14
Relative purity, speed of fluctuations, and bounds on equilibration times Diego Paiva Pires 1 and Thiago R. de Oliveira 2 1 Departamento de F´ ısica, Universidade Federal do Maranh˜ ao, Campus Universit´ ario do Bacanga, 65080-805, S˜ ao Lu´ ıs, Maranh˜ ao, Brazil 2 Instituto de F´ ısica, Universidade Federal Fluminense, Av. Gal. Milton Tavares de Souza s/n, Gragoat´ a, 24210-346, Niter´ oi, RJ, Brazil We discuss the local equilibration of closed systems using the relative purity, a paradigmatic information-theoretic distinguishability measure that finds applications ranging from quantum metrology to quantum speed limits. First we obtain an upper bound on the average size of the fluctuations on the relative purity: it depends on the effective dimension resembling the bound obtained with the trace distance. Second, we investigate the dynamics of relative purity and its rate of change as a probe of the speed of fluctuations around the equilibrium. In turn, such speed captures the notion of how fast some nonequilibrium state approaches the steady state under the local nonunitary dynamics, somehow giving the information of the quantum speed limit towards the equilibration. We show that the size of fluctuations depends on the quantum coherences of the initial state with respect to the eigenbasis of the Hamiltonian, also addressing the role played by the correlations between system and reservoir into the averaged speed. Finally, we have derived a family of lower bounds on the time of evolution between these states, thus obtaining an estimate for the equilibration time at the local level. These results could be of interest to the subjects of equilibration, quantum speed limits, and also quantum metrology. I. INTRODUCTION Both the subjects of equilibration and thermalization fall into the very basic foundational questions of sta- tistical mechanics: how to derive the macroscopic laws of thermodynamics from the microscopic many-particle laws. To do so, one can consider a quantum system ini- tialized in a well defined pure state and, under some mi- nimal assumptions, still conclude the system behaves as if it were described by an equilibrium ensemble [1]. This argument has been made technically rigorous, and also numerically confirmed by simulating several interacting quantum many-body systems [2, 3]. Indeed, the problem of equilibration has attracted much interest in the last decades from both the theoretical [4] and experimental communities [510]. Overall, probing the mechanism of local equilibration of a closed quantum system requires answering whether and how some of its nonequilibrium states equilibrate, even if such states do not belong to a Gibbs-like statis- tical ensemble. The isolated system is initialized in a pure state and undergoes a unitary evolution governed by the time-independent Hamiltonian H. At the local level, the notion of equilibration involves monitoring the nonunitary dynamics of some reduced state of a small subregion of the isolated system, and quantifying how far apart it is from an equilibrium state [11]. In turn, the task of distinguishing such states can be accomplished by means of a suitable distance measure on the Hilbert space, for example the Schatten 1-norm [12, 13]. So far, while there are plenty of rigorous results showing that equilibration should occur under very general conditions for small subsystems, there are a few results about the time scales involved in such physical process [14]. Here we will study the relative purity as a figure of merit for equilibration and show under which conditions the subsystem equilibrates. The relative purity stands as a versatile information-theoretic quantifier for distin- guishing two quantum states, also being an experimental friendly measure since it relies on the overlap of den- sity matrices [15]. We also consider the rate of change of relative purity as signaling the speed of the fluctua- tions, thus deriving upper bounds on such velocity in a similar fashion to the well-known discussion of quantum speed limits (QSLs). Noteworthy, from these inequalities we obtain lower bounds on the equilibration time of the subsystem, thus connecting both the subjects of equili- bration and quantum speed limits. The paper is organized as follows. In Sec. II we re- view basic concepts on the subject of equilibration, also introducing the relative purity as a figure of merit to sig- nal equilibration. In Sec. III we discuss the fluctuations around the equilibrium of a bipartite closed quantum sys- tem (S + B), thus analyzing the dynamics of relative pu- rity between a marginal state of subsystem S and some steady state. We proved a set of lower bounds on the equilibration time that are fully characterized by the ini- tial state and the Hamiltonian of the closed system [see Secs. III A, III B, III C, III D, and III E]. In Sec. IV we illustrate our findings by means of two paradigmatic spin models, namely, the transverse field Ising model and the nonintegrable XXZ model. Finally, in Sec. V we summa- rize our conclusions. II. RELATIVE PURITY AND EQUILIBRATION Let us consider a quantum system described by a finite-dimensional Hilbert space H = H S ⊗H B , with d = dim H, thus being split into a subsystem S of dimen- sion d S = dim H S , and its complement B of dimension d B = dim H B . The whole system S + B evolves unitarily arXiv:2108.05936v2 [quant-ph] 27 Nov 2021

arXiv:2108.05936v1 [quant-ph] 12 Aug 2021

  • Upload
    others

  • View
    5

  • Download
    0

Embed Size (px)

Citation preview

Page 1: arXiv:2108.05936v1 [quant-ph] 12 Aug 2021

Relative purity, speed of fluctuations, and bounds on equilibration times

Diego Paiva Pires1 and Thiago R. de Oliveira2

1Departamento de Fısica, Universidade Federal do Maranhao,Campus Universitario do Bacanga, 65080-805, Sao Luıs, Maranhao, Brazil

2Instituto de Fısica, Universidade Federal Fluminense,Av. Gal. Milton Tavares de Souza s/n, Gragoata, 24210-346, Niteroi, RJ, Brazil

We discuss the local equilibration of closed systems using the relative purity, a paradigmaticinformation-theoretic distinguishability measure that finds applications ranging from quantummetrology to quantum speed limits. First we obtain an upper bound on the average size of thefluctuations on the relative purity: it depends on the effective dimension resembling the boundobtained with the trace distance. Second, we investigate the dynamics of relative purity and itsrate of change as a probe of the speed of fluctuations around the equilibrium. In turn, such speedcaptures the notion of how fast some nonequilibrium state approaches the steady state under thelocal nonunitary dynamics, somehow giving the information of the quantum speed limit towardsthe equilibration. We show that the size of fluctuations depends on the quantum coherences of theinitial state with respect to the eigenbasis of the Hamiltonian, also addressing the role played bythe correlations between system and reservoir into the averaged speed. Finally, we have derived afamily of lower bounds on the time of evolution between these states, thus obtaining an estimatefor the equilibration time at the local level. These results could be of interest to the subjects ofequilibration, quantum speed limits, and also quantum metrology.

I. INTRODUCTION

Both the subjects of equilibration and thermalizationfall into the very basic foundational questions of sta-tistical mechanics: how to derive the macroscopic lawsof thermodynamics from the microscopic many-particlelaws. To do so, one can consider a quantum system ini-tialized in a well defined pure state and, under some mi-nimal assumptions, still conclude the system behaves asif it were described by an equilibrium ensemble [1]. Thisargument has been made technically rigorous, and alsonumerically confirmed by simulating several interactingquantum many-body systems [2, 3]. Indeed, the problemof equilibration has attracted much interest in the lastdecades from both the theoretical [4] and experimentalcommunities [5–10].

Overall, probing the mechanism of local equilibrationof a closed quantum system requires answering whetherand how some of its nonequilibrium states equilibrate,even if such states do not belong to a Gibbs-like statis-tical ensemble. The isolated system is initialized in apure state and undergoes a unitary evolution governedby the time-independent Hamiltonian H. At the locallevel, the notion of equilibration involves monitoring thenonunitary dynamics of some reduced state of a smallsubregion of the isolated system, and quantifying howfar apart it is from an equilibrium state [11]. In turn, thetask of distinguishing such states can be accomplishedby means of a suitable distance measure on the Hilbertspace, for example the Schatten 1-norm [12, 13]. So far,while there are plenty of rigorous results showing thatequilibration should occur under very general conditionsfor small subsystems, there are a few results about thetime scales involved in such physical process [14].

Here we will study the relative purity as a figure ofmerit for equilibration and show under which conditions

the subsystem equilibrates. The relative purity standsas a versatile information-theoretic quantifier for distin-guishing two quantum states, also being an experimentalfriendly measure since it relies on the overlap of den-sity matrices [15]. We also consider the rate of changeof relative purity as signaling the speed of the fluctua-tions, thus deriving upper bounds on such velocity in asimilar fashion to the well-known discussion of quantumspeed limits (QSLs). Noteworthy, from these inequalitieswe obtain lower bounds on the equilibration time of thesubsystem, thus connecting both the subjects of equili-bration and quantum speed limits.

The paper is organized as follows. In Sec. II we re-view basic concepts on the subject of equilibration, alsointroducing the relative purity as a figure of merit to sig-nal equilibration. In Sec. III we discuss the fluctuationsaround the equilibrium of a bipartite closed quantum sys-tem (S + B), thus analyzing the dynamics of relative pu-rity between a marginal state of subsystem S and somesteady state. We proved a set of lower bounds on theequilibration time that are fully characterized by the ini-tial state and the Hamiltonian of the closed system [seeSecs. III A, III B, III C, III D, and III E]. In Sec. IV weillustrate our findings by means of two paradigmatic spinmodels, namely, the transverse field Ising model and thenonintegrable XXZ model. Finally, in Sec. V we summa-rize our conclusions.

II. RELATIVE PURITY AND EQUILIBRATION

Let us consider a quantum system described by afinite-dimensional Hilbert space H = HS ⊗ HB, withd = dimH, thus being split into a subsystem S of dimen-sion dS = dimHS, and its complement B of dimensiondB = dimHB. The whole system S + B evolves unitarily

arX

iv:2

108.

0593

6v2

[qu

ant-

ph]

27

Nov

202

1

Page 2: arXiv:2108.05936v1 [quant-ph] 12 Aug 2021

2

under the time-independent Hamiltonian

H = HS ⊗ IB + IS ⊗HB +HSB . (1)

The Hamiltonian is chosen to be nondegenerate and dis-plays the spectral decomposition H =

∑nEn|En〉〈En|,

where {|En〉}n=1,...,dE spans an orthonormal eigenbasisrelated to dE distinct energy levels. In addition, we as-sume the energy gaps of the system are nondegenerated,i.e., Ei − Ej 6= Ek − El for i 6= k and j 6= l. It isworth mentioning that such assumptions may be relaxedtowards degenerate systems [16]. The initial state of thesystem is pure, ρ(0) = |Ψ(0)〉〈Ψ(0)|, and thus the instan-taneous state ρ(t) = U(t)ρ(0)U(t)† will remain as a purestate for all times, with U(t) = e−itH . For simplicity,from now on we set ~ = 1. Therefore to have some kindof equilibration we have to consider a subsystem S; therest, B, play the role of a bath. The marginal states ofthe quantum system are given by ρS(B)(t) = TrB(S)(ρ(t)),also written as

ρS(B)(t) =∑j,l

〈Ej |ρ(0)|El〉 e−it(Ej−El) TrB(S)(|Ej〉〈El|) .

