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

Quantum concentration inequalities

Giacomo De Palma [email protected] Scuola Normale Superiore, 56126 Pisa, Italy Cambyse Rouzé [email protected] Technische Universität München, 85748 Garching, Germany
Abstract

We establish Transportation Cost Inequalities (TCIs) with respect to the quantum Wasserstein distance by introducing quantum extensions of well-known classical methods: First, we generalize the Dobrushin uniqueness condition to prove that Gibbs states of 1D commuting Hamiltonians satisfy a TCI at any positive temperature and provide conditions under which this first result can be extended to non-commuting Hamiltonians. Next, using a non-commutative version of Ollivier’s coarse Ricci curvature, we prove that high temperature Gibbs states of commuting Hamiltonians on arbitrary hypergraphs H=(V,E)H=(V,E) satisfy a TCI with constant scaling as O(|V|)O(|V|). Third, we argue that the temperature range for which the TCI holds can be enlarged by relating it to recently established modified logarithmic Sobolev inequalities. Fourth, we prove that the inequality still holds for fixed points of arbitrary reversible local quantum Markov semigroups on regular lattices, albeit with slightly worsened constants, under a seemingly weaker condition of local indistinguishability of the fixed points. Finally, we use our framework to prove Gaussian concentration bounds for the distribution of eigenvalues of quasi-local observables and argue the usefulness of the TCI in proving the equivalence of the canonical and microcanonical ensembles and an exponential improvement over the weak Eigenstate Thermalization Hypothesis.

1 Introduction

Given a random variable XX of law μ\mu taking values on a metric space (Ω,dΩ)(\Omega,d_{\Omega}) and a function f:Ωf:\Omega\to\mathbb{R}, a concentration of measure inequality quantifies the probability that the random variable f(X)f(X) deviates from its mean or its median. Since the early age of the theory, concentration inequalities have seen many new methods, refinements and exciting applications to various areas of mathematics [1, 2, 3]. Among the different classes of concentration inequalities, Gaussian concentration is arguably the most standard one: the measure μ\mu is said to be sub-Gaussian if there exist constants K,κ>0K,\kappa>0 such that, for all AΩA\subseteq\Omega with μ(A)1/2\mu(A)\geq 1/2 we have for any r0r\geq 0

μ({xΩ:dΩ(x,A)<r})1Keκr2.\displaystyle\mu\left(\{x\in\Omega:\,d_{\Omega}(x,A)<r\}\right)\geq 1-Ke^{-\kappa r^{2}}\,. (1)

In her seminal work [4], Marton made the beautiful observation that the above behavior can be obtained as a consequence of a transportation cost inequality: if there exists c>0c>0 such that, for any probability measure ν<<μ\nu<\!<\mu,

W1(μ,ν)cS(νμ),\displaystyle W_{1}(\mu,\nu)\leq\sqrt{c\,S(\nu\|\mu)}\,, (TC(c)\operatorname{TC}(c))

then (1) holds with constants κ=1c\kappa=\frac{1}{c} and K=1K=1 for all r>cln2r>\sqrt{c\ln 2}. Here, S(νμ)S(\nu\|\mu) refers to the relative entropy between the measures ν\nu and μ\mu, whereas the quantity W1(μ,ν)W_{1}(\mu,\nu) in (TC(c)\operatorname{TC}(c)) is the Wasserstein distance between the two measures μ,ν\mu,\nu:

W1(μ,ν):=supfL1|𝔼μ[f]𝔼ν[f]|.\displaystyle W_{1}(\mu,\nu):=\sup_{\|f\|_{L}\leq 1}\,\big{|}\mathbb{E}_{\mu}[f]-\mathbb{E}_{\nu}[f]\big{|}\,.

Later, [5] proved that transportation cost inequalities are in fact equivalent to the property of sub-Gaussianity: more precisely, (TC(c)\operatorname{TC}(c)) holds if and only if for all Lipschitz functions f:Ωf:\Omega\to\mathbb{R},

μ(|f(X)𝔼μ[f(X)]|>t)2et2c,t0.\displaystyle\mathbb{P}_{\mu}\Big{(}\big{|}f(X)-\mathbb{E}_{\mu}[f(X)]\big{|}>t\Big{)}\leq 2\,e^{-\frac{t^{2}}{c}}\,,\forall t\geq 0\,.

One of the main advantages of transportation cost inequalities is their tensorization property: assume that μ\mu satisfies TC(c)\operatorname{TC}(c), then μn\mu^{\otimes n} satisfies TC(nc)\operatorname{TC}(nc) for all nn\in\mathbb{N}, where the set Ωn\Omega^{n} is provided with the metric

dn(xn,yn):=i=1ndΩ(xi,yi).\displaystyle d_{n}(x^{n},y^{n}):=\sum_{i=1}^{n}\,d_{\Omega}(x_{i},y_{i})\,.

Perhaps the simplest example of that sort is given by taking Ωn=[d]n\Omega^{n}=[d]^{n} endowed with the Hamming distance dHd_{H}. In the case n=1n=1, the corresponding Wasserstein distance reduces to the total variation, and TC(1/2)\operatorname{TC}(1/2) holds for any measure μ\mu, since it simply reduces to Pinsker’s inequality. For n1n\geq 1, μn\mu^{\otimes n} satisfies TC(n/2)\operatorname{TC}(n/2).

While the theory of concentration inequalities for i.i.d. random variables is by now well understood, things become more challenging when the random variables are allowed to depend on each other [3, 6]. One way to extend concentration bounds to weakly dependent random variables is to assume that their joint law μ\mu satisfies the so-called Dobrushin uniqueness condition [6]. Dobrushin’s uniqueness condition plays an important role in the study of Gibbs measures in the one-phase region, however it often turns out to be a very strong requirement on the measure μ\mu. More recently, Marton gave an attempt at extending the i.i.d. theory beyond the mere Gibbs setting [7]. Her main result consists in a logarithmic Sobolev inequality for a generic measure μ\mu - well known to imply transportation cost inequalities - under the so-called Dobrushin–Shlosman mixing condition [8], the latter condition being weaker than Dobrushin’s uniqueness condition. As mentioned in [9], such paths to establish Gaussian concentration suffer from the difficulty of deriving explicit constants. Moreover, the result of [7] also relies on the crucial assumption that the measure μ\mu has full support.

Recently, concentration inequalities have attracted much attention in the communities of random matrix theory, quantum information theory and operator algebras [10, 11, 12, 13, 14, 15, 16, 17, 18]. In [14], a quantum Wasserstein distance of order 1 (or quantum W1W_{1} distance) was defined on the set of the quantum states of nn qudits with the property that it strictly reduces to the classical Wasserstein distance on [d]n[d]^{n} for states that are diagonal in the computational basis. This quantum generalization of the Wasserstein distance is based on the notion of neighboring states. Two quantum states of nn qudits are neighboring if they differ only in one qudit, i.e., if they coincide after that qudit is discarded. The quantum W1W_{1} distance is then that induced by the maximum norm that assigns distance at most one to every couple of neighboring states [14, Definition 4]. Such norm is called quantum W1W_{1} norm and is denoted with W1\left\|\cdot\right\|_{W_{1}}. The quantum W1W_{1} norm proposed in Ref. [14] admits a dual formulation in terms of a quantum generalization of the Lipschitz constant. Denoting with 𝒪n\mathcal{O}_{n} the set of the observables of nn qudits, the Lipschitz constant of the observable H𝒪nH\in\mathcal{O}_{n} is defined as [14, Section V]

HL:=2maxi[n]min{HH(i):H(i)𝒪n does not act on the i-th qudit}.\displaystyle\|H\|_{L}:=2\max_{i\in[n]}\min\left\{\left\|H-H^{(i)}\right\|_{\infty}:H^{(i)}\in\mathcal{O}_{n}\text{ does not act on the $i$-th qudit}\right\}\,. (2)

Then, the quantum W1W_{1} distance between the states ρ\rho and ω\omega can also be expressed as [14, Section V]

ρωW1=max{Tr[(ρω)H]:H𝒪n,HL1}.\displaystyle\left\|\rho-\omega\right\|_{W_{1}}=\max\left\{\mathrm{Tr}\left[\left(\rho-\omega\right)H\right]:\,H\in\mathcal{O}_{n},\,\|H\|_{L}\leq 1\right\}\,. (3)

Moreover, in [14] it was showed that TC(n/2)\operatorname{TC}(n/2) holds for any tensor product ω=ω1ωn\omega=\omega_{1}\otimes\ldots\otimes\omega_{n} of quantum states, hence extending Marton’s original inequality with the exact same constant: for any state ρ\rho of nn qudits,

ρω1ωnW1n2S(ρω1ωn),\displaystyle\left\|\rho-\omega_{1}\otimes\ldots\otimes\omega_{n}\right\|_{W_{1}}\leq\sqrt{\frac{n}{2}\,S(\rho\|\omega_{1}\otimes\ldots\otimes\omega_{n})}\,,

where S(ρω):=Tr[ρ(lnρlnω)]S(\rho\|\omega):=\mathrm{Tr}[\rho\,(\ln\rho-\ln\omega)] denotes Umegaki’s relative entropy between the states ρ\rho and ω\omega.

Main results:

In this paper, we prove that any of the following conditions implies a transportation cost inequality:

  • (i)

    A non-commutative Dobrushin uniqueness condition (section 3);

  • (ii)

    A generalization of Ollivier’s coarse Ricci curvature bound (section 4);

  • (iii)

    A modified logarithmic Sobolev inequality condition (section 5);

  • (iv)

    A condition of local indistinguishability of the state (section 6).

Each of these methods comes with its strengths and weaknesses:

  • (i)

    The non-commutative Dobrushin uniqueness condition implies a nontrivial TCI at any temperature (see Remark 3), but the scaling of the constant cc with the number of subsystems is optimal only in one dimension (see Remark 2).

  • (ii)

    The coarse Ricci curvature bound provides TC inequalities for essentially any geometry, but it is only valid above a threshold temperature that depends on the locality of the Hamiltonian. Furthermore, such threshold temperature is in practice strictly larger than the true critical temperature (see Proposition 9).

  • (iii)

    Quantum modified logarithmic Sobolev inequalities are typically more difficult to prove than their classical counterparts, and are currently only proven to hold in specific cases. However, for one-dimensional systems, a recently derived modified logarithmic Sobolev inequality [19, 20] provides us with TC (up to polylogarithmic overhead) at any positive temperature.

  • (iv)

    The condition of local indistinguishability of the state for regular lattices. Although the condition can be checked for classical systems, we do not yet have a way to prove it in the quantum setting.

We conclude the article with two natural applications of our bounds. First, we derive Gaussian concentration bounds for a large class of Lipschitz observables whenever the state ω\omega is that of a commuting Hamiltonian at large enough temperature (section 7). Second, we argue on the use of the transportation cost inequality in proving the equivalence between the microcanonical and the canonical ensembles and an exponential improvement over the weak Eigenstate Thermalization Hypothesis (section 8).

2 Notations and basic definitions

Given a finite set VV, we denote by V=vVv\mathcal{H}_{V}=\bigotimes_{v\in V}\mathcal{H}_{v} the Hilbert space of n=|V|n=|V| qudits (i.e., vd\mathcal{H}_{v}\equiv\mathbb{C}^{d} for all vVv\in V) and by V\mathcal{B}_{V} the algebra of linear operators on V\mathcal{H}_{V}. 𝒪V\mathcal{O}_{V} corresponds to the space of self-adjoint linear operators on V\mathcal{H}_{V}, whereas 𝒪VT𝒪V\mathcal{O}^{T}_{V}\subset\mathcal{O}_{V} is the subspace of traceless self-adjoint linear operators. 𝒪V+\mathcal{O}_{V}^{+} denotes the cone of positive semidefinite linear operators on V\mathcal{H}_{V} and 𝒮V𝒪V+\mathcal{S}_{V}\subset\mathcal{O}_{V}^{+} denotes the set of quantum states. We denote by 𝒫V\mathcal{P}_{V} the set of probability measures on [d]V[d]^{V}. For any subset AVA\subseteq V, we use the standard notations 𝒪A,𝒮A\mathcal{O}_{A},\mathcal{S}_{A}\ldots for the corresponding objects defined on subsystem AA. Given a state ρ𝒮V\rho\in\mathcal{S}_{V}, we denote by ρA\rho_{A} its marginal onto the subsystem AA. For any X𝒪VX\in\mathcal{O}_{V}, we denote by X1\|X\|_{1} its trace norm. The identity on 𝒪v\mathcal{O}_{v}, vVv\in V, is denoted by 𝕀v\mathbb{I}_{v}.

Given two states ρ,ω𝒮V\rho,\,\omega\in\mathcal{S}_{V} such that supp(ρ)supp(ω)\operatorname{supp}(\rho)\subseteq\operatorname{supp}(\omega), their quantum relative entropy is defined as [21, 22, 23]

S(ρω)=Tr[ρ(lnρlnω)].\displaystyle S(\rho\|\omega)=\mathrm{Tr}\big{[}\rho\,(\ln\rho-\ln\omega)\big{]}\,. (4)

Whenever ρ=ρAB\rho=\rho_{AB} is a bipartite state and ω=ρAρB\omega=\rho_{A}\otimes\rho_{B}, their relative entropy reduces to the mutual information

I(A;B)ρ:=S(ρABρAρB).\displaystyle I(A;B)_{\rho}:=S(\rho_{AB}\|\rho_{A}\otimes\rho_{B})\,. (5)

In the next sections, we also utilize the measured relative entropy [24, 25, 26, 27]

S𝕄(ρω):=sup(𝒳,M)S(Pρ,MPσ,M),\displaystyle S_{\mathbb{M}}(\rho\|\omega):=\sup_{(\mathcal{X},M)}\,S(P_{\rho,M}\|P_{\sigma,M})\,, (6)

where the supremum above is over all positive operator valued measures MM that map the input quantum state to a probability distribution on a finite set 𝒳\mathcal{X} with probability mass function given by Pρ,M(x)=TrρM(x)P_{\rho,M}(x)=\mathrm{Tr}\rho M(x).

In this paper, we study inequalities relating the W1W_{1} distance between two states to their relative entropy. More precisely, for a fixed state ω𝒮V\omega\in\mathcal{S}_{V}, we are interested in upper bounding the best constant C(ω)>0C(\omega)>0 such that, for all ρ𝒮V\rho\in\mathcal{S}_{V} with supp(ρ)supp(ω)\operatorname{supp}(\rho)\subseteq\operatorname{supp}(\omega),

ρωW1C(ω)S(ρω).\displaystyle\left\|\rho-\omega\right\|_{W_{1}}\leq\,\sqrt{C(\omega)\,S(\rho\|\omega)}\,. (7)

In general, given a constant cC(ω)c\geq C(\omega), we refer to the above inequality for C(ω)C(\omega) replaced by cc as a transportation cost inequality, denoted by TC(c)\operatorname{TC}(c). As mentioned in the introduction, the following holds [14, Theorem 2]:

Proposition 1.

For any product state ω𝒮V\omega\in\mathcal{S}_{V},

C(ω)|V|2.C(\omega)\leq\frac{|V|}{2}\,. (8)

In the next sections, we aim at recovering the linear dependence of the constant C(ω)C(\omega) on the size n=|V|n=|V| of the system under various measures of independence.

We will need the following properties of the quantum W1W_{1} distance:

Proposition 2 ([14, Proposition 2]).

The quantum W1W_{1} distance coincides with the trace distance for quantum states that differ in only one site, i.e., for any X𝒪VTX\in\mathcal{O}_{V}^{T} such that TrvX=0\mathrm{Tr}_{v}X=0 for some vVv\in V we have

XW1=12X1.\left\|X\right\|_{W_{1}}=\frac{1}{2}\left\|X\right\|_{1}\,. (9)
Proposition 3 ([14, Proposition 5]).

The quantum W1W_{1} distance between two quantum states that differ only in the region AVA\subseteq V is at most 2|A|2\left|A\right| times their trace distance, i.e., for any X𝒪VTX\in\mathcal{O}_{V}^{T} such that TrAX=0\mathrm{Tr}_{A}X=0 we have

XW1|A|X1.\left\|X\right\|_{W_{1}}\leq\left|A\right|\left\|X\right\|_{1}\,. (10)
Proposition 4 (Tensorization [14, Proposition 4]).

The quantum W1W_{1} distance is additive with respect to the tensor product, i.e., let A,BA,\,B be disjoint subsets of VV. Then, for any ρA,σA𝒮A\rho_{A},\,\sigma_{A}\in\mathcal{S}_{A} and any ρB,σB𝒮B\rho_{B},\,\sigma_{B}\in\mathcal{S}_{B} we have

ρAρBσAσBW1=ρAσAW1+ρBσBW1.\left\|\rho_{A}\otimes\rho_{B}-\sigma_{A}\otimes\sigma_{B}\right\|_{W_{1}}=\left\|\rho_{A}-\sigma_{A}\right\|_{W_{1}}+\left\|\rho_{B}-\sigma_{B}\right\|_{W_{1}}\,. (11)
Proposition 5 ([14, Proposition 13]).

Let Φ:𝒪V𝒪V\Phi:\mathcal{O}_{V}\to\mathcal{O}_{V} be a quantum channel. For any vVv\in V, let AvVA_{v}\subseteq V be the light-cone of the site vv, i.e., the minimum subset of VV such that TrAiΦ(X)=0\mathrm{Tr}_{A_{i}}\Phi(X)=0 for any X𝒪VX\in\mathcal{O}_{V} such that TrvX=0\mathrm{Tr}_{v}X=0. Then, Φ\Phi can expand the quantum W1W_{1} distance by at most twice the size of the largest light-cone, i.e., for any X𝒪VTX\in\mathcal{O}_{V}^{T} we have

Φ(X)W12maxvV|Av|XW1.\left\|\Phi(X)\right\|_{W_{1}}\leq 2\max_{v\in V}\left|A_{v}\right|\left\|X\right\|_{W_{1}}\,. (12)
Proposition 6 ([14, Proposition 15]).

For any H𝒪VH\in\mathcal{O}_{V} and any vVv\in V we have

H𝕀v1dTrvHHL.\left\|H-\mathbb{I}_{v}\otimes\frac{1}{d}\,\mathrm{Tr}_{v}H\right\|_{\infty}\leq\left\|H\right\|_{L}\,. (13)
Proposition 7 ([14, Corollary 1]).

For any ρ,σ𝒮V\rho,\,\sigma\in\mathcal{S}_{V},

ρσW112vVρvσv1,\left\|\rho-\sigma\right\|_{W_{1}}\geq\frac{1}{2}\sum_{v\in V}\left\|\rho_{v}-\sigma_{v}\right\|_{1}\,, (14)

and equality holds whenever both ρ\rho and σ\sigma are product states.

Theorem 1 (W1W_{1} continuity of the entropy [14, Theorem 1]).

For any ρ,σ𝒮V\rho,\,\sigma\in\mathcal{S}_{V} we have

|S(ρ)S(σ)|g(ρσW1)+ρσW1ln(d2|V|),\left|S(\rho)-S(\sigma)\right|\leq g\left(\left\|\rho-\sigma\right\|_{W_{1}}\right)+\left\|\rho-\sigma\right\|_{W_{1}}\ln\left(d^{2}\left|V\right|\right)\,, (15)

where for any t0t\geq 0

g(t)=(t+1)ln(t+1)tlnt.g(t)=\left(t+1\right)\ln\left(t+1\right)-t\ln t\,. (16)

3 Dobrushin uniqueness condition

In this section, we consider a spin chain and prove the transportation cost inequality under a quantum generalization of Dobrushin’s uniqueness condition [6]. Such condition is formulated in terms of the conditional probability distributions of the state of a subset of VV conditioned on the state of a second disjoint subset of VV. Therefore, formulating a quantum version of Dobrushin’s uniqueness condition requires a quantum counterpart of the conditional probability distribution. In the classical setting, given two random variables XX and YY taking values in finite sets and with joint probability distribution ωXY\omega_{XY}, the conditional probability distribution ωY|X\omega_{Y|X} of YY given XX with probability mass function

ωY|X=x(y)=ωXY(x,y)ωX(x)\omega_{Y|X=x}(y)=\frac{\omega_{XY}(x,y)}{\omega_{X}(x)} (17)

represents the knowledge that we have on YY when we know only the value of XX. We can associate to such conditional distribution the stochastic map ΦXXY\Phi_{X\to XY} that has as input a probability distribution pXp_{X} for XX and as output the joint probability distribution of XYXY with probability mass function given by

ΦXXY(pX)(x,y)=ωY|X=x(y)pX(x)=ωXY(x,y)ωX(x)pX(x).\Phi_{X\to XY}(p_{X})(x,y)=\omega_{Y|X=x}(y)\,p_{X}(x)=\frac{\omega_{XY}(x,y)}{\omega_{X}(x)}\,p_{X}(x)\,. (18)

In the quantum setting, we consider a bipartite quantum system ABAB and a joint quantum state ωAB\omega_{AB} of ABAB. The quantum counterpart of the stochastic map (18) is called quantum recovery map [28, 29] and its action on a quantum state ρA\rho_{A} of AA is

ΦAAB(ρA)=ωAB1it2ωAit12ρAωA1+it2ωAB1+it2𝑑μ0(t),\Phi_{A\to AB}(\rho_{A})=\int_{\mathbb{R}}\omega_{AB}^{\frac{1-it}{2}}\,\omega_{A}^{\frac{it-1}{2}}\,\rho_{A}\,\omega_{A}^{-\frac{1+it}{2}}\,\omega_{AB}^{\frac{1+it}{2}}\,d\mu_{0}(t)\,, (19)

where μ0\mu_{0} is the probability distrbution on \mathbb{R} with density

dμ0(t)=πdt2(cosh(πt)+1).d\mu_{0}(t)=\frac{\pi\,dt}{2\left(\cosh(\pi t)+1\right)}\,. (20)

We stress that (19) reduces to (18) whenever ρA\rho_{A}, ωA\omega_{A} and ωAB\omega_{AB} commute. If AA is in the state ωA\omega_{A}, the recovery map ΦAAB\Phi_{A\to AB} recovers the joint state ωAB\omega_{AB}, i.e., ΦAAB(ωA)=ωAB\Phi_{A\to AB}(\omega_{A})=\omega_{AB}. The relevance of the recovery map comes from the recoverability theorem [29], which states that ΦAAB\Phi_{A\to AB} can recover a generic joint state ρAB\rho_{AB} from its marginal ρA\rho_{A} if removing the subsystem BB does not significantly decrease the relative entropy between ρ\rho and ω\omega. More precisely, for any quantum state ρAB\rho_{AB} of ABAB we have

S(ρABωAB)S(ρAωA)S𝕄(ρABΦAAB(ρA)).S(\rho_{AB}\|\omega_{AB})-S(\rho_{A}\|\omega_{A})\geq S_{\mathbb{M}}(\rho_{AB}\|\Phi_{A\to AB}(\rho_{A}))\,. (21)

We consider the setting where VV is partitioned as

V=A1Am.V=A_{1}\sqcup\ldots\sqcup A_{m}\,. (22)

For any i[m]i\in[m], we denote with A1iA_{1}^{i} the union A1AiA_{1}\sqcup\ldots\sqcup A_{i}. The recoverability theorem implies the following Lemma 1, which we will employ several times:

Lemma 1.

