Location via proxy:   [ UP ]  
[Report a bug]   [Manage cookies]                

Asymmetries of thermal processes in open quantum systems

Álvaro Tejero atejero@onsager.ugr.es Electromagnetism and Condensed Matter Department, Universidad de Granada, 18071 Granada, Spain Instituto Carlos I de Física Teórica y Computacional, Universidad de Granada, 18071 Granada, Spain    Rafael Sánchez Departamento de Física Teórica de la Materia Condensada, Universidad Autónoma de Madrid, 28049 Madrid, Spain Condensed Matter Physics Center (IFIMAC), Universidad Autónoma de Madrid, 28049 Madrid, Spain Instituto Nicolás Cabrera, Universidad Autónoma de Madrid, 28049 Madrid, Spain    Laiachi El Kaoutit Departamento de Álgebra, Universidad de Granada, 18071 Granada, Spain    Daniel Manzano manzano@onsager.ugr.es Electromagnetism and Condensed Matter Department, Universidad de Granada, 18071 Granada, Spain Instituto Carlos I de Física Teórica y Computacional, Universidad de Granada, 18071 Granada, Spain    Antonio Lasanta alasanta@ugr.es Instituto Carlos I de Física Teórica y Computacional, Universidad de Granada, 18071 Granada, Spain Departamento de Álgebra, Facultad de Educación, Economía y Tecnología de Ceuta, Universidad de Granada, Cortadura del Valle, s/n, 51001 Ceuta, Spain Nanoparticles Trapping Laboratory, Universidad de Granada, Granada, Spain
(June 28, 2024)
Abstract

An intriguing phenomenon in non-equilibrium quantum thermodynamics is the asymmetry of thermal processes. Relaxation to thermal equilibrium is the most important dissipative process, being a key concept for the design of heat engines and refrigerators, contributing to the study of foundational questions of thermodynamics, and being relevant for quantum computing through the process of algorithmic cooling. Despite the importance of this kind of processes, their dynamics are far from being understood. We show that the free relaxation to thermal equilibrium follows intrinsically different paths depending on whether the temperature of the system increases (heating up) or decreases (cooling down), being faster in the first case. Our theory is exemplified using the recently developed thermal kinematics based on information geometry theory, utilizing three prototypical examples: a quantum two-level system, the quantum harmonic oscillator, and a trapped quantum Brownian particle, including both analytic results and numerical simulations. For this, we have extended the thermal kinematic approach to open quantum systems. Additionally, we offer a simple theoretical explanation in the case of a two level system and a more general picture for the other two systems based on the spectral decomposition of the Liouvillian and the spectral gap of reciprocal processes.

I Introduction

When a system is pushed far from equilibrium, its evolution may follow anomalous paths. A series of seminal works done during the past century [1, 2, 3, 4, 5, 6, 7] has provided essential advances in studying transitory phenomena in the linear regime associated with fluctuations, except for some particular cases [8, 9] where predictions can extend beyond equilibrium. Despite this progress, we still lack a general theory beyond linear response and fluctuation theorems to decipher the dynamics and behavior of transient regimes of a freely evolving system between two desired states [10, 11]. This problem is of particular interest for quantum thermodynamic processes [12], finite-time quantum heat engines [13, 14, 15, 16] and establishing speed limit bounds [17, 18, 19, 20]. Recent progress in unraveling anomalous shortcuts during relaxation processes in out-of-equilibrium systems points in this direction [21].