(2)For finite dimensional systems, it follows that ρS(t)

never equilibrates since it never stops to evolve; therewill be recurrences. However, it may be very closeto some steady state ωS for most of the time. LetD(x, y) be a suitable information-theoretic distinguisha-bility measure of quantum states. We say subsystemequilibration has taken place at time τ when, for someε > 0, it follows 〈D(ρS(t), ωS)〉T ≤ ε, for all T > τ ,

with 〈h(t)〉T := 1T

∫ T0dt h(t) being the time-average [13].

Note that, if the time average is small, then D(ρS(t), ωS)should be small most of the time, and in this sense wesay the system does equilibrate. We stress that recur-rences will occur but they should be rare, and its timescale should increase with the system size.

If the equilibration process really occurs, the equilib-rium state of the closed system S+B is given by the infi-nite time-averaged state ω := 〈ρ(t)〉∞ = limT→∞〈ρ(t)〉T ,and would be identical to the dephased state1

ω = ∆(ρ(0)) , (3)

where ∆(•) =∑j 〈Ej | • |Ej〉|Ej〉〈Ej | stands for the

fully dephasing operator with respect to the eigenbasisof the Hamiltonian. In this setting, the steady state ofsubsystem S(B) is given by the marginal state ωS(B) =TrB(S)(ω). Remarkably, if one uses the Schatten 1-norm

D(x, y) := 12‖x− y‖1 as a bona fide distance measure

1 Here we are not interested if the equilibrium state is a thermalor Gibbs state, which is necessary to claim that the system ther-malizes. To have thermalization one need further conditions, theeigenstate thermalization hypothesis (ETH) being the most usedone.

M

f(0)

f(τ)

ρS(t)ρS(0)

ωSρS(τ

′)ρS(τ′′)

ρS(τ)

FIG. 1. (Color online) Depiction of the physical setting. Theinitial state ρS(0) undergoes a local nonunitary evolution go-verned by the Hamiltonian H [see Eq. (2)]. The evolved stateρS(t) follows a path (blue dashed curve) over the manifoldof marginal states M (gray surface). In practice, this statenever equilibrates since it keeps evolving nonunitarily, andfor instance can reach states ρS(τ), ρS(τ ′), or ρS(τ ′′) that arearbitrarily close to the equilibrium state ωS. The relative pu-rity f(τ) (black dotted curve) captures the distinguishabilitybetween ρS(τ) and the equilibrium state ωS of the subsys-tem, while f(0) (red solid curve) signals how distinguishablethe initial state ρS(0) is compared to such equilibrium state.Note that, while not being a distance measure in the formalsense, the relative purity f(t) provides insightful informationabout how far apart is the subsystem S from ωS.

over the space of quantum states, it has been proved that

limT→∞

〈D(ρS(t), ωS)〉T ≤1

2

√d2

S

deff(ω), (4)

where deff(ω) := 1/Tr(ω2) is the so-called effective di-mension [12, 17]. More precisely, the larger the effectivedimension deff(ω) compared to the subsystem dimensiondS, the closer the system to some steady state. Notethat the effective dimension measures the number of en-ergy eigenstates that contribute to the superposition ofthe initial state. It can be argued that for many-particlesystems with local interactions this is typically the case,since the distance between the energy levels becomes ex-ponentially small and it is very hard to prepare an initialstate with only a few levels [17].

The aforementioned criterion for equilibration is basedon the closeness of states ρS(t) and ωS measured by theSchatten 1-norm, i.e., a geometric distance that signalsthe distinguishability between two quantum states. How-ever, we emphasize that there are several ways to char-acterize such a distance [18]. Indeed, the convex space ofquantum states is equipped with a plethora of bona fidedistances [19–22], and this nonuniqueness has been of rel-evance for several investigations in quantum informationprocessing, quantum thermodynamics [23, 24], quantumspeed limits [25, 26], and quantum metrology [27, 28], toname a few.

Page 3: arXiv:2108.05936v1 [quant-ph] 12 Aug 2021

3

Here we will consider the relative purity f(ρ, %) :=Tr(ρ%) as a natural distinguishability measure of twoquantum states [29]. Importantly, such an information-theoretic quantifier has been useful in the study of quan-tum speed limits [30–33], information scrambling andLoschmidt echoes [34, 35], and also for probing quantumcoherence from photonic metrological setups [36]. Whilenot being a distance in the stringent sense, the relati-ve purity is symmetric, f(ρ, %) = f(%, ρ), non-negative,f(ρ, %) ≥ 0, for all states ρ and %, and vanishes forthe case in which the states are maximally distingui-shable. Moreover, for ρ = % it recovers the quantumpurity: f(ρ, ρ) = Tr(ρ2) ≤ 1. Noteworthy, for pure statesρ = |ψ〉〈ψ| and % = |φ〉〈φ|, relative purity reduces to thepairwise fidelity, f(ρ, %) := |〈ψ|φ〉|2, which in turn wasalready investigated for understanding the equilibrationof isolated thermodynamic systems [37].

Overall, while one can have equilibration at the locallevel, the whole system never equilibrates since it evolvesunitarily. In fact, in Appendix A we verify that therelative purity of states ρ(t) and ω remains as a con-stant of motion of the dynamics, in turn being equalto the quantum fidelity of these states2. Hence, as forthe Schatten 1-norm, we consider the relative purityf(t) := TrS(ωS ρS(t)) of subsystem S [see Fig. 1]. FromEqs. (2) and (3) it is straightforward to show that

f(t) =∑j,l

〈Ej |ρ(0)|El〉 e−it(Ej−El) 〈El|(ωS ⊗ IB)|Ej〉 .

(5)For an initial nonequilibrium state, i.e., when states ρS(t)and ωS have nonoverlapping supports, one should expectthe relative purity to take small values at the earlier timesof the dynamics, while it approaches the infinite time-averaged value 〈f(t)〉∞ = TrS(ω2

S) as the system evolvesin time and equilibrates. In turn, the latter is nothing butthe quantum purity of the steady state ωS of subsystemS. In this way, we now introduce the following figure ofmerit for equilibration

g(t) := |f(t)− 〈f(t)〉∞|2 . (6)

Hence, the closer the system is to a given steady state,the smaller the figure of merit in Eq. (6), i.e., g(t) ≈ 0for t > τ , with τ being the equilibration time. Similarlyto what happens to the trace distance, it is possible toupper bound the time average of g(t). In Appendix B,we proved that the time average of the figure of merit inEq. (6) is upper bounded as

〈g(t)〉∞ ≤‖ωS‖2∞deff(ω)

. (7)

2 Instead of looking at the state of the system, which can onlyequilibrate locally, we can look at some observables and also showthey equilibrate under some general conditions. Note that evenglobal observables, as the total magnetization, can equilibrateand in this sense one can have global equilibration.

Equation (7) is one of the main results of the pa-per. It shows that the effective dimension plays a fun-damental role on the equilibration process when it ismonitored by the relative purity. Importantly, this re-sult somehow agrees with the aforementioned case inwhich the trace distance is the distinguishability mea-sure. The subsystem S approaches the equilibriumwhenever the effective dimension deff(ω) of the globalsteady state is much larger than the operator norm ofthe dephased marginal state ωS. In fact, given that‖ωS‖∞ = λmax(ωS) ≤ 1, where λmax(•) sets the maxi-mum eigenvalue of the density matrix, it is reasonableto expect that ‖ωS‖2∞/deff(ω) � 1 since the effective di-mension of the equilibrium state ω typically takes largevalues. In addition, note that ‖ωS‖2∞ ≤ ‖ωS‖22, with‖ωS‖22 = TrS(ω2

S) = 1/deff(ωS), and thus the bound canbe recast as 〈g(t)〉∞ ≤ 1/[deff(ωS) deff(ω)]. It is worthmentioning that, regardless of its simplicity, we point outthe latter inequality might be less tight than the boundin Eq. (7).

III. DYNAMICS OF RELATIVE PURITY

How fast does a given quantum system fluctuatearound the equilibrium? So far, this problem has beenaddressed in a few works showing that such speed wouldbe extremely small for almost all times in typical ther-modynamic cases. Indeed, it can be proved that theinfinity time-averaged speed of state ρS(t) quantifiedby the Schatten 1-norm reads as 〈‖dρS(t)/dt‖1〉∞ ≤2 ‖HS ⊗ IB +HSB‖∞

√d3

S/deff(ω) [38], thus depending

on the effective dimension and the size of the interactingterm HSB [see Eq. (1)]. Furthermore, it has been shownthat the speed of fluctuations around the equilibrium canbe signaled by means of quantum purity, also unveilingthe role of correlations between system and environmentin the equilibration process [39].

Despite these remarkable theoretical achievements,much less is known about the time scales involved inthe equilibration process [14, 40, 41]. In this section wewill investigate the dynamical behavior of relative purity[see Eq. (5)], thus bounding its rate of change in termsof fundamental quantities such as the initial state andthe Hamiltonian of the system. Furthermore, we providebounds on the equilibration time τ in a similar fashionto the framework of quantum speed limits, i.e., the verybasic question of how fast a quantum system evolves be-tween two states.

Here we focus on the time derivative of relative purityas probing the speed of fluctuations around the equilib-rium. We shall begin noticing that, since the dynamicsof the subsystem S is fully encoded in the differentialequation dρS(t)/dt = iTrB([ρ(t), H]), with the Hamil-tonian H defined in Eq. (1), the absolute value of the

Page 4: arXiv:2108.05936v1 [quant-ph] 12 Aug 2021

4

time-derivative of relative purity becomes∣∣∣∣ ddtf(t)

∣∣∣∣ = |iTrSB((ωS ⊗ IB)[ρ(t), H])| . (8)

Importantly, from Eq. (8) we are able to derive boundson the rate of change of purity, which signals the fluctua-tions of ρS(t) around the fixed point ωS of the dynamics.From this we also derive bounds for the time in whichthe system approaches the equilibrium. In the followingwe will discuss in detail such issues.