For any ρ,ω𝒮V\rho,\,\omega\in\mathcal{S}_{V} we have

S(ρω)12m(i=1mρA1iΦA1i1A1i(ρA1i)1)2,\displaystyle S(\rho\|\omega)\geq\frac{1}{2m}\left(\sum_{i=1}^{m}\left\|\rho_{A_{1}^{i}}-\Phi_{A_{1}^{i-1}\to A_{1}^{i}}(\rho_{A_{1}^{i}})\right\|_{1}\right)^{2}\,, (23)
(12mi=1mρA1iΦA1i1A1i(ρA1i)1)21exp(S(ρω)m),\displaystyle\left(\frac{1}{2m}\sum_{i=1}^{m}\left\|\rho_{A_{1}^{i}}-\Phi_{A_{1}^{i-1}\to A_{1}^{i}}(\rho_{A_{1}^{i}})\right\|_{1}\right)^{2}\leq 1-\exp\left(-\frac{S(\rho\|\omega)}{m}\right)\,, (24)

where ΦA1i1A1i\Phi_{A_{1}^{i-1}\to A_{1}^{i}} are the recovery maps associated to ω\omega.

Proof.

Eq. (19) and Pinsker’s inequality imply for any i[m]i\in[m]

S(ρA1iωA1i)S(ρA1i1ωA1i1)S𝕄(ρA1iΦA1i1A1i(ρA1i1))12ρA1iΦA1i1A1i(ρA1i1)12.S(\rho_{A_{1}^{i}}\|\omega_{A_{1}^{i}})-S(\rho_{A_{1}^{i-1}}\|\omega_{A_{1}^{i-1}})\geq S_{\mathbb{M}}\left(\rho_{A_{1}^{i}}\left\|\Phi_{A_{1}^{i-1}\to A_{1}^{i}}(\rho_{A_{1}^{i-1}})\right.\right)\geq\frac{1}{2}\left\|\rho_{A_{1}^{i}}-\Phi_{A_{1}^{i-1}\to A_{1}^{i}}(\rho_{A_{1}^{i-1}})\right\|_{1}^{2}\,. (25)

Summing (25) over ii and using the convexity of the square function yields

S(ρω)12i=1mρA1iΦA1i1A1i(ρA1i1)1212m(i=1mρA1iΦA1i1A1i(ρA1i)1)2.S(\rho\|\omega)\geq\frac{1}{2}\sum_{i=1}^{m}\left\|\rho_{A_{1}^{i}}-\Phi_{A_{1}^{i-1}\to A_{1}^{i}}(\rho_{A_{1}^{i-1}})\right\|_{1}^{2}\geq\frac{1}{2m}\left(\sum_{i=1}^{m}\left\|\rho_{A_{1}^{i}}-\Phi_{A_{1}^{i-1}\to A_{1}^{i}}(\rho_{A_{1}^{i}})\right\|_{1}\right)^{2}\,. (26)

The claim (23) follows.

With an analogous proof, applying the improved Pinsker’s inequality

14στ121eS𝕄(στ)\frac{1}{4}\left\|\sigma-\tau\right\|_{1}^{2}\leq 1-e^{-S_{\mathbb{M}}(\sigma\|\tau)} (27)

and the convexity of the function tln(1t2)t\mapsto-\ln\left(1-t^{2}\right) we get

S(ρω)mln(1(12mi=1mρA1iΦA1i1A1i(ρA1i)1)2).S(\rho\|\omega)\geq-m\ln\left(1-\left(\frac{1}{2m}\sum_{i=1}^{m}\left\|\rho_{A_{1}^{i}}-\Phi_{A_{1}^{i-1}\to A_{1}^{i}}(\rho_{A_{1}^{i}})\right\|_{1}\right)^{2}\right)\,. (28)

The claim (24) follows. ∎

The following property of the recovery map will be fundamental:

Lemma 2.

Let ωABC\omega_{ABC} be a joint state of the tripartite quantum system ABCABC. Let us assume that ωABC\omega_{ABC} is Markovian, i.e.,

I(A;C|B)ω=0.I(A;C|B)_{\omega}=0\,. (29)

Then, the recovery map ΦABABC\Phi_{AB\to ABC} associated to ωABC\omega_{ABC} does not act on the subsystem AA.

Proof.

From the characterization of the states that saturate the strong subadditivity [30], the Hilbert space B\mathcal{H}_{B} of BB has a decomposition

B=i=1kBiLBiR,\mathcal{H}_{B}=\bigoplus_{i=1}^{k}\mathcal{H}_{B_{i}^{L}}\otimes\mathcal{H}_{B_{i}^{R}}\,, (30)

where the Hilbert spaces {BiL}i=1k\left\{\mathcal{H}_{B_{i}^{L}}\right\}_{i=1}^{k} and {BiL}i=1k\left\{\mathcal{H}_{B_{i}^{L}}\right\}_{i=1}^{k} are pairwise orthogonal, and ωABC\omega_{ABC} can be expressed as

ωABC=i=1kpiωABiL(i)ωBiRC(i),\omega_{ABC}=\bigoplus_{i=1}^{k}p_{i}\,\omega_{AB_{i}^{L}}^{(i)}\otimes\omega_{B_{i}^{R}C}^{(i)}\,, (31)

where pp is a probability distribution on [k][k], and each ωABiL(i)\omega_{AB_{i}^{L}}^{(i)} or ωBiRC(i)\omega_{B_{i}^{R}C}^{(i)} is a quantum state with support in the corresponding ABiL\mathcal{H}_{A}\otimes\mathcal{H}_{B_{i}^{L}} or BiRC\mathcal{H}_{B_{i}^{R}}\otimes\mathcal{H}_{C}. We have for any quantum state ρAB\rho_{AB} of ABAB

ΦABABC(ρAB)=ωABC1it2ωABit12ρABωAB1+it2ωABC1+it2𝑑μ0(t).\Phi_{AB\to ABC}(\rho_{AB})=\int_{\mathbb{R}}\omega_{ABC}^{\frac{1-it}{2}}\,\omega_{AB}^{\frac{it-1}{2}}\,\rho_{AB}\,\omega_{AB}^{-\frac{1+it}{2}}\,\omega_{ABC}^{\frac{1+it}{2}}\,d\mu_{0}(t)\,. (32)

We have for any zz\in\mathbb{C} that

ωABCzωABz=i=1k(ωBiRC(i))z(ωBiR(i))z\omega_{ABC}^{z}\,\omega_{AB}^{-z}=\bigoplus_{i=1}^{k}\left(\omega_{B_{i}^{R}C}^{(i)}\right)^{z}\left(\omega_{B_{i}^{R}}^{(i)}\right)^{-z} (33)

does not act on AA, and the claim follows choosing z=(1it)/2z=\left(1-it\right)/2. ∎

3.1 Markovian case

In this subsection, we assume that ω𝒮V\omega\in\mathcal{S}_{V} is a one-dimensional quantum Markov state. More preciesly, let {A1,,Am}\left\{A_{1},\,\ldots,\,A_{m}\right\} be a partition of VV and let K=max(|A1|,,|Am|)K=\max\left(\left|A_{1}\right|,\,\ldots,\,\left|A_{m}\right|\right). Then, we assume that

I(Ai;A1i2|Ai1)ω=0I(A_{i};A_{1}^{i-2}|A_{i-1})_{\omega}=0 (34)

for any i=3,,mi=3,\,\ldots,\,m. For any i[m]i\in[m], let Φi\Phi_{i} be the recovery map (19) associated to ωA1i\omega_{A_{1}^{i}} that recovers AiA_{i} from A1i1A_{1}^{i-1}. From Lemma 2, Φi\Phi_{i} acts only on Ai1A_{i-1}, i.e., it is a map Φi:𝒪Ai1𝒪Ai1Ai\Phi_{i}:\mathcal{O}_{A_{i-1}}\to\mathcal{O}_{A_{i-1}A_{i}}. We also define

Φ~i=TrAi1Φi:𝒪Ai1𝒪Ai.\tilde{\Phi}_{i}=\mathrm{Tr}_{A_{i-1}}\circ\Phi_{i}:\mathcal{O}_{A_{i-1}}\to\mathcal{O}_{A_{i}}\,. (35)

We can now state the main result of this Section:

Theorem 2.

Let us assume that for any i[m]i\in[m], Φ~i\tilde{\Phi}_{i} is a contraction with respect to the trace norm for all the couples of quantum states of A1i1A_{1}^{i-1} that differ only on the subsystem Ai1A_{i-1}, i.e., that coincide after discarding Ai1A_{i-1}. More precisely, we assume that there exists 0η<10\leq\eta<1 such that for any i[m]i\in[m] and any X𝒪A1i1TX\in\mathcal{O}_{A_{1}^{i-1}}^{T} with TrAi1X=0\mathrm{Tr}_{A_{i-1}}X=0 we have

Φ~i(X)1ηX1.\left\|\tilde{\Phi}_{i}(X)\right\|_{1}\leq\eta\left\|X\right\|_{1}\,. (36)

Then, we have

C(ω)2mK2(11η+1)2.C(\omega)\leq 2m\,K^{2}\left(\frac{1}{1-\eta}+1\right)^{2}\,. (37)

Furthermore, for any ρ𝒮V\rho\in\mathcal{S}_{V} we have

ρωW1K(11η+1)2m1eS(ρω)m.\left\|\rho-\omega\right\|_{W_{1}}\leq K\left(\frac{1}{1-\eta}+1\right)2m\sqrt{1-e^{-\frac{S(\rho\|\omega)}{m}}}\,. (38)
Proof.

Let ρ𝒮V\rho\in\mathcal{S}_{V}. On the one hand, we have from Lemma 3

ρωW1\displaystyle\left\|\rho-\omega\right\|_{W_{1}} =i=1m(ΦmΦi+1)(ρA1iΦi(ρA1i1))W1\displaystyle=\left\|\sum_{i=1}^{m}(\Phi_{m}\circ\ldots\circ\Phi_{i+1})(\rho_{A_{1}^{i}}-\Phi_{i}(\rho_{A_{1}^{i-1}}))\right\|_{W_{1}}
i=1m(ΦmΦi+1)(ρA1iΦi(ρA1i1))W1\displaystyle\leq\sum_{i=1}^{m}\left\|(\Phi_{m}\circ\ldots\circ\Phi_{i+1})(\rho_{A_{1}^{i}}-\Phi_{i}(\rho_{A_{1}^{i-1}}))\right\|_{W_{1}}
K(11η+1)i=1mρA1iΦi(ρA1i1)1.\displaystyle\leq K\left(\frac{1}{1-\eta}+1\right)\sum_{i=1}^{m}\left\|\rho_{A_{1}^{i}}-\Phi_{i}(\rho_{A_{1}^{i-1}})\right\|_{1}\,. (39)

On the other hand, we have from (23) of Lemma 1

S(ρω)12m(i=1mρA1iΦi(ρA1i1)1)2,S(\rho\|\omega)\geq\frac{1}{2m}\left(\sum_{i=1}^{m}\left\|\rho_{A_{1}^{i}}-\Phi_{i}(\rho_{A_{1}^{i-1}})\right\|_{1}\right)^{2}\,, (40)

and the claim (37) follows. The claim (38) follows by employing (24) in place of (23). ∎

Remark 1.

Condition (36) holds for some η<1\eta<1 iff Φ~i\tilde{\Phi}_{i} strictly decreases the trace distance between any two quantum states that differ only in the subsystem AiA_{i} on which Φ~i\tilde{\Phi}_{i} acts, i.e., that coincide after discarding AiA_{i}. We expect this condition to hold for any strictly positive temperature.

Remark 2.

An example of quantum state satisfying (34) is a Gibbs state of a nearest-neighbor Hamiltonian on the DD-dimensional cubic lattice Λ=[L]D\Lambda=[L]^{D}, where x,yΛx,\,y\in\Lambda are neighbors iff xy1=1\left\|x-y\right\|_{1}=1. We can then choose m=L+1m=L+1 and

Ai={xΛ:x1=i},i=0,,L,A_{i}=\left\{x\in\Lambda:x_{1}=i\right\}\,,\qquad i=0,\,\ldots,\,L\,, (41)

with

K=(L+1)D1,K=\left(L+1\right)^{D-1}\,, (42)

and get from Theorem 2

C(ω)2(L+1)2D1(11η+1)2=2|V|2D1D(11η+1)2.C(\omega)\leq 2\left(L+1\right)^{2D-1}\left(\frac{1}{1-\eta}+1\right)^{2}=2\left|V\right|^{\frac{2D-1}{D}}\left(\frac{1}{1-\eta}+1\right)^{2}\,. (43)

We stress that, assuming that η\eta remains bounded away from 11, we get C(ω)=O(|V|)C(\omega)=O(|V|) iff D=1D=1, i.e., for one-dimensional systems.

Remark 3.

We can choose

η\displaystyle\eta =max{Φ~i(X)ωAiTrAi1X1:i[m],X𝒪A1i1,X1=1}\displaystyle=\max\left\{\left\|\tilde{\Phi}_{i}(X)-\omega_{A_{i}}\otimes\mathrm{Tr}_{A_{i-1}}X\right\|_{1}:i\in[m]\,,\;X\in\mathcal{O}_{A_{1}^{i-1}}\,,\;\left\|X\right\|_{1}=1\right\}
maxi[m]Φ~iωAiTrAi1,\displaystyle\leq\max_{i\in[m]}\left\|\tilde{\Phi}_{i}-\omega_{A_{i}}\otimes\mathrm{Tr}_{A_{i-1}}\right\|_{\diamond}\,, (44)

where ωAiTrAi1:𝒪Ai1𝒪Ai\omega_{A_{i}}\otimes\mathrm{Tr}_{A_{i-1}}:\mathcal{O}_{A_{i-1}}\to\mathcal{O}_{A_{i}} is the quantum channel that replaces the input quantum state with ωAi\omega_{A_{i}} and

Φ=sup{(Φ𝕀())(X)1:X(2),X1=1}\left\|\Phi\right\|_{\diamond}=\sup\left\{\left\|(\Phi\otimes\mathbb{I}_{\mathcal{B}(\mathcal{H})})(X)\right\|_{1}:X\in\mathcal{B}(\mathcal{H}^{\otimes 2})\,,\;\left\|X\right\|_{1}=1\right\} (45)

denotes the diamond norm of the linear map Φ\Phi on ()\mathcal{B}(\mathcal{H}).

Proposition 8.

Let ω𝒮V\omega\in\mathcal{S}_{V} satisfy (34), and assume

a=maxi[m1]S(ωAiωAi+1ωAiAi+1)<12,a=\max_{i\in[m-1]}S_{\infty}(\omega_{A_{i}}\otimes\omega_{A_{i+1}}\|\omega_{A_{i}A_{i+1}})<\frac{1}{2}\,, (46)

where

S(ρσ)=lninf{λ:ρλσ}S_{\infty}(\rho\|\sigma)=\ln\inf\left\{\lambda\in\mathbb{R}:\rho\leq\lambda\,\sigma\right\} (47)

denotes the quantum max-divergence [31] between the quantum states ρ\rho and σ\sigma. Then, we can choose in (36)

η=2a.\eta=\sqrt{2\,a}\,. (48)
Proof.

From Remark 3, we can choose

η=maxi[m]max|ψiΦ~i(|ψiψi|)ωAiTrAi1|ψiψi|1maxi[m]max|ψiΦi(|ψiψi|)ωAi|ψiψi|1,\eta=\max_{i\in[m]}\max_{|\psi_{i}\rangle}\left\|\tilde{\Phi}_{i}(|\psi_{i}\rangle\langle\psi_{i}|)-\omega_{A_{i}}\otimes\mathrm{Tr}_{A_{i-1}}|\psi_{i}\rangle\langle\psi_{i}|\right\|_{1}\leq\max_{i\in[m]}\max_{|\psi_{i}\rangle}\left\|\Phi_{i}(|\psi_{i}\rangle\langle\psi_{i}|)-\omega_{A_{i}}\otimes|\psi_{i}\rangle\langle\psi_{i}|\right\|_{1}\,, (49)

where each |ψi|\psi_{i}\rangle is a unit vector in A1i1\mathcal{H}_{A_{1}^{i-1}}. We have from Pinsker’s inequality

Φi(|ψiψi|)|ψiψi|ωAi12S𝕄(|ψiψi|ωAiΦi(|ψiψi|)).\left\|\Phi_{i}(|\psi_{i}\rangle\langle\psi_{i}|)-|\psi_{i}\rangle\langle\psi_{i}|\otimes\omega_{A_{i}}\right\|_{1}\leq\sqrt{2\,S_{\mathbb{M}}(|\psi_{i}\rangle\langle\psi_{i}|\otimes\omega_{A_{i}}\|\Phi_{i}(|\psi_{i}\rangle\langle\psi_{i}|))}\,. (50)

(46) implies

lnωAi1AilnωAi1+lnωAia.\ln\omega_{A_{i-1}A_{i}}\geq\ln\omega_{A_{i-1}}+\ln\omega_{A_{i}}-a\,. (51)

From the characterization of the states that saturate the strong subadditivity [30] we get

lnωA1i1+lnωAi1Ai=lnωAi1+lnωA1i,\ln\omega_{A_{1}^{i-1}}+\ln\omega_{A_{i-1}A_{i}}=\ln\omega_{A_{i-1}}+\ln\omega_{A_{1}^{i}}\,, (52)

therefore, (51) can be rewritten as

lnωA1ilnωA1i1+lnωAia.\ln\omega_{A_{1}^{i}}\geq\ln\omega_{A_{1}^{i-1}}+\ln\omega_{A_{i}}-a\,. (53)

Choosing in (25) ρA1i=|ψiψi|ωAi\rho_{A_{1}^{i}}=|\psi_{i}\rangle\langle\psi_{i}|\otimes\omega_{A_{i}} we get with the help of (53)

S𝕄(|ψiψi|ωAiΦi(|ψiψi|))ψi|(lnωA1i1TrAi[ωAilnωA1i])|ψiS(Ai)ωa,S_{\mathbb{M}}(|\psi_{i}\rangle\langle\psi_{i}|\otimes\omega_{A_{i}}\|\Phi_{i}(|\psi_{i}\rangle\langle\psi_{i}|))\leq\langle\psi_{i}|\left(\ln\omega_{A_{1}^{i-1}}-\mathrm{Tr}_{A_{i}}\left[\omega_{A_{i}}\ln\omega_{A_{1}^{i}}\right]\right)|\psi_{i}\rangle-S(A_{i})_{\omega}\leq a\,, (54)

and the claim follows. ∎

Remark 4.

Condition (36) is reminiscent of the so-called Dobrushin uniqueness condition (see [6, Theorem 4]).

3.2 Non-Markovian states

Here, we prove an alternative version of Theorem 2 where the Markov condition (34) is replaced by exponential decay of correlations.

Theorem 3.

Let V=[n]V=[n] be a one-dimensional lattice, and let ω𝒮V\omega\in\mathcal{S}_{V}. For any i[n]i\in[n], let Φi\Phi_{i} be the recovery map associated to ω1i\omega_{1\ldots i} that recovers the site ii from the sites 1i11\ldots i-1. We assume that ω\omega has exponentially decaying correlations, in the sense that there exist C0C\geq 0 and 0η<10\leq\eta<1 such that for any i[n]i\in[n], any k=0,,max(i,ni)k=0,\,\ldots,\,\max(i,\,n-i) and any τ𝒮1i\tau\in\mathcal{S}_{1\ldots i},

Trik+1i+k(ΦnΦi+1)(τ1i)τ1ikωi+k+1n1Cηk.\left\|\mathrm{Tr}_{i-k+1\ldots i+k}(\Phi_{n}\circ\ldots\circ\Phi_{i+1})(\tau_{1\ldots i})-\tau_{1\ldots i-k}\otimes\omega_{i+k+1\ldots n}\right\|_{1}\leq C\,\eta^{k}\,. (55)

We also assume that for any i[n]i\in[n], any k=0,,i1k=0,\,\ldots,\,i-1 and any τ𝒮1i1\tau\in\mathcal{S}_{1\ldots i-1}

TrikiΦi(τ1i1)τ1ik11Cηk.\left\|\mathrm{Tr}_{i-k\ldots i}\Phi_{i}(\tau_{1\ldots i-1})-\tau_{1\ldots i-k-1}\right\|_{1}\leq C\,\eta^{k}\,. (56)

Then,

C(ω)8n(2+C+11ηln(C2n)2lnη)2.C(\omega)\leq 8\,n\left(2+\frac{C+1}{1-\eta}-\frac{\ln\left(C^{2}\,n\right)}{2\ln\eta}\right)^{2}\,. (57)
Proof.