A remarkable example of a possible counter-intuitive behavior of a system is the Mpemba-like effect (ME) [22, 23, 24]. Namely, put two identical systems at different initial temperatures in contact with a reservoir at a hotter or colder temperature than those of the two systems. The ME occurs when the initially hotter/colder system cools down/heats up faster than the system that was initially closer to the final temperature. In the case of cooling, the effect is called normal ME, and for heating, it is called inverse ME [25, 26]. In Markovian systems, the ME can be well understood using a spectral decomposition of the decay modes, diminishing (weak ME), or canceling slow-decaying modes (strong ME) to enhance the fast ones, making it possible to control the speed of the relaxation. In this way, up to an exponential acceleration is achievable [27]. This phenomenon has been realized both in classical [25, 27, 28, 29, 30, 31, 32, 33] and open quantum systems [34, 35, 36, 37, 38]. Additionally, a generalization of the ME to quantum entangled configurations has been very recently proposed [39, 40, 41, 42]. Note that a strong relation exists between exceptional points and speed up relaxation in open quantum systems [36, 43].

Alternatively, when spectral methods are not applicable, other strategies can be used to understand anomalous evolution using macroscopic observables depending on the system of interest. The origin of anomalous relaxation is associated with energy non-equipartition in water and granular gases composed by rough hard spheres [44, 45], a particular condition in kurtosis also in the former with smooth hard spheres [26, 46], and correlation length in spin glasses [47]. Furthermore, the strategy of employing several sudden changes in temperature has been probed useful for shortening relaxation times, such as preheating protocols [48]. This approach takes advantage of the slow growth of magnetic domains near phase transitions in systems where time-scale separation is not possible [49], or through different control techniques [50, 51, 52].

Refer to caption
Figure 1: (a) Asymmetric cooling and heating relaxation to an equidistant stationary state at temperature TWsubscript𝑇𝑊T_{W}italic_T start_POSTSUBSCRIPT italic_W end_POSTSUBSCRIPT. It takes longer when the system is initially hot (thermalized with a bath at THsubscript𝑇𝐻T_{H}italic_T start_POSTSUBSCRIPT italic_H end_POSTSUBSCRIPT which is decoupled at t=0𝑡0t=0italic_t = 0) than when it is cold (at TCsubscript𝑇𝐶T_{C}italic_T start_POSTSUBSCRIPT italic_C end_POSTSUBSCRIPT), with TC<TW<THsubscript𝑇𝐶subscript𝑇𝑊subscript𝑇𝐻T_{C}<T_{W}<T_{H}italic_T start_POSTSUBSCRIPT italic_C end_POSTSUBSCRIPT < italic_T start_POSTSUBSCRIPT italic_W end_POSTSUBSCRIPT < italic_T start_POSTSUBSCRIPT italic_H end_POSTSUBSCRIPT. (b) Asymmetric cooling and heating evolution between two states at temperatures TCsubscript𝑇𝐶T_{C}italic_T start_POSTSUBSCRIPT italic_C end_POSTSUBSCRIPT and THsubscript𝑇𝐻T_{H}italic_T start_POSTSUBSCRIPT italic_H end_POSTSUBSCRIPT, with TC<THsubscript𝑇𝐶subscript𝑇𝐻T_{C}<T_{H}italic_T start_POSTSUBSCRIPT italic_C end_POSTSUBSCRIPT < italic_T start_POSTSUBSCRIPT italic_H end_POSTSUBSCRIPT. The evolution from hot to temperature (cooling) is slower than from cold to hot (heating).

A fundamental question, illustrated in Fig. 1, is whether free cooling and heating processes after a sudden change of the environment temperature are identical or follow intrinsically different paths, see Fig. 1(a). In classical systems heating and cooling can show an asymmetry that has been verified both theoretically and experimentally far from equilibrium  [53, 54]. An even more emphatic result is that the asymmetry is revealed when relaxation processes occur between two fixed temperatures [54], see Fig. 1(b). This has been successfully explained mathematically by using the so-called thermal kinematics [54], based on information geometrical arguments [55, 56]. In this paper we focus on that question, that is, unraveling the mechanism of the heating and cooling processes in the realm of open quantum systems. In order to do this we develop an extension of the thermal kinematics theory to the thermodynamics of open quantum systems. We analyze whether a relaxation process far from equilibrium, say from an initially hot to a colder thermal state, is equally fast than its reverse, from the colder to the hotter, and relate it to the properties of the spectral gap [57, 58, 59, 60, 61, 62]. To showcase this, we use simple models based on a thermal qubit, a quantum harmonic oscillator, and a quantum Brownian particle.