A. Relative purity and Schatten 2-norm

Here we will show that the speed in Eq. (8) satisfies anupper bound that is related to the Schatten 2-norm andthe variance of the Hamiltonian H. Let |Tr(A1A2)| ≤‖A1‖2‖A2‖2 be the Cauchy-Schwarz inequality for ope-rators {Aj}j=1,2. In this case, Eq. (8) gives rise to thefollowing inequality∣∣∣∣ ddtf(t)

∣∣∣∣ ≤ dB ‖ωS‖2 ‖[ρ(t), H]‖2 , (9)

where we have also used that ‖ωS ⊗ IB‖2 = dB‖ωS‖2.Since the time-independent Hamiltonian H commuteswith the evolution operator U(t) = e−itH , it is straight-forward to conclude that [ρ(t), H] = U(t)[ρ(0), H]U(t)†.Hence, due to the unitary invariance of the Schatten 2-norm, it follows that ‖[ρ(t), H]‖2 = ‖[ρ(0), H]‖2, and thetime average of Eq. (9) over the interval t ∈ [0, τ ] thusbecomes⟨∣∣∣∣ ddtf(t)

∣∣∣∣⟩τ

≤ 2 dB ‖ωS‖2√IL(ρ(0), H) , (10)

where

IL(ρ(0), H) := −1

4Tr([ρ(0), H]2) =

1

4‖[ρ(0), H]‖22 .

(11)We point out that the right-hand side of Eq. (10) is timeindependent and depends on the Hamiltonian, the initialstate of system S + B, the steady state ωS, and the di-mension of subsystem B that plays the role of a bath.Importantly, the quantifier IL has been already intro-duced in the context of quantum coherence characteri-zation, in turn defining a lower bound on the so-calledWigner-Yanase skew information [42]. In this sense, themore commuting both the Hamiltonian and the initialstate ρ(0), the smaller the fluctuations on the speed〈|df(t)/dt|〉τ .

In particular, for the pure state ρ(0) = |Ψ(0)〉〈Ψ(0)|of the global system S + B, the quantity IL reduces fur-ther to half of the variance of H, i.e., IL(ρ(0), H) =(1/2)(∆H)2 := (1/2)(〈Ψ(0)|H2|Ψ(0)〉−〈Ψ(0)|H|Ψ(0)〉2).Next, applying the inequality

∫dx|g(x)| ≥ |

∫dxg(x)|

into Eq. (10), one obtains the lower bound τ ≥ τ (1), with

τ (1) :=|TrS(ωS ρS(τ))− TrS(ωS ρS(0))|√

2 dB‖ωS‖2 ∆H. (12)

Noteworthy, the bound in Eq. (12) fits into theMandelstam-Tamm class of quantum speed limit timesfor closed systems, i.e., the minimum evolution time isinversely proportional to the variance of the generatorH [26, 43, 44]. If the system equilibrates at time τeq suchas the relative purity collapses into the purity of the de-phased state, i.e., TrS(ωS ρS(τeq)) ≈ TrS(ω2

S), the lowerbound in Eq. (12) yields the estimation for the equilibra-

tion time as τeq ≥ τ (1)eq , where

τ (1)eq :=

‖ωS‖2√2 dB ∆H

∣∣∣∣1− TrS(ωS ρS(0))

TrS(ω2S)

∣∣∣∣ . (13)

In particular, when ρS(0) and ωS are maximallydistinguishable states, the orthogonality conditionTrS(ωS ρS(0)) = 0 implies the equilibration time will re-

duce to the case τ(1)eq ≈ ‖ωS‖2/(

√2 dB ∆H), which will

be smaller the higher the dimension of the subsystem B.

B. Relative purity and `1-norm of coherence

Now we will present a bound on the speed in Eq. (8)that is related to the Schatten 1-norm. Let ω be thesteady state of system S+B that is written in terms of theeigenbasis of the Hamiltonian [see Eq. (3)]. In this case,since ω and H are commuting operators, i.e., [ω,H] = 0,we thus have that [ρ(t), H] = U(t)[ρ(0) − ω,H]U(t)†,where we have used the fact that U(t)†HU(t) = H, andalso that U(t)†ωU(t) = ω is a fixed point of the unitarydynamics. Hence, inserting this result into Eq. (8) andtaking its time average over the interval t ∈ [0, τ ], onegets⟨∣∣∣∣ ddtf(t)

∣∣∣∣⟩τ

≤ 2 ‖ωS‖∞‖H‖∞‖ρ(0)− ω‖1 . (14)

where we have invoked the inequality |Tr(A1[A2, A3])| ≤‖A1‖∞‖[A2, A3]‖1 ≤ 2 ‖A1‖∞‖A2‖1‖A3‖∞ [45, 46], andemployed the unitary invariance of the Schatten 1-norm,‖[ρ(t)− ω,H]‖1 = ‖[ρ(0)− ω,H]‖1, also using the iden-tity ‖ωS ⊗ IB‖∞ = ‖ωS‖∞.

Equation (14) means that the speed of fluctuationsaround the equilibrium is upper bounded by the productof maximum eigenvalues of both the Hamiltonian H andthe steady state ωS. Importantly, the bound depends onthe coherences of the initial state of the system. In fact,since the dephased state ω is a fully diagonal matrix thatcomprises the populations of ρ(0) in the energy eigenba-sis of H, the Schatten 1-norm ‖ρ(0)− ω‖1 plays the roleof the `1-norm of coherence of ρ(0) respective to sucheigenbasis, thus quantifying how far apart it is from theincoherent state ω [47]. Overall, the more incoherent the

Page 5: arXiv:2108.05936v1 [quant-ph] 12 Aug 2021

5

initial state with respect to the steady state, the smallerthe speed of the fluctuations.

Next, applying the inequality∫dx|g(x)| ≥ |

∫dxg(x)|

into Eq. (14), one gets the lower bound τ ≥ τ (2), with

τ (2) :=|TrS(ωS ρS(τ))− TrS(ωS ρS(0))|

2 ‖ωS‖∞‖H‖∞‖ρ(0)− ω‖1. (15)

Suppose the system equilibrates at time τeq, with the rel-ative purity TrS(ωS ρS(τeq)) ≈ TrS(ω2

S) = ‖ωS‖2 recover-ing the purity of the steady state. In this case, given that‖ωS‖2 ≥ ‖ωS‖∞, the lower bound in Eq. (15) implies that

τeq ≥ τ (2)eq , with

τ (2)eq :=

1

2 ‖H‖∞‖ρ(0)− ω‖1

∣∣∣∣1− TrS(ωS ρS(0))

TrS(ω2S)

∣∣∣∣ , (16)

We stress that, for the case of two states overlapping

to zero as TrS(ωS ρS(0)) = 0, Eq. (16) reduces to τ(2)eq ≈

1/(2 ‖H‖∞‖ρ(0)− ω‖1). In words, the more coherent thestate ρ(0) in the eigenbasis of H, the smaller would be

τ(2)eq , thus showing that quantum coherence of the probe

state plays a role on the equilibration time.

C. Relative purity and Quantum FisherInformation

We shall point out that one may arrive at a slightlydifferent upper bound on the speed of fluctuations byexploiting another set of inequalities. To proceed, byinvoking both the inequality |Tr(A1A2)| ≤ ‖A1‖∞‖A2‖1and the unitary invariance ‖[ρ(t), H]‖1 = ‖[ρ(0), H]‖1 ofthe Schatten 1-norm, Eq. (8) is written as∣∣∣∣ ddtf(t)

∣∣∣∣ ≤ ‖ωS‖∞‖[ρ(0), H]‖1 . (17)

Interestingly, the right-hand side of Eq. (17) can be upperbounded via the inequality ‖[%,H]‖21 ≤ 4FQ(%,H) [48,49], where FQ(%,H) is the so-called quantum Fisher in-formation (QFI) and reads as

FQ(%,H) =1

2

∑k,l

(λk − λl)2

λk + λl|〈k|H|l〉|2 , (18)

where {λl, |l〉}l are the eigenvalues and eigenvectors ofsome mixed state %. Hence, by inserting such boundinto Eq. (17), and also taking the time average over theinterval t ∈ [0, τ ], it yields⟨∣∣∣∣ ddtf(t)

∣∣∣∣⟩τ

≤ 2 ‖ωS‖∞√FQ(ρ(0), H) . (19)

Equation (19) means that the speed of fluctuationsaround the equilibrium is upper bounded by the QFI,a paradigmatic quantity in quantum metrology that iswidely applied for enhance phase estimation [28, 50, 51].

In particular, for the pure state ρ(0) = |Ψ(0)〉〈Ψ(0)|,QFI reduces further to the variance of the generatorH, i.e., FQ(ρ(0), H) = (∆H)2 := 〈Ψ(0)|H2|Ψ(0)〉 −〈Ψ(0)|H|Ψ(0)〉2 [52]. It can be proved that Eq. (19) im-plies the lower bound τ ≥ τ (3), where

τ (3) :=|TrS(ωS ρS(τ))− TrS(ωS ρS(0))|

2 ‖ωS‖∞√FQ(ρ(0), H)

, (20)

where we used that∫dx|g(x)| ≥ |

∫dxg(x)|. At the equi-

librium, using that TrS(ωS ρS(τeq)) ≈ TrS(ω2S) = ‖ωS‖2,

and ‖ωS‖2 ≥ ‖ωS‖∞, the lower bound in Eq. (20) givesrise to the following time scale for equilibration as τeq ≥τ

(3)eq , where we define

τ (3)eq :=

1

2∆H

∣∣∣∣1− TrS(ωS ρS(0))

TrS(ω2S)

∣∣∣∣ , (21)

which in turn will reduce to the simplest case τ(3)eq ≈

1/∆H for two maximally distinguishable states ρ(0) andωS.