From (23) of Lemma 1, we have for any ρ𝒮V\rho\in\mathcal{S}_{V}

S(ρω)12n(i=1nρ1iΦi(ρ1i1)1)2.S(\rho\|\omega)\geq\frac{1}{2n}\left(\sum_{i=1}^{n}\left\|\rho_{1\ldots i}-\Phi_{i}(\rho_{1\ldots i-1})\right\|_{1}\right)^{2}\,. (58)

We have

ρωW1\displaystyle\left\|\rho-\omega\right\|_{W_{1}} =i=1n(ΦnΦi+1)(ρ1iΦi(ρ1i1))W1\displaystyle=\left\|\sum_{i=1}^{n}(\Phi_{n}\circ\ldots\circ\Phi_{i+1})(\rho_{1\ldots i}-\Phi_{i}(\rho_{1\ldots i-1}))\right\|_{W_{1}}
i=1n(ΦnΦi+1)(ρ1iΦi(ρ1i1))W1.\displaystyle\leq\sum_{i=1}^{n}\left\|(\Phi_{n}\circ\ldots\circ\Phi_{i+1})(\rho_{1\ldots i}-\Phi_{i}(\rho_{1\ldots i-1}))\right\|_{W_{1}}\,. (59)

For any i[n]i\in[n], we have from Lemma 5

(ΦnΦi+1)(ρ1iΦi(ρ1i1))W1\displaystyle\left\|(\Phi_{n}\circ\ldots\circ\Phi_{i+1})(\rho_{1\ldots i}-\Phi_{i}(\rho_{1\ldots i-1}))\right\|_{W_{1}}
2k=0max{i,ni}Trik+1i+k(ΦnΦi+1)(ρ1iΦi(ρ1i1))1\displaystyle\leq 2\sum_{k=0}^{\max\{i,\,n-i\}}\left\|\mathrm{Tr}_{i-k+1\ldots i+k}(\Phi_{n}\circ\ldots\circ\Phi_{i+1})(\rho_{1\ldots i}-\Phi_{i}(\rho_{1\ldots i-1}))\right\|_{1}
2k=0max{i,ni}(ρ1ikTrik+1iΦi(ρ1i1))1+Cηkρ1iΦi(ρ1i1))1).\displaystyle\leq 2\sum_{k=0}^{\max\{i,\,n-i\}}\left(\left\|\rho_{1\ldots i-k}-\mathrm{Tr}_{i-k+1\ldots i}\Phi_{i}(\rho_{1\ldots i-1}))\right\|_{1}+C\,\eta^{k}\left\|\rho_{1\ldots i}-\Phi_{i}(\rho_{1\ldots i-1}))\right\|_{1}\right)\,. (60)

We have for any k0{0,,i1}k_{0}\in\left\{0,\,\ldots,\,i-1\right\}

k=0i1ρ1ikTrik+1iΦi(ρ1i1)1\displaystyle\sum_{k=0}^{i-1}\left\|\rho_{1\ldots i-k}-\mathrm{Tr}_{i-k+1\ldots i}\Phi_{i}(\rho_{1\ldots i-1})\right\|_{1}
k=0k0ρ1iΦi(ρ1i1)1+k=k0+1i1ρ1ikω1ikTrik+1iΦi(ρ1ikω1ik)1\displaystyle\leq\sum_{k=0}^{k_{0}}\left\|\rho_{1\ldots i}-\Phi_{i}(\rho_{1\ldots i-1})\right\|_{1}+\sum_{k=k_{0}+1}^{i-1}\left\|\rho_{1\ldots i-k}-\omega_{1\ldots i-k}-\mathrm{Tr}_{i-k+1\ldots i}\Phi_{i}(\rho_{1\ldots i-k}-\omega_{1\ldots i-k})\right\|_{1}
(k0+1)ρ1iΦi(ρ1i1)1+Ck=k0+1i1ηk1ρ1ikω1ik1\displaystyle\leq\left(k_{0}+1\right)\left\|\rho_{1\ldots i}-\Phi_{i}(\rho_{1\ldots i-1})\right\|_{1}+C\sum_{k=k_{0}+1}^{i-1}\eta^{k-1}\left\|\rho_{1\ldots i-k}-\omega_{1\ldots i-k}\right\|_{1}
(k0+1)ρ1iΦi(ρ1i1)1+Cηk01ηρω1,\displaystyle\leq\left(k_{0}+1\right)\left\|\rho_{1\ldots i}-\Phi_{i}(\rho_{1\ldots i-1})\right\|_{1}+\frac{C\,\eta^{k_{0}}}{1-\eta}\left\|\rho-\omega\right\|_{1}\,, (61)

therefore

(ΦnΦi+1)(ρ1iΦi(ρ1i1))W1\displaystyle\left\|(\Phi_{n}\circ\ldots\circ\Phi_{i+1})(\rho_{1\ldots i}-\Phi_{i}(\rho_{1\ldots i-1}))\right\|_{W_{1}}
2((k0+1+C1η)ρ1iΦi(ρ1i1)1+Cηk01ηρω1),\displaystyle\leq 2\left(\left(k_{0}+1+\frac{C}{1-\eta}\right)\left\|\rho_{1\ldots i}-\Phi_{i}(\rho_{1\ldots i-1})\right\|_{1}+\frac{C\,\eta^{k_{0}}}{1-\eta}\left\|\rho-\omega\right\|_{1}\right)\,, (62)

and

ρωW1\displaystyle\left\|\rho-\omega\right\|_{W_{1}} 2((k0+1+C1η)i=1nρ1iΦi(ρ1i1)1+nCηk01ηρω1)\displaystyle\leq 2\left(\left(k_{0}+1+\frac{C}{1-\eta}\right)\sum_{i=1}^{n}\left\|\rho_{1\ldots i}-\Phi_{i}(\rho_{1\ldots i-1})\right\|_{1}+\frac{n\,C\,\eta^{k_{0}}}{1-\eta}\left\|\rho-\omega\right\|_{1}\right)
2(k0+1+C1+ηk0n1η)2nS(ρω),\displaystyle\leq 2\left(k_{0}+1+C\,\frac{1+\eta^{k_{0}}\sqrt{n}}{1-\eta}\right)\sqrt{2\,n\,S(\rho\|\omega)}\,, (63)

and the claim follows choosing

k0=ln(C2n)2lnη.k_{0}=\left\lceil-\frac{\ln\left(C^{2}\,n\right)}{2\ln\eta}\right\rceil\,. (64)

3.3 Auxiliary lemmas

Lemma 3.

Under the same hypotheses of Theorem 2, for any i[m]i\in[m] and any X𝒪A1iTX\in\mathcal{O}_{A_{1}^{i}}^{T} such that

TrAi1AiX=0\mathrm{Tr}_{A_{i-1}A_{i}}X=0 (65)

we have

(ΦmΦi+1)(X)W1K(11η+1)X1.\left\|(\Phi_{m}\circ\ldots\circ\Phi_{i+1})(X)\right\|_{W_{1}}\leq K\left(\frac{1}{1-\eta}+1\right)\left\|X\right\|_{1}\,. (66)
Proof.

We have from Lemma 4 and from the contractivity of the trace distance

(ΦmΦi+1)(X)W1\displaystyle\left\|(\Phi_{m}\circ\ldots\circ\Phi_{i+1})(X)\right\|_{W_{1}} |Ai1|(ΦmΦi+1)(X)1+(ΦmΦi+1)(TrAi1X)W1\displaystyle\leq\left|A_{i-1}\right|\left\|(\Phi_{m}\circ\ldots\circ\Phi_{i+1})(X)\right\|_{1}+\left\|(\Phi_{m}\circ\ldots\circ\Phi_{i+1})\left(\mathrm{Tr}_{A_{i-1}}X\right)\right\|_{W_{1}}
KX1+(ΦmΦi+1)(TrAi1X)W1.\displaystyle\leq K\left\|X\right\|_{1}+\left\|(\Phi_{m}\circ\ldots\circ\Phi_{i+1})\left(\mathrm{Tr}_{A_{i-1}}X\right)\right\|_{W_{1}}\,. (67)

We have

(ΦmΦi+1)(TrAi1X)W1\displaystyle\left\|(\Phi_{m}\circ\ldots\circ\Phi_{i+1})\left(\mathrm{Tr}_{A_{i-1}}X\right)\right\|_{W_{1}}
|Ai|(ΦmΦi+1)(TrAi1X)1+TrAi(ΦmΦi+1)(TrAi1X)W1\displaystyle\leq\left|A_{i}\right|\left\|(\Phi_{m}\circ\ldots\circ\Phi_{i+1})\left(\mathrm{Tr}_{A_{i-1}}X\right)\right\|_{1}+\left\|\mathrm{Tr}_{A_{i}}(\Phi_{m}\circ\ldots\circ\Phi_{i+1})\left(\mathrm{Tr}_{A_{i-1}}X\right)\right\|_{W_{1}}
KTrAi1X1+(ΦmΦi+2Φ~i+1)(TrAi1X)W1.\displaystyle\leq K\left\|\mathrm{Tr}_{A_{i-1}}X\right\|_{1}+\left\|(\Phi_{m}\circ\ldots\circ\Phi_{i+2}\circ\tilde{\Phi}_{i+1})\left(\mathrm{Tr}_{A_{i-1}}X\right)\right\|_{W_{1}}\,. (68)

Iterating the procedure we get

(ΦmΦi+1)(TrAi1X)W1\displaystyle\left\|(\Phi_{m}\circ\ldots\circ\Phi_{i+1})\left(\mathrm{Tr}_{A_{i-1}}X\right)\right\|_{W_{1}}
K(TrAi1X1+Φ~i+1(TrAi1X)1++(Φ~mΦ~i+1)(TrAi1X)1)\displaystyle\leq K\left(\left\|\mathrm{Tr}_{A_{i-1}}X\right\|_{1}+\left\|\tilde{\Phi}_{i+1}\left(\mathrm{Tr}_{A_{i-1}}X\right)\right\|_{1}+\ldots+\left\|(\tilde{\Phi}_{m}\circ\ldots\circ\tilde{\Phi}_{i+1})\left(\mathrm{Tr}_{A_{i-1}}X\right)\right\|_{1}\right)
+TrAm(Φ~mΦ~i+1)(TrAi1X)W1\displaystyle\phantom{\leq}+\left\|\mathrm{Tr}_{A_{m}}(\tilde{\Phi}_{m}\circ\ldots\circ\tilde{\Phi}_{i+1})\left(\mathrm{Tr}_{A_{i-1}}X\right)\right\|_{W_{1}}
K(1+η++ηmi)TrAi1X1+TrAi1AiX1\displaystyle\leq K\left(1+\eta+\ldots+\eta^{m-i}\right)\left\|\mathrm{Tr}_{A_{i-1}}X\right\|_{1}+\left\|\mathrm{Tr}_{A_{i-1}A_{i}}X\right\|_{1}
K1ηTrAi1X1K1ηX1,\displaystyle\leq\frac{K}{1-\eta}\left\|\mathrm{Tr}_{A_{i-1}}X\right\|_{1}\leq\frac{K}{1-\eta}\left\|X\right\|_{1}\,, (69)

where the last two inequalities follow from (36) and (65), respectively. The claim follows. ∎

Lemma 4.

For any X𝒪VTX\in\mathcal{O}_{V}^{T} and any AVA\subseteq V,

XW1|A|X1+TrAXW1.\left\|X\right\|_{W_{1}}\leq\left|A\right|\left\|X\right\|_{1}+\left\|\mathrm{Tr}_{A}X\right\|_{W_{1}}\,. (70)
Proof.

Without loss of generality, we can assume that V=[n]V=[n] and A=[k]A=[k] for some k[n]k\in[n]. We have

XW1\displaystyle\left\|X\right\|_{W_{1}} X𝕀dTr1XW1+𝕀dTr1XW1=12X𝕀dTr1X1+Tr1XW1\displaystyle\leq\left\|X-\frac{\mathbb{I}}{d}\otimes\mathrm{Tr}_{1}X\right\|_{W_{1}}+\left\|\frac{\mathbb{I}}{d}\otimes\mathrm{Tr}_{1}X\right\|_{W_{1}}=\frac{1}{2}\left\|X-\frac{\mathbb{I}}{d}\otimes\mathrm{Tr}_{1}X\right\|_{1}+\left\|\mathrm{Tr}_{1}X\right\|_{W_{1}}
X1+Tr1XW1,\displaystyle\leq\left\|X\right\|_{1}+\left\|\mathrm{Tr}_{1}X\right\|_{W_{1}}\,, (71)

where the equality follows from Proposition 2 and Proposition 4 and the last inequality follows from the triangle inequality for the trace norm and its contractivity with respect to partial traces. By induction we get

XW1\displaystyle\left\|X\right\|_{W_{1}} (X1++Tr1k1X1)+Tr1kXW1\displaystyle\leq\left(\left\|X\right\|_{1}+\ldots+\left\|\mathrm{Tr}_{1\ldots k-1}X\right\|_{1}\right)+\left\|\mathrm{Tr}_{1\ldots k}X\right\|_{W_{1}}
kX1+Tr1kXW1,\displaystyle\leq k\left\|X\right\|_{1}+\left\|\mathrm{Tr}_{1\ldots k}X\right\|_{W_{1}}\,, (72)

and the claim follows. ∎

Lemma 5.

Let V=[n]V=[n]. Then, for any X𝒪VTX\in\mathcal{O}_{V}^{T},

XW1X1+Tr1X1++Tr1n1X1.\left\|X\right\|_{W_{1}}\leq\left\|X\right\|_{1}+\left\|\mathrm{Tr}_{1}X\right\|_{1}+\ldots+\left\|\mathrm{Tr}_{1\ldots n-1}X\right\|_{1}\,. (73)
Proof.

Follows from Lemma 4. ∎

4 Curvature bound

In the seminal paper [32], Ollivier introduced a generalization of the notion of curvature to generic, possibly discrete, metric spaces. In his framework, the curvature of a metric space (Ω,d)(\Omega,d) endowed with a classical stochastic map PP acting on the probability measures on Ω\Omega is defined as the following contraction property of the Wasserstein distance W1W_{1}: for any two probability measures μ1,μ2\mu_{1},\mu_{2},

W1(P(μ1),P(μ2))(1κ)W1(μ1,μ2).\displaystyle W_{1}(P(\mu_{1}),P(\mu_{2}))\leq\big{(}1-\kappa\big{)}\,W_{1}(\mu_{1},\mu_{2})\,. (74)

The constant κ>0\kappa>0 is called the coarse Ricci curvature of the triple (Ω,d,P)(\Omega,d,P). In particular, it is easy to verify that the existence of a positive coarse Ricci curvature induces the uniqueness of the invariant measure ν\nu for the Markov kernel PP. Moreover, it was recently proven in [33] that Ollivier’s coarse Ricci curvature provides an upper bound on the transportation cost inequality for the measure ν\nu, hence recovering the results from the smooth Riemannian setting.

Here, inspired by the works of [32] and [33], we prove that a contraction of the Lipschitz constant under a certain quantum channel constructed from the Petz recovery maps of the Gibbs state ω\omega can be used to conclude that ω\omega satisfies a transportation cost inequality. In particular, we do not need to assume that the underlying graph is \mathbb{Z}, in contrast with section 3. Let G=(V,E)G=(V,E) be a hypergraph with n=|V|n=|V|, and let H:=AEhAH:=\sum_{A\in E}h_{A} be a Hamiltonian whose local terms hAh_{A} pairwise commute and are supported on the hyperedges AEA\in E. For a given site vVv\in V, we recall the composition of the partial trace Trv\mathrm{Tr}_{v} on vv with the rotated Petz recovery map of vv:

Ψv(ρ)=ΦvTrv(ρ)=ω1it2ωvc1+it2(ρvcIv)ωvc1it2ω1+it2𝑑μ0(t)\displaystyle\Psi_{v}(\rho)=\Phi_{v}\circ\mathrm{Tr}_{v}(\rho)=\int_{\mathbb{R}}\omega^{\frac{1-it}{2}}\omega_{v^{c}}^{\frac{-1+it}{2}}\,(\rho_{v^{c}}\otimes I_{v})\,\omega_{v^{c}}^{\frac{-1-it}{2}}\omega^{\frac{1+it}{2}}\,d\mu_{0}(t) (75)

for the probability density μ0(t):=π2(cosh(πt)+1)1\mu_{0}(t):=\frac{\pi}{2}\big{(}\cosh(\pi t)+1\big{)}^{-1}. Note that since we assumed ω\omega to be the Gibbs state of a commuting Hamiltonian, the map Ψv\Psi_{v} acts non-trivially on the neighborhood of vv

Nv:={AE:vA}.N_{v}:=\bigcup\left\{A\in E:v\in A\right\}\,. (76)

We also introduce the quantum channel

Ψ=1nvVΨv.\Psi=\frac{1}{n}\sum_{v\in V}\Psi_{v}\,. (77)

We assume that Ψ\Psi is a contraction with respect to the W1W_{1} norm, i.e., that

ΨW1W1=maxΔ𝒪VTΨ(Δ)W1ΔW11κn\displaystyle\left\|\Psi\right\|_{W_{1}\to W_{1}}=\max_{\Delta\in\mathcal{O}_{V}^{T}}\frac{\left\|\Psi(\Delta)\right\|_{W_{1}}}{\left\|\Delta\right\|_{W_{1}}}\leq 1-\frac{\kappa}{n} (78)

for some κ>0\kappa>0, in analogy with (74). This contraction property was already derived in Ollivier’s original article [32] as a generalization of Dobrushin’s uniqueness condition. Here, we first prove that this condition implies the transportation cost inequality for the Gibbs state ωωβ:=eβH/TreβH\omega\equiv\omega_{\beta}:=e^{-\beta H}/\mathrm{Tr}\,e^{-\beta H}:

Theorem 4.

With the conditions of the previous paragraph, we have

C(ωβ)2nN2(1eκ)2,\displaystyle C(\omega_{\beta})\leq 2n\,\frac{N^{2}}{\left(1-e^{-\kappa}\right)^{2}}\,, (79)

where N:=maxvV|Nv|N:=\max_{v\in V}|N_{v}|.

Proof.

We have for any state ρ𝒮V\rho\in\mathcal{S}_{V}

ρωβW1i=1nΨi1(ρ)Ψi(ρ)W1+Ψn(ρ)ωβW1.\displaystyle\left\|\rho-\omega_{\beta}\right\|_{W_{1}}\leq\sum_{i=1}^{n}\left\|\Psi^{i-1}(\rho)-\Psi^{i}(\rho)\right\|_{W_{1}}+\left\|\Psi^{n}(\rho)-\omega_{\beta}\right\|_{W_{1}}\,. (80)

The last term can be controlled by ρωβW1\left\|\rho-\omega_{\beta}\right\|_{W_{1}} thanks to the contraction (78):

Ψn(ρ)ωβW1(1κn)nρωβW1eκρωβW1.\displaystyle\left\|\Psi^{n}(\rho)-\omega_{\beta}\right\|_{W_{1}}\leq\left(1-\frac{\kappa}{n}\right)^{n}\left\|\rho-\omega_{\beta}\right\|_{W_{1}}\leq e^{-\kappa}\left\|\rho-\omega_{\beta}\right\|_{W_{1}}\,. (81)

On the other hand, the sum on the right-hand side of (80) can be controlled as follows:

i=1nΨi1(ρ)Ψi(ρ)W1\displaystyle\sum_{i=1}^{n}\left\|\Psi^{i-1}(\rho)-\Psi^{i}(\rho)\right\|_{W_{1}} 1ni=1nvVΨv(Ψi1(ρ))Ψi1(ρ)W1\displaystyle\leq\frac{1}{n}\sum_{i=1}^{n}\sum_{v\in V}\left\|\Psi_{v}(\Psi^{i-1}(\rho))-\Psi^{i-1}(\rho)\right\|_{W_{1}}
Nni=1nvVΨv(Ψi1(ρ))Ψi1(ρ)1,\displaystyle\leq\frac{N}{n}\sum_{i=1}^{n}\sum_{v\in V}\left\|\Psi_{v}(\Psi^{i-1}(\rho))-\Psi^{i-1}(\rho)\right\|_{1}\,, (82)

where the last inequality follows by Proposition 3. Proceeding as in the proof of Lemma 1, by the joint use of Pinsker’s inequality with the recoverability bound followed by the data processing inequality we can further bound the trace distances above so that

i=1nΨi1(ρ)Ψi(ρ)W1\displaystyle\sum_{i=1}^{n}\left\|\Psi^{i-1}(\rho)-\Psi^{i}(\rho)\right\|_{W_{1}} Nni=1nvV2S𝕄(Ψi1(ρ)Ψv(Ψi1(ρ)))\displaystyle\leq\frac{N}{n}\,\sum_{i=1}^{n}\sum_{v\in V}\sqrt{2\,S_{\mathbb{M}}(\Psi^{i-1}(\rho)\|\Psi_{v}(\Psi^{i-1}(\rho)))}
N2i=1nvVS𝕄(Ψi1(ρ)Ψv(Ψi1(ρ)))\displaystyle\leq{N}\sqrt{2\sum_{i=1}^{n}\sum_{v\in V}{S_{\mathbb{M}}(\Psi^{i-1}(\rho)\|\Psi_{v}(\Psi^{i-1}(\rho)))}}
N2i=1nvV(S(Ψi1(ρ)ωβ)S(TrvΨi1(ρ)Trvωβ))\displaystyle\leq N\sqrt{2\sum_{i=1}^{n}\sum_{v\in V}\left(S(\Psi^{i-1}(\rho)\|\omega_{\beta})-S(\mathrm{Tr}_{v}\Psi^{i-1}(\rho)\|\mathrm{Tr}_{v}\omega_{\beta})\right)}
N2i=1nvV(S(Ψi1(ρ)ωβ)S(Ψv(Ψi1(ρ))ωβ))\displaystyle\leq N\sqrt{2\sum_{i=1}^{n}\sum_{v\in V}\left(S(\Psi^{i-1}(\rho)\|\omega_{\beta})-S(\Psi_{v}(\Psi^{i-1}(\rho))\|\omega_{\beta})\right)}
(1)N2ni=1n(S(Ψi1(ρ)ωβ)S(Ψ(Ψi1(ρ))ωβ))\displaystyle\overset{(1)}{\leq}N\sqrt{2n\sum_{i=1}^{n}\left(S(\Psi^{i-1}(\rho)\|\omega_{\beta})-S(\Psi(\Psi^{i-1}(\rho))\|\omega_{\beta})\right)}
N2nS(ρωβ).\displaystyle\leq N\sqrt{2n\,S(\rho\|\omega_{\beta})}\,. (83)

Inequality (1) above uses the concavity of the entropy, so that for any state ρ\rho

1nvV(S(ρωβ)S(Ψv(ρ)ωβ))\displaystyle\frac{1}{n}\sum_{v\in V}\left(S(\rho\|\omega_{\beta})-S(\Psi_{v}(\rho)\|\omega_{\beta})\right) =S(ρωβ)+1nvVS(Ψv(ρ))+1nvVTr[Ψv(ρ)lnωβ]\displaystyle=S(\rho\|\omega_{\beta})+\frac{1}{n}\sum_{v\in V}S(\Psi_{v}(\rho))+\frac{1}{n}\sum_{v\in V}\mathrm{Tr}\left[\Psi_{v}(\rho)\ln\omega_{\beta}\right]
S(ρωβ)+S(Ψ(ρ))+Tr[Ψ(ρ)lnωβ]\displaystyle\leq S(\rho\|\omega_{\beta})+S(\Psi(\rho))+\mathrm{Tr}\left[\Psi(\rho)\ln\omega_{\beta}\right]
=S(ρωβ)S(Ψ(ρ)ωβ).\displaystyle=S(\rho\|\omega_{\beta})-S(\Psi(\rho)\|\omega_{\beta})\,. (84)

Plugging (81) and (83) onto (80), the result follows. ∎

It remains to prove that (78) is satisfied at high enough temperature.