The heat properties of such simple quantum systems have recently become accessible experimentally. Solid state realizations of qubits coupled to fermionic or bosonic reservoirs allow to control the spectral properties, couplings and temperatures externally [63, 64, 65]. This is the case of quantum dot systems [66, 67], which can selectively be (un)connected to different reservoirs with gate voltages [68] and whose distribution can be measured via charge detectors [69, 70, 71, 72], or of superconducting circuits coupled to resistors acting as thermal baths via tunable resonators [73, 74, 75, 76, 77, 78]. Furthermore, the qubit state can be monitored [79, 80, 81, 82, 83]. Improvements in high frequency thermometry even allows to detect single temperature fluctuations [84]. These ingredients make the detection of relaxation paths in quantum information systems possible.

The recent measurement of asymmetric relaxation of a classical particle in a harmonic trap [54] motivates us to treat this problem from a quantum perspective. To do so, we investigate the thermalization of a quantum Brownian particle, a model that has successfully been applied to describe a plethora of quantum effects, such as quantum dissipation [85, 86], harmonic oscillators [87], macroscopic quantum tunneling [88, 89, 90], metastable states [91], single-electron transistors [92], the spin-boson problem [93], or impurity dynamics in Luttinger liquids [94] and ultracold atomic gases [95]. We hence emphasize that understanding the relaxation processes is of importance for quantum thermodynamics and for the physics of driven nanoscale devices [12, 11].

This manuscript is organized as follows. In Secs. II and III, we present the theoretical framework based on the master equation for open quantum systems and the measures of thermodynamic distances. In Sec. IV, the definitions of the different protocols are provided. We then apply these methods to three different systems. In Sec. V we consider a the simplest case of two level system coupled to a thermal bath, which can be solved analytically. Then Sec. VI considers more complex systems, namely the harmonic oscillator and the quantum Brownian particle. Additionally, this section includes numerical simulations for the non-analytically solvable systems. Section VII provides a theoretical justification for all the phenomena based on the spectrum of the Liouvillian and the influence of the considered initial state on the evolution. Finally, the conclusions are drawn in Sec. VIII.

II Theoretical framework: Markovian Open Quantum Systems

The state of the quantum system, weakly coupled to the environment, is described by its reduced density matrix ρ(t)𝜌𝑡\rho(t)italic_ρ ( italic_t ), whose evolution is governed by the Gorini–Kossakowski–Sudarshan–Lindblad (GKSL) quantum master equation [96, 97, 98, 99, 100, 101, 102, 103, 104]

ρ˙(t)=[ρ(t)],˙𝜌𝑡delimited-[]𝜌𝑡\dot{\rho}(t)=\mathcal{L}[\rho(t)],over˙ start_ARG italic_ρ end_ARG ( italic_t ) = caligraphic_L [ italic_ρ ( italic_t ) ] , (1)

where \mathcal{L}caligraphic_L is the Liouvillian superoperator

[ρ(t)]=i[H,ρ(t)]+i=1N(Liρ(t)Li12{LiLi,ρ(t)}),delimited-[]𝜌𝑡𝑖Planck-constant-over-2-pi𝐻𝜌𝑡superscriptsubscript𝑖1𝑁subscript𝐿𝑖𝜌𝑡subscriptsuperscript𝐿𝑖12subscriptsuperscript𝐿𝑖subscript𝐿𝑖𝜌𝑡\mathcal{L}[\rho(t)]=-\dfrac{i}{\hbar}[H,\rho(t)]+\sum_{i=1}^{N}\left(L_{i}% \rho(t)L^{\dagger}_{i}-\frac{1}{2}\left\{L^{\dagger}_{i}L_{i},\rho(t)\right\}% \right),caligraphic_L [ italic_ρ ( italic_t ) ] = - divide start_ARG italic_i end_ARG start_ARG roman_ℏ end_ARG [ italic_H , italic_ρ ( italic_t ) ] + ∑ start_POSTSUBSCRIPT italic_i = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_N end_POSTSUPERSCRIPT ( italic_L start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT italic_ρ ( italic_t ) italic_L start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT - divide start_ARG 1 end_ARG start_ARG 2 end_ARG { italic_L start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT italic_L start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT , italic_ρ ( italic_t ) } ) , (2)