D. Relative purity and mutual information

Lastly, we obtain an upper bound that depends on thecorrelations between S and B. In order to do so, we willintroduce the traceless operator

ρ(t) := ρ(t)− ωS ⊗ ωB , (22)

which in turn leads to an infinity time-averaged oper-ator 〈ρ(t)〉∞ = ω − ωS ⊗ ωB that is a fully correlatedstate. The operator in Eq. (22) also implies the trace-less marginal states ρS(t) = TrB(ρ(t)) = ρS(t) − ωS andρB(t) = TrS(ρ(t)) = ρB(t) − ωB, which are also zero-valued operators under the infinity time average, i.e.,〈ρS(t)〉∞ = 〈ρB(t)〉∞ = 0. Next, by exploiting the cyclicproperty of the trace, Tr(A1[A2, A3]) = Tr([A1, A2]A3),it is straightforward to show that TrSB((ωS ⊗ IB)[ωS ⊗ωB, H]) = 0, which immediately implies the identity

TrSB((ωS ⊗ IB)[ρ(t), H]) = TrSB((ωS ⊗ IB)[ρ(t), H]) .(23)

Inserting Eq. (23) into Eq. (8), also noting thatTrSB((ωS ⊗ IB)[ρ(t), IS ⊗HB]) = 0, one obtains∣∣∣∣ ddtf(t)

∣∣∣∣ = |iTrSB((ωS ⊗ IB)[ρ(t), HS ⊗ IB +HSB])|

≤ 2 ‖ωS‖∞‖ρ(t)‖1‖HS ⊗ IB +HSB‖∞ , (24)

where we have used the inequalities |Tr(A1[A2, A3])| ≤‖A1‖∞‖[A2, A3]‖1, ‖[A2, A3]‖1 ≤ 2 ‖A2‖1‖A3‖∞ [45, 46],and the fact that ‖ωS ⊗ IB‖∞ = ‖ωS‖∞. We pointout that, from Pinsker’s inequality, the trace norm ofthe traceless operator ρ(t) in Eq. (22) is upper boundedas [53]

‖ρ(t)‖1 ≤√

2S(ρ(t)‖ωS ⊗ ωB) , (25)

Page 6: arXiv:2108.05936v1 [quant-ph] 12 Aug 2021

6

with the relative entropy defined as S(x‖y) = −S(x) −TrSB(x ln y), and S(x) = −TrSB(x lnx) being the vonNeumann entropy. In particular, it can be proved thatthe relative entropy is written as S(ρ(t)‖ωS ⊗ ωB) =ISB(ρ(t)) + S(ρS(t)‖ωS) + S(ρB(t)‖ωB), and thus it de-pends on the correlations of the system measured by themutual information, ISB(ρ(t)) := S(ρS(t)) + S(ρB(t)) −

S(ρ(t)). This also means the relative entropy depends onthe distance S(ρS,B(t)‖ωS,B) between the marginal statesρS,B(t) and ωS,B, thus assigning a geometric perspectiveto the bound in Eq. (25). Inserting Eq. (25) into Eq. (24)and taking the time average over the interval t ∈ [0, τ ]yields

⟨∣∣∣∣ ddtf(t)

∣∣∣∣⟩τ

≤ 2√

2 ‖ωS‖∞‖HS ⊗ IB +HSB‖∞√〈S(ρ(t)‖ωS ⊗ ωB)〉τ , (26)

where we have exploited the concavity of the square-rootfunction.

Importantly, Eq. (26) means that the speed of fluctu-ations is upper bounded by the relative entropy whichdistinguishes the instantaneous state of the whole sys-tem from its uncorrelated steady state, also being a func-tion of the maximum eigenvalue of the marginal dephasedstate ωS and the operator norm of HS⊗IB +HSB. In thisregard, it is worth noting that if the interacting Hamil-tonian HSB couples the system to only a few degrees offreedom of the bath, which is typically the case of spinmodels with nearest-neighbor couplings, thus the upperbound in Eq. (26) will be mostly independent of the sizeof subsystem B. In Appendix C we discussed a simi-lar bound to the speed of fluctuations for the quantumpurity for subsystem S.

In the limit τ → ∞, the infinity time average ofthe relative entropy is written as 〈S(ρ(t)‖ωS ⊗ ωB)〉∞ =S(ωS) + S(ωB) = 2S(ωS). In detail, this comes fromthe fact that the von Neumann entropy S(ρ(t)) =S(ρ(0)) = 0 remains unchanged for a pure state ρ(t) =U(t)ρ(0)U†(t) evolving unitarily, and also that the twodephased marginal states ωS,B of the bipartite systemstore the same amount of information, i.e., S(ωS) =S(ωB). In this case, one readily gets

⟨∣∣∣∣ ddtf(t)

∣∣∣∣⟩∞≤ 4 ‖ωS‖∞‖HS ⊗ IB +HSB‖∞

√ln dS ,

(27)where we have used that 0 ≤ S(ωS) ≤ ln dS.

Next, we can make use of Eq. (26) to bound the timeevolution τ of the closed quantum system. Indeed, weobtain the bound τ ≥ τ (4), where

τ (4) :=(1/2√

2) |TrS(ωS ρS(τ))− TrS(ωS ρS(0))|‖ωS‖∞‖HS ⊗ IB +HSB‖∞

√〈S(ρ(t)‖ωS ⊗ ωB)〉τ

,

(28)where we have applied the inequality |

∫dxg(x)| ≤∫

dx|g(x)|. If the system approaches the equilibrium attime τeq, with TrS(ωS ρS(τeq)) ≈ TrS(ω2

S) = ‖ωS‖2, it

follows that τeq ≥ τ (4)eq , with

τ (4)eq :=

12√

2

∣∣∣1− TrS(ωS ρS(0))TrS(ω2

S)

∣∣∣‖HS ⊗ IB +HSB‖∞

√ln[1/λmin(ωS ⊗ ωB)]

.

(29)where we used that ‖ωS‖2 ≥ ‖ωS‖∞, and also invokedthe inequality S(ρ(t)‖ωS⊗ωB) ≤ ln[1/λmin(ωS⊗ωB)] [54,55], with λmin(•) setting the minimum eigenvalue of thedensity matrix.

E. Discussion

In the previous sections, we presented a set of lowerbounds on the speed of evolution and the equilibrationtime for the subsystem S. The relative purity signalsthe distinguishability between the reduced state ρS(t)and the steady state ωS, thus indicating how far apartthey are in the sense of witnessing whether both thestates have zero or nonzero overlapping supports. FromEqs (10), and (14), we see that the average speed of thefluctuations depends on the coherences of the pure initialstate ρ(0) of the bipartite system.

On the one hand, the more commuting ρ(0) and H,the smaller the fluctuations on the speed [see Eq. (10)].However, this bound seems to be looser since it dependson the dimension dB of the subsystem B, which in turncan be large. On the other hand, the bound in Eq. (14)shows that the fluctuations on the speed are constrainedto the quantum fluctuations ofH captured by its varianceregarding ρ(0). Noteworthy, this bound sounds more ap-pealing since it grows with the maximum eigenvalue ofthe steady state, while being of interest for metrologicalpurposes due to its connection with the quantum Fisherinformation. From Eq. (19), note that the fluctuationson the speed will decrease as ρ(0) approaches the equilib-rium state ω, which in turn is a fully incoherent state intothe eigenbasis of H. Opposite to these results, Eq. (26)depends on the correlations of the bipartite system viathe existing link between relative entropy and mutual in-formation. It is worth noting that its right-hand side is afunction of time τ , while the previous upper bounds arefully time independent.

Page 7: arXiv:2108.05936v1 [quant-ph] 12 Aug 2021

7

The speed of the relative purity captures the notion ofhow fast some nonequilibrium state of subsystem S ap-proaches its steady state under the local nonunitary dy-namics, somehow giving the information of the QSL to-wards the equilibration. In turn, such set of speeds limitsimplies a family of lower bounds on the time of evolutionbetween these states. Indeed, from Eqs. (12), (15), (20)and (28), the minimum time of evolution displays theQSL time as

τQSL := max{τ (1), τ (2), τ (3), τ (4)} . (30)

The QSL time is a quantity that fully characterizes thedynamics of the set of eigenstates of the Hamiltoniangoverning the dynamics. Note that τQSL ≡ τQSL(τ) isa time-dependent quantity, which is expected since weare comparing states ρS(τ) and ωS. We point out thatmost of the bounds on the time evolution of differentphysical systems that have appeared in the literature ad-dress time-dependent QSLs (see, for example, Ref. [44]and references therein). This “caveat” in the QSLs isnot well discussed in the literature, and we are followingthe aforementioned standard procedure. From the QSLtime, fixing ρS(τ) ≈ ωS , we also obtain a time scale forequilibration at the local level (this one time indepen-dent). Hence, from Eqs. (13), (16), (21) and (29), it ispossible to concatenate the previous results into a unifiedestimation for the equilibration time yields

τeq := max{τ (1)eq , τ

(2)eq , τ

(3)eq , τ

(4)eq } . (31)

We point out that the bounds in Eqs. (13) and (21) areinversely proportional to the variance of the HamiltonianH, thus resembling the well-known class of QSLs a laMandelstam-Tamm [44].

IV. EXAMPLES

In the following we will illustrate our findings by focus-ing on two prototypical quantum many-body systems.The first is the transverse field Ising model with localfields:

HIsing = J

L−1∑j=1

σxj σxj+1 +

L∑j=1

(hxσ

xj + hzσ

zj

), (32)

with parameters J = 1, hx = 0.5, hz = −1.05 [40]. Theequilibration properties of this model have already beennumerically investigated, particularly identifying initialstates and sets of parameters for which equilibration oc-curs rapidly [56], or even never takes place [57]. The se-cond is the nonintegrable XXZ model with next-nearest-neighbor hopping:

HXXZ = J

L−1∑j=1

(σxj σ

xj+1 + σyj σ

yj+1

)+ U

L∑j=1

σzjσzj+1

+ Jnnn

L−2∑j=1

(σxj σ

zj+1σ

xj+2 + σyj σ

zj+1σ

yj+2

), (33)

0

0.2

0.4

0.6

0.8

1

0 2 4 6 8 10

0

0.2

0.4

0.6

0.8

1

0 2 4 6 8 10

0

0.02

0.04

0 1 2 3 4 5

g(t)/g(0)

t

(a)

〈g(t)〉 τ

/g(0)

τ

L = 4 L = 6 L = 8 L = 10

(b)

δ τ

τ

FIG. 2. (Color online) Plot of the figure of merit for the Isingmodel, with parameters J = 1, hx = 0.5, hz = −1.05 [seeEq. (32)]. Here, we set the system sizes L = {4, 6, 8, 10},with open boundary conditions, while LS = 1, and LB ={3, 5, 7, 9}. The system is initialized at the charge-density-wave-like state |Ψ(0)〉 = |1, 0, 1, 0, . . . , 0, 1〉, with |0〉 and |1〉denoting the spin-up and -down state, respectively.

where we set the input configuration J = 1, U = 2,and Jnnn = 0.2 [40]. The two spin models are initial-ized in a charge-density-wave-like state, i.e., |Ψ(0)〉 =|1, 0, 1, 0, . . . , 0, 1〉, with |0〉 and |1〉 denoting the spin-upand -down state, respectively. Here we will investigatethe role played by the figure of merit g(t) [see Eq. (6)]for signaling the equilibration process in both many-bodyquantum systems. We will also discuss the tightness ofthe bound in Eq. (7) by introducing the relative error

δτ :=‖ωS‖2∞deff(ω)

− 〈g(t)〉τ . (34)

Overall, the smaller the relative error, the tighter thebound on the fluctuations of the relative purity capturedby the function g(t). In general, it is reasonable to expectthat limτ→∞ δτ ≈ 0 as we increase the system size Lof the system. From now on we set the system sizesL = {4, 6, 8, 10}, where LS = 1, and LB = {3, 5, 7, 9}.

In Fig. 2, we show plots of the figure of merit g(t) [seeEq. (6)] for the Ising model with open boundary condi-tions. In Fig. 2(a), we plot the normalized time signalg(t)/g(0) as a function of time. Noteworthy, the recur-rences exhibited in the signal are mostly suppressed aswe increase the system size L. In other words, we ex-pect that the fluctuations tend to decrease in the limit

Page 8: arXiv:2108.05936v1 [quant-ph] 12 Aug 2021

8

0

0.2

0.4

0.6

0.8

1

0 1 2 3 4 5 6

0

0.2

0.4

0.6

0.8

1

0 1 2 3 4 5 6

0

0.05

0.1

0 1 2 3 4 5

g(t)/g(0)

t

(a)〈g(t)〉 τ

/g(0)

τ

L = 4 L = 6 L = 8 L = 10

(b)

δ τ

τ

FIG. 3. (Color online) Plot of the normalized figure of meritfor the non-integrable XXZ model, with parameters J = 1,U = 2, Jnnn = 0.2 [see Eq. (33)]. Here, we set the systemsizes L = {4, 6, 8, 10}, with open boundary conditions, whileLS = 1, and LB = {3, 5, 7, 9}. The system is initialized at thecharge-density-wave-like state |Ψ(0)〉 = |1, 0, 1, 0, . . . , 0, 1〉,with |0〉 and |1〉 denoting the spin-up and -down state, re-spectively.

of larger system sizes. In Fig. 2(b), we plot the finitetime average of the figure of merit, 〈g(t)〉τ/g(0). Next, inFig. 3 we show our results for the XXZ model with openboundary conditions. In Fig. 3(a), we show the plotsof the normalized time signal g(t)/g(0), while Fig. 3(b)show the plot of 〈g(t)〉τ/g(0). The results are quite sim-ilar to the case of the Ising model. Overall, note the sizeof fluctuations in g(t) decreases as we increase the systemsize L, thus signaling the system equilibrates.

The insets in Figs. 2(b) and 3(b) show the plot of therelative error δτ [see Eq. (34)] as a function of time τ .In agreement with Eq. (7), the relative error satisfies thecondition δτ ≥ 0 for all τ ≥ 0. In addition, the am-plitude of the relative error decreases as the system sizeincreases. We see that each of the plots saturates at fixedvalues for all times. To see this in detail, we first notethat 〈g(t)〉τ is a time-dependent function, while the quan-tity ‖ωS‖2∞/deff(ω) is time independent and stands as aconstant for a given system size L. We point out that thetime-averaged quantity 〈g(t)〉τ is smaller than the ratio‖ωS‖2∞/deff(ω) by some orders of magnitude, for all τ ≥ 0and system size L. In other words, the time oscillationsof 〈g(t)〉τ are negligible when compared to the constantvalue of ‖ωS‖2∞/deff(ω).

0

0.005

0.01

0.015

0 2 4 6 8 10

0

0.002

0.004

0.006

0 2 4 6 8 10

τ QSL

τ

(a)

τ QSL

τL = 4 L = 6 L = 8

(b)

FIG. 4. (Color online) Plot of the QSL time τQSL [see Eq. (30)]for (a) the Ising model, with parameters J = 1, hx = 0.5, hz =−1.05 [see Eq. (32)], and (b) the nonintegrable XXZ model,with parameters J = 1, U = 2, Jnnn = 0.2 [see Eq. (33)].Here, we set the system sizes L = {4, 6, 8}, with open bounda-ry conditions, while LS = 1, and LB = {3, 5, 7}. The systemis initialized at the charge-density-wave-like state |Ψ(0)〉 =|1, 0, 1, 0, . . . , 0, 1〉, with |0〉 and |1〉 denoting the spin-up and-down state, respectively.

In Fig. 4, we plot the QSL time in Eq. (30) for both theIsing model [see Fig. 4(a)] and XXZ model [see Fig. 4(b)].Note that the QSL time exhibits nonperiodic oscilla-tions the amplitudes of which are suppressed as we in-crease the system size. From Figs. 4(a) and 4(b), wesee that the larger the system size L, the smaller theamplitude of the QSL time. We find that, regardless ofthe time-dependent behavior of τ (1), τ (2), and τ (4), itfollows that τ (3) dominates the numerical maximizationindicated in Eq. (30). We refer to Appendix D for moredetails on the set of QSL times. Note that τ (3) is in-versely proportional to the variance ∆H [see Eq. (20)].Given the initial pure state |Ψ(0)〉 = |1, 0, 1, 0, . . . , 0, 1〉,the variance of the Ising Hamiltonian in Eq. (32) read

as ∆HIsing =√

(L− 1)J2 + Lh2x, while one gets the

variance ∆HXXZ = 2J√L− 1 for the XXZ model in

Eq. (33) with respect to such initial state. For largersystem size L, these variances become ∆H ∼ L1/2, andin this case the QSL time will behave as τQSL = τ (3) ∼(1/2)L−1/2 ‖ωS‖−1

∞ |TrS(ωS ρS(τ))− TrS(ωS ρS(0))|. Weexpect this result should hold in the limit of larger va-lues of L, but it can already be seen in Fig. 4 that the

Page 9: arXiv:2108.05936v1 [quant-ph] 12 Aug 2021

9

amplitude of the QSL time decreases as the system sizeL grows.

Next, we comment on the equilibration time τeq

in Eq. (31). For both the aforementioned spin sys-

tems, we find that τeq = τ(3)eq = (1/2)(∆H)−1|1 −

TrS(ωS ρS(0))[TrS(ω2S)]−1|. For the transverse field Ising

model we find the equilibration time τeq ≈ 9 × 10−3

(L = 4 and 6), and τeq ≈ 10−3 (L = 8). For the non-integrable XXZ model, the equilibration time reads asτeq ≈ 3 × 10−3 (L = 4 and 6), and τeq ≈ 4 × 10−4

(L = 8). In both cases, τeq has units of ~/J . In thelimit of large system size L, since the variances behaveas ∆H ∼ L1/2, thus Eqs. (21) and (31) imply that

τeq = τ(3)eq ∼ (1/2)L−1/2

∣∣1− TrS(ωS ρS(0))[TrS(ω2S)]−1

∣∣.Hence, again we expect the lower bound on the equili-bration time to scale with 1/

√L for large L, but already

see evidence of such decay for the small values of L weused.

We close this section discussing the equilibration timeand the lower bound obtained on both spin systems. InFigs. 2 and 3, we see that the system starts to equilibrateat a time τ such that the figure of merit g(τ) approaches azero value. On the one hand, for the transverse field Isingmodel, Fig. 2(a) shows that this time is of order τ ≈ 1for most of the system sizes, while one gets τeq ≈ 10−3.On the other hand, for the nonintegrable XXZ model,Fig. 2(b) shows that τ ≈ 0.5, but we have found theearlier times 10−4 . τeq . 10−3 for the referred systemsizes. In both cases, we see that the bound is fulfilled(τ ≥ τeq), but is not tight for these two spin models andthe initial state we choose. While the bound may be tightfor other models for larger L, we offer some reasons forit not being tight.

The QSL bound is obtained bounding from above thetime derivative of the relative purity f(t); we are tryingto obtain the minimum time for a change in f(t) assum-ing it always varies at its maximum rate [see Sec. III]. Inthis sense, the bound is expected to not be tight if f(t)strongly oscillates. In fact, the tightness of the QSL isrelated to the distinguishability measure between quan-tum states. Tighter QSL bounds have been discussedfor information-theoretic quantifiers such as the quan-tum Fisher information [58, 59], Wigner-Yanase skew in-formation [26], and also geometric measures based on theBures angle [60]. Furthermore, we have invoked severalinequalities to derive those lower bounds. In spite of sim-plifying the calculations, applying such inequalities mayhave compromised the tightness of the bounds.

Finally, although the bound does not provide quantita-tive information about the equilibration time for the spinmodels and initial state used as examples, they still pro-vide insight into the physical properties involved in theequilibration process [see Sec. III E]. They are also of in-terest, since they allow one to connect both the subjectsof equilibration and speed limits. Lastly, we also showedthat the relative purity is a useful witness for equilibra-tion, and it has the advantage of being more amenableto analytical calculation and experimental measure than

other figures of merit as the trace distance, for example.We emphasize that the lower bounds on the equilibra-tion time could be tightened by invoking some minimalamount of inequalities, and also verifying other distin-guishability measures. Indeed, this is an issue that wehope to address in further investigations.

V. CONCLUSIONS

In conclusion, we have discussed the local equilibra-tion of closed quantum systems and the speed of fluctu-ations around the equilibrium. We provided a criterionfor witnessing equilibration at the local level by introduc-ing a figure of merit that is rooted in relative purity [seeEq. (7)]. In turn, the latter stands as a distinguishabil-ity measure of quantum states, particularly quantifyingthe overlap between a nonequilibrium state of a smallsubsystem and a given steady state. We show that therelative purity is a useful witness for equilibration, and ithas the advantage of being more amenable to analyticalcalculation and experimental measure than other figuresof merit as the trace distance, for example. We haveproved an upper bound on such figure of merit that de-pends on the effective dimension of the equilibrium stateof the closed system. We find that the larger the effectivedimension, the smaller the size of fluctuations around thesystem. Indeed, this somehow agrees with previous re-sults reported in the literature where the authors haveconsidered the Schatten 1-norm as a bona fide measurefor equilibration.

We have analyzed the dynamics of relative purity andits rate of change as a probe of the speed of fluctua-tions around equilibrium. Indeed, we have proved a setof upper bounds on such averaged speed that dependson the initial state and the Hamiltonian of the isolatedsystem [see Eqs. (10), (14), (19), and (27)]. Overall, thebounds show that such fluctuations depend on the co-herences of the pure initial state regarding the referenceeigenbasis of the Hamiltonian. Furthermore, we showthe averaged speed also depends on the correlations ofthe bipartite system that are quantified via the relativeentropy and mutual information. We find that if the in-teracting Hamiltonian HSB couples the system to only afew degrees of freedom of the bath, which is the case ofspin models with nearest-neighbor couplings, such upperbound does not scale with the size of the subsystem B.

From these speeds we have derived a family of lowerbounds on the time of evolution between such states, thusobtaining an estimate for the equilibration time at the lo-cal level. Importantly, regarding the equilibration time,we follow the same procedure as the one that is com-monly applied in the derivation of quantum speed limits[see Eqs. (13), (16), (21), (29), and (31)]. We have verifiedthe bounds on the equilibration time are not tight for thespin models and initial state used as examples, but theystill provide insight in the physical properties involvedin the equilibration process. Indeed, some of the bounds

Page 10: arXiv:2108.05936v1 [quant-ph] 12 Aug 2021

10

fit into the Mandelstam-Tamm class of QSLs due to itsdependence on the inverse of the variance of the Hamil-tonian. Hence, our results somehow may bridge both thesubjects of QSLs and equilibration. Finally, we believethat our results may find applications in the study ofequilibration of many-body quantum systems and quan-tum speed limits, also being useful for discussing the en-hancing of phase estimation in quantum systems aroundequilibrium that is of interest to quantum metrology.

ACKNOWLEDGMENTS

This work is supported by the Brazilian National In-stitute of Science and Technology for Quantum Informa-tion (INCT-IQ) Grant No. 465469/2014-0, and by theAir Force Office of Scientific Research under Grant No.FA9550-19-1-0361.

APPENDIX

A. RELATIVE PURITY AND UHLMANNFIDELITY

In this Appendix we will show that both the relativepurity and Uhlmann fidelity stand as constants of mo-tions with respect to the global unitary evolution, alsotaking identical values for states ρ(t) and ωS. Let ρ(0) =|ψ(0)〉〈ψ(0)| be the initial state of the system S + B thatundergoes the unitary evolution ρ(t) = U(t)ρ(0)U†(t),with U(t) = e−itH being the evolution operator, andH =

∑j Ej |Ej〉〈Ej | being the time-independent Hamil-

tonian of the system. In turn, ω = 〈ρ(t)〉∞ is the infinitetime-averaged state of the full system, also written in theenergy eigenbasis of H as

ω =∑j

〈Ej |ρ(0)|Ej〉|Ej〉〈Ej | . (A1)

From Eq. (A1), note that both the Hamiltonian H andthe dephased state ω are commuting operators, thus im-plying that U†(t)ωU(t) = ω. In this case, it is straight-forward to conclude the relative purity F (ρ(t), ω) =Tr(ωρ(t)) of such states is given by

F (ρ(t), ω) = Tr(ωU(t)ρ(0)U†(t))

= Tr(U†(t)ωU(t)ρ(0))

= Tr(ωρ(0)) . (A2)

Clearly, Eq. (A2) shows that F (ρ(t), ω) = F (ρ(0), ω) =Tr(ωρ(0)) stands as a time-independent quantity that de-pends on the initial state of the full system and also itssteady state.

Next, the Uhlmann fidelity of states ρ(t) and ω is writ-ten as [22]

F (ρ(t), ω) =

(Tr

[√√ρ(t)ω

√ρ(t)

])2

. (A3)

We stress that Uhlmann fidelity is a positive quantity forall quantum states, also being a symmetric function over

its entries, i.e., F (ρ(t), ω) = F (ω, ρ(t)). Invoking the

identity√ρ(t) = U(t)

√ρ(0)U†(t), which holds for any

density matrix undergoing a given unitary evolution [61,62], and since U†(t)ωU(t) = ω, one gets√

ρ(t)ω√ρ(t) = U(t)

√ρ(0)ω

√ρ(0)U†(t)

= 〈ψ(0)|ω|ψ(0)〉 ρ(t) , (A4)

where we have also used the fact that√ρ(0) = ρ(0) =

|ψ(0)〉〈ψ(0)| since the initial state is pure. Hence, byplugging Eq. (A4) into Eq. (A3), one readily gets

F (ρ(t), ω) = 〈ψ(0)|ω|ψ(0)〉 . (A5)

Analogously to the case of relative purity, this meansthe Uhlmann fidelity stands as a time-independent quan-tity, also being a function of both the initial and equili-brated states. Indeed, from Eq. (A2), we point out that

F (ρ(t), ω) is nothing but the relative purity F (ρ(0), ω),and it yields

F (ρ(t), ω) = Tr(ω |ψ(0)〉〈ψ(0)|)= Tr(ωρ(0))

= F (ρ(t), ω) . (A6)

As a final remark, we emphasize the identities presentedin Eqs. (A2) and (A5) come from the fact that ω is afixed point of the unitary dynamics of the system, thusimplying that the relative purity and Uhlmann fidelitycharacterize such a constant of motion.

B. BOUND ON THE FIGURE OF MERIT

In this Appendix we will present in detail the deriva-tion of the inequality presented in Eq. (7), which in turnstands as an upper bound on the figure of merit for equi-libration given by

g(t) = |TrS [(ρS(t)− ωS)ωS]|2 . (B1)

From Eqs. (3) and (6), it is straightforward to concludethat

ρS(t)− ωS =∑k 6=l

ckc∗l e−it(Ek−El) TrB(|Ek〉〈El|) , (B2)

where we have defined cj := 〈Ej |ψ(0)〉. From Eq. (B2),the time average of the figure of merit in Eq. (B1) be-comes

Page 11: arXiv:2108.05936v1 [quant-ph] 12 Aug 2021

11

〈g(t)〉∞ =∑k 6=l

∑m 6=n

ckc∗l cmc

∗n

⟨e−it(Ek−El+Em−En)

⟩∞

TrSB [|Ek〉〈El|(ωS ⊗ IB)] TrSB [|Em〉〈En|(ωS ⊗ IB)] . (B3)

To evaluate the time average in the right-hand side of Eq. (B3), we will use the fact that the Hamiltonian H hasnondegenerate energy gaps, and thus the double summation will only include terms where k 6= l and m 6= n [12]. Inthis case, we find the only nonzero terms are those matrix elements labeled as n = k and m = l, thus implying that

〈g(t)〉∞ =∑k 6=l|ck|2|cl|2 TrSB [|Ek〉〈El|(ωS ⊗ IB)] TrSB [|El〉〈Ek|(ωS ⊗ IB)]

=∑k 6=l|ck|2|cl|2 〈El|ωS ⊗ IB|Ek〉〈Ek|ωS ⊗ IB|El〉 , (B4)

where from the first to the second line we have used the cyclic property of trace. We point out that the sum inEq. (B4) can be recast as

∑k 6=l =

∑k,l −

∑k=l, and thus one gets

〈g(t)〉∞ =∑k,l

|ck|2|cl|2 〈El|ωS ⊗ IB|Ek〉〈Ek|ωS ⊗ IB|El〉 −∑k

|ck|4 〈Ek|ωS ⊗ IB|Ek〉2

= TrSB [ω(ωS ⊗ IB)ω(ωS ⊗ IB)]−∑k

|ck|4 〈Ek|ωS ⊗ IB|Ek〉2

≤ TrSB [ω(ωS ⊗ IB)ω(ωS ⊗ IB)] , (B5)

where we have recognized the equilibrium state ω =∑j |cj |2|Ej〉〈Ej | [see Eq. (3)], and also used that∑k |ck|4 〈Ek|ωS ⊗ IB|Ek〉2 ≥ 0. Invoking the Cauchy-

Scharwz inequality for operators, i.e., |Tr(AB)| ≤‖A‖2‖B‖2, and choosing A = B = ω(ωS ⊗ IB), one mayverify Eq. (B5) implies that

〈g(t)〉∞ ≤ TrSB

[(ωS ⊗ IB)ω2(ωS ⊗ IB)

]. (B6)

Next, given that Tr(PQ) ≤ ‖P‖∞Tr(Q) for two positiveoperators P and Q, it follows that

〈g(t)〉∞ ≤ ‖ωS ⊗ IB‖∞TrSB

[ω2(ωS ⊗ IB)

]≤ ‖ωS ⊗ IB‖2∞TrSB

(ω2). (B7)

Note that Eq. (B7) can be recast by using that‖ωS ⊗ IB‖∞ = ‖ωS‖∞, and also recognizing the effec-tive dimension deff(ω) = 1/TrSB

(ω2). Hence, the time

average of the figure of merit is upper bounded as

〈g(t)〉∞ ≤‖ωS‖2∞deff(ω)

. (B8)

C. FLUCTUATIONS OF THE PURITY OF THEREDUCED STATE

In this Appendix we follow Ref. [39] and investigatethe role of quantum purity as a figure of merit for equili-bration of subsystem S, for which we evaluate the pu-rity pS(t) = TrS(ρS(t)2), where ρS(t) = TrB(ρ(t)) isthe reduced density matrix, with ρ(t) = |Ψ(t)〉〈Ψ(t)|.The dynamics of the subsystem S is governed by theequation dρS(t)/dt = iTrB([ρ(t), H]), where H = HS ⊗

IB + IS ⊗ HB + HSB is the Hamiltonian of the sys-tem. In particular, by exploiting the cyclic property ofthe trace, Tr(A1[A2, A3]) = Tr([A1, A2]A3), also usingthe identities TrSB((ρS(t) ⊗ IB)[ρ(t), HS ⊗ IB]) = 0 andTrSB((ρS(t)⊗ IB)[ρ(t), IS ⊗HB]) = 0, one may prove thetime derivative of the purity can be written as

d

dtpS(t) = 2iTrSB((ρS(t)⊗ IB)[ρ(t), HSB]) . (C1)

Next, let ρcor(t) = ρ(t)−ρS(t)⊗ρB(t) be the traceless cor-relation operator, i.e., TrS(ρcor(t)) = TrB(ρcor(t)) = 0.This operator is identically zero when the global stateρ(t) is fully uncorrelated, also satisfying the identityTrSB((ρS(t)⊗ IB)[ρS(t)⊗ ρB(t), HSB]) = 0, which impliesthat

d

dtpS(t) = 2iTrSB((ρS(t)⊗ IB)[ρcor(t), HSB]) . (C2)

In the following, we will derive an upper bound to the ratedpS(t)/dt exploiting the role of quantum correlations ofthe bipartite system S + B. To do so, note the absolutevalue of both sides of Eq. (C2) can be recast as∣∣∣∣ ddtpS(t)

∣∣∣∣ ≤ 2 ‖(ρS(t)⊗ IB)[ρcor(t), HSB]‖1

≤ 2 ‖ρS(t)⊗ IB‖∞ ‖[ρcor(t), HSB]‖1 , (C3)

where we have used the inequalities |Tr(A1A2)| ≤‖A1A2‖1 ≤ ‖A1‖∞‖A2‖1, with ‖ρS(t)⊗ IB‖∞ =‖ρS(t)‖∞. Moreover, invoking the triangle inequal-ity ‖A1 +A2‖1 ≤ ‖A1‖1 + ‖A2‖1 and also apply-ing the upper bound ‖A1A2‖1 ≤ ‖A1‖∞‖A2‖1, itfollows that ‖[ρcor(t), HSB]‖1 ≤ 2 ‖ρcor(t)HSB‖1 ≤

Page 12: arXiv:2108.05936v1 [quant-ph] 12 Aug 2021

12

0

0.005

0.01

0.015

0 2 4 6 8 100

0.004

0.008

0.012

0 2 4 6 8 100

0.0005

0.001

0.0015

0.002

0 2 4 6 8 10

0

0.002

0.004

0.006

0 2 4 6 8 100

0.001

0.002

0.003

0.004

0 2 4 6 8 100

0.0002

0.0004

0.0006

0.0008

0 2 4 6 8 10

{τ(k) }

k=1,...,4

τ

(a)

{τ(k) }

k=1,...,4

τ

(b)

{τ(k) }

k=1,...,4

τ

(c)

{τ(k) }

k=1,...,4

τ

(d)

{τ(k) }

k=1,...,4

τ

(e)

{τ(k) }

k=1,...,4

τ

τ (1) τ (2) τ (3) τ (4)

(f)

FIG. 5. (Color online) Plot of the set of QSL times {τ (k)}k=1,...,4 in Eqs. (12), (15), (20) and (28). The panels 5(a), 5(b), 5(c)refer to QSL times for the transverse field Ising model [see Eq. (32), with J = 1, hx = 0.5, hz = −1.05], and panels 5(d), 5(e), 5(f)describe the QSL times for the nonintegrable XXZ model with next-nearest-neighbor hopping [see Eq. (33), with J = 1, U = 2,Jnnn = 0.2]. The system is initialized at the Neel state |Ψ(0)〉 = |1, 0, 1, 0, . . . , 0, 1〉, with open boundary conditions, for thesystem sizes L = 4 [Figs. 5(a), 5(d)], L = 6 [Figs. 5(b), 5(e)], and L = 8 [Figs. 5(c), 5(f)].

2 ‖ρcor(t)‖1‖HSB‖∞. Inserting such results into Eq. (C3),one gets∣∣∣∣ ddtpS(t)

∣∣∣∣ ≤ 4 ‖ρS(t)‖∞ ‖ρcor(t)‖1‖HSB‖∞ . (C4)

Interestingly, note the Schatten 1-norm of the correlatedoperator is upper bounded according to Pinsker’s in-equality as [53]

‖ρcor(t)‖1 ≤√

2 ISB(ρ(t)) , (C5)

where the mutual information is defined

ISB(ρ(t)) := S(ρS(t)) + S(ρB(t))− S(ρ(t)) , (C6)

with S(%) = −Tr(% ln %) being the von Neumann entropy.Hence, by combining Eqs. (C4) and (C5), it yields∣∣∣∣ ddtpS(t)

∣∣∣∣ ≤ 4√

2ISB(ρ(t)) ‖ρS(t)‖∞ ‖HSB‖∞ . (C7)

Importantly, Eq. (C7) relates the fluctuations of pu-rity of the subsystem S with the quantum correlationsof the whole bipartite system captured by the mu-tual information. On the one hand, since ‖ρS(t)‖∞ ≤1, Eq. (C7) implies the upper bound |dpS(t)/dt| ≤4√

2ISB(ρ(t)) ‖ρS(t)‖∞ ‖HSB‖∞, which was already pre-sented in Ref. [63]. On the other hand, since ‖ρS(t)‖∞ ≤‖ρS(t)‖2 =

√TrS(ρS(t)2) =

√pS(t), we thus obtain the

tighter upper bound∣∣∣∣ ddtpS(t)

∣∣∣∣ ≤ 4√

2ISB(ρ(t))√pS(t) ‖HSB‖∞ . (C8)

We point out that, since the bipartite system is initial-ized in a pure state, i.e., Tr(ρ(0)2) = Tr(ρ(0)) = 1, itsinstantaneous state ρ(t) = U(t)ρ(0)U†(t) will be purefor all time t, i.e., Tr(ρ(t)2) = Tr(ρ(t)) = 1. As a con-sequence, the mutual information of the input state tri-vially collapses into the sum of the von Neumann entropyof the marginal states, ISB(ρ(t)) = S(ρS(t))+S(ρB(t)) =2S(ρS(t)) = 2S(ρB(t)). In this case, Eq. (C8) can berecast as∣∣∣∣ ddtpS(t)

∣∣∣∣ ≤ 8√S(ρS(t))

√pS(t) ‖HSB‖∞ , (C9)

with the von Neumann entropy satisfying the inequality0 ≤ S(ρS(t)) ≤ ln(dS). Hence, by integrating Eq. (C9)over the range 0 ≤ t ≤ τ , it follows that

τ ≥

∣∣∣√pS(τ)−√pS(0)

∣∣∣4√

ln(dS) ‖HSB‖∞. (C10)

where we have used that∫dx|g(x)| ≥ |

∫dxg(x)|. Here

τ stands for the time to reach the equilibrium purity,i.e., pS(τ) ≡ peq

S . Particularly, if the initial state is fullyuncorrelated, i.e., ρ(0) = ρS(0) ⊗ ρB(0), while ρS(0) is apure state with pS(0) = Tr(ρS(0)) = 1, we finally arriveat the bound

τ ≥

∣∣∣√peqS − 1

∣∣∣4√

ln(dS) ‖HSB‖∞. (C11)

We point out Eq. (C11) is slightly different from the re-sult presented in Ref. [39], where here one finds the equi-libration time depends on the square root of the equilib-rium purity.

Page 13: arXiv:2108.05936v1 [quant-ph] 12 Aug 2021

13

D. DETAILS ON THE SET OF QSL TIMES

In this Appendix we provide details on the lowerbounds {τ (k)}k=1,...,4 for both the spin models discussedin Sec. IV. In Figure 5, we show the set of QSL timesin Eqs. (12), (15), (20) and (28) as a function of time τ ,for the transverse field Ising model [see Figs. 5(a), 5(b),and 5(c)], and the nonintegrable XXZ model with next-nearest-neighbor hopping [see Figs. 5(d), 5(e), and 5(f)].In both cases, the spin system is initialized at the Neelstate |Ψ(0)〉 = |1, 0, 1, 0, . . . , 0, 1〉, with open bound-ary conditions. Here we set the system sizes L = 4[Figs. 5(a), 5(d)], L = 6 [Figs. 5(b), 5(e)], and L = 8

[Figs. 5(c), 5(f)]. We see that, for all τ ≥ 0 andsystem size L, the curve of τ (3) is always above theQSL times τ (1), τ (2), and τ (4), regardless of the spinsystem. This clearly illustrate the fact that τQSL =

max{τ (1), τ (2), τ (3), τ (4)} = τ (3), i.e., the QSL time inEq. (30) is given by the lower bound τ (3), the latter be-ing inversely proportional to the variance of the Hamil-tonian governing the dynamics [see Eq. (20)]. The quan-tity τ (1) stands as a lower bound to the set of QSLtimes {τ (2), τ (3), τ (4)}, and approaches small values asthe size L grows. In addition, we see the amplitude of{τ (k)}k=1,...,4 decreases as we increase the system size L.

[1] A. Polkovnikov, K. Sengupta, A. Silva, and M. Vengalat-tore, “Colloquium: Nonequilibrium dynamics of closedinteracting quantum systems,” Rev. Mod. Phys. 83, 863(2011).

[2] L. D’Alessio, Y. Kafri, A. Polkovnikov, and M. Rigol,“From quantum chaos and eigenstate thermalization tostatistical mechanics and thermodynamics,” Adv. Phys.65, 239 (2016).

[3] J.-S. Caux and F. H. L. Essler, “Time Evolution of LocalObservables After Quenching to an Integrable Model,”Phys. Rev. Lett. 110, 257203 (2013).

[4] C. Gogolin and J. Eisert, “Equilibration, thermalisa-tion, and the emergence of statistical mechanics in closedquantum systems,” Rep. Prog. Phys. 79, 056001 (2016).

[5] L. Guidoni, E. Beaurepaire, and J.-Y. Bigot, “Magneto-optics in the Ultrafast Regime: Thermalization of SpinPopulations in Ferromagnetic Films,” Phys. Rev. Lett.89, 017401 (2002).

[6] I. Bloch, J. Dalibard, and S. Nascimbene, “Quantumsimulations with ultracold quantum gases,” Nat. Phys.8, 267 (2012).

[7] R. Blatt and C. F. Roos, “Quantum simulations withtrapped ions,” Nat. Phys. 8, 277 (2012).

[8] S. Trotzky, Y.-A. Chen, A. Flesch, I. P. McCulloch,U. Schollwock, J. Eisert, and I. Bloch, “Probing therelaxation towards equilibrium in an isolated stronglycorrelated one-dimensional Bose gas,” Nat. Phys. 8, 325(2012).

[9] M. Schreiber, S. S. Hodgman, P. Bordia, H. P. Luschen,M. H. Fischer, R. Vosk, E. Altman, U. Schneider, andI. Bloch, “Observation of many-body localization of in-teracting fermions in a quasirandom optical lattice,” Sci-ence 349, 842 (2015).

[10] A. M. Kaufman, M. E. Tai, A. Lukin, M. Rispoli,R. Schittko, P. M. Preiss, and M. Greiner, “Quan-tum thermalization through entanglement in an isolatedmany-body system,” Science 353, 794 (2016).

[11] A. J. Short, “Equilibration of quantum systems and sub-systems,” New J. Phys. 13, 053009 (2011).

[12] N. Linden, S. Popescu, A. J. Short, and A. Winter,“Quantum mechanical evolution towards thermal equi-librium,” Phys. Rev. E 79, 061103 (2009).

[13] A. S. L. Malabarba, L. P. Garcıa-Pintos, N. Linden, T. C.Farrelly, and A. J. Short, “Quantum systems equilibraterapidly for most observables,” Phys. Rev. E 90, 012121

(2014).[14] T. R. Oliveira, C. Charalambous, D. Jonathan,

M. Lewenstein, and A. Riera, “Equilibration time scalesin closed many-body quantum systems,” New. J. Phys.20, 033032 (2018).

[15] T. Giordani, D. J. Brod, C. Esposito, N. Viggianiello,M. Romano, F. Flamini, G. Carvacho, N. Spagnolo,E. F. Galvao, and F. Sciarrino, “Experimental quantifi-cation of four-photon indistinguishability,” N. J. Phys.22, 043001 (2020).

[16] P. Reimann and M. Kastner, “Equilibration of isolatedmacroscopic quantum systems,” New J. Phys. 14, 043020(2012).

[17] P. Reimann, “Foundation of Statistical Mechanics underExperimentally Realistic Conditions,” Phys. Rev. Lett.101, 190403 (2008).

[18] S. L. Braunstein and C. M. Caves, “Statistical Distanceand the Geometry of Quantum States,” Phys. Rev. Lett.72, 3439 (1994).

[19] E. A. Morozova and N. N. Cencov, “Markov invariantgeometry on state manifolds,” J. Sov. Math. 36, 2648(1991).

[20] D. Petz, “Monotone metrics on matrix spaces,” Lin. Al-geb. Appl. 244, 81 (1996).

[21] D. Petz and C. Sudar, “Geometries of quantum states,”J. Math. Phys. 37, 2662 (1996).

[22] I. Bengtsson and K. Zyczkowski, Geometry of Quan-tum States: An Introduction to Quantum Entanglement(Cambridge University Press, England, 2006).

[23] P. Salamon, J. D. Nulton, and R. S. Berry, “Length instatistical thermodynamics,” J. Chem. Phys. 82, 2433(1985).

[24] H. J. D. Miller and M. Mehboudi, “Geometry of WorkFluctuations versus Efficiency in Microscopic ThermalMachines,” Phys. Rev. Lett. 125, 260602 (2020).

[25] V. Giovannetti, S. Lloyd, and L. Maccone, “Quantumlimits to dynamical evolution,” Phys. Rev. A 67, 052109(2003).

[26] D. P. Pires, M. Cianciaruso, L. C. Celeri, G. Adesso,and D. O. Soares-Pinto, “Generalized geometric quan-tum speed limits,” Phys. Rev. X 6, 021031 (2016).

[27] M. Jarzyna and J. Ko lodynski, “Geometric approach toquantum statistical inference,” IEEE J. Sel. Areas Inf.Theor. 1, 367 (2020).

Page 14: arXiv:2108.05936v1 [quant-ph] 12 Aug 2021

14

[28] J. S. Sidhu and P. Kok, “Geometric perspective onquantum parameter estimation,” AVS Quantum Sci. 2,014701 (2020).

[29] P. E. M. F. Mendonca, R. d. J. Napolitano, M. A. Mar-chiolli, C. J. Foster, and Y.-C. Liang, “Alternative fi-delity measure between quantum states,” Phys. Rev. A78, 052330 (2008).

[30] A. del Campo, I. L. Egusquiza, M. B. Plenio, and S. F.Huelga, “Quantum Speed Limits in Open System Dy-namics,” Phys. Rev. Lett. 110, 050403 (2013).

[31] J. Jing, L. A. Wu, and A. del Campo, “FundamentalSpeed Limits to the Generation of Quantumness,” Sci.Rep. 6, 38149 (2016).

[32] S.-X. Wu and C.-S. Yu, “Quantum speed limit for amixed initial state,” Phys. Rev. A 98, 042132 (2018).

[33] M. Gessner and A. Smerzi, “Statistical speed of quan-tum states: Generalized quantum Fisher information andSchatten speed,” Phys. Rev. A 97, 022109 (2018).

[34] B. Yan, L. Cincio, and W. H. Zurek, “InformationScrambling and Loschmidt Echo,” Phys. Rev. Lett. 124,160603 (2020).

[35] C. M. Sanchez, A. K. Chattah, K. X. Wei, L. Buljubasich,P. Cappellaro, and H. M. Pastawski, “Perturbation In-dependent Decay of the Loschmidt Echo in a Many-BodySystem,” Phys. Rev. Lett. 124, 030601 (2020).

[36] T. Giordani, C. Esposito, F. Hoch, G. Carvacho, D. J.Brod, E. F. Galvao, N. Spagnolo, and F. Sciarrino, “Wit-nesses of coherence and dimension from multiphoton in-distinguishability tests,” Phys. Rev. Research 3, 023031(2021).

[37] T. Monnai, “General Relaxation Time of the Fidelity forIsolated Quantum Thermodynamic Systems,” J. Phys.Soc. Japan 83, 064001 (2014).

[38] N. Linden, S. Popescu, A. J. Short, and A. Winter, “Onthe speed of fluctuations around thermodynamic equilib-rium,” New J. Phys. 12, 055021 (2010).

[39] C. Gogolin, “Pure State Quantum Statistical Mecha-nics,” arXiv:1003.5058.

[40] H. Wilming, M. Goihl, C. Krumnow, and J. Eisert, “To-wards local equilibration in closed interacting quantummany-body systems,” arXiv:1704.06291.

[41] L. P. Garcıa-Pintos, N. Linden, A. S. L. Malabarba,A. J. Short, and A. Winter, “Equilibration Time Scalesof Physically Relevant Observables,” Phys. Rev. X 7,031027 (2017).

[42] D. Girolami, “Observable Measure of Quantum Coher-ence in Finite Dimensional Systems,” Phys. Rev. Lett.113, 170401 (2014).

[43] L. Mandelstam and I. Tamm, “The uncertainty relationbetween energy and time in non-relativistic quantum me-chanics,” in Selected Papers, edited by B. M. Bolotovskii,V. Y. Frenkel, and R. Peierls (Springer, Berlin, Heidel-berg, 1991) p. 115.

[44] S. Deffner and S. Campbell, “Quantum speed limits:from Heisenberg’s uncertainty principle to optimal quan-tum control,” J. Phys. A: Math. Theor. 50, 453001(2017).

[45] A. Bottcher and D. Wenzel, “The Frobenius norm andthe commutator,” Lin. Alg. Appl. 429, 1864 (2008).

[46] K. M. R. Audenaert, “Variance bounds, with an applica-tion to norm bounds for commutators,” Lin. Alg. Appl.432, 1126 (2010).

[47] A. Streltsov, G. Adesso, and M. B. Plenio, “Colloquium:Quantum coherence as a resource,” Rev. Mod. Phys. 89,041003 (2017).

[48] R. Augusiak, J. Ko lodynski, A. Streltsov, M. N. Bera,A. Acın, and M. Lewenstein, “Asymptotic role of en-tanglement in quantum metrology,” Phys. Rev. A 94,012339 (2016).

[49] M. Oszmaniec, R. Augusiak, C. Gogolin, J. Ko lodynski,A. Acın, and M. Lewenstein, “Random Bosonic Statesfor Robust Quantum Metrology,” Phys. Rev. X 6, 041044(2016).

[50] L. Pezze and A. Smerzi, “Entanglement, Nonlinear Dy-namics, and the Heisenberg Limit,” Phys. Rev. Lett.102, 100401 (2009).

[51] L. Pezze and A. Smerzi, “Quantum theory of phase esti-mation,” in Atom Interferometry, Proceedings of the In-ternational School of Physics “Enrico Fermi”, Vol. 188,edited by G. M. Tino and M. A. Kasevich (2014) pp.691–741.

[52] D. Girolami and B. Yadin, “Witnessing Multipartite En-tanglement by Detecting Asymmetry,” Entropy 19(3),124 (2017).

[53] E. A. Carlen and E. H. Lieb, “Remainder terms forsome quantum entropy inequalities,” J. Math. Phys. 55,042201 (2014).

[54] K. M. R. Audenaert and J. Eisert, “Continuity boundson the quantum relative entropy,” J. Math. Phys. 46,102104 (2005).

[55] K. M. R. Audenaert and J. Eisert, “Continuity boundson the quantum relative entropy – II,” J. Math. Phys.52, 112201 (2011).

[56] E. Leviatan, F. Pollmann, J. H. Bardarson, D. A. Huse,and E. Altman, “Quantum thermalization dynamics withMatrix-Product States,” arXiv:1702.08894.

[57] M. C. Banuls, J. I. Cirac, and M. B. Hastings,“Strong and Weak Thermalization of Infinite Noninte-grable Quantum Systems,” Phys. Rev. Lett. 106, 050405(2011).

[58] M. M. Taddei, B. M. Escher, L. Davidovich, and R. L.de Matos Filho, “Quantum Speed Limit for Physical Pro-cesses,” Phys. Rev. Lett. 110, 050402 (2013).

[59] L. P. Garcıa-Pintos, S. Nicholson, J. R. Green, A. delCampo, and A. V. Gorshkov, “Unifying Quan-tum and Classical Speed Limits on Observables,”arXiv:2108.04261.

[60] F. Campaioli, F. A. Pollock, F. C. Binder, and K. Modi,“Tightening Quantum Speed Limits for Almost AllStates,” Phys. Rev. Lett. 120, 060409 (2018).

[61] R. Bathia, Matrix Analysis (Springer-Verlag, New York,1997).

[62] D. P. Pires, L. C. Celeri, and D. O. Soares-Pinto, “Geo-metric lower bound for a quantum coherence measure,”Phys. Rev. A 91, 042330 (2015).

[63] G. Kimura, H. Ohno, and H. Hayashi, “How to detecta possible correlation from the information of a subsys-tem in quantum-mechanical systems,” Phys. Rev. A 76,042123 (2007)