Proposition 9.

There exists an inverse temperature βc>0\beta_{c}>0 such that for all β<βc\beta<\beta_{c}, (78) holds for some constant κ(β)>0\kappa(\beta)>0. In particular, whenever N>1N>1, one can choose

βc=(5NmaxAEhA)1W(116d3),\displaystyle\beta_{c}=(5N\max_{A\in E}\|h_{A}\|_{\infty})^{-1}W\Big{(}\frac{1}{16d^{3}}\Big{)}\,, (85)

where WW denotes the Lambert function and is defined as the inverse of xxexx\mapsto xe^{x}.

Proof.

We have

ΨW1W1=maxΔ𝒪VTΨ(Δ)W1ΔW1.\left\|\Psi\right\|_{W_{1}\to W_{1}}=\max_{\Delta\in\mathcal{O}_{V}^{T}}\frac{\left\|\Psi(\Delta)\right\|_{W_{1}}}{\left\|\Delta\right\|_{W_{1}}}\,. (86)

Any Δ𝒪VT\Delta\in\mathcal{O}_{V}^{T} can be expressed as [14, Section III]

Δ=vVΔv\Delta=\sum_{v\in V}\Delta_{v} (87)

such that for any vVv\in V, Δv𝒪VT\Delta_{v}\in\mathcal{O}_{V}^{T} satisfies TrvΔv=0\mathrm{Tr}_{v}\Delta_{v}=0 and

ΔW1=vVΔvW1=12vVΔv1.\left\|\Delta\right\|_{W_{1}}=\sum_{v\in V}\left\|\Delta_{v}\right\|_{W_{1}}=\frac{1}{2}\sum_{v\in V}\left\|\Delta_{v}\right\|_{1}\,. (88)

Therefore, we have

ΨW1W1=maxvVmax{Ψ(Δv)W1:Δv𝒪VT,TrvΔv=0,Δv1=2}.\left\|\Psi\right\|_{W_{1}\to W_{1}}=\max_{v\in V}\max\left\{\left\|\Psi(\Delta_{v})\right\|_{W_{1}}:\Delta_{v}\in\mathcal{O}_{V}^{T},\,\mathrm{Tr}_{v}\Delta_{v}=0,\,\left\|\Delta_{v}\right\|_{1}=2\right\}\,. (89)

We have

Ψ(Δv)W1\displaystyle\left\|\Psi(\Delta_{v})\right\|_{W_{1}} Ψ(Δv)𝕀vdTrvΨ(Δv)W1+𝕀vdTrvΨ(Δv)W1\displaystyle\leq\left\|\Psi(\Delta_{v})-\frac{\mathbb{I}_{v}}{d}\otimes\mathrm{Tr}_{v}\Psi(\Delta_{v})\right\|_{W_{1}}+\left\|\frac{\mathbb{I}_{v}}{d}\otimes\mathrm{Tr}_{v}\Psi(\Delta_{v})\right\|_{W_{1}}
=12Ψ(Δv)𝕀vdTrvΨ(Δv)1+TrvΨ(Δv)W1\displaystyle=\frac{1}{2}\left\|\Psi(\Delta_{v})-\frac{\mathbb{I}_{v}}{d}\otimes\mathrm{Tr}_{v}\Psi(\Delta_{v})\right\|_{1}+\left\|\mathrm{Tr}_{v}\Psi(\Delta_{v})\right\|_{W_{1}}
12Ψ(Δv)1+12TrvΨ(Δv)1+TrvΨ(Δv)W1,\displaystyle\leq\frac{1}{2}\left\|\Psi(\Delta_{v})\right\|_{1}+\frac{1}{2}\left\|\mathrm{Tr}_{v}\Psi(\Delta_{v})\right\|_{1}+\left\|\mathrm{Tr}_{v}\Psi(\Delta_{v})\right\|_{W_{1}}\,, (90)

where the equality follows from Proposition 2 and Proposition 4. Since TrvΔv=0\mathrm{Tr}_{v}\Delta_{v}=0, we have Ψv(Δv)=0\Psi_{v}(\Delta_{v})=0, and

12Ψ(Δv)112nwVvΨw(Δv)111n.\frac{1}{2}\left\|\Psi(\Delta_{v})\right\|_{1}\leq\frac{1}{2n}\sum_{w\in V\setminus v}\left\|\Psi_{w}(\Delta_{v})\right\|_{1}\leq 1-\frac{1}{n}\,. (91)

For any wVNvw\in V\setminus N_{v} we have

TrvΨw(Δv)=Ψw(TrvΔv)=0.\mathrm{Tr}_{v}\Psi_{w}(\Delta_{v})=\Psi_{w}(\mathrm{Tr}_{v}\Delta_{v})=0\,. (92)

Then,

TrvΨ(Δv)=1nwNvvTrvΨw(Δv).\mathrm{Tr}_{v}\Psi(\Delta_{v})=\frac{1}{n}\sum_{w\in N_{v}\setminus v}\mathrm{Tr}_{v}\Psi_{w}(\Delta_{v})\,. (93)

We have for any wNvvw\in N_{v}\setminus v, recalling that vNwv\in N_{w},

TrNwvTrvΨw(Δv)=TrNwΨw(Δv)=TrNwΔv=0,\mathrm{Tr}_{N_{w}\setminus v}\mathrm{Tr}_{v}\Psi_{w}(\Delta_{v})=\mathrm{Tr}_{N_{w}}\Psi_{w}(\Delta_{v})=\mathrm{Tr}_{N_{w}}\Delta_{v}=0\,, (94)

therefore,

TrvΨw(Δv)W1(N1)TrvΨw(Δv)1,\left\|\mathrm{Tr}_{v}\Psi_{w}(\Delta_{v})\right\|_{W_{1}}\leq\left(N-1\right)\left\|\mathrm{Tr}_{v}\Psi_{w}(\Delta_{v})\right\|_{1}\,, (95)

and

TrvΨ(Δv)W1N1nwNvvTrvΨw(Δv)1.\left\|\mathrm{Tr}_{v}\Psi(\Delta_{v})\right\|_{W_{1}}\leq\frac{N-1}{n}\sum_{w\in N_{v}\setminus v}\left\|\mathrm{Tr}_{v}\Psi_{w}(\Delta_{v})\right\|_{1}\,. (96)

Moreover,

TrvΨ(Δv)11nwNvvTrvΨw(Δv)1.\left\|\mathrm{Tr}_{v}\Psi(\Delta_{v})\right\|_{1}\leq\frac{1}{n}\sum_{w\in N_{v}\setminus v}\left\|\mathrm{Tr}_{v}\Psi_{w}(\Delta_{v})\right\|_{1}\,. (97)

Putting together (4), (91), (96) and (97), we get

Ψ(Δv)W1\displaystyle\left\|\Psi(\Delta_{v})\right\|_{W_{1}} 11n+N12nwNvvTrvΨw(Δv)1\displaystyle\leq 1-\frac{1}{n}+\frac{N-\frac{1}{2}}{n}\sum_{w\in N_{v}\setminus v}\left\|\mathrm{Tr}_{v}\Psi_{w}(\Delta_{v})\right\|_{1}
11n+N12nwNvv(ωwTrvwΔv1+2ΨwωwTrw)\displaystyle\leq 1-\frac{1}{n}+\frac{N-\frac{1}{2}}{n}\sum_{w\in N_{v}\setminus v}\left(\left\|\omega_{w}\otimes\mathrm{Tr}_{vw}\Delta_{v}\right\|_{1}+2\left\|\Psi_{w}-\omega_{w}\otimes\mathrm{Tr}_{w}\right\|_{\diamond}\right)
=11n+2N1nwNvvΨwωwTrw,\displaystyle=1-\frac{1}{n}+\frac{2N-1}{n}\sum_{w\in N_{v}\setminus v}\left\|\Psi_{w}-\omega_{w}\otimes\mathrm{Tr}_{w}\right\|_{\diamond}\,, (98)

where ωwTrw\omega_{w}\otimes\mathrm{Tr}_{w} is the quantum channel that replaces with ωw\omega_{w} the state of the site ww. We then have

ΨW1W111n+2N1nwNvvΨwωwTrw.\left\|\Psi\right\|_{W_{1}\to W_{1}}\leq 1-\frac{1}{n}+\frac{2N-1}{n}\sum_{w\in N_{v}\setminus v}\left\|\Psi_{w}-\omega_{w}\otimes\mathrm{Tr}_{w}\right\|_{\diamond}\,. (99)

We have

ΨwωwTrwω1it2ωwcit12(𝕀wTrw[])ωwcit+12ω1+it2ωwTrw[]𝑑μ0(t).\left\|\Psi_{w}-\omega_{w}\otimes\mathrm{Tr}_{w}\right\|_{\diamond}\leq\int_{\mathbb{R}}\left\|\omega^{\frac{1-it}{2}}\,\omega_{w^{c}}^{\frac{it-1}{2}}\left(\mathbb{I}_{w}\otimes\mathrm{Tr}_{w}\left[\cdot\right]\right)\omega_{w^{c}}^{-\frac{it+1}{2}}\,\omega^{\frac{1+it}{2}}-\omega_{w}\otimes\mathrm{Tr}_{w}\left[\cdot\right]\right\|_{\diamond}d\mu_{0}(t)\,. (100)

Since the Hamiltonian terms hAh_{A} commute we have that, given Hv:=AvhAH_{v}:=\sum_{A\ni v}h_{A},

ω1it2ωvcit12=eβ1it2Hv(Trv[eβHv])it12.\displaystyle\omega^{\frac{1-it}{2}}\,\omega_{v^{c}}^{\frac{it-1}{2}}=e^{-\beta\frac{1-it}{2}H_{v}}\Big{(}\mathrm{Tr}_{v}\big{[}e^{-\beta H_{v}}\big{]}\Big{)}^{\frac{it-1}{2}}\,. (101)

Now,

ω1it2ωvcit12dit12𝕀\displaystyle\Big{\|}\omega^{\frac{1-it}{2}}\omega_{v^{c}}^{\frac{it-1}{2}}-d^{\frac{it-1}{2}}\mathbb{I}\Big{\|}_{\infty} eβ1it2Hv𝕀(Trv[eβHv])it12\displaystyle\leq\Big{\|}e^{-\beta\frac{1-it}{2}H_{v}}-\mathbb{I}\Big{\|}_{\infty}\,\Big{\|}\Big{(}\mathrm{Tr}_{v}\big{[}e^{-\beta H_{v}}\big{]}\Big{)}^{\frac{it-1}{2}}\Big{\|}_{\infty}
+(Trv[eβHv])it12dit12𝕀\displaystyle\quad+\,\Big{\|}\Big{(}\mathrm{Tr}_{v}\big{[}e^{-\beta H_{v}}\big{]}\Big{)}^{\frac{it-1}{2}}-d^{\frac{it-1}{2}}\mathbb{I}\Big{\|}_{\infty}
(1)β1+t22Hveβ1+t22Hvd12eβ2Hv\displaystyle\overset{(1)}{\leq}\beta\,\frac{\sqrt{1+t^{2}}}{2}\|H_{v}\|_{\infty}\,e^{\beta\frac{\sqrt{1+t^{2}}}{2}\|H_{v}\|_{\infty}}d^{-\frac{1}{2}}\,e^{\frac{\beta}{2}\|H_{v}\|_{\infty}}
+1+t22M1+1+t22dd1Trv[eβHv]𝕀\displaystyle\quad+\frac{\sqrt{1+t^{2}}}{2}\,M^{1+\frac{\sqrt{1+t^{2}}}{2}}\,d\Big{\|}d^{-1}\mathrm{Tr}_{v}\big{[}e^{-\beta H_{v}}\big{]}-\mathbb{I}\Big{\|}_{\infty}
β1+t2Hvd1+1+t22eβHv(2+1+t22)\displaystyle\leq{\beta\,\sqrt{1+t^{2}}\|H_{v}\|_{\infty}}\,d^{1+\frac{\sqrt{1+t^{2}}}{2}}e^{\beta\|H_{v}\|_{\infty}\big{(}2+\frac{\sqrt{1+t^{2}}}{2}\big{)}}
fv(β,t).\displaystyle\equiv f_{v}(\beta,t)\,. (102)

Inequality (1) above follows from the operator convexity of xx12x\mapsto x^{-\frac{1}{2}} as well as Lemma 6, where M:=max{Trv[eβHv],Trv[eβHv]1,d}deβHvM:=\max\{\|\mathrm{Tr}_{v}\big{[}e^{-\beta H_{v}}\big{]}\|_{\infty},\|\mathrm{Tr}_{v}\big{[}e^{-\beta H_{v}}\big{]}^{-1}\|_{\infty},d\}\leq d\,e^{\beta\|H_{v}\|_{\infty}}. Moreover,

e2βHvd1𝕀ωve2βHvd1𝕀ωvd1𝕀12βHve2βHv.\displaystyle e^{-2\beta\|H_{v}\|_{\infty}}d^{-1}\mathbb{I}\leq\omega_{v}\leq e^{2\beta\|H_{v}\|_{\infty}}d^{-1}\mathbb{I}\quad\Rightarrow\quad\|\omega_{v}-d^{-1}\mathbb{I}\|_{1}\leq 2\beta\|H_{v}\|_{\infty}\,e^{2\beta\|H_{v}\|_{\infty}}\,. (103)

Therefore,

ω1it2ωvcit12(𝕀vTrv[])ωvcit+12ω1+it2ωvTrv[]\displaystyle\left\|\omega^{\frac{1-it}{2}}\,\omega_{v^{c}}^{\frac{it-1}{2}}\left(\mathbb{I}_{v}\otimes\mathrm{Tr}_{v}\left[\cdot\right]\right)\omega_{v^{c}}^{-\frac{it+1}{2}}\,\omega^{\frac{1+it}{2}}-\omega_{v}\otimes\mathrm{Tr}_{v}\left[\cdot\right]\right\|_{\diamond}
d12(eβHv+1)fv(β,t)+d1𝕀ωv\displaystyle\qquad\qquad\leq d^{\frac{1}{2}}\big{(}e^{\beta\|H_{v}\|_{\infty}}+1\big{)}\,f_{v}(\beta,t)+\|d^{-1}\mathbb{I}-\omega_{v}\|_{\infty}
d12(eβHv+1)fv(β,t)+2βHve2βHv,\displaystyle\qquad\qquad\leq d^{\frac{1}{2}}\big{(}e^{\beta\|H_{v}\|_{\infty}}+1\big{)}\,f_{v}(\beta,t)+2\beta\|H_{v}\|_{\infty}\,e^{2\beta\|H_{v}\|_{\infty}}\,, (104)

and the integrand in (100) tends to zero pointwise for β0\beta\to 0. On the other hand, we have for any tt\in\mathbb{R}

ω1it2ωvcit12(𝕀vTrv[])ωvcit+12ω1+it2ωvTrv[]\displaystyle\left\|\omega^{\frac{1-it}{2}}\,\omega_{v^{c}}^{\frac{it-1}{2}}\left(\mathbb{I}_{v}\otimes\mathrm{Tr}_{v}\left[\cdot\right]\right)\omega_{v^{c}}^{-\frac{it+1}{2}}\,\omega^{\frac{1+it}{2}}-\omega_{v}\otimes\mathrm{Tr}_{v}\left[\cdot\right]\right\|_{\diamond}
ω1it2ωvcit12(𝕀vTrv[])ωvcit+12ω1+it2+ωvTrv[]\displaystyle\leq\left\|\omega^{\frac{1-it}{2}}\,\omega_{v^{c}}^{\frac{it-1}{2}}\left(\mathbb{I}_{v}\otimes\mathrm{Tr}_{v}\left[\cdot\right]\right)\omega_{v^{c}}^{-\frac{it+1}{2}}\,\omega^{\frac{1+it}{2}}\right\|_{\diamond}+\left\|\omega_{v}\otimes\mathrm{Tr}_{v}\left[\cdot\right]\right\|_{\diamond}
2,\displaystyle\leq 2\,, (105)

therefore the integrand in (100) is uniformly bounded. Then, we get for all t+t\in\mathbb{R}_{+} that

ΨvωvTrvd12(eβHv+1)fv(β,t)+2βHve2βHv+2μ0([t,t]c)\displaystyle\left\|\Psi_{v}-\omega_{v}\otimes\mathrm{Tr}_{v}\right\|_{\diamond}\leq d^{\frac{1}{2}}\big{(}e^{\beta\|H_{v}\|_{\infty}}+1\big{)}\,f_{v}(\beta,t)+2\beta\|H_{v}\|_{\infty}\,e^{2\beta\|H_{v}\|_{\infty}}+2\mu_{0}([-t,t]^{c}) (106)

Therefore, for any 0<κ<10<\kappa<1 there exists β(κ)>0\beta(\kappa)>0 such that condition (78) is satisfied for all 0ββ(κ)0\leq\beta\leq\beta(\kappa). More precisely, in view of (106) and (4), it is sufficient that

4β1+t2Cd3+1+t22eβC(3+1+t22)+2μ0([t,t]c)N12N1\displaystyle 4\,{\beta\,\sqrt{1+t^{2}}C}\,d^{\frac{3+\sqrt{1+t^{2}}}{2}}e^{\beta C\big{(}3+\frac{\sqrt{1+t^{2}}}{2}\big{)}}+2\mu_{0}([-t,t]^{c})\leq\frac{N-1}{2N-1} (107)

where C:=supvHvC:=\sup_{v}\|H_{v}\|_{\infty}. Moreover, it is clear that μ0([t,t]c)2eπt\mu_{0}([-t,t]^{c})\leq 2e^{-\pi t}. The result follows after choosing tt so that the exponentially decaying term 4eπt4e^{-\pi t} counts for at most half the upper bound and solving (107) for βc\beta_{c}, up to some numerical simplifications. ∎

Remark 5.

The lower bound (85) can be compared to that in the classical setting [32, Example 17] (see also [34]): there, the author showed that for a Hamiltonian of the form U(S):=xyGS(x)S(y)HxS(x)U(S):=-\sum_{x\sim y\in G}S(x)S(y)-H\sum_{x}S(x), where S(x){1,1}S(x)\in\{-1,1\} denotes the spin configuration at the site xx of a graph GG, i.e. d=2d=2,

βc12ln(N+1N1)N1N,\displaystyle\beta_{c}\geq\frac{1}{2}\,\ln\Big{(}\frac{N+1}{N-1}\Big{)}\sim_{N\to\infty}\frac{1}{N}\,,

which shows asymptotic optimality of our result, up to numerical multiplicative constants. For comparison, the exact value of βc\beta_{c} for the Ising model on the regular infinite tree with degree NN is known to be equal to 12ln(NN2)\frac{1}{2}\ln\big{(}\frac{N}{N-2}\big{)}.

4.1 Auxiliary lemma

Lemma 6.

For any positive, definite matrices A,BA,B and all zz\in\mathbb{C},

AzBz|z|max{A,A1,B,B1}1+|Re(z)|AB,\displaystyle\|A^{z}-B^{z}\|_{\infty}\leq|z|\,\max\{\|A\|_{\infty},\|A^{-1}\|_{\infty},\|B\|_{\infty},\|B^{-1}\|_{\infty}\}^{1+|\operatorname{Re}(z)|}\,\|A-B\|_{\infty}\,, (108)
Proof.

It suffices to use a linear interpolation between AA and BB: A(s):=sA+(1s)BA(s):=sA+(1-s)B. We have

AzBz\displaystyle A^{z}-B^{z} =01ddsA(s)z𝑑s\displaystyle=\int_{0}^{1}\,\frac{d}{ds}A(s)^{z}\,ds
=z[0,1]2A(s)zuddsln(A(s))Az(1u)𝑑s𝑑u\displaystyle=z\,\iint_{[0,1]^{2}}\,A(s)^{zu}\,\frac{d}{ds}\ln(A(s))\,A^{z(1-u)}\,dsdu
=z[0,1]20A(s)zu(A(s)+v)1(AB)(A(s)+v)1A(s)z(1u)𝑑v𝑑u𝑑s.\displaystyle=z\iint_{[0,1]^{2}}\int_{0}^{\infty}\,A(s)^{zu}(A(s)+v)^{-1}\,(A-B)(A(s)+v)^{-1}A(s)^{z(1-u)}\,dvduds\,. (109)

Then,

AzBz\displaystyle\|A^{z}-B^{z}\|_{\infty} |z|010A(s)Re(z)(A(s)+v)12AB𝑑v𝑑s\displaystyle\leq|z|\,\int_{0}^{1}\int_{0}^{\infty}\,\|A(s)^{\operatorname{Re}(z)}\|_{\infty}\,\|(A(s)+v)^{-1}\|_{\infty}^{2}\,\|A-B\|_{\infty}\,dvds
|z|ABM(z)01A(s)1𝑑s\displaystyle\leq|z|\,\|A-B\|_{\infty}\,M(z)\,\int_{0}^{1}\|A(s)^{-1}\|_{\infty}\,ds
|z|M(z)MAB.\displaystyle\leq|z|\,M(z)\cdot M^{\prime}\,\|A-B\|_{\infty}\,. (110)

by the operator convexity of xx1x\mapsto x^{-1} where M(z):=maxs[0,1](sA+(1s)B)Re(z)M(z):=\max_{s\in[0,1]}\|(sA+(1-s)B)^{\operatorname{Re}(z)}\|_{\infty} and M:=max{A1,B1}M^{\prime}:=\max\{\|A^{-1}\|_{\infty},\|B^{-1}\|_{\infty}\}. The result follows by operator convexity of the inverse function and further simple estimates. ∎

5 Modified logarithmic Sobolev inequalities

In this section, we pursue a different approach to prove transportation cost inequalities for W1W_{1}, namely through the existence of a non-commutative entropic inequality known as the modified logarithmic Sobolev inequality [35, 36]. In order to introduce our main result, we need a variation of the Lipschitz constant that was introduced in [13]. This definition departs from a noncommutative differential structure, which we define below (see [37]):

Definition 1 (Differential structure).

A set of operators Lk𝒪VL_{k}\in\mathcal{O}_{V} and constants ωk\omega_{k}\in\mathbb{R} define a differential structure {Lk,ωk}k𝒦\{L_{k},\omega_{k}\}_{k\in\mathcal{K}} for a full rank state ω𝒮V\omega\in\mathcal{S}_{V} if

  • 1

    {Lk}k𝒦={Lk}k𝒦\{L_{k}\}_{k\in\mathcal{K}}=\{L_{k}^{\dagger}\}_{k\in\mathcal{K}};

  • 2

    {Lk}k𝒦\{L_{k}\}_{k\in\mathcal{K}} consists of eigenvectors of the modular operator Δω(X):=ωXω1\Delta_{\omega}(X):=\omega X\omega^{-1} with

    Δω(Lk)=eωkLk.\displaystyle\Delta_{\omega}(L_{k})=e^{-\omega_{k}}L_{k}\,. (111)

Such a differential structure can be used to provide the set of matrices with a Lipschitz constant that is tailored to ω\omega, see e.g. [13, 37] for more on this. In order to distinguish that constant from .L\|.\|_{L}, we refer to it as the differential Lipschitz constant and denote it by |X|{\left|\kern-1.07639pt\left|\kern-1.07639pt\left|\nabla X\right|\kern-1.07639pt\right|\kern-1.07639pt\right|}. It is defined as:

|X|:=(k𝒦(eωk/2+eωk/2)kX2)1/2,\displaystyle{\left|\kern-1.07639pt\left|\kern-1.07639pt\left|\nabla X\right|\kern-1.07639pt\right|\kern-1.07639pt\right|}:=\left(\sum_{k\in\mathcal{K}}(e^{-\omega_{k}/2}+e^{\omega_{k}/2})\|\partial_{k}X\|_{\infty}^{2}\right)^{1/2}\,, (112)

where kX[Lk,X]\partial_{k}X\equiv[L_{k},X]. For ease of notations, we will denote the differential structure by the couple (,ω)(\nabla,\omega). The notion of a differential structure is also intimately connected to that of the generator of a quantum dynamical semigroup converging to ω\omega [37], and properties of that semigroup immediately translate to properties of the metric. This is because the differential structure can be used to define an operator that behaves in an analogous way to the Laplacian on a smoth manifold, which in turn induces a heat semigroup. We refer to [37, 13] for more details on this connection and interpretation.

When the state ω\omega is a quantum Gibbs state corresponding to a local, commuting Hamiltonian associated to a uniformly bounded interaction defined on a lattice VDV\subset\subset\mathbb{Z}^{D}, the differential structure (,ω)(\nabla,\omega) can be chosen as local. This means that the operators LkLi,αL_{k}\equiv L_{i,\alpha} are indexed by a site iVi\in V and an index α\alpha of a set Γ\Gamma whose cardinality only depends on the local dimension dd and the locality κ\kappa of ω\omega. Moreover, we assume that the operators Li,αL_{i,\alpha} are supported on a neighborhood 𝒩i\mathcal{N}_{i} of site ii of diameter rr(κ)r\equiv r(\kappa) and the corresponding constants ωi,α\omega_{i,\alpha} are uniformly bounded: supiDmaxαΓ|ωi,α|Ω<\sup_{i\in\mathbb{Z}^{D}}\max_{\alpha\in\Gamma}|\omega_{i,\alpha}|\equiv\Omega<\infty. The definition in Eq. (112) yields a metric on states by duality:

W(ρ,ω):=supX=X,|X|1|Tr(X(ρω))|.\displaystyle W_{\nabla}(\rho,\omega):=\sup\limits_{X=X^{\dagger},\,{\left|\kern-0.75346pt\left|\kern-0.75346pt\left|\nabla X\right|\kern-0.75346pt\right|\kern-0.75346pt\right|}\leq 1}\left|\operatorname{Tr}\left(X(\rho-\omega)\right)\right|.
Proposition 10.

Given the Gibbs state ω\omega of a local commuting Hamiltonian HH on VDV\subset\subset\mathbb{Z}^{D} with |V|=n|V|=n and associated local differential structure (,ω)(\nabla,\omega), the following bound holds for all ρ𝒮V\rho\in\mathcal{S}_{V}:

ρωW1CnW(ρ,ω),\displaystyle\left\|\rho-\omega\right\|_{W_{1}}\leq C\,\sqrt{n}\,W_{\nabla}(\rho,\omega)\,,

for some constant CC independent of nn.

Proof.

By duality, it is equivalent to prove that for all H𝒪VH\in\mathcal{O}_{V}

|H|CnHL.\displaystyle{\left|\kern-1.07639pt\left|\kern-1.07639pt\left|\nabla H\right|\kern-1.07639pt\right|\kern-1.07639pt\right|}\leq C\,\sqrt{n}\,\|H\|_{L}\,.

First, we have

|H|\displaystyle{\left|\kern-1.07639pt\left|\kern-1.07639pt\left|\nabla H\right|\kern-1.07639pt\right|\kern-1.07639pt\right|} =(iVαΓ(eωi,α/2+eωi,α/2)[Li,α,H]2)12\displaystyle=\Big{(}\sum_{i\in V}\sum_{\alpha\in\Gamma}(e^{-\omega_{i,\alpha}/2}+e^{\omega_{i,\alpha}/2})\,\|[L_{i,\alpha},H]\|_{\infty}^{2}\Big{)}^{\frac{1}{2}} (113)
n|Γ|2eΩ/2maxiVmaxαΓ[Li,α,H].\displaystyle\leq\sqrt{n|\Gamma|}\,\sqrt{2\,e^{\Omega/2}}\,\max_{i\in V}\,\max_{\alpha\in\Gamma}\,\|[L_{i,\alpha},H]\|_{\infty}\,. (114)

Now, since for each pair (i,α)(i,\alpha), Li,αL_{i,\alpha} is supported on a neighborhood 𝒩i\mathcal{N}_{i} of site iVi\in V,

[Li,α,H]\displaystyle\|[L_{i,\alpha},H]\|_{\infty} =[Li,α,H𝕀𝒩id|𝒩i|Tr𝒩i(H)]\displaystyle=\big{\|}[L_{i,\alpha},H-\frac{\mathbb{I}_{\mathcal{N}_{i}}}{d^{|\mathcal{N}_{i}|}}\otimes\mathrm{Tr}_{\mathcal{N}_{i}}(H)]\big{\|}_{\infty} (115)
2Li,αH𝕀𝒩id|𝒩i|Tr𝒩i(H).\displaystyle\leq 2\,\|L_{i,\alpha}\|_{\infty}\,\big{\|}H-\frac{\mathbb{I}_{\mathcal{N}_{i}}}{d^{|\mathcal{N}_{i}|}}\otimes\mathrm{Tr}_{\mathcal{N}_{i}}(H)\big{\|}_{\infty}\,. (116)

Next, by a telescopic sum argument, we can further control the last infinity norm on the right hand side above as follows: given an arbitrary ordering of the region 𝒩i\mathcal{N}_{i},

H𝕀𝒩id|𝒩i|Tr𝒩i(H)\displaystyle\big{\|}H-\frac{\mathbb{I}_{\mathcal{N}_{i}}}{d^{|\mathcal{N}_{i}|}}\otimes\mathrm{Tr}_{\mathcal{N}_{i}}(H)\big{\|}_{\infty} (1)j=1|𝒩i|(𝕀1j1dj1Tr1j1𝕀1jdjTr1j)(H)\displaystyle\overset{(1)}{\leq}\sum_{j=1}^{|\mathcal{N}_{i}|}\Big{\|}\Big{(}\frac{\mathbb{I}_{1\ldots j-1}}{d^{j-1}}\otimes\mathrm{Tr}_{1\ldots j-1}-\frac{\mathbb{I}_{1\ldots j}}{d^{j}}\otimes\mathrm{Tr}_{1\ldots j}\Big{)}(H)\Big{\|}_{\infty} (117)
(2)|𝒩i|maxj𝒩iH𝕀jdTrj(H),\displaystyle\overset{(2)}{\leq}|\mathcal{N}_{i}|\max_{j\in\mathcal{N}_{i}}\,\|H-\frac{\mathbb{I}_{j}}{d}\otimes\mathrm{Tr}_{j}(H)\|_{\infty}\,, (118)

where (1)(1) follows from the triangle inequality whereas (2)(2) follows from the fact that the maps 𝕀1j1dj1Tr1j1\frac{\mathbb{I}_{1\ldots j-1}}{d^{j-1}}\otimes\mathrm{Tr}_{1\ldots j-1} are completely positive and unital, and therefore contract the operator norm. All in all, we have derived the following bound on the differential Lipschitz constant of HH:

|H|\displaystyle{\left|\kern-1.07639pt\left|\kern-1.07639pt\left|\nabla{H}\right|\kern-1.07639pt\right|\kern-1.07639pt\right|} n|Γ| 22eΩ/2maxiVmaxαΓLi,α|𝒩i|maxj𝒩iH𝕀jdTrj(H)\displaystyle\leq\sqrt{n|\Gamma|}\,2\sqrt{2\,e^{\Omega/2}}\,\max_{i\in V}\,\max_{\alpha\in\Gamma}\,\|L_{i,\alpha}\|_{\infty}\,|\mathcal{N}_{i}|\,\max_{j\in\mathcal{N}_{i}}\,\|H-\frac{\mathbb{I}_{j}}{d}\otimes\mathrm{Tr}_{j}(H)\|_{\infty} (119)
(3)d21d2n|Γ| 22eΩ/2maxiVmaxαΓLi,α|𝒩i|HL\displaystyle\overset{(3)}{\leq}\frac{d^{2}-1}{d^{2}}\sqrt{n|\Gamma|}\,2\sqrt{2\,e^{\Omega/2}}\,\max_{i\in V}\,\max_{\alpha\in\Gamma}\,\|L_{i,\alpha}\|_{\infty}\,|\mathcal{N}_{i}|\,\|H\|_{L} (120)
CnmaxiVH𝕀jdTrj(H),\displaystyle\equiv C\,\sqrt{n}\,\max_{i\in V}\|H-\frac{\mathbb{I}_{j}}{d}\otimes\mathrm{Tr}_{j}(H)\|_{\infty}\,, (121)

for some constant CC independent of nn, and where (3)(3) follows from Proposition 6. ∎

The advantage of W1W_{1} as compared to WW_{\nabla} is that it does not depend on the state ω\omega. On the other hand, the bound derived in Proposition 10 can be used in conjunction with recently proved transportation cost inequalities for WW_{\nabla} through the proof of the existence of a modified logarithmic Sobolev inequality in order to get analogous inequalities for W1W_{1} (see [38] for more details):

Theorem 5.

Let ω\omega be the Gibbs state of a local commuting Hamiltonian HH at inverse temperature β\beta on VDV\subset\subset\mathbb{Z}^{D}. Then, there exists a critical inverse temperature βc\beta_{c} such that C(ωβ)CnC(\omega_{\beta})\leq C\,n for some constant CC independent of n=|V|n=|V| whenever β<βc\beta<\beta_{c} if any of the two conditions below is satisfied:

  • (i)\operatorname{(i)}

    HH is classical;

  • (ii)\operatorname{(ii)}

    HH is a nearest neighbour Hamiltonian.

Moreover, we can drop the assumption of 22-locality in the 11D case, where βc=0\beta_{c}=0 at the cost of getting a slightly worsened constant C(ωβ)Cnpolylog(n)C(\omega_{\beta})\leq Cn\operatorname{polylog}(n), so that we recover the result of Theorem 2.

Proof.

In [38, 19, 20] the existence of local differential structures associated to ω\omega that satisfy the so-called modified logarithmic Sobolev inequality was proved under the conditions of the theorem. Moreover, the modified logarithmic Sobolev inequality implies the transportation cost inequality for the differential Wasserstein distance [13]: there exists a constant CC^{\prime} independent of nn such that

W(ρ,ω)CS(ρω)\displaystyle W_{\nabla}(\rho,\omega)\leq\sqrt{C^{\prime}\,S(\rho\|\omega)} (122)

for all state ρ𝒮V\rho\in\mathcal{S}_{V}. This fact in conjunction with Proposition 10 allows us to conclude. ∎

6 Local indistinguishability

In this section, we provide a transportation cost inequality under a condition of local indistinguishability [39, 40, 41]. In the classical setting, this condition constitues a weakening of Dobrushin Shlosman’s mixing condition [8] recently considered by Marton [7]. Moreover, as opposed to the latter, our technique has the benefit of not requiring the local specifications of the state to be uniformly lower bounded by a positive number, at the cost of getting a slightly worsened constant.

6.1 Transportation cost from local indistinguishability

We start by proving our general result in the quantum setting. Here, we assume that the n=(2m+1)Dn=(2m+1)^{D} qudits are arranged on a DD-dimensional regular lattice V:=[m,m]DV:=[-m,m]^{D}. Before we state our main result, we need to introduce the notion of a non-commutative conditional expectation.

Definition 2 (Conditional expectations).

Let 𝒩V\mathcal{N}\subseteq\mathcal{B}_{V} be a von Neumann subalgebra111We recall that a finite dimensional von Neumann algebra is a matrix algebra that is close under taking the adjoint. of V\mathcal{B}_{V}. A conditional expectation onto 𝒩\mathcal{N} is a completely positive unital map E𝒩:V𝒩E_{\mathcal{N}}^{\dagger}:\mathcal{B}_{V}\to\mathcal{N} satisfying

  1. i)

    for all X𝒩X\in\mathcal{N}, E𝒩(X)=XE^{\dagger}_{\mathcal{N}}(X)=X;

  2. ii)

    for all a,b𝒩,XVa,b\in\mathcal{N},X\in\mathcal{B}_{V}, E𝒩(aXb)=aE𝒩(X)bE^{\dagger}_{\mathcal{N}}(aXb)=aE_{\mathcal{N}}(X)b.

We denote by E𝒩E_{\mathcal{N}} its adjoint map with respect to the trace inner product, i.e.

Tr(E𝒩(X)Y)=Tr(XE𝒩(Y)).\displaystyle\mathrm{Tr}(E_{\mathcal{N}}(X)Y)=\mathrm{Tr}(XE_{\mathcal{N}}^{\dagger}(Y))\,.

As a simple example, we consider a full-rank state σ𝒮V\sigma\in\mathcal{S}_{V} and let (et)t0(e^{t\mathcal{L}})_{t\geq 0} be a quantum Markov semigroup. Under the following detailed balance condition, the limit limtet=E𝒩\lim_{t\to\infty}e^{t\mathcal{L}^{\dagger}}=E^{\dagger}_{\mathcal{N}} is a conditional expectation onto the algebra 𝒩\mathcal{N} of fixed points of the semigroup:

X,YV,Tr(σX(Y))=Tr(σ(X)Y).\displaystyle\forall X,Y\in\mathcal{B}_{V},\quad\mathrm{Tr}\big{(}\sigma\,X^{\dagger}\mathcal{L}^{\dagger}(Y)\big{)}=\mathrm{Tr}\big{(}\sigma\,\mathcal{L}^{\dagger}(X)^{\dagger}Y\big{)}\,.

Next, for a state ρ\rho, the relative entropy with respect to 𝒩\mathcal{N} is defined as follows

S(ρ𝒩):=S(ρE𝒩(ρ))=infE𝒩(σ)=σS(ρσ),S(\rho\|\mathcal{N}):=S(\rho\|E_{\mathcal{N}}(\rho))=\inf_{E_{\mathcal{N}}(\sigma)=\sigma}S(\rho\|\sigma)\,,

where the infimum is always attained by σ=E𝒩(ρ)\sigma=E_{\mathcal{N}}(\rho). Indeed, for any σ\sigma satisfying E𝒩(σ)=σE_{\mathcal{N}}(\sigma)=\sigma, we have the following chain rule (see [42, Lemma 3.4])

S(ρσ)=S(ρE𝒩(ρ))+S(E𝒩(ρ)σ).\displaystyle S(\rho\|\sigma)=S(\rho\|E_{\mathcal{N}}(\rho))+S(E_{\mathcal{N}}(\rho)\|\sigma)\,. (123)

Hence the infimum is attained if and only if S(E𝒩(ρ)σ)=0S(E_{\mathcal{N}}(\rho)\|\sigma)=0.

Definition 3 (Local indistinguishability).

Let {𝒩C}CV\{\mathcal{N}_{C}\}_{C\subseteq V} be a set of subalgebras of V\mathcal{B}_{V} such that Cc𝒩C\mathcal{B}_{{C}^{c}}\subset\mathcal{N}_{C} and 𝐄:={EC}CV\mathbf{E}:=\{E_{C}\}_{C\subseteq V} be a set of compatible conditional expectations EC:V𝒩CE^{\dagger}_{C}:\mathcal{B}_{V}\to\mathcal{N}_{C} acting non-trivially on region CC, i.e., they satisfy the property that for any CCC\subseteq C^{\prime}, ECEC=ECEC=ECE_{C}\circ E_{C^{\prime}}=E_{C^{\prime}}\circ E_{C}=E_{C^{\prime}}. Then, we say that 𝐄\mathbf{E} satisfies local indistinguishability if there exists a fast decaying function φ:\varphi:\mathbb{N}\to\mathbb{R} independent of VV such that for every regions XYZVXYZ\subset V with dist(X,Z)\operatorname{dist}(X,Z)\geq\ell, and for all states ρ𝒮V\rho\in\mathcal{S}_{V},

EYZ(EXYZEXY)(ρ)1|XYZ|φ(),\displaystyle\|E_{YZ}\circ(E_{XYZ}-E_{XY})(\rho)\|_{1}\leq\,|XYZ|\,\varphi(\ell)\,,

For instance, take a product state ω𝒮V\omega\in\mathcal{S}_{V} and for each region CVC\subseteq V, denote EC(ρ)=TrC(ρ)ωCE_{C}(\rho)=\mathrm{Tr}_{C}(\rho)\otimes\omega_{C}. One can easily verify that the maps ECE_{C} are conditional expectations and satisfy the local indistinguishability condition with φ=0\varphi=0. We are now ready to state and prove the main theorem of this section. For a strictly decreasing function φ:+\varphi:\mathbb{N}\to\mathbb{R}_{+} and a positive real number a>0a>0, we denote by φ1(a):=min{:φ()a}\varphi^{-1}(a):=\min\{\ell\in\mathbb{N}:\varphi(\ell)\leq a\}.

Theorem 6.

Let 𝐄\mathbf{E} be a set of compatible conditional expectations satisfying local indistinguishability with fast decaying function φ\varphi. Then for all hypercubes V0VV_{0}\subset V and all ρ𝒮V\rho\in\mathcal{S}_{V},

ρEV0(ρ)W1220(7φ1(|V0|3/2))D|V0|S(ρEV0(ρ)),\displaystyle\left\|\rho-E_{V_{0}}(\rho)\right\|_{W_{1}}\leq 2\sqrt{20}\,(7\varphi^{-1}(|V_{0}|^{-3/2}))^{D}\sqrt{|V_{0}|\,S(\rho\|E_{V_{0}}(\rho))}\,, (124)

for some fixed constant cc of order 11. In particular, whenever EV(ρ)=ωV𝒮VE_{V}(\rho)=\omega_{V}\in\mathcal{S}_{V} for all states ρ\rho, and assuming the exponential clustering function φ():=κe/ξ\varphi(\ell):=\kappa e^{-\ell/\xi}, the state ωV\omega_{V} satisfies TC(c)\operatorname{TC}(c) with c=O(npolylog(n))c=O({n}\operatorname{polylog}(n)).

Refer to caption
Figure 1: Geometry of the lattice in the proof of Theorem 6 with tiling by regions A+:=iA+,iA_{+}:=\cup_{i}A_{+,i}, B+:=iB+,iB_{+}:=\cup_{i}B_{+,i} and C:=CiC:=\cup C_{i}.
Proof.

For sake of clarity, we provide the proof for D=2D=2 only, although the general case follows similarly. First, we partition the hypercube V0V_{0} into regions AA, B+B_{+} and C+C_{+} in the same way as done in [41] (see also Figure 1). Then, by triangle inequality

ρEV0(ρ)W1\displaystyle\|\rho-E_{V_{0}}(\rho)\|_{W_{1}}
ρEA+EBC(ρ)W1+EA+EBC(ρ)EV0(ρ)W1\displaystyle\leavevmode\nobreak\ \leavevmode\nobreak\ \leq\|\rho-E_{A_{+}}E_{B_{-}C}(\rho)\|_{W_{1}}+\|E_{A_{+}}E_{B_{-}C}(\rho)-E_{V_{0}}(\rho)\|_{W_{1}}\,
(idEA+EB+EC)(ρ)W1+EA+(EB+ECEBC)(ρ)W1+(EA+EBCEV0)(ρ)W1\displaystyle\leavevmode\nobreak\ \leavevmode\nobreak\ \leq\|(\mathrm{id}-E_{A+}E_{B_{+}}E_{C})(\rho)\|_{W_{1}}+\|E_{A+}(E_{B_{+}}E_{C}-E_{B_{-}C})(\rho)\|_{W_{1}}+\|(E_{A_{+}}E_{B_{-}C}-E_{V_{0}})(\rho)\|_{W_{1}}\,
(I)+2|A+max|(II)+(III),\displaystyle\leavevmode\nobreak\ \leavevmode\nobreak\ \leq(I)+2|A_{+}^{\max}|\,(II)+(III)\,, (125)

where

(I):=(idEA+EB+EC)(ρ)W1\displaystyle(I):=\|(\mathrm{id}-E_{A+}E_{B_{+}}E_{C})(\rho)\|_{W_{1}}
(II):=(EB+ECEBC)(ρ)W1\displaystyle(II):=\|(E_{B_{+}}E_{C}-E_{B_{-}C})(\rho)\|_{W_{1}}
(III):=(EA+EBCEV0)(ρ)W1\displaystyle(III):=\|(E_{A_{+}}E_{B_{-}C}-E_{V_{0}})(\rho)\|_{W_{1}}

and where the last bound in (125) follows from Proposition 5 with |A+max|:=maxi|A+,i||A_{+}^{\max}|:=\max_{i}|A_{+,i}|. Now, we control each of the norms on the right-hand side of (125) separately. First, we denote by EC(0):=idE_{C^{(0)}}:=\mathrm{id}, C(i):=jiCjC^{(i)}:=\cup_{j\leq i}C_{j} given an arbitrary ordering of the connected subregions in CC, and similarly for the other regions A+A_{+} and B+B_{+}. Then,

(I)\displaystyle(I) (idEC)(ρ)W1+(ECEB+EC)(ρ)W1+(EB+ECEA+EB+EC)(ρ)W1\displaystyle\leq\|(\mathrm{id}-E_{C})(\rho)\|_{W_{1}}+\|(E_{C}-E_{B_{+}}E_{C})(\rho)\|_{W_{1}}+\|(E_{B_{+}}E_{C}-E_{A_{+}}E_{B+}E_{C})(\rho)\|_{W_{1}} (126)
|Cmax|iC(EC(i1)EC(i))(ρ)1+|B+max|iB+(EB+(i1)EB+(i))(EC(ρ))1\displaystyle\leq|C^{\max}|\,\sum_{i\in\mathcal{I}_{C}}\|(E_{C^{(i-1)}}-E_{C^{(i)}})(\rho)\|_{1}+|B_{+}^{\max}|\sum_{i\in\mathcal{I}_{B+}}\|(E_{B_{+}^{(i-1)}}-E_{B_{+}^{(i)}})(E_{C}(\rho))\|_{1}
+|A+max|iA+(EA+(i1)EA+(i))EB+EC(ρ)1\displaystyle\leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ +|A_{+}^{\max}|\sum_{i\in\mathcal{I}_{A+}}\|(E_{A_{+}^{(i-1)}}-E_{A_{+}^{(i)}})E_{B_{+}}E_{C}(\rho)\|_{1} (127)
|V0max|{2|C|iCS(EC(i1)(ρ)EC(i)(ρ))\displaystyle\leq|V_{0}^{\max}|\,\Big{\{}\,\sqrt{2|\mathcal{I}_{C}|}\,\sqrt{\sum_{i\in\mathcal{I}_{C}}\,S(E_{C^{(i-1)}}(\rho)\|E_{C^{(i)}}(\rho))}
+2|B+|iB+S(EB+(i1)(EC(ρ))EB+(i)(EC(ρ)))\displaystyle\leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ +\,\sqrt{2|\mathcal{I}_{B_{+}}|}\,\sqrt{\sum_{i\in\mathcal{I}_{B+}}\,S(E_{B_{+}^{(i-1)}}(E_{C}(\rho))\|E_{B_{+}^{(i)}}(E_{C}(\rho)))}
+2|A+|iA+S(EA+(i1)EB+EC(ρ)EA+(i)EB+EC(ρ))}\displaystyle\leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ +\sqrt{2|\mathcal{I}_{A_{+}}|}\,\sqrt{\sum_{i\in\mathcal{I}_{A_{+}}}\,S(E_{A_{+}^{(i-1)}}E_{B_{+}}E_{C}(\rho)\|E_{A_{+}^{(i)}}E_{B_{+}}E_{C}(\rho))}\Big{\}} (128)
|V0max|2|max|{S(ρEC(ρ))+S(EC(ρ)EB+EC(ρ))\displaystyle\leq|V_{0}^{\max}|\,\sqrt{2|\mathcal{I}_{\max}|}\,\Big{\{}\sqrt{S(\rho\|E_{C}(\rho))}+\sqrt{S(E_{C}(\rho)\|E_{B_{+}}E_{C}(\rho))}
+S(EB+EC(ρ)EA+EB+EC(ρ))}\displaystyle\leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ +\sqrt{S(E_{B_{+}}E_{C}(\rho)\|E_{A_{+}}E_{B_{+}}E_{C}(\rho))}\Big{\}} (129)

with

|Cmax|:=maxi|Ci|\displaystyle|C^{\max}|:=\max_{i}|C_{i}|
|V0max|:=max{|A+max|,|B+max|,|Cmax|}\displaystyle|V_{0}^{\max}|:=\max\{|A_{+}^{\max}|,|B_{+}^{\max}|,|C^{\max}|\}
|max|:=max{|A+|,|B+|,|C|}\displaystyle|\mathcal{I}_{\max}|:=\max\{|\mathcal{I}_{A_{+}}|,|\mathcal{I}_{B_{+}}|,|\mathcal{I}_{C}|\}

where |C||\mathcal{I}_{C}| denotes the number of connected components in CC, and similarly for the other sets. Above, (126) follows by the triangle inequality, (127) by the triangle inequality and Proposition 3, and (128) by Pinsker’s inequality as well as Jensen’s inequality for xx12x\mapsto x^{\frac{1}{2}}. (129) follows from the chain rule in (123), and the fact that the regions CiC_{i}, resp. A+,iA_{+,i}, resp. B+,iB_{+,i}, do not overlap, so that for instance EA+,iEA+,j=EA+,jEA+,i=EA+,jA+,iE_{A_{+,i}}E_{A_{+,j}}=E_{A_{+,j}}E_{A_{+,i}}=E_{A_{+,j}\cup A_{+,i}}. Next, we control the second norm on the right-hand side of (125): using Proposition 3 , we have

(II)\displaystyle(II) 2|B+|(EB+ECEBC)(ρ)1\displaystyle\leq 2|B_{+}|\,\|(E_{B_{+}}E_{C}-E_{B_{-}C})(\rho)\|_{1}
=2|B+|(EB+ECEBC)(ECEBC)(ρ)1\displaystyle=2|B_{+}|\,\|(E_{B+}E_{C}-E_{B_{-}C})(E_{C}-E_{B_{-}C})(\rho)\|_{1} (130)
2|B+|EB+ECEBC11(ECEBC)(ρ)1\displaystyle\leq 2|B_{+}|\,\|E_{B+}E_{C}-E_{B_{-}C}\|_{1\to 1}\,\|(E_{C}-E_{B_{-}C})(\rho)\|_{1}
2|B+|maxρ𝒮VEB(ECEBC)(ρ)1EC(ρ)EBC(ρ)1\displaystyle\leq 2|B_{+}|\,\max_{\rho^{\prime}\in\mathcal{S}_{V}}\,\|E_{B}(E_{C}-E_{B_{-}C})(\rho^{\prime})\|_{1}\,\|E_{C}(\rho)-E_{B_{-}C}(\rho)\|_{1} (131)
2|B+||BC|φ()EC(ρ)EBC(ρ)1\displaystyle\leq 2\,|B_{+}|\,|B_{-}C|\,\varphi(\ell)\,\|E_{C}(\rho)-E_{B_{-}C}(\rho)\|_{1} (132)
2|B+||BC|φ()2S(EC(ρ)EBC(ρ)).\displaystyle\leq 2\,|B_{+}|\,|B_{-}C|\,\varphi(\ell)\,\sqrt{2\,S(E_{C}(\rho)\|E_{B_{-}C}(\rho))}\,. (133)

Above, (130) follows form the compatibility of the conditional expectations, and (131) from EB+EB=EB+E_{B_{+}}\circ E_{B}=E_{B_{+}} and the monotonicity of the trace-distance under such CPTP map. (132) follows from the condition of local indistinguishability when taking X=C\BX=C\backslash B, Y=B\BY=B\backslash B_{-} and Z=BZ=B_{-}, and assuming that dist(X,Z)\operatorname{dist}(X,Z)\geq\ell. Finally, (133) follows from an application of Pinsker’s inequality. Similarly, we find

(III)2|A+||V0|φ()2S(EBC(ρ)EV0(ρ))(III)\leq 2\,|A_{+}|\,|V_{0}|\,\varphi(\ell)\,\sqrt{2\,S(E_{B_{-}C}(\rho)\|E_{V_{0}}(\rho))} (134)

Then, by inserting (129), (133) and (134) into (125), we have

ρEV0(ρ)W1\displaystyle\|\rho-E_{V_{0}}(\rho)\|_{W_{1}}
22max{|V0max||max|,|A+||V0|φ(),|B+||BC|φ()}\displaystyle\leq 2\sqrt{2}\max\big{\{}|V_{0}^{\max}|\,\sqrt{|\mathcal{I}_{\max}|},|A_{+}|\,|V_{0}|\,\varphi(\ell),|B_{+}|\,|B_{-}C|\varphi(\ell)\big{\}}
{S(ρEC(ρ))+S(EC(ρ)EB+EC(ρ))+S(EB+EC(ρ)EA+EB+EC(ρ))\displaystyle\,\Big{\{}\sqrt{S(\rho\|E_{C}(\rho))}+\sqrt{S(E_{C}(\rho)\|E_{B_{+}}E_{C}(\rho))}+\sqrt{S(E_{B_{+}}E_{C}(\rho)\|E_{A_{+}}E_{B_{+}}E_{C}(\rho))}
+S(EBC(ρ)EV0(ρ))+S(EC(ρ)EBC(ρ))}\displaystyle\leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ +\sqrt{S(E_{B_{-}C}(\rho)\|E_{V_{0}}(\rho))}+\sqrt{S(E_{C}(\rho)\|E_{B_{-}C}(\rho))}\Big{\}}
210max{|V0max||max|,|A+||V0|φ(),|B+||BC|φ()}\displaystyle\leq 2\sqrt{10}\max\big{\{}|V_{0}^{\max}|\,\sqrt{|\mathcal{I}_{\max}|},|A_{+}|\,|V_{0}|\,\varphi(\ell),|B_{+}|\,|B_{-}C|\varphi(\ell)\big{\}}
(S(ρEC(ρ))+S(EC(ρ)EB+EC(ρ))+S(EB+EC(ρ)EA+EB+EC(ρ))\displaystyle\Big{(}S(\rho\|E_{C}(\rho))+S(E_{C}(\rho)\|E_{B_{+}}E_{C}(\rho))+S(E_{B_{+}}E_{C}(\rho)\|E_{A_{+}}E_{B_{+}}E_{C}(\rho))
+S(EC(ρ)EBC(ρ))+S(EBC(ρ)EV0(ρ)))12\displaystyle\leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ +S(E_{C}(\rho)\|E_{B_{-}C}(\rho))+S(E_{B_{-}C}(\rho)\|E_{V_{0}}(\rho))\Big{)}^{\frac{1}{2}} (135)
220max{|V0max||max|,|A+||V0|φ(),|B+||BC|φ()}S(ρEV0(ρ))12\displaystyle\leq 2\sqrt{20}\max\big{\{}|V_{0}^{\max}|\,\sqrt{|\mathcal{I}_{\max}|},|A_{+}|\,|V_{0}|\,\varphi(\ell),|B_{+}|\,|B_{-}C|\varphi(\ell)\big{\}}S(\rho\|E_{V_{0}}(\rho))^{\frac{1}{2}} (136)

where (135) is another directly application of Jensen’s inequality for xx12x\mapsto x^{\frac{1}{2}}, whereas (136) follows from two uses of the chain rule (123) after adding the positive term S(EA+EB+EC(ρ)EV0(ρ))S(E_{A_{+}}E_{B_{+}}E_{C}(\rho)\|E_{V_{0}}(\rho)) to the square root and a final use of the data processing inequality. The result then follows after choosing the length :=φ1(|V0|3/2)\ell:=\varphi^{-1}(|V_{0}|^{-3/2}) so that

max{|A+||V0|,|B+||BC|}φ()|V0|2φ()|V0max||max|.\displaystyle\max\big{\{}|A_{+}|\,|V_{0}|,\,|B_{+}|\,|B_{-}C|\big{\}}\,\varphi(\ell)\leq|V_{0}|^{2}\varphi(\ell)\leq|V_{0}^{\max}|\,\sqrt{|\mathcal{I}_{\max}|}\,.

With this choice, and estimating |V0max|(7)D|V_{0}^{\max}|\leq(7\ell)^{D} the bound found in (136) can be further controlled by

ρEV0(ρ)W1\displaystyle\|\rho-E_{V_{0}}(\rho)\|_{W_{1}} 220(7)D|V0|S(ρEV0(ρ))\displaystyle\leq 2\sqrt{20}\,(7\ell)^{D}\,\sqrt{|V_{0}|\,S(\rho\|E_{V_{0}}(\rho))} (137)
220(7φ1(|V0|3/2))D|V0|S(ρEV0(ρ)).\displaystyle\leq 2\sqrt{20}\,(7\varphi^{-1}(|V_{0}|^{-3/2}))^{D}\sqrt{|V_{0}|\,S(\rho\|E_{V_{0}}(\rho))}\,. (138)

6.2 Classical case

In this section, we restrict our analysis to classical conditional expectations and probability measures. In this setting, it is easy to see that the property of local indistinguishability is implied by the following condition. Here, with a slight abuse of notations, we will use the same symbol for a probability measure μ\mu on the Borel sets of [d]V[d]^{V} and its corresponding probability mass function.

Definition 4 (Local indistinguishability, classical case).

Let μ\mu be a probability measure on [d]V[d]^{V}, and {μC}CV\{\mu_{C}\}_{C\subseteq V} be a set of compatible conditional probability measures μC(.|xCc)\mu_{C}(.|x_{{C}^{c}}) acting on the sets [d]C[d]^{C}, i.e. they satisfy the property that for any CCC\subseteq C^{\prime}, 𝔼μC𝔼μC=𝔼μC𝔼μC=𝔼μC\mathbb{E}_{\mu_{C}}\circ\mathbb{E}_{\mu_{C^{\prime}}}=\mathbb{E}_{\mu_{C^{\prime}}}\circ\mathbb{E}_{\mu_{C}}=\mathbb{E}_{\mu_{C^{\prime}}}. Then, we say that the measure μ\mu satisfies local indistinguishability if there exists a fast decaying function φ:+\varphi:\mathbb{N}\to\mathbb{R}_{+} such that for every regions V=XYZVV^{\prime}=XYZ\subset V such that dist(i,j)\operatorname{dist}(i,j)\geq\ell for any iXi\in X and jZj\in Z,

maxxX[d]XmaxxVc[d]VcyV|μY|XVc(yY|xXxVc)μXY|ZVc(yXY|yZxVc)μV|Vc(yV|xVc)||V|φ(),\displaystyle\max_{x_{X}\in[d]^{X}}\,\max_{x_{{V^{\prime}}^{c}}\in[d]^{{V^{\prime}}^{c}}}\,\,\sum_{y_{V^{\prime}}}\big{|}\mu_{Y|X{V^{\prime}}^{c}}(y_{Y}|x_{X}x_{{V^{\prime}}^{c}})\mu_{XY|ZV^{\prime c}}(y_{XY}|y_{Z}x_{V^{\prime c}})-\mu_{V^{\prime}|{V^{\prime}}^{c}}(y_{V^{\prime}}|x_{{V^{\prime}}^{c}})\big{|}\leq|V^{\prime}|\varphi(\ell)\,,

where Z\partial Z denotes the boundary of ZZ.

Corollary 1.

Let μ\mu be a probability measure on [d]V[d]^{V} satisfying local indistinguishability with fast decaying function φ\varphi. Then for all ν<<μ\nu<\!<\mu,

W1(ν,μ)220(7φ1(n3/2))DnS(νμ).\displaystyle W_{1}(\nu,\mu)\leq 2\sqrt{20}\,(7\varphi^{-1}(n^{-3/2}))^{D}\sqrt{n\,S(\nu\|\mu)}\,. (139)

Equivalently, the measure μ\mu satisfies the following sub-Gaussian tail: for any function ff such that fL1\|f\|_{L}\leq 1,

μ(|f(X)𝔼μ[f(X)]|>t)2exp(t280n(7φ1(n3/2))2D).\displaystyle\mathbb{P}_{\mu}\Big{(}|f(X)-\mathbb{E}_{\mu}[f(X)]|>t\Big{)}\leq 2\,\exp\left(-\frac{t^{2}}{80n(7\varphi^{-1}(n^{-3/2}))^{2D}}\right)\,. (140)

7 Gaussian concentration

As mentioned before, the classical transportation cost inequalities for a measure μ\mu is equivalent to the sub-Gaussian bounds on the tail probability of any Lipschitz function ff of a random variable XX drawn according to μ\mu. One way to see this is by using the variational formulation of the relative entropy in order to bound the Laplace transform of f(X)f(X). In the non-commutative setting, this leads to the following characterization of the transportation cost constant C(ω)C(\omega):

Proposition 11.

For any ω𝒮V\omega\in\mathcal{S}_{V},

C(ω)=4supK𝒪VlnTrexp(K+lnω)Tr[ωK]KL2,C(\omega)=4\sup_{K\in\mathcal{O}_{V}}\frac{\ln\mathrm{Tr}\exp\left(K+\ln\omega\right)-\mathrm{Tr}\left[\omega\,K\right]}{\left\|K\right\|_{L}^{2}}\,, (141)

and the sup\sup can be restricted to K𝒪VK\in\mathcal{O}_{V} such that Tr[ωK]=0\mathrm{Tr}\left[\omega\,K\right]=0.

Proof.

Let C~(ω)\tilde{C}(\omega) be the right-hand side of (141). On the one hand, let K𝒪VK\in\mathcal{O}_{V} satisfy Tr[ωK]=0\mathrm{Tr}\left[\omega\,K\right]=0, and let

ρ=exp(K+lnω)Trexp(K+lnω)𝒮V.\rho=\frac{\exp\left(K+\ln\omega\right)}{\mathrm{Tr}\exp\left(K+\ln\omega\right)}\in\mathcal{S}_{V}\,. (142)

We have

lnTrexp(K+lnω)=Tr[ρK]S(ρω)ρωW1KLρωW12C(ω)C(ω)KL24,\ln\mathrm{Tr}\exp\left(K+\ln\omega\right)=\mathrm{Tr}\left[\rho\,K\right]-S(\rho\|\omega)\leq\left\|\rho-\omega\right\|_{W_{1}}\left\|K\right\|_{L}-\frac{\left\|\rho-\omega\right\|_{W_{1}}^{2}}{C(\omega)}\leq\frac{C(\omega)\left\|K\right\|_{L}^{2}}{4}\,, (143)

therefore C~(ω)C(ω)\tilde{C}(\omega)\leq C(\omega).

On the other hand, let ρ𝒮V\rho\in\mathcal{S}_{V}, and let K𝒪VK\in\mathcal{O}_{V} such that

KL=2ρωW1C~(ω),Tr[ωK]=0,Tr[ρK]=2ρωW12C~(ω).\left\|K\right\|_{L}=\frac{2\left\|\rho-\omega\right\|_{W_{1}}}{\tilde{C}(\omega)}\,,\qquad\mathrm{Tr}\left[\omega\,K\right]=0\,,\qquad\mathrm{Tr}\left[\rho\,K\right]=\frac{2\left\|\rho-\omega\right\|_{W_{1}}^{2}}{\tilde{C}(\omega)}\,. (144)

We have

S(ρω)Tr[ρK]lnTrexp(K+lnω)ρωW12C~(ω),S(\rho\|\omega)\geq\mathrm{Tr}\left[\rho\,K\right]-\ln\mathrm{Tr}\exp\left(K+\ln\omega\right)\geq\frac{\left\|\rho-\omega\right\|_{W_{1}}^{2}}{\tilde{C}(\omega)}\,, (145)

where the last inequality follows from the definition of C~(ω)\tilde{C}(\omega), therefore C(ω)C~(ω)C(\omega)\leq\tilde{C}(\omega), and the claim follows. ∎

In the tracial setting [10], and more generally whenever [K,ω]=0[K,\omega]=0 the quantity Trexp(K+lnω)\mathrm{Tr}\exp(K+\ln\omega) can be interpreted as the Laplace transform of KK in the state ω\omega, and therefore the equivalence between Gaussian concentration and the transportation cost inequality holds. However, this is no longer true when KK and ω\omega do not commute, and the following bound can turn out to be strictly stronger to the transportation cost inequality as a consequence of the Golden-Thompson inequality: for any K𝒪VK\in\mathcal{O}_{V} such that Tr[ωK]=0\mathrm{Tr}[\omega K]=0,

lnTr[ωeK]C(ω)4KL2.\displaystyle\ln\mathrm{Tr}\left[\omega\,e^{K}\right]\leq\,\frac{C^{\prime}(\omega)}{4}\|K\|_{L}^{2}\,. (146)

In other words, C(ω)C(ω)C(\omega)^{\prime}\geq C(\omega). Recently, bounds of the form of (146) were obtained for some subclasses of Lipschitz observables KK (typically local observables) when the ω\omega is the Gibbs state of a (possibly non-commuting) quasi-local Hamiltonian HH [43] using cluster expansion techniques. However, the existence of the Gaussian concentration inequality for general Lipschitz observables was left open.

Here instead, we pursue a different approach using our transportation cost inequality. In particular, we prove that (146) can be approximately recovered for Gibbs states of commuting Hamiltonians for a larger class of Lipschitz observables than those considered in [43]. For this, we adapt the result of [13, Theorem 8] which was written for W1,W_{1,\nabla} to the case of W1W_{1}. In this section, we denote by XRX_{R}, respectively XIX_{I} the real, respectively imaginary parts of an operator XVX\in\mathcal{B}_{V}. Given an observable O𝒪VO\in\mathcal{O}_{V} with spectral decomposition O:=λλPλO:=\sum_{\lambda}\lambda P_{\lambda}, a state ω𝒮V\omega\in\mathcal{S}_{V} and a real number rr\in\mathbb{R}, we denote by

ω(Or):=λrTr(ωPλ)\displaystyle\mathbb{P}_{\omega}(O\geq r):=\sum_{\lambda\geq r}\mathrm{Tr}(\omega P_{\lambda}) (147)

the probability of getting an eigenvalue λr\lambda\geq r when measuring OO on the state ω\omega.

Theorem 7.

Assume that the full-rank state ω𝒮V\omega\in\mathcal{S}_{V} satisfies TC(c)\operatorname{TC}(c) for some c>0c>0. Then, for any observable O𝒪VO\in\mathcal{O}_{V},

ω(|OTr(ωO)𝕀|r)2exp(r24cmax{(ω12Oω12)RL2,(ω12Oω12)IL2}).\displaystyle\mathbb{P}_{\omega}\big{(}|O-\mathrm{Tr}({\omega O})\,\mathbb{I}|\geq r\big{)}\leq 2\,\exp\left(-\frac{r^{2}}{4c\max\big{\{}\|(\omega^{-\frac{1}{2}}O\omega^{\frac{1}{2}})_{R}\|_{L}^{2},\|(\omega^{-\frac{1}{2}}O\omega^{\frac{1}{2}})_{I}\|_{L}^{2}\big{\}}}\right)\,. (148)

Whenever [O,ω]=0[O,\omega]=0 the bound can be tightened into

ω(|OTr(ωO)𝕀|r)2exp(r2cOL2).\displaystyle\mathbb{P}_{\omega}\big{(}|O-\mathrm{Tr}({\omega O})\,\mathbb{I}|\geq r\big{)}\leq 2\,\exp\left(-\frac{r^{2}}{c\|O\|_{L}^{2}}\right)\,. (149)

Therefore, whenever ω𝒮V\omega\in\mathcal{S}_{V} corresponds to the Gibbs state of a local commuting Hamiltonian on a hypergraph at inverse temperature β\beta, the above bounds hold as long as 0<β<βc0<\beta<\beta_{c} where βc\beta_{c} is defined in (85).

Proof.

Given XVX\in\mathcal{B}_{V}, we denote by X:=XR+iXIX:=X_{R}+iX_{I} its decomposition onto real and imaginary parts. We also assume that Tr(ωX)=0\mathrm{Tr}(\omega X)=0 and XRL,XIL1\|X_{R}\|_{L},\|X_{I}\|_{L}\leq 1. By assumption, we have that for any ρ𝒮V\rho\in\mathcal{S}_{V}

|Tr(ρX)||Tr(ρXR)|+|Tr(ρXI)|2ρωW12cS(ρω).\displaystyle\left|\mathrm{Tr}(\rho X)\right|\leq\left|\mathrm{Tr}(\rho X_{R})\right|+\left|\mathrm{Tr}(\rho X_{I})\right|\leq 2\left\|\rho-\omega\right\|_{W_{1}}\leq 2\sqrt{c\,S(\rho\|\omega)}\,. (150)

Then, since infθ>0(aθ+bθ2)=2ab\inf_{\theta>0}\Big{(}\frac{a}{\theta}+\frac{b\theta}{2}\Big{)}=\sqrt{2ab} for any a,b0a,b\geq 0, we have that for all θ>0\theta>0:

|Tr(ρX)|2S(ρω)θ+cθ2θ|Tr(ρX)|c2θ22S(ρω).\displaystyle\big{|}\mathrm{Tr}(\rho X)\big{|}\leq\sqrt{2}\,\frac{S(\rho\|\omega)}{\theta}+\frac{c\,\theta}{\sqrt{2}}\qquad\Leftrightarrow\qquad\theta\big{|}\mathrm{Tr}(\rho X)\big{|}-\frac{c}{\sqrt{2}}\,\theta^{2}\leq\sqrt{2}\,S(\rho\|\omega)\,. (151)

Next, we further upper bound the relative entropy in terms of the maximal divergence S^(ρω):=Tr[ω(ω12ρω12)ln(ω12ρω12)]\widehat{S}(\rho\|\omega):=\mathrm{Tr}\big{[}\omega\big{(}\omega^{-\frac{1}{2}}\rho\omega^{-\frac{1}{2}}\big{)}\ln(\omega^{-\frac{1}{2}}\rho\omega^{-\frac{1}{2}})\big{]} [44]. Choosing ρ=ω12eθOω12/Tr(ωeθO)\rho=\omega^{\frac{1}{2}}e^{\theta O}\omega^{\frac{1}{2}}/\mathrm{Tr}(\omega e^{\theta O}) for some observable O𝒪VO\in\mathcal{O}_{V}, we arrive at

θ|Tr(ρX)|c2θ22θTr(ωeθOO)Tr(ωeθO)2ln(Tr(ωeθO)).\displaystyle\theta\big{|}\mathrm{Tr}(\rho X)\big{|}-\frac{c}{\sqrt{2}}\,\theta^{2}\leq\sqrt{2}\,\theta\frac{\mathrm{Tr}(\omega e^{\theta O}O)}{\mathrm{Tr}(\omega e^{\theta O})}-\sqrt{2}\ln\big{(}\mathrm{Tr}(\omega e^{\theta O})\big{)}\,. (152)

Next, we choose X=2ω12Oω12X=\sqrt{2}\omega^{-\frac{1}{2}}O\omega^{\frac{1}{2}}, so that the previous inequality reduces to

ln(Tr(ωeθO))c2θ2.\displaystyle\ln\big{(}\mathrm{Tr}(\omega e^{\theta O})\big{)}\leq\frac{c}{2}\,\theta^{2}\,. (153)

The above inequality can be interpreted as a bound on the log-Laplace transform of the non-commutative variable OO in the state ω\omega. By a use of Markov’s inequality followed by an optimization over the variable θ>0\theta>0, we finally get

ω(|O|r)2er22c.\displaystyle\mathbb{P}_{\omega}\big{(}\big{|}O\big{|}\geq r\big{)}\leq 2e^{-\frac{r^{2}}{2c}}\,. (154)

The result follows after simple rescalings. The tightening in the case of an observable commuting with ω\omega can be found by following the same steps as the ones above. ∎

In general, there is no way to precisely relate the Lipschitz constants of the real and imaginary parts of ω12Oω12\omega^{-\frac{1}{2}}O\omega^{\frac{1}{2}} to the Lipschitz constant of OO when [O,ω]0[O,\omega]\neq 0. In the next result, we however prove that the constants have similar scalings in the case of a commuting Gibbs measure ω\omega of a local Hamiltonian.

Lemma 7.

Let O=AVλAOA𝕀AcO=\sum_{A\subseteq V}\lambda_{A}\,O_{A}\otimes\mathbb{I}_{A^{c}} be the decomposition of an observable OO in 𝒪V\mathcal{O}_{V}, where for each subregion AA, OAO_{A} is exactly supported in AA with OA1\|O_{A}\|_{\infty}\leq 1, and λA\lambda_{A}\in\mathbb{R}. Let further ω\omega be the Gibbs state of a geometrically kk-local, commuting Hamiltonian HV:=BVhB𝕀BcH_{V}:=\sum_{B\subset V}h_{B}\otimes\mathbb{I}_{B^{c}} at inverse temperature β\beta. Then,

(ω12Oω12)RL,(ω12Oω12)IL4maxiVAVAki|λA|exp(βBAhB),\displaystyle\|(\omega^{-\frac{1}{2}}O\omega^{\frac{1}{2}})_{R}\|_{L}\,,\leavevmode\nobreak\ \leavevmode\nobreak\ \|(\omega^{-\frac{1}{2}}O\omega^{\frac{1}{2}})_{I}\|_{L}\leq 4\max_{i\in V}\,\sum_{\begin{subarray}{c}A\subset V\\ A_{\partial k}\ni i\end{subarray}}\,|\lambda_{A}|\,\exp\Big{(}\beta\sum_{B\cap A\neq\emptyset}\|h_{B}\|_{\infty}\Big{)}\,, (155)

where Ak:={jV:dist(j,A)k}A_{\partial k}:=\{j\in V:\,\operatorname{dist}(j,A)\leq k\} denotes the kk-enlargement of AA. In particular, whenever the state ω\omega satisfies TC(c)\operatorname{TC}(c) with c=O(n)c=O(n), any local observable OO gives rise to a sub-Gaussian random variable variance O(n)O(n) when measured in the state ω\omega.

Proof.

We prove the bound for the real part of (ω12Oω12)R(\omega^{-\frac{1}{2}}O\omega^{\frac{1}{2}})_{R} since the proof for the imaginary part follows the exact same reasoning. First, by Proposition 6, since for any AVA\subset V, (σ12OAσ12)R(\sigma^{-\frac{1}{2}}O_{A}\sigma^{\frac{1}{2}})_{R} is supported in region AkA_{\partial k}, we have that

(ω12Oω12)RL\displaystyle\|(\omega^{-\frac{1}{2}}O\omega^{\frac{1}{2}})_{R}\|_{L} 2maxiVAVAkiλA[(ω12OAω12)RTri[(ω12OAω12)R]𝕀dd]\displaystyle\leq 2\,\max_{i\in V}\,\Big{\|}\sum_{\begin{subarray}{c}A\subset V\\ A_{\partial k}\ni i\end{subarray}}\lambda_{A}\,\Big{[}(\omega^{-\frac{1}{2}}O_{A}\omega^{\frac{1}{2}})_{R}-\mathrm{Tr}_{i}\big{[}(\omega^{-\frac{1}{2}}O_{A}\omega^{\frac{1}{2}})_{R}\big{]}\otimes\frac{\mathbb{I}_{d}}{d}\Big{]}\Big{\|}_{\infty} (156)
4maxiVAVAki|λA|ω12OAω12\displaystyle\leq 4\,\max_{i\in V}\sum_{\begin{subarray}{c}A\subset V\\ A_{\partial k\ni i}\end{subarray}}\,|\lambda_{A}|\,\|\omega^{-\frac{1}{2}}O_{A}\omega^{\frac{1}{2}}\|_{\infty} (157)
4maxiVAVAki|λA|eβBAhBOAeβBAhB\displaystyle\leq 4\,\max_{i\in V}\sum_{\begin{subarray}{c}A\subset V\\ A_{\partial k\ni i}\end{subarray}}\,|\lambda_{A}|\,\|e^{\beta\sum_{B\cap A\neq\emptyset}h_{B}}O_{A}e^{-\beta\sum_{B\cap A\neq\emptyset}h_{B}}\|_{\infty} (158)
4maxiVAVAki|λA|eβBAhB.\displaystyle\leq 4\,\max_{i\in V}\sum_{\begin{subarray}{c}A\subset V\\ A_{\partial k\ni i}\end{subarray}}\,|\lambda_{A}|\,e^{\beta\sum_{B\cap A\neq\emptyset}\|h_{B}\|_{\infty}}\,. (159)

The result follows. ∎

7.1 Comparison to previous tail bounds

Our main result can be compared to other recently derived concentration bounds for quantum Gibbs states: in [45, Corollary 5.4], the authors consider a product state ρ=vVρv𝒮V\rho=\bigotimes_{v\in V}\rho_{v}\in\mathcal{S}_{V} as well as a Hamiltonian H=AEk,mhAH=\sum_{A\in E_{k,m}}h_{A}, where the set Ek,mE_{k,m} of subsets of VV has the following properties: for any AEk,mA\in E_{k,m},

  • (i)

    |A|k|A|\leq k;

  • (ii)

    |{AEk,m:AA}|m\big{|}\{A^{\prime}\in E_{k,m}:\,A\cap A^{\prime}\neq\emptyset\}\big{|}\leq m.

With these conditions, he was able to prove that

ρ(|HTr(ρH)|r)2er24eN3kn,\displaystyle\mathbb{P}_{\rho}\big{(}|H-\mathrm{Tr}(\rho H)|\geq r\big{)}\leq 2e^{-\frac{r^{2}}{4eN^{3}kn}}\,,

where number N:=maxvVAEk,m|vA 1N:=\max_{v\in V}\sum_{A\in E_{k,m}|v\in A}\,1 is the number of local terms acting non-trivially on spin vv. A similar bound was previously derived by Kuwahara [46, Theorem 7], under a notion of gg-extensivity: a local Hamiltonian HH is said to be gg-extensive if for every spin vv, AEk,m|vAhAg\sum_{A\in E_{k,m}|\,v\in A}\|h_{A}\|_{\infty}\leq g. Under this condition, he shows that

ρ(|HTr(ρH)|r)=𝒪(1)er2cnlog(rn),\displaystyle\mathbb{P}_{\rho}\big{(}|H-\mathrm{Tr}(\rho H)|\geq r\big{)}=\mathcal{O}(1)\,e^{-\frac{r^{2}}{cn\log(\frac{r}{\sqrt{n}})}}\,,

where cc is a 𝒪(1)\mathcal{O}(1) constant which depends only on kk and gg. Although these results recover the Gaussian tails of our Theorem 7 (up to logarithmic overheads), they only work for tensor product states and a subclass of Lipschitz observables. In particular, the tails become trivial whenever the Hamiltonian is a sum of terms acting on non-intersecting regions AA of arbitrary size. In contrast, our bound is still non-trivial for this class of observables, since their Lipschitz constant is still 𝒪(1)\mathcal{O}(1).

More recently, Kuwahara and Saito derived new concentration bounds for Gibbs states of interacting Hamiltonians in order to study the problem of equivalence of quantum statistical ensembles [47, 43] (see section 8): in [47] first, the authors consider a Gibbs state ω\omega of a local Hamiltonian on a DD-dimensional regular lattice D\mathbb{Z}^{D}. They further assume the following (r0,ξ)(r_{0},\xi) clustering: for any operators OA,OBO_{A},O_{B} supported on the subsets AA and BB,

|Tr(ωOAOB)Tr(ωOA)Tr(ωOB)|OAOBedist(A,B)/ξ,\displaystyle|\mathrm{Tr}(\omega O_{A}O_{B})-\mathrm{Tr}(\omega O_{A})\mathrm{Tr}(\omega O_{B})|\leq\|O_{A}\|_{\infty}\,\|O_{B}\|_{\infty}\,e^{-\operatorname{dist}(A,B)/\xi}, (160)

whenever dist(A,B)r0\operatorname{dist}(A,B)\geq r_{0}. Under this condition, they were able to show in Equation (S.17) (see also [45, Theorem 4.2] for a similar bound)

ω(|OTr(ωO)|r)min{1,(e+3eξ)max(e(r2/(cn))1D+1,er2cDn)},\displaystyle\mathbb{P}_{\omega}\big{(}|O-\mathrm{Tr}(\omega O)|\geq r\big{)}\leq\min\big{\{}1,(e+3e\xi)\max\big{(}e^{-(r^{2}/(cn))^{\frac{1}{D+1}}},\,e^{-\frac{r^{2}}{c^{\prime}\ell^{D}n}}\big{)}\big{\}}\,,

for some constants c,cc,c^{\prime} further depending on ξ\xi and DD, and where \ell denotes the locality of the observable OO. Therefore, and although the clustering of correlations is known to hold at high enough temperature [48], the bound is suboptimal for two reasons: firstly, whenever rr is small enough, the exponent has the worse scaling r2/(D+1)r^{2/(D+1)}. Secondly, the bounds baddly dependence on the locality of OO, and becomes trivial whenever OO is a sum of highly non-local terms. This second limitation also holds for the Gaussian concentration bound found in [43, Corollary 1] for high-temperature Gibbs states of Hamiltonians with long-range interactions. In comparison to the works cited above, our bound always provides better dependence of the tail on the locality of the observable, albeit under the condition that the Hamiltonian is made of local commuting terms.

8 Equivalence of statistical mechanical ensembles

The three main ensembles employed in quantum statistical mechanics to compute the equilibrium properties of quantum systems are the canonical ensemble, the microcanonical ensemble and the diagonal ensemble. The quantum state associated to the canonical ensemble is the Gibbs state, which describes the physics of a system that is at thermal equilibrium with a large bath at a given temperature. The diagonal and microcanonical ensembles both describe the physics of an isolated quantum system, and the associated states are convex combinations of the eigenstates of the Hamiltonian. The microcanonical ensemble assumes a uniform probability distribution for the energy in a given energy shell. The diagonal ensemble includes all the states that are diagonal in the eigenbasis of the Hamiltonian, and in particular it includes the eigenstates themselves.

For many quantum systems, the microcanonical and canonical ensembles give the same expectation values for local observables if the corresponding states have approximately the same average energy. A lot of effort has been devoted to determining conditions under which the two ensembles are equivalent [49, 50, 51, 52]. The most prominent among such conditions are short ranged interactions and a finite correlation length, but analytical proofs can be obtained only in the case of regular lattices [52]. The situation is more complex for the diagonal ensemble. The condition under which this ensemble is equivalent to the microcanonical and canonical ensembles is called Eigenstate Thermalization Hypothesis (ETH) [53, 54, 55, 56, 57], and states that the expectation values of local observables on the eigenstates of the Hamiltonian are a smooth function of the energy, i.e., for any given local observable, any two eigenstates with approximately the same energy yield approximately the same expectation value. The ETH is an extremely strong condition on the Hamiltonian and several quantum systems, including all integrable systems, do not satisfy it. A weak version of the ETH has been formulated [58, 47], stating that for any given local observable, most eigenstates in an energy shell yield approximately the same expectation value, or, more precisely, that the fraction of eigenstates yielding expectation values far from the Gibbs state with the same average energy vanishes in the thermodynamical limit. The weak ETH implies the equivalence between the canonical and microcanonical ensembles, but is not sufficient to prove their equivalence with the diagonal ensemble. Under the hypothesis of finite correlation length in the Gibbs state, an analytical proof of the weak ETH is available only for regular lattices [47].

A connection between a transportation cost inequality and the ETH was made by one of the authors in the case of a regular lattice and a nearest neighbour Hamiltonian [38]. Here we look at the general problem of the equivalence of the statistical mechanical ensembles and of the weak ETH from the perspective of optimal mass transport, and show that such equivalence can be formulated as closeness of the respective states in the W1W_{1} distance. The closeness in the W1W_{1} distance implies closeness of the expectation values of all Lipschitz observables, which constitute a significantly larger class than local observables. Therefore, the perspective of optimal mass transport can significantly extend the previous results. Moreover, we will show that the equivalence of the ensembles is intimately linked to the constant of the transportation cost inequality for the Gibbs states.

As in the rest of the paper, we consider a quantum system made by nn qudits located at the vertices of a graph with vertex set VV. Let us assume that a Gibbs state ω𝒮V\omega\in\mathcal{S}_{V} satisfies the transportation cost inequality with a constant

C(ω)nC,C(\omega)\leq n\,C\,, (161)

where CC does not depend on nn. This condition is satisfied under the hypotheses of Theorem 2, Theorem 4 or Theorem 5. We stress that, contrarily to the results of Refs. [52, 47], the condition does not require us to restrict to regular lattices, since Theorem 4 does not need this hypothesis. The following Proposition 12 implies that any state ρ𝒮V\rho\in\mathcal{S}_{V} is close in W1W_{1} distance to the Gibbs state ω\omega with the same average energy, provided that ρ\rho and ω\omega have approximately the same entropy, i.e.,

S(ω)S(ρ)n.S(\omega)-S(\rho)\ll n\,. (162)

Moreover, under the same hypothesis, the average reduced states over one qudit of ρ\rho and ω\omega are close in trace distance.

Proposition 12.

Let ω𝒮V\omega\in\mathcal{S}_{V} be a Gibbs state for the Hamiltonian H𝒪VH\in\mathcal{O}_{V}. Then, any quantum state ρ𝒮V\rho\in\mathcal{S}_{V} with the same average energy as ω\omega satisfies

1nρωW1CS(ω)S(ρ)n.\frac{1}{n}\left\|\rho-\omega\right\|_{W_{1}}\leq\sqrt{C\,\frac{S(\omega)-S(\rho)}{n}}\,. (163)

Moreover, let Λ:𝒪V𝒪(d)\Lambda:\mathcal{O}_{V}\to\mathcal{O}(\mathbb{C}^{d}) be the quantum channel that computes the average marginal state over one qudit, i.e., for any ρ𝒮V\rho\in\mathcal{S}_{V},

Λ(ρ)=1nvVρv.\Lambda(\rho)=\frac{1}{n}\sum_{v\in V}\rho_{v}\,. (164)

Then,

Λ(ρ)Λ(ω)12CS(ω)S(ρ)n.\left\|\Lambda(\rho)-\Lambda(\omega)\right\|_{1}\leq 2\sqrt{C\,\frac{S(\omega)-S(\rho)}{n}}\,. (165)
Proof.

We have from the transportation cost inequality

ρωW1C(ω)S(ρω)=C(ω)(S(ω)S(ρ)),\left\|\rho-\omega\right\|_{W_{1}}\leq\sqrt{C(\omega)\,S(\rho\|\omega)}=\sqrt{C(\omega)\left(S(\omega)-S(\rho)\right)}\,, (166)

where the last equality follows since Tr[ρlnω]=Tr[ωlnω]\mathrm{Tr}\left[\rho\ln\omega\right]=\mathrm{Tr}\left[\omega\ln\omega\right].

We have from Proposition 7

Λ(ρ)Λ(ω)11nvVρvωv12nρωW12nC(ω)(S(ω)S(ρ)),\displaystyle\left\|\Lambda(\rho)-\Lambda(\omega)\right\|_{1}\leq\frac{1}{n}\sum_{v\in V}\left\|\rho_{v}-\omega_{v}\right\|_{1}\leq\frac{2}{n}\left\|\rho-\omega\right\|_{W_{1}}\leq\frac{2}{n}\sqrt{C(\omega)\left(S(\omega)-S(\rho)\right)}\,, (167)

and the claim follows. ∎

Choosing ρ\rho to be diagonal in the eigenbasis of the Hamiltonian, Proposition 12 implies that any convex combination of a sufficiently large number of eigenstates is close in W1W_{1} distance to the Gibbs state with the same average energy. Such number of eigenstates can even be an exponentially small fraction of the total number of eigenstates appearing in a microcanonical state, since the uniform superposition of a fraction enϵe^{-n\epsilon} of the eigenstates decreases the entropy by ϵn\epsilon\,n. Therefore, Proposition 12 constitutes an exponential improvement over the weak ETH.

A natural question is whether also the strong ETH can be captured by the W1W_{1} distance. Unfortunately the answer is negative. Indeed, proving the strong ETH via optimal mass transport would mean to prove that all the eigenstates of the Hamiltonian are close in W1W_{1} distance to the Gibbs states with the corresponding average energy. However, Theorem 1 implies that any state with low entropy, and in particular any pure state, is far from any state with large entropy, and in particular from a Gibbs state with temperature Ω(1)\Omega(1). More precisely, for any two states ρ,ω𝒮V\rho,\,\omega\in\mathcal{S}_{V},

ρωW1S(ω)S(ρ)ln(n+1)1ln(d2n).\left\|\rho-\omega\right\|_{W_{1}}\geq\frac{S(\omega)-S(\rho)-\ln\left(n+1\right)-1}{\ln\left(d^{2}n\right)}\,. (168)

Equation (168) also implies that any quantum state which is close in W1W_{1} distance to the Gibbs state with the same average energy must have approximately also the same entropy, and in this sense Proposition 12 is optimal.

8.1 Comparison with previous results

To make our result more easily comparable to the literature, let us introduce more formally the microcanonical ensemble: given the decomposition H=EEP(E)H=\sum_{E}EP(E), we define the microcanonical ensemble state

ωE,Δ:=P(E,Δ)Tr(P(E,Δ)),\displaystyle\omega_{E,\Delta}:=\frac{P(E,\Delta)}{\mathrm{Tr}(P(E,\Delta))}\,,

where P(E,Δ)P(E,\Delta) corresponds to the projection onto the subspace spanned by the eigenvectors whose energy belongs to the interval (EΔ,E](E-\Delta,E].

Corollary 2.

Assume the Gibbs state ω\omega satisfies C(ω)CnC(\omega)\leq Cn. Then for any Lipschitz observable OO,

1n|Tr(ωO)Tr(ωE,ΔO)|OLon(1).\displaystyle\frac{1}{n}\,\big{|}\mathrm{Tr}(\omega O)-\mathrm{Tr}(\omega_{E,\Delta}O)\big{|}\leq\|O\|_{L}\,o_{n\to\infty}(1)\,.
Proof.

In view of Proposition 12, it suffices to control the relative entropy between the microcanonical and canonical ensemble states. Then,

S(ωE,Δω)=lnTr(eβH)Tr(P(E,Δ))+βTr[HP(E,Δ)Tr(P(E,Δ))]\displaystyle S(\omega_{E,\Delta}\|\omega)=\ln\frac{\mathrm{Tr}(e^{-\beta H})}{\mathrm{Tr}(P(E,\Delta))}+\beta\mathrm{Tr}\Big{[}H\,\frac{P(E,\Delta)}{\mathrm{Tr}(P(E,\Delta))}\Big{]} βE+ln[Tr(eβH)Tr(P(E,Δ))].\displaystyle\leq\beta E+\ln\Big{[}\frac{\mathrm{Tr}(e^{-\beta H})}{\mathrm{Tr}(P(E,\Delta))}\Big{]}\,. (169)

Next, we control the ration Tr(eβH)Tr(P(E,Δ))\frac{\mathrm{Tr}(e^{-\beta H})}{\mathrm{Tr}(P(E,\Delta))}. For this, we use an argument which was already used in [47, Equation (S.56)]: First, we have found in (149) that

ω(|HTr(ωH)𝕀|r)2er2CnHL2.\displaystyle\mathbb{P}_{\omega}(|H-\mathrm{Tr}(\omega H)\mathbb{I}|\geq r)\leq 2\,e^{-\frac{r^{2}}{Cn\|H\|_{L}^{2}}}\,.

Therefore, choosing the interval Δ~:=(Tr(ωH)Cnln(4)HL,Tr(ωH)+Cnln(4)HL]\tilde{\Delta}:=(\mathrm{Tr}(\omega H)-\sqrt{Cn\ln(4)}\|H\|_{L},\mathrm{Tr}(\omega H)+\sqrt{Cn\ln(4)}\|H\|_{L}], we have

Tr[EΔ~eβETr(eβH)P(E)]12\displaystyle\mathrm{Tr}\Big{[}\sum_{E\in\tilde{\Delta}}\,\frac{e^{-\beta E}}{\mathrm{Tr}(e^{-\beta H})}P(E)\Big{]}\geq\frac{1}{2}

Next, we define

Z~:=Tr[EΔ~eβEP(E)]Tr(eβH)2\displaystyle\tilde{Z}:=\mathrm{Tr}\Big{[}\sum_{E\in\tilde{\Delta}}e^{-\beta E}P(E)\Big{]}\geq\frac{\mathrm{Tr}(e^{-\beta H})}{2} (170)

Choosing a slightly extended interval Δ~:=(Tr(ωH)Cnln(4)HLΔ,Tr(ωH)+Cnln(4)HL+Δ]\tilde{\Delta}^{\prime}:=(\mathrm{Tr}(\omega H)-\sqrt{Cn\ln(4)}\|H\|_{L}-\Delta,\mathrm{Tr}(\omega H)+\sqrt{Cn\ln(4)}\|H\|_{L}+\Delta], we have

Z~ν:νΔΔ~Tr(P(νΔ,Δ))eβΔ(ν1)\displaystyle\tilde{Z}\leq\sum_{\nu\in\mathbb{Z}:\,\nu\Delta\in\tilde{\Delta}^{\prime}}\mathrm{Tr}(P(\nu\Delta,\Delta))\,e^{-\beta\Delta(\nu-1)} (171)

Now, for E:=argmaxeβETr(P(E,Δ))E:=\operatorname{argmax}e^{-\beta E}\mathrm{Tr}(P(E,\Delta)), we have

Tr(νΔ,Δ)eβδ(ν1)eβΔeβETr(P(E,Δ)).\displaystyle\mathrm{Tr}(\nu\Delta,\Delta)e^{-\beta\delta(\nu-1)}\leq e^{\beta\Delta}e^{-\beta E}\,\mathrm{Tr}(P(E,\Delta))\,.

Replacing in (171), we have that

Z~eβΔ(2+2Cnln(4)HL/Δ)eβETr(P(E,Δ))\displaystyle\tilde{Z}\leq e^{\beta\Delta}(2+2\sqrt{Cn\ln(4)}\|H\|_{L}/\Delta)e^{-\beta E}\mathrm{Tr}(P(E,\Delta))

Finally, using the lower bound (170), we have that

ln[Tr(eβH)Tr(P(E,Δ))]β(ΔE)+ln(4+4Cnln(4)HL/Δ)\displaystyle\ln\Big{[}\frac{\mathrm{Tr}(e^{-\beta H})}{\mathrm{Tr}(P(E,\Delta))}\Big{]}\leq\beta(\Delta-E)+\ln(4+4\sqrt{Cn\ln(4)}\|H\|_{L}/\Delta)

Therefore, plugging this last bound into (169), we have found that

S(ωE,Δω)βΔ+ln(4+4Cnln(4)HL/Δ).\displaystyle S(\omega_{E,\Delta}\|\omega)\leq\beta\Delta+\ln(4+4\sqrt{Cn\ln(4)}\|H\|_{L}/\Delta)\,.

Therefore, S(ωE,Δω)=o(n)S(\omega_{E,\Delta}\|\omega)=o(n) whenever Δ=eo(n)\Delta=e^{-o(n)}, and the result follows.

In [47, Theorem 2], it is showed that, under the (r0,ξ)(r_{0},\xi)-clustering of correlations (160), for any observable O:=vOvO:=\sum_{v}O_{v} where each OvO_{v} acts on spin vv as well as other spins ww with dist(v,w)\operatorname{dist}(v,w)\leq\ell and has Ov1\|O_{v}\|_{\infty}\leq 1,

1n|Tr(ωE,ΔO)Tr(ωO)|1nmax(c1B1,c2B2),\displaystyle\frac{1}{n}\,\big{|}\mathrm{Tr}(\omega_{E,\Delta}O)-\mathrm{Tr}(\omega O)\big{|}\leq\frac{1}{\sqrt{n}}\,\max(c_{1}B_{1},c_{2}B_{2}),

where B1:=log(n/Δ)d+12B_{1}:=\log(\sqrt{n}/\Delta)^{\frac{d+1}{2}}, B2:=(Dlog(n/Δ))12B_{2}:=(\ell^{D}\log(\sqrt{n}/\Delta))^{\frac{1}{2}}, and the constants c1c_{1} and c2c_{2} depend on D,r0,ξD,r_{0},\xi and the locality kk of HH. Therefore, as long as the energy shell Δ\Delta is chosen as Δe𝒪(n1D+1)\Delta\sim e^{-\mathcal{O}(n^{\frac{1}{D+1}})} the averages of the operator density On\frac{O}{n} in the canonical and microcanonical ensemble states converge to the same number as nn\to\infty. Similar bounds were also derived in [43, Corollary 3] for larger classes of non-local Hamiltonians and observables above some threshold temperature. Corollary 2 constitutes an improvement over these results in two senses: Firstly, it applies to a more general class of Lipschitz observables. Secondly, it allows for a smaller energy shell Δ=eo(n)\Delta=e^{-o(n)}. However, the condition TC(Cn)\operatorname{TC}(Cn) is currently only known to hold for the smaller class of local commuting Hamiltonians.

Acknowledgements

The research of CR has been supported by project QTraj (ANR-20-CE40-0024-01) of the French National Research Agency (ANR) and by a Junior Researcher START Fellowship from the MCQST.

References

  • [1] Maxim Raginsky and Igal Sason. Concentration of measure inequalities in information theory, communications and coding. arXiv preprint arXiv:1212.4663, 2012.
  • [2] Stéphane Boucheron, Gábor Lugosi, and Pascal Massart. Concentration inequalities: A nonasymptotic theory of independence. Oxford university press, 2013.
  • [3] Aryeh Kontorovich and Maxim Raginsky. Concentration of measure without independence: a unified approach via the martingale method. In Convexity and Concentration, pages 183–210. Springer, 2017.
  • [4] Katalin Marton. A simple proof of the blowing-up lemma (corresp.). IEEE Transactions on Information Theory, 32(3):445–446, 1986.
  • [5] Sergej G Bobkov and Friedrich Götze. Exponential integrability and transportation cost related to logarithmic sobolev inequalities. Journal of Functional Analysis, 163(1):1–28, 1999.
  • [6] Roland L Dobrushin. Prescribing a system of random variables by conditional distributions. Theory of Probability & Its Applications, 15(3):458–486, 1970.
  • [7] Katalin Marton. Logarithmic sobolev inequalities in discrete product spaces. Combinatorics, Probability & Computing, 28(6):919–935, 2019.
  • [8] Roland L Dobrushin and Senya B Shlosman. Completely analytical interactions: constructive description. Journal of Statistical Physics, 46(5):983–1014, 1987.
  • [9] Christof Külske. Concentration inequalities for functions of gibbs fields with application to diffraction and random gibbs measures. Communications in mathematical physics, 239(1):29–51, 2003.
  • [10] Marius Junge and Qiang Zeng. Noncommutative martingale deviation and poincaré type inequalities with applications. Probability Theory and Related Fields, 161(3-4):449–507, February 2014.
  • [11] Joel A. Tropp. An introduction to matrix concentration inequalities. Foundations and Trends® in Machine Learning, 8(1-2):1–230, 2015.
  • [12] De Huang and Joel A. Tropp. From poincaré inequalities to nonlinear matrix concentration. Bernoulli, 27(3), May 2021.
  • [13] Cambyse Rouzé and Nilanjana Datta. Concentration of quantum states from quantum functional and transportation cost inequalities. Journal of Mathematical Physics, 60(1):012202, January 2019.
  • [14] Giacomo De Palma, Milad Marvian, Dario Trevisan, and Seth Lloyd. The Quantum Wasserstein Distance of Order 1. IEEE Transactions on Information Theory, 67(10):6627–6643, 2021.
  • [15] Giacomo De Palma and Dario Trevisan. Quantum Optimal Transport with Quantum Channels. Annales Henri Poincaré, 22:3199–3234, 2021.
  • [16] François Golse, Clément Mouhot, and Thierry Paul. On the mean field and classical limits of quantum mechanics. Communications in Mathematical Physics, 343(1):165–205, January 2016.
  • [17] Sam Cole, Michał Eckstein, Shmuel Friedland, and Karol Życzkowski. Quantum optimal transport. arXiv preprint arXiv:2105.06922, 2021.
  • [18] Eric A Carlen and Jan Maas. Gradient flow and entropy inequalities for quantum markov semigroups with detailed balance. Journal of Functional Analysis, 273(5):1810–1869, 2017.
  • [19] Ivan Bardet, Ángela Capel, Li Gao, Angelo Lucia, David Pérez-García, and Cambyse Rouzé. Entropy decay for davies semigroups of a one dimensional quantum lattice, 2021.
  • [20] Ivan Bardet, Ángela Capel, Li Gao, Angelo Lucia, David Pérez-García, and Cambyse Rouzé. Rapid thermalization of spin chain commuting hamiltonians, 2021.
  • [21] Michael A Nielsen and Isaac Chuang. Quantum computation and quantum information, 2002.
  • [22] Mark M Wilde. Quantum information theory. Cambridge University Press, 2013.
  • [23] Alexander S Holevo. Quantum systems, channels, information: a mathematical introduction, volume 16. Walter de Gruyter, 2012.
  • [24] Matthew J Donald. On the relative entropy. Communications in Mathematical Physics, 105(1):13–34, 1986.
  • [25] Dénes Petz. Sufficient subalgebras and the relative entropy of states of a von neumann algebra. Communications in mathematical physics, 105(1):123–131, 1986.
  • [26] Fumio Hiai and Dénes Petz. The proper formula for relative entropy and its asymptotics in quantum probability. Communications in mathematical physics, 143(1):99–114, 1991.
  • [27] Mario Berta, Omar Fawzi, and Marco Tomamichel. On variational expressions for quantum relative entropies. Letters in Mathematical Physics, 107(12):2239–2265, 2017.
  • [28] Marius Junge, Renato Renner, David Sutter, Mark M Wilde, and Andreas Winter. Universal recovery maps and approximate sufficiency of quantum relative entropy. In Annales Henri Poincaré, volume 19, pages 2955–2978. Springer, 2018.
  • [29] David Sutter, Mario Berta, and Marco Tomamichel. Multivariate trace inequalities. Communications in Mathematical Physics, 352(1):37–58, October 2016.
  • [30] Patrick Hayden, Richard Jozsa, Denes Petz, and Andreas Winter. Structure of states which satisfy strong subadditivity of quantum entropy with equality. Communications in mathematical physics, 246(2):359–374, 2004.
  • [31] M. Tomamichel. Quantum Information Processing with Finite Resources: Mathematical Foundations. SpringerBriefs in Mathematical Physics. Springer International Publishing, 2015.
  • [32] Yann Ollivier. Ricci curvature of Markov chains on metric spaces. Journal of Functional Analysis, 256(3):810–864, February 2009.
  • [33] Ronen Eldan, James R. Lee, and Joseph Lehec. Transport-entropy inequalities and curvature in discrete-space markov chains. In A Journey Through Discrete Mathematics, pages 391–406. Springer International Publishing, 2017.
  • [34] Robert B. Griffiths. Correlations in ising ferromagnets. III. Communications in Mathematical Physics, 6(2):121–127, June 1967.
  • [35] Michael J. Kastoryano and Kristan Temme. Quantum logarithmic sobolev inequalities and rapid mixing. Journal of Mathematical Physics, 54(5):052202, May 2013.
  • [36] Li Gao and Cambyse Rouzé. Complete entropic inequalities for quantum Markov chains. arXiv preprint arXiv:2102.04146, 2021.
  • [37] Eric A. Carlen and Jan Maas. Non-commutative calculus, optimal transport and functional inequalities in dissipative quantum systems. Journal of Statistical Physics, 178(2):319–378, nov 2019.
  • [38] Ángela Capel, Cambyse Rouzé, and Daniel Stilck França. The modified logarithmic Sobolev inequality for quantum spin systems: classical and commuting nearest neighbour interactions. arXiv preprint arXiv:2009.11817, 2020.
  • [39] Toby S Cubitt, Angelo Lucia, Spyridon Michalakis, and David Perez-Garcia. Stability of local quantum dissipative systems. Communications in Mathematical Physics, 337:1275–1315, 2015.
  • [40] Michael J Kastoryano and Fernando GSL Brandao. Quantum gibbs samplers: the commuting case. Communications in Mathematical Physics, 344(3):915–957, 2016.
  • [41] Fernando GSL Brandão and Michael J Kastoryano. Finite correlation length implies efficient preparation of quantum thermal states. Communications in Mathematical Physics, 365(1):1–16, 2019.
  • [42] Marius Junge, Nicholas LaRacuente, and Cambyse Rouzé. Stability of logarithmic sobolev inequalities under a noncommutative change of measure. arXiv preprint arXiv:1911.08533, 2019.
  • [43] Tomotaka Kuwahara and Keiji Saito. Gaussian concentration bound and ensemble equivalence in generic quantum many-body systems including long-range interactions. Annals of Physics, 421:168278, October 2020.
  • [44] Keiji Matsumoto. A new quantum version of f-divergence. In Nagoya Winter Workshop: Reality and Measurement in Algebraic Quantum Theory, pages 229–273. Springer, 2015.
  • [45] Anurag Anshu. Concentration bounds for quantum states with finite correlation length on quantum spin lattice systems. New Journal of Physics, 18(8):083011, August 2016.
  • [46] Tomotaka Kuwahara. Connecting the probability distributions of different operators and generalization of the chernoff–hoeffding inequality. Journal of Statistical Mechanics: Theory and Experiment, 2016(11):113103, November 2016.
  • [47] Tomotaka Kuwahara and Keiji Saito. Eigenstate thermalization from the clustering property of correlation. Physical review letters, 124(20):200604, 2020.
  • [48] M. Kliesch, C. Gogolin, M. J. Kastoryano, A. Riera, and J. Eisert. Locality of temperature. Phys. Rev. X, 4:031019, Jul 2014.
  • [49] R Lima. Equivalence of ensembles in quantum lattice systems. Annales de l’IHP Physique théorique, 15(1):61–68, 1971.
  • [50] R Lima et al. Equivalence of ensembles in quantum lattice systems: states. Communications in Mathematical Physics, 24(3):180–192, 1972.
  • [51] Markus P Müller, Emily Adlam, Lluís Masanes, and Nathan Wiebe. Thermalization and canonical typicality in translation-invariant quantum lattice systems. Communications in Mathematical Physics, 340(2):499–561, 2015.
  • [52] Fernando G. S. L. Brandão and Marcus Cramer. Equivalence of Statistical Mechanical Ensembles for Non-Critical Quantum Systems. arXiv:1502.03263 [cond-mat, physics:quant-ph], February 2015. arXiv: 1502.03263.
  • [53] Josh M Deutsch. Quantum statistical mechanics in a closed system. Physical Review A, 43(4):2046, 1991.
  • [54] Mark Srednicki. Chaos and quantum thermalization. Physical Review E, 50(2):888–901, August 1994.
  • [55] Christian Gogolin and Jens Eisert. Equilibration, thermalisation, and the emergence of statistical mechanics in closed quantum systems. Reports on Progress in Physics, 79(5):056001, 2016.
  • [56] Peter Reimann. Dynamical typicality approach to eigenstate thermalization. Physical review letters, 120(23):230601, 2018.
  • [57] Giacomo De Palma, Alessio Serafini, Vittorio Giovannetti, and Marcus Cramer. Necessity of eigenstate thermalization. Physical Review Letters, 115(22):220401, 2015.
  • [58] Giulio Biroli, Corinna Kollath, and Andreas M Läuchli. Effect of rare fluctuations on the thermalization of isolated quantum systems. Physical review letters, 105(25):250401, 2010.