being H𝐻Hitalic_H the Hamiltonian of the system, describing its coherent dynamics. The N𝑁Nitalic_N jump operators Lisubscript𝐿𝑖L_{i}italic_L start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT describe the dissipative effects due to the presence of an environment. The Liouvillian superoperator \mathcal{L}caligraphic_L preserves the trace, i.e. Tr([ρ(t)])=0tracedelimited-[]𝜌𝑡0\Tr\left(\mathcal{L}[\rho(t)]\right)=0roman_Tr ( caligraphic_L [ italic_ρ ( italic_t ) ] ) = 0, hermiticity, i.e. ([ρ(t)])=[ρ(t)]superscriptdelimited-[]𝜌𝑡delimited-[]superscript𝜌𝑡\left(\mathcal{L}[\rho(t)]\right)^{\dagger}=\mathcal{L}[\rho^{\dagger}(t)]( caligraphic_L [ italic_ρ ( italic_t ) ] ) start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT = caligraphic_L [ italic_ρ start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT ( italic_t ) ], ρ(t)for-all𝜌𝑡\forall\rho(t)∀ italic_ρ ( italic_t ), and complete positivity.

The general solution to Eq. (1) can be directly obtained as ρ(t)=et[ρ(0)]𝜌𝑡superscript𝑒𝑡delimited-[]𝜌0\rho(t)=e^{t\mathcal{L}}[\rho(0)]italic_ρ ( italic_t ) = italic_e start_POSTSUPERSCRIPT italic_t caligraphic_L end_POSTSUPERSCRIPT [ italic_ρ ( 0 ) ], where the superoperator etsuperscript𝑒𝑡e^{t\mathcal{L}}italic_e start_POSTSUPERSCRIPT italic_t caligraphic_L end_POSTSUPERSCRIPT is defined by its power expansion. Assuming the generator to be diagonalizable, one can find the right eigenmatrices, ΛkrsubscriptsuperscriptΛ𝑟𝑘\Lambda^{r}_{k}roman_Λ start_POSTSUPERSCRIPT italic_r end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT, such that

[Λkr]=λkΛkr.delimited-[]subscriptsuperscriptΛ𝑟𝑘subscript𝜆𝑘subscriptsuperscriptΛ𝑟𝑘\mathcal{L}[\Lambda^{r}_{k}]=\lambda_{k}\Lambda^{r}_{k}.caligraphic_L [ roman_Λ start_POSTSUPERSCRIPT italic_r end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT ] = italic_λ start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT roman_Λ start_POSTSUPERSCRIPT italic_r end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT . (3)

The complex numbers λksubscript𝜆𝑘\lambda_{k}italic_λ start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT are the eigenvalues of the Liouvillian. Note that, due to the hermiticity-preservation of \mathcal{L}caligraphic_L, if λksubscript𝜆𝑘\lambda_{k}italic_λ start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT is a complex eigenvalue, then λksuperscriptsubscript𝜆𝑘\lambda_{k}^{*}italic_λ start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT must also be an eigenvalue. For the same reason, one can also show that if λksubscript𝜆𝑘\lambda_{k}italic_λ start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT is real, then ΛkrsubscriptsuperscriptΛ𝑟𝑘\Lambda^{r}_{k}roman_Λ start_POSTSUPERSCRIPT italic_r end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT can be chosen to be Hermitian. Associated with the map defined in Eq. (2), there is a dual map, also called the adjoint Lindblad map, which implements the evolution of observables:

[O]=i[H,O]+i=1N(LiOLi12{O,LiLi}).superscriptdelimited-[]𝑂𝑖Planck-constant-over-2-pi𝐻𝑂superscriptsubscript𝑖1𝑁subscriptsuperscript𝐿𝑖𝑂subscript𝐿𝑖12𝑂subscriptsuperscript𝐿𝑖subscript𝐿𝑖\mathcal{L}^{\dagger}[O]=\dfrac{i}{\hbar}[H,O]+\sum_{i=1}^{N}\left(L^{\dagger}% _{i}OL_{i}-\frac{1}{2}\left\{O,L^{\dagger}_{i}L_{i}\right\}\right).caligraphic_L start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT [ italic_O ] = divide start_ARG italic_i end_ARG start_ARG roman_ℏ end_ARG [ italic_H , italic_O ] + ∑ start_POSTSUBSCRIPT italic_i = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_N end_POSTSUPERSCRIPT ( italic_L start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT italic_O italic_L start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT - divide start_ARG 1 end_ARG start_ARG 2 end_ARG { italic_O , italic_L start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT italic_L start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT } ) . (4)

This dual map, superscript\mathcal{L}^{\dagger}caligraphic_L start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT, is diagonalized by the left eigenmatrices ΛksubscriptsuperscriptΛ𝑘\Lambda^{\ell}_{k}roman_Λ start_POSTSUPERSCRIPT roman_ℓ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT,

[Λk]=λkΛk.superscriptdelimited-[]subscriptsuperscriptΛ𝑘subscript𝜆𝑘subscriptsuperscriptΛ𝑘\mathcal{L}^{\dagger}[\Lambda^{\ell}_{k}]=\lambda_{k}\Lambda^{\ell}_{k}.caligraphic_L start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT [ roman_Λ start_POSTSUPERSCRIPT roman_ℓ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT ] = italic_λ start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT roman_Λ start_POSTSUPERSCRIPT roman_ℓ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT . (5)

The matrices ΛksubscriptsuperscriptΛ𝑘\Lambda^{\ell}_{k}roman_Λ start_POSTSUPERSCRIPT roman_ℓ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT are in principle different from the matrices ΛkrsubscriptsuperscriptΛ𝑟𝑘\Lambda^{r}_{k}roman_Λ start_POSTSUPERSCRIPT italic_r end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT in Eq. (3). However, ΛksubscriptsuperscriptΛ𝑘\Lambda^{\ell}_{k}roman_Λ start_POSTSUPERSCRIPT roman_ℓ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT and ΛkrsubscriptsuperscriptΛ𝑟𝑘\Lambda^{r}_{k}roman_Λ start_POSTSUPERSCRIPT italic_r end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT still form a bi-orthogonal basis for the space of matrices and can always be defined fulfilling the property Tr(ΛkΛhr)=δkhtracesubscriptsuperscriptΛ𝑘subscriptsuperscriptΛ𝑟subscript𝛿𝑘\Tr\left(\Lambda^{\ell}_{k}\Lambda^{r}_{h}\right)=\delta_{kh}roman_Tr ( roman_Λ start_POSTSUPERSCRIPT roman_ℓ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT roman_Λ start_POSTSUPERSCRIPT italic_r end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_h end_POSTSUBSCRIPT ) = italic_δ start_POSTSUBSCRIPT italic_k italic_h end_POSTSUBSCRIPT.

Since the dynamics generated by \mathcal{L}caligraphic_L is completely positive, the eigenvalues of the Liouvillian superoperator all have a non-positive real part, Re(λk)0Resubscript𝜆𝑘0{\rm Re}\left(\lambda_{k}\right)\leq 0roman_Re ( italic_λ start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT ) ≤ 0. Furthermore, for bounded systems, Evan’s Theorem [105] enforces that at least one eigenvalue is zero, λ1=0subscript𝜆10\lambda_{1}=0italic_λ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT = 0, and this is also the case for many unbounded systems. Assuming that the null eigenvalue is non-degenerate, the asymptotic stationary state of the open quantum system is directly related to its associated eigenmatrix [106, 107],

ρss=limtρ(t)=Λ1r.subscript𝜌sssubscript𝑡𝜌𝑡superscriptsubscriptΛ1𝑟\rho_{\rm ss}=\lim_{t\to\infty}\rho(t)=\Lambda_{1}^{r}.italic_ρ start_POSTSUBSCRIPT roman_ss end_POSTSUBSCRIPT = roman_lim start_POSTSUBSCRIPT italic_t → ∞ end_POSTSUBSCRIPT italic_ρ ( italic_t ) = roman_Λ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_r end_POSTSUPERSCRIPT . (6)

Integrating Eq. (2), the spectral decomposition of \mathcal{L}caligraphic_L allows us to write the dynamics of any initial density matrix as

ρ(t)=et[ρ0]=Λ1r+k=2d2etλkTr(Λkρ0)Λkr,𝜌𝑡superscript𝑒𝑡delimited-[]subscript𝜌0subscriptsuperscriptΛ𝑟1superscriptsubscript𝑘2superscript𝑑2superscript𝑒𝑡subscript𝜆𝑘TrsubscriptsuperscriptΛ𝑘subscript𝜌0subscriptsuperscriptΛ𝑟𝑘\rho(t)=e^{t\mathcal{L}}\left[\rho_{0}\right]=\Lambda^{r}_{1}+\sum_{k=2}^{d^{2% }}e^{t\lambda_{k}}{\rm Tr}\left(\Lambda^{\ell}_{k}\rho_{0}\right)\Lambda^{r}_{% k},italic_ρ ( italic_t ) = italic_e start_POSTSUPERSCRIPT italic_t caligraphic_L end_POSTSUPERSCRIPT [ italic_ρ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ] = roman_Λ start_POSTSUPERSCRIPT italic_r end_POSTSUPERSCRIPT start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT + ∑ start_POSTSUBSCRIPT italic_k = 2 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_d start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_POSTSUPERSCRIPT italic_e start_POSTSUPERSCRIPT italic_t italic_λ start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT end_POSTSUPERSCRIPT roman_Tr ( roman_Λ start_POSTSUPERSCRIPT roman_ℓ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT italic_ρ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ) roman_Λ start_POSTSUPERSCRIPT italic_r end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT , (7)

where d𝑑ditalic_d is the dimension of the Hilbert space of the system. This decomposition shows that the matrices ΛkrsuperscriptsubscriptΛ𝑘𝑟\Lambda_{k}^{r}roman_Λ start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_r end_POSTSUPERSCRIPT are nothing but the excitation modes of the system, each one characterized by a decay rate |Re(λk)|Resubscript𝜆𝑘|{\rm Re}(\lambda_{k})|| roman_Re ( italic_λ start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT ) |. For long times, the relevant terms are those related to the λksubscript𝜆𝑘\lambda_{k}italic_λ start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT with the smallest real part in modulus and finite overlap with the initial state. To study the time-evolution of our systems we order the eigenvalues λksubscript𝜆𝑘\lambda_{k}italic_λ start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT in such a way that |Re(λ2)||Re(λ3)||Re(λm)|Resubscript𝜆2Resubscript𝜆3Resubscript𝜆𝑚|{\rm Re}\left(\lambda_{2}\right)|\leq|{\rm Re}\left(\lambda_{3}\right)|\leq% \ldots\leq|{\rm Re}\left(\lambda_{m}\right)|| roman_Re ( italic_λ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ) | ≤ | roman_Re ( italic_λ start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT ) | ≤ … ≤ | roman_Re ( italic_λ start_POSTSUBSCRIPT italic_m end_POSTSUBSCRIPT ) |. The overlap between the ilimit-from𝑖i-italic_i -th eigenmatrix and the initial state, ρ0subscript𝜌0\rho_{0}italic_ρ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT, is determined by

ξi=Tr(Λiρ0),subscript𝜉𝑖tracesuperscriