d) are finite-dimensional versions of the HWI-inequality, modified log-Sobolev inequality (MLSI) and Talagrand inequality, respectively. The infinite-dimensional results (i.e. N=N=\infty) were obtained in [CM17, CM20, DR20].

Proposition 5.2.

Fix the logarithmic mean Λ=Λlog\Lambda=\Lambda_{\log}. Let KK\in\mathbb{R} and N>0N>0. If (Pt)(P_{t}) satisfies GE(K,N)\mathrm{GE}(K,N), then the following functional inequalities hold:

  1. (a)

    𝒲\mathcal{W}-expansion bound:

    sK/N(12𝒲(Ptρ0,Psρ1))2\displaystyle s_{K/N}\left(\frac{1}{2}\mathcal{W}(P_{t}\rho_{0},P_{s}\rho_{1})\right)^{2}
    eK(s+t)sK/N(12𝒲(ρ0,ρ1))2+NK(1eK(s+t))(ts)22(s+t)\displaystyle\qquad\leq e^{-K(s+t)}s_{K/N}\left(\frac{1}{2}\mathcal{W}(\rho_{0},\rho_{1})\right)^{2}+\frac{N}{K}\left(1-e^{-K(s+t)}\right)\frac{(\sqrt{t}-\sqrt{s})^{2}}{2(s+t)}

    for ρ0,ρ1𝒮+()\rho_{0},\rho_{1}\in\mathcal{S}_{+}(\mathcal{M}) and s,t0s,t\geq 0.

  2. (b)

    NN-HWI inequality:

    UN(ρ1)UN(ρ0)cK/N(𝒲(ρ0,ρ1))+1NsK/N(𝒲(ρ0,ρ1))(ρ0),\frac{U_{N}(\rho_{1})}{U_{N}(\rho_{0})}\leq c_{K/N}(\mathcal{W}(\rho_{0},\rho_{1}))+\frac{1}{N}s_{K/N}(\mathcal{W}(\rho_{0},\rho_{1}))\sqrt{\mathcal{I}(\rho_{0})},

    for ρ0,ρ1𝒮+()\rho_{0},\rho_{1}\in\mathcal{S}_{+}(\mathcal{M}) and s,t0s,t\geq 0.

If K>0K>0, then additionally the following functional inequalities hold:

  1. (c)

    NN-MLSI:

    KN(UN(ρ)21)(ρ),KN\left(U_{N}(\rho)^{-2}-1\right)\leq\mathcal{I}(\rho),

    for ρ𝒮+()\rho\in\mathcal{S}_{+}(\mathcal{M}).

  2. (d)

    NN-Talagrand inequality:

    Ent(ρ)Nlogcos(KN𝒲(ρ,𝟏)),\operatorname{Ent}(\rho)\geq-N\log\cos\left(\sqrt{\frac{K}{N}}\mathcal{W}(\rho,\mathbf{1})\right),

    for ρ𝒮+()\rho\in\mathcal{S}_{+}(\mathcal{M}).

Proof.

The proofs of Theorems 2.19, 3.28 and Corollaries 3.29, 3.31 from [EKS15] can be easily adapted to our setting. ∎

5.3. Concavity of entropy power

Let us now move on to the concavity of entropy power:

tUN(Ptρ)2=exp(2NEnt(Ptρ)).t\mapsto U_{N}(P_{t}\rho)^{2}=\exp\left(-\frac{2}{N}\operatorname{Ent}(P_{t}\rho)\right).

For the heat semigroup on n\mathbb{R}^{n}, the concavity of entropy power along the heat flow was first proved by Costa in [Cos85]. In [Vil00] Villani gave a short proof and remarked that this can be proved using Γ2\Gamma_{2}-calculus. Recently Li and Li [LL20] considered this problem on the Riemannian manifold under the curvature-dimension condition CD(K,NK,N). Here we show that the geodesic concavity of the entropy power follows from the (K,N)(K,N)-convexity of the entropy.

Theorem 5.3.

Let KK\in\mathbb{R} and N>0N>0. If (Pt)(P_{t}) satisfies GE(K,N)\mathrm{GE}(K,N) for logarithmic mean, then

d2dt2UN(Ptρ)22KddtUN(Ptρ)2,t0.\frac{d^{2}}{dt^{2}}U_{N}(P_{t}\rho)^{2}\leq-2K\frac{d}{dt}U_{N}(P_{t}\rho)^{2},\leavevmode\nobreak\ \leavevmode\nobreak\ t\geq 0.

In particular, if K0K\geq 0, then d2dt2UN(Ptρ)20\frac{d^{2}}{dt^{2}}U_{N}(P_{t}\rho)^{2}\leq 0. This implies the concavity of the entropy power tUN(Ptρ)2t\mapsto U_{N}(P_{t}\rho)^{2}.

Proof.

Let ρt=Ptρ\rho_{t}=P_{t}\rho. Since Ent\operatorname{Ent} is (K,N)(K,N)-convex by Theorem 5.1 and (Pt)(P_{t}) is a gradient flow of Ent\operatorname{Ent} in (𝒮+(),g)(\mathcal{S}_{+}(\mathcal{M}),g) by our choice of the operator mean, we have

d2dt2UN(ρt)2\displaystyle\frac{d^{2}}{dt^{2}}U_{N}(\rho_{t})^{2} =ddt(2NgEnt(ρt),ρ˙tUN(xt))\displaystyle=\frac{d}{dt}\left(-\frac{2}{N}\langle\nabla_{g}\operatorname{Ent}(\rho_{t}),\dot{\rho}_{t}\rangle U_{N}(x_{t})\right)
=ddt(2Nρ˙t,ρ˙tUN(ρt)2)\displaystyle=\frac{d}{dt}\left(\frac{2}{N}\langle\dot{\rho}_{t},\dot{\rho}_{t}\rangle U_{N}(\rho_{t})^{2}\right)
=(4Nρ˙t,ρ˙tρ˙t+4N2ρ˙t,ρ˙t2)UN(ρt)2\displaystyle=\left(\frac{4}{N}\langle\dot{\rho}_{t},\nabla_{\dot{\rho}_{t}}\dot{\rho}_{t}\rangle+\frac{4}{N^{2}}\langle\dot{\rho}_{t},\dot{\rho}_{t}\rangle^{2}\right)U_{N}(\rho_{t})^{2}
=4N(HessEnt(ρt)[ρ˙t,ρ˙t]+1NEnt(ρt),ρ˙t2)UN(ρt)2\displaystyle=\frac{4}{N}\left(-\operatorname{Hess}\operatorname{Ent}(\rho_{t})[\dot{\rho}_{t},\dot{\rho}_{t}]+\frac{1}{N}\langle\nabla\operatorname{Ent}(\rho_{t}),\dot{\rho}_{t}\rangle^{2}\right)U_{N}(\rho_{t})^{2}
4KNρ˙t,ρ˙tUN(ρt)2\displaystyle\leq-\frac{4K}{N}\langle\dot{\rho}_{t},\dot{\rho}_{t}\rangle U_{N}(\rho_{t})^{2}
=2KddtUN(ρt)2.\displaystyle=-2K\frac{d}{dt}U_{N}(\rho_{t})^{2}.\qed
Remark 5.4.

The same proof applies abstractly whenever FF is a (K,N)(K,N)-convex functional on a Riemannian manifold and (ρt)(\rho_{t}) is a gradient flow curve of FF.

The following proof is closer to the spirit of Villani.

Another proof of Theorem 5.3.

Put φ(t):=Ent(ρt)=τ(ρtlogρt)\varphi(t):=-\operatorname{Ent}(\rho_{t})=-\tau(\rho_{t}\log\rho_{t}) with ρt=Ptρ\rho_{t}=P_{t}\rho. Recall the chain rule

ρ=ρ^logρ.\partial\rho=\hat{\rho}\partial\log\rho.

Thus

φ(t)=ρt,logρt=ρt^logρt),logρt.\varphi^{\prime}(t)=\langle\mathcal{L}\rho_{t},\log\rho_{t}\rangle=\langle\hat{\rho_{t}}\partial\log\rho_{t}),\partial\log\rho_{t}\rangle. (5.3)

This allows to give two forms of φ′′\varphi^{\prime\prime}:

φ′′(t)=ddtρt,logρt=ρt,ddtlogρtρt,logρt=:I,\varphi^{\prime\prime}(t)=\frac{d}{dt}\langle\mathcal{L}\rho_{t},\log\rho_{t}\rangle=\langle\mathcal{L}\rho_{t},\frac{d}{dt}\log\rho_{t}\rangle-\langle\mathcal{L}\rho_{t},\mathcal{L}\log\rho_{t}\rangle=:\mathrm{I}, (5.4)

and

φ′′(t)\displaystyle\varphi^{\prime\prime}(t) =ddtρt^logρt,logρt\displaystyle=\frac{d}{dt}\langle\widehat{\rho_{t}}\partial\log\rho_{t},\partial\log\rho_{t}\rangle
=2ρt^logρt,ddr|r=tlogρr+ddr|r=tρr^logρt,logρt\displaystyle=2\langle\widehat{\rho_{t}}\partial\log\rho_{t},\partial\frac{d}{dr}\big{|}_{r=t}\log\rho_{r}\rangle+\langle\frac{d}{dr}\big{|}_{r=t}\widehat{\rho_{r}}\partial\log\rho_{t},\partial\log\rho_{t}\rangle
=2ρt,ddtlogρt+ddr|r=tρr^logρt,logρt=:II.\displaystyle=2\langle\mathcal{L}\rho_{t},\frac{d}{dt}\log\rho_{t}\rangle+\langle\frac{d}{dr}\big{|}_{r=t}\widehat{\rho_{r}}\partial\log\rho_{t},\partial\log\rho_{t}\rangle=:\mathrm{II}. (5.5)

From (5.4) and (5.5) we deduce that

φ′′(t)=2III=2ρt,logρtddr|r=tρr^logρt,logρt.\varphi^{\prime\prime}(t)=2\mathrm{I}-\mathrm{II}=-2\langle\mathcal{L}\rho_{t},\mathcal{L}\log\rho_{t}\rangle-\langle\frac{d}{dr}\big{|}_{r=t}\widehat{\rho_{r}}\partial\log\rho_{t},\partial\log\rho_{t}\rangle. (5.6)

Since (Pt)(P_{t}) satisfies GE(K,N)\mathrm{GE}(K,N) we have by Proposition 4.4 that

ρt^logρt,logρt+12ddr|r=tρr^logρt,logρtKlogρtρ2+1N|(logρt,ρt)|2,\langle\hat{\rho_{t}}\partial\mathcal{L}\log\rho_{t},\partial\log\rho_{t}\rangle+\frac{1}{2}\langle\frac{d}{dr}\big{|}_{r=t}\widehat{\rho_{r}}\partial\log\rho_{t},\partial\log\rho_{t}\rangle\geq K\lVert\partial\log\rho_{t}\rVert_{\rho}^{2}+\frac{1}{N}\lvert\mathcal{E}(\log\rho_{t},\rho_{t})\rvert^{2},

that is,

2logρt,ρt+ddr|r=tρr^logρt,logρt2Klogρtρ2+2N|(logρt,ρt)|2.2\langle\mathcal{L}\log\rho_{t},\mathcal{L}\rho_{t}\rangle+\langle\frac{d}{dr}\big{|}_{r=t}\widehat{\rho_{r}}\partial\log\rho_{t},\partial\log\rho_{t}\rangle\geq 2K\lVert\partial\log\rho_{t}\rVert_{\rho}^{2}+\frac{2}{N}\lvert\mathcal{E}(\log\rho_{t},\rho_{t})\rvert^{2}.

This, together with (5.3) and (5.6), yields

φ′′(t)2Klogρtρt22N|(logρt,ρt)|2=2Kφ(t)2Nφ(t)2.\varphi^{\prime\prime}(t)\leq-2K\|\partial\log\rho_{t}\|_{\rho_{t}}^{2}-\frac{2}{N}|\mathcal{E}(\log\rho_{t},\rho_{t})|^{2}=-2K\varphi^{\prime}(t)-\frac{2}{N}\varphi^{\prime}(t)^{2}. (5.7)

A direct computation gives

ddtUN(Ptρ)2=2NUN(Ptρ)2φ(t),\frac{d}{dt}U_{N}(P_{t}\rho)^{2}=\frac{2}{N}U_{N}(P_{t}\rho)^{2}\varphi^{\prime}(t),

and

d2dt2UN(Ptρ)2=2NUN(Ptρ)2(2Nφ(t)2+φ′′(t)).\frac{d^{2}}{dt^{2}}U_{N}(P_{t}\rho)^{2}=\frac{2}{N}U_{N}(P_{t}\rho)^{2}\left(\frac{2}{N}\varphi^{\prime}(t)^{2}+\varphi^{\prime\prime}(t)\right).

So by (5.7) we get

d2dt2UN(Ptρ)24KNUN(Ptρ)2φ(t)=2KddtUN(Ptρ)2.\frac{d^{2}}{dt^{2}}U_{N}(P_{t}\rho)^{2}\leq-\frac{4K}{N}U_{N}(P_{t}\rho)^{2}\varphi^{\prime}(t)=-2K\frac{d}{dt}U_{N}(P_{t}\rho)^{2}.\qed
Remark 5.5.

Here we used the fact that I=II,\mathrm{I}=\mathrm{II}, or equivalently,

ρt,ddtlogρt+ρt,logρt+ddr|r=tρr^logρt,logρt=0.\langle\mathcal{L}\rho_{t},\frac{d}{dt}\log\rho_{t}\rangle+\langle\mathcal{L}\rho_{t},\mathcal{L}\log\rho_{t}\rangle+\langle\frac{d}{dr}\big{|}_{r=t}\widehat{\rho_{r}}\partial\log\rho_{t},\partial\log\rho_{t}\rangle=0.

If we consider the heat semigroup Pt=etΔP_{t}=e^{t\Delta} on n\mathbb{R}^{n}, then this follows from the elementary identity

Δff=Δ(logf)+|(logf)|2,\frac{\Delta f}{f}=\Delta(\log f)+|\nabla(\log f)|^{2},

as used in Villani’s proof [Vil00].

6. Examples

In this section we present several classes of examples of quantum Markov semigroups satisfying BE(K,N)\mathrm{BE}(K,N) and GE(K,N)\mathrm{GE}(K,N). The verification of these examples relies crucially on the criteria from Proposition 3.7 and Theorem 4.7.

6.1. Schur multipliers over matrix algebras

A Schur multiplier AA over the n×nn\times n matrix algebra Mn()M_{n}(\mathbb{C}) is a linear map of the form:

Aeij:=aijeij,Ae_{ij}:=a_{ij}e_{ij},

where aija_{ij}\in\mathbb{C} and {eij}i,j=1n\{e_{ij}\}_{i,j=1}^{n} are the matrix units. By Schoenberg’s theorem (see for example [BO08, Appendix D]),

Pt[xij]=etA[xij]=[etaijxij],t0,P_{t}[x_{ij}]=e^{-tA}[x_{ij}]=[e^{-ta_{ij}}x_{ij}],\leavevmode\nobreak\ \leavevmode\nobreak\ t\geq 0,

defines a symmetric quantum Markov semigroup over Mn()M_{n}(\mathbb{C}) if and only if

  1. (a)

    aii=0a_{ii}=0 for all 1in1\leq i\leq n,

  2. (b)

    aij=aji0a_{ij}=a_{ji}\geq 0 for all 1i,jn1\leq i,j\leq n,

  3. (c)

    [aij][a_{ij}] is conditionally negative definite:

    i,j=1nαi¯αjaij0,\sum_{i,j=1}^{n}\overline{\alpha_{i}}\alpha_{j}a_{ij}\leq 0,

    whenever α1,,αn\alpha_{1},\dots,\alpha_{n} are complex numbers such that j=1nαj=0\sum_{j=1}^{n}\alpha_{j}=0.

If this is the case, then there exists a real Hilbert space HH and elements a(j)Ha(j)\in H, 1jn1\leq j\leq n, such that

aij=a(i)a(j)2, 1i,jn.a_{ij}=\|a(i)-a(j)\|^{2},\leavevmode\nobreak\ \leavevmode\nobreak\ 1\leq i,j\leq n.

Suppose that (ek)1kd(e_{k})_{1\leq k\leq d} is an orthonormal basis of HH. Define for each 1kd1\leq k\leq d

vk:=j=1na(j),ekejjMn().v_{k}:=\sum_{j=1}^{n}\langle a(j),e_{k}\rangle e_{jj}\in M_{n}(\mathbb{C}).

Then for any 1i,jn1\leq i,j\leq n:

[vk,eij]=vkeijeijvk=a(i)a(j),ekeij,[v_{k},e_{ij}]=v_{k}e_{ij}-e_{ij}v_{k}=\langle a(i)-a(j),e_{k}\rangle e_{ij},

and

[vk,[vk,eij]]=|a(i)a(j),ek|2eij.[v_{k},[v_{k},e_{ij}]]=|\langle a(i)-a(j),e_{k}\rangle|^{2}e_{ij}.

By the choice of (ek)(e_{k}), we have

k=1d[vk,[vk,eij]]=a(i)a(j)2eij=aijeij.\sum_{k=1}^{d}[v_{k},[v_{k},e_{ij}]]=\|a(i)-a(j)\|^{2}e_{ij}=a_{ij}e_{ij}.

Therefore,

A=k=1[vk,[vk,]],A=\sum_{k=1}[v_{k},[v_{k},\cdot]],

and it is easy to see that [vk,A]=A[vk,][v_{k},A\cdot]=A[v_{k},\cdot] for each kk. So by Propositions 3.13 and 4.10 we have CBE(0,d)\mathrm{CBE}(0,d) and CGE(0,d)\mathrm{CGE}(0,d) for any operator mean.

6.2. Herz-Schur multipliers over group algebras

Let GG be a finite group. Suppose that λ\lambda is the left-regular representation, i.e. for gGg\in G,

λg:2(G)2(G),λg𝟙h=𝟙gh,\lambda_{g}\colon\ell_{2}(G)\to\ell_{2}(G),\,\lambda_{g}\mathds{1}_{h}=\mathds{1}_{gh},

where 𝟙h\mathds{1}_{h} is the delta function at hh. The group algebra of GG is then the (complex) linear span of {λggG}\{\lambda_{g}\mid g\in G\}, denoted by [G]\mathbb{C}[G]. It carries a canonical tracial state τ\tau given by τ(x)=x𝟙e,𝟙e\tau(x)=\langle x\mathds{1}_{e},\mathds{1}_{e}\rangle, where ee is the unit element of GG.

We say that :G[0,)\ell\colon G\to[0,\infty) is a conditionally negative definite length function if (e)=0\ell(e)=0, (g1)=(g)\ell(g^{-1})=\ell(g) for all gGg\in G and

gGα¯gαh(g1h)0\sum_{g\in G}\bar{\alpha}_{g}\alpha_{h}\ell(g^{-1}h)\leq 0

whenever αg\alpha_{g}, gGg\in G, are complex numbers such that gGαg=0\sum_{g\in G}\alpha_{g}=0. By Schoenberg’s theorem (see for example [BO08, Appendix D]), there exists a 1-cocycle (H,π,b)(H,\pi,b) consisting of a real Hilbert space HH of dimension dimH|G|1\dim H\leq\lvert G\rvert-1, a unitary representation π:GB(H)\pi\colon G\to B(H) and a map b:GHb\colon G\to H satisfying the cocycle condition

b(gh)=b(g)+π(g)b(h)b(gh)=b(g)+\pi(g)b(h)

for g,hGg,h\in G such that (g)=b(g)2\ell(g)=\lVert b(g)\rVert^{2}.

Every conditionally negative definite length function \ell on GG induces a τ\tau-symmetric quantum Markov semigroup (Pt)(P_{t}) on [G]\mathbb{C}[G] characterized by Ptλg=et(g)λgP_{t}\lambda_{g}=e^{-t\ell(g)}\lambda_{g} for gGg\in G. Let e1,,ede_{1},\dots,e_{d} be an orthonormal basis of HH. As argued in [WZ20] (or similar to the Schur multipliers case), the generator \mathcal{L} of (Pt)(P_{t}) can be written as

=j=1d[vj,[vj,]]\mathcal{L}=\sum_{j=1}^{d}[v_{j},[v_{j},\cdot\,]]

with d=dimHd=\dim H and

vj:2(G)2(G),vj𝟙h=b(h),ej𝟙h.v_{j}\colon\ell_{2}(G)\to\ell_{2}(G),\,v_{j}\mathds{1}_{h}=\langle b(h),e_{j}\rangle\mathds{1}_{h}.

The operators vjv_{j} are not contained in [G]\mathbb{C}[G] in general, but one can extend \mathcal{L} to a linear operator on B(2(G))B(\ell_{2}(G)) by the same formula, and a direct computation shows [vj,]=[vj,][v_{j},\mathcal{L}\,\cdot\,]=\mathcal{L}[v_{j},\cdot\,]. By Propositions 3.13 and 4.10, (Pt)(P_{t}) satisfies CBE(0,d)\mathrm{CBE}(0,d) and CGE(0,d)\mathrm{CGE}(0,d) for any operator mean.

Example 6.1.

The cyclic group n={0,1,,n1}\mathbb{Z}_{n}=\{0,1,\dots,n-1\}; see [JZ15a, Example 5.9] or [WZ20, Example 5.7]: Its group (von Neumann) algebra is spanned by λk,0kn1\lambda_{k},0\leq k\leq n-1. One can embed n\mathbb{Z}_{n} to 2n\mathbb{Z}_{2n}, so let us assume that nn is even. The word length of knk\in\mathbb{Z}_{n} is given by (k)=min{k,nk}\ell(k)=\min\{k,n-k\}. The associated 1-cocycle can be chosen with H=n2H=\mathbb{R}^{\frac{n}{2}} and b:nn2b\colon\mathbb{Z}_{n}\to\mathbb{R}^{\frac{n}{2}} via

b(k)={0,k=0,j=1kej,1kn2,j=kn2+1n2ej,n2+1kn1,b(k)=\begin{cases}0,&k=0,\\ \sum_{j=1}^{k}e_{j},&1\leq k\leq\frac{n}{2},\\ \sum_{j=k-\frac{n}{2}+1}^{\frac{n}{2}}e_{j},&\frac{n}{2}+1\leq k\leq n-1,\end{cases}

where (ej)1jn2(e_{j})_{1\leq j\leq\frac{n}{2}} is an orthonormal basis of n2\mathbb{R}^{\frac{n}{2}}. Thus the quantum Markov semigroup associated with \ell satisfies CBE(0,n/2)\mathrm{CBE}(0,n/2) and CGE(0,n/2)\mathrm{CGE}(0,n/2) for any operator mean.

Example 6.2.

The symmetric group SnS_{n}; see [WZ20, Example 5.8]: Let \ell be the length function induced by the (non-normalized) Hamming metric, that is, (σ)=#{j:σ(j)j}\ell(\sigma)=\#\{j:\sigma(j)\neq j\}. Let AσMn()A_{\sigma}\in M_{n}(\mathbb{R}) be the permutation matrix associated with σ\sigma, i.e., Aσδj=δσ(j)A_{\sigma}\delta_{j}=\delta_{\sigma(j)}. Then the associated cocycle is given by H=L2(Mn(),12tr)H=L^{2}(M_{n}(\mathbb{R}),\frac{1}{2}\mathrm{tr}), b(σ)=Aσ1b(\sigma)=A_{\sigma}-1 and π(σ)=Aσ\pi(\sigma)=A_{\sigma}. Thus the quantum Markov semigroup associated with \ell satisfies CBE(0,d)\mathrm{CBE}(0,d) and CGE(0,d)\mathrm{CGE}(0,d) for any operator mean with d=min{|Sn|1,n2}d=\min\{|S_{n}|-1,n^{2}\}.

6.3. Depolarizing Semigroup

Let τ\tau be the normalized trace on =Md()\mathcal{M}=M_{d}(\mathbb{C}). The depolarizing semigroup Md()M_{d}(\mathbb{C}) is defined by

Pt:Md()Md(),Pta=eta+(1et)τ(a)𝟏.P_{t}\colon M_{d}(\mathbb{C})\to M_{d}(\mathbb{C}),\,P_{t}a=e^{-t}a+(1-e^{-t})\tau(a)\mathbf{1}.

Its generator is given by a=aτ(a)𝟏\mathcal{L}a=a-\tau(a)\mathbf{1}. We will show that (Pt)(P_{t}) satisfies BE(1/2,2d)\mathrm{BE}(1/2,2d) and GE(1/2,2d)\mathrm{GE}(1/2,2d) for any operator mean Λ\Lambda.

Recall that the generator admits a Lindblad form:

=j𝒥jj.\mathcal{L}=\sum_{j\in\mathcal{J}}\partial_{j}^{\dagger}\partial_{j}.

Since j(𝟏)=0\partial_{j}(\mathbf{1})=0, we have jPt=etj.\partial_{j}P_{t}=e^{-t}\partial_{j}. Then

Ptaρ2=e2taρ2.\lVert\partial P_{t}a\rVert_{\rho}^{2}=e^{-2t}\lVert\partial a\rVert_{\rho}^{2}.

Fix aa\in\mathcal{M} and ρ+\rho\in\mathcal{M}_{+} with τ(ρ)=1\tau(\rho)=1. By positive homogeneity (Lemma 2.1 (a)), concavity (Lemma 2.1 (b)) and the definition of operator mean Λ\Lambda, we get

jaPtρ2=\displaystyle\|\partial_{j}a\|^{2}_{P_{t}\rho}= Λ(etL(ρ)+(1et)L(𝟏),etR(ρ)+(1et)R(𝟏))ja,ja\displaystyle\langle\Lambda(e^{-t}L(\rho)+(1-e^{-t})L(\mathbf{1}),e^{-t}R(\rho)+(1-e^{-t})R(\mathbf{1}))\partial_{j}a,\partial_{j}a\rangle
\displaystyle\geq etΛ(L(ρ),R(ρ))ja,ja+(1et)ja,ja.\displaystyle e^{-t}\langle\Lambda(L(\rho),R(\rho))\partial_{j}a,\partial_{j}a\rangle+(1-e^{-t})\langle\partial_{j}a,\partial_{j}a\rangle.

So

aPtρ2etaρ2+(1et)a,a.\|\partial a\|^{2}_{P_{t}\rho}\geq e^{-t}\|\partial a\|_{\rho}^{2}+(1-e^{-t})\langle\mathcal{L}a,a\rangle.

Note that 2=\mathcal{L}^{2}=\mathcal{L} and Ptρd𝟏P_{t}\rho\leq d\mathbf{1}. All combined, and using Cauchy-Schwarz inequality for τ(ρ)\tau(\rho\cdot), we obtain

aPtρ2\displaystyle\lVert\partial a\rVert^{2}_{P_{t}\rho} etaρ2+(1et)a,a\displaystyle\geq e^{-t}\lVert\partial a\rVert_{\rho}^{2}+(1-e^{-t})\langle\mathcal{L}a,\mathcal{L}a\rangle
etaρ2+1etdτ(|a|2Ptρ)\displaystyle\geq e^{-t}\lVert\partial a\rVert_{\rho}^{2}+\frac{1-e^{-t}}{d}\tau(\lvert\mathcal{L}a\rvert^{2}P_{t}\rho)
etaρ2+1etd|τ((a)Ptρ)|2\displaystyle\geq e^{-t}\lVert\partial a\rVert_{\rho}^{2}+\frac{1-e^{-t}}{d}\lvert\tau((\mathcal{L}a)P_{t}\rho)\rvert^{2}
=etPtaρ2+1etd|(a,Ptρ)|2,\displaystyle=e^{t}\lVert\partial P_{t}a\rVert_{\rho}^{2}+\frac{1-e^{-t}}{d}\lvert\mathcal{E}(a,P_{t}\rho)\rvert^{2},

or equivalently,

Ptaρ2etaPtρ2f(t)|(a,Ptρ)|2.\lVert\partial P_{t}a\rVert_{\rho}^{2}\leq e^{-t}\lVert\partial a\rVert^{2}_{P_{t}\rho}-f(t)\lvert\mathcal{E}(a,P_{t}\rho)\rvert^{2}.

Here f(t)=(ete2t)/df(t)=(e^{-t}-e^{-2t})/d, and it is easy to see f(0)=0f(0)=0 and f(0)=1/df^{\prime}(0)=1/d. By Remark 4.5, (Pt)(P_{t}) satisfies GE(1/2,2d)\mathrm{GE}(1/2,2d).

Choosing Λ\Lambda as the left trivial mean, we actually proved (without using Cauchy-Schwarz inequality):

τ(Pt(|a|2)ρ)etτ(|a|2ρ)+1etdτ(|a|2Ptρ).\tau(P_{t}(|\partial a|^{2})\rho)\geq e^{-t}\tau(|\partial a|^{2}\rho)+\frac{1-e^{-t}}{d}\tau(|\mathcal{L}a|^{2}P_{t}\rho).

Since both sides agree at t=0t=0, we obtain by taking derivative at t=0t=0 that

Γ2(a)12Γ(a)+12d|a|2.\Gamma_{2}(a)\geq\frac{1}{2}\Gamma(a)+\frac{1}{2d}|\mathcal{L}a|^{2}.

So (Pt)(P_{t}) satisfies BE(1/2,2d)\mathrm{BE}(1/2,2d).

7. Curvature-dimension conditions without assuming tracial symmetry

In plenty of applications one encounters quantum Markov semigroups that are not necessarily tracially symmetric, but only satisfy the detailed balance condition σ\sigma-DBC (with σ𝟏\sigma\neq\mathbf{1}) we mentioned in Section 2. Many of the results from this article still apply in this case, with one important caveat, as we will discuss here.

The definition of the Bakry–Émery gradient estimate BE(K,N)\mathrm{BE}(K,N) makes sense for arbitrary quantum Markov semigroups on matrix algebras without any change, and all the consequences we proved stay valid in this more general setting with essentially the same proofs.

The gradient estimate GE(K,N)\mathrm{GE}(K,N) relies on the Lindblad form of the generator of the semigroup. By Alicki’s theorem, a similar Lindblad form exists for generators of quantum Markov semigroups satisfying the σ\sigma-DBC, and the norms ξρ\lVert\xi\rVert_{\rho} have been defined in this setting in [CM17, CM20] – in fact, instead of a single operator mean one can choose a family of operator connections that depends on the index jj. With this norm, one can formulate GE(K,N)\mathrm{GE}(K,N) as

Ptaρ2e2KtaPtρ21e2KtKN|τ((Pta)ρ)|2,\lVert\partial P_{t}a\rVert_{\rho}^{2}\leq e^{-2Kt}\lVert\partial a\rVert^{2}_{P_{t}^{\dagger}\rho}-\frac{1-e^{-2Kt}}{KN}\lvert\tau((\mathcal{L}P_{t}a)\rho)\rvert^{2},

where one now has to distinguish between PtP_{t} and PtP_{t}^{\dagger} because of the lack of tracial symmetry.

The connection between a generalization of the metric 𝒲\mathcal{W}, the semigroup (Pt)(P_{t}) and the relative entropy still remains true in this more general setting with an appropriate modification of the definition of 𝒲\mathcal{W} [CM17, CM20], so that the identification of GE(K,N)\mathrm{GE}(K,N) with the (K,N)(K,N)-convexity condition for an entropy functional from Theorem 5.1 along with its applications also has an appropriate analog for quantum Markov semigroups satisfying the σ\sigma-DBC.

However, the criteria from Proposition 3.7 and Theorem 4.7, which actually allow us to verify BE(K,N)\mathrm{BE}(K,N) and GE(K,N)\mathrm{GE}(K,N) in concrete examples, rely crucially on the Lindblad form of generators of tracially symmetric quantum Markov semigroups and do not immediately carry over to the σ\sigma-detailed balance case. Thus the question of proving BE(K,N)\mathrm{BE}(K,N) and GE(K,N)\mathrm{GE}(K,N) for not necessarily tracially symmetric quantum Markov semigroups remains open, so its usefulness in this case is still to be proven.

References

  • [AB20] N. Javier Buitrago Aza and Diego Andrés Barbosa Trujillo. Entropy Power Inequality in Fermionic Quantum Computation. arXiv e-prints, arXiv:2008.05532, 2020.
  • [AGS14a] Luigi Ambrosio, Nicola Gigli, and Giuseppe Savaré. Calculus and heat flow in metric measure spaces and applications to spaces with Ricci bounds from below. Invent. Math., 195(2):289–391, 2014.
  • [AGS14b] Luigi Ambrosio, Nicola Gigli, and Giuseppe Savaré. Metric measure spaces with Riemannian Ricci curvature bounded from below. Duke Math. J., 163(7):1405–1490, 2014.
  • [AGS15] Luigi Ambrosio, Nicola Gigli, and Giuseppe Savaré. Bakry-Émery curvature-dimension condition and Riemannian Ricci curvature bounds. Ann. Probab., 43(1):339–404, 2015.
  • [Ali76] Robert Alicki. On the detailed balance condition for non-Hamiltonian systems. Rep. Math. Phys., 10(2):249–258, 1976.
  • [BÉ85] Dominique Bakry and Michel Émery. Diffusions hypercontractives. In Séminaire de probabilités, XIX, 1983/84, volume 1123 of Lecture Notes in Math., pages 177–206. Springer, Berlin, 1985.
  • [BGJ20a] Michael Brannan, Li Gao, and Marius Junge. Complete Logarithmic Sobolev inequalities via Ricci curvature bounded below. arXiv e-prints, arXiv:2007.06138, 2020.
  • [BGJ20b] Michael Brannan, Li Gao, and Marius Junge. Complete Logarithmic Sobolev Inequalities via Ricci Curvature Bounded Below II. arXiv e-prints, arXiv:2008.12038, 2020.
  • [BO08] Nathanial P. Brown and Narutaka Ozawa. CC^{*}-algebras and finite-dimensional approximations, volume 88 of Graduate Studies in Mathematics. American Mathematical Society, Providence, RI, 2008.
  • [CM14] Eric A. Carlen and Jan Maas. An analog of the 2-Wasserstein metric in non-commutative probability under which the fermionic Fokker-Planck equation is gradient flow for the entropy. Comm. Math. Phys., 331(3):887–926, 2014.
  • [CM17] Eric A. Carlen and Jan Maas. Gradient flow and entropy inequalities for quantum Markov semigroups with detailed balance. J. Funct. Anal., 273(5):1810–1869, 2017.
  • [CM20] Eric A. Carlen and Jan Maas. Non-commutative Calculus, Optimal Transport and Functional Inequalities in Dissipative Quantum Systems. J. Stat. Phys., 178(2):319–378, 2020.
  • [Cos85] Max H. M. Costa. A new entropy power inequality. IEEE Trans. Inform. Theory, 31(6):751–760, 1985.
  • [DPT18] Giacomo De Palma and Dario Trevisan. The conditional entropy power inequality for bosonic quantum systems. Comm. Math. Phys., 360(2):639–662, 2018.
  • [DR20] Nilanjana Datta and Cambyse Rouzé. Relating Relative Entropy, Optimal Transport and Fisher Information: A Quantum HWI Inequality. Ann. Henri Poincaré, 2020.
  • [EKS15] Matthias Erbar, Kazumasa Kuwada, and Karl-Theodor Sturm. On the equivalence of the entropic curvature-dimension condition and Bochner’s inequality on metric measure spaces. Invent. Math., 201(3):993–1071, 2015.
  • [EM12] Matthias Erbar and Jan Maas. Ricci curvature of finite Markov chains via convexity of the entropy. Arch. Ration. Mech. Anal., 206(3):997–1038, 2012.
  • [HKV17] Stefan Huber, Robert König, and Anna Vershynina. Geometric inequalities from phase space translations. J. Math. Phys., 58(1):012206, 28, 2017.
  • [JM10] Marius Junge and Tao Mei. Noncommutative Riesz transforms—a probabilistic approach. Amer. J. Math., 132(3):611–680, 2010.
  • [JZ15a] Marius Junge and Qiang Zeng. Noncommutative martingale deviation and Poincaré type inequalities with applications. Probab. Theory Related Fields, 161(3-4):449–507, 2015.
  • [JZ15b] Marius Junge and Qiang Zeng. Subgaussian 1-cocycles on discrete groups. J. Lond. Math. Soc. (2), 92(2):242–264, 2015.
  • [KA80] Fumio Kubo and Tsuyoshi Ando. Means of positive linear operators. Math. Ann., 246(3):205–224, 1980.
  • [KS14] Robert König and Graeme Smith. The entropy power inequality for quantum systems. IEEE Trans. Inform. Theory, 60(3):1536–1548, 2014.
  • [Li20] Haojian Li. Complete Sobolev Type Inequalities. arXiv e-prints, arXiv:2008.09278, 2020.
  • [LJL20] Haojian Li, Marius Junge, and Nicholas LaRacuente. Graph Hörmander Systems. arXiv e-prints, arXiv:2006.14578, 2020.
  • [LL20] Songzi Li and Xiang-Dong Li. On the Shannon entropy power on Riemannian manifolds and Ricci flow. arXiv e-prints, arXiv:2001.00410, 2020.
  • [LMP18] Shiping Liu, Florentin Münch, and Norbert Peyerimhoff. Bakry-Émery curvature and diameter bounds on graphs. Calc. Var. Partial Differential Equations, 57(2):Paper No. 67, 9, 2018.
  • [LV09] John Lott and Cédric Villani. Ricci curvature for metric-measure spaces via optimal transport. Ann. of Math. (2), 169(3):903–991, 2009.
  • [Maa11] Jan Maas. Gradient flows of the entropy for finite Markov chains. J. Funct. Anal., 261(8):2250–2292, 2011.
  • [Mie13] Alexander Mielke. Geodesic convexity of the relative entropy in reversible Markov chains. Calc. Var. Partial Differential Equations, 48(1-2):1–31, 2013.
  • [MM17] Markus Mittnenzweig and Alexander Mielke. An entropic gradient structure for Lindblad equations and couplings of quantum systems to macroscopic models. J. Stat. Phys., 167(2):205–233, 2017.
  • [Stu06a] Karl-Theodor Sturm. On the geometry of metric measure spaces. I. Acta Math., 196(1):65–131, 2006.
  • [Stu06b] Karl-Theodor Sturm. On the geometry of metric measure spaces. II. Acta Math., 196(1):133–177, 2006.
  • [Vil00] Cédric Villani. A short proof of the “concavity of entropy power”. IEEE Trans. Inform. Theory, 46(4):1695–1696, 2000.
  • [vRS05] Max-Konstantin von Renesse and Karl-Theodor Sturm. Transport inequalities, gradient estimates, entropy, and Ricci curvature. Comm. Pure Appl. Math., 58(7):923–940, 2005.
  • [Wir18] Melchior Wirth. A Noncommutative Transport Metric and Symmetric Quantum Markov Semigroups as Gradient Flows of the Entropy. arXiv e-prints, arXiv:1808.05419, 2018.
  • [WZ20] Melchior Wirth and Haonan Zhang. Complete gradient estimates of quantum Markov semigroups. arXiv e-prints, arXiv:2007.13506, 2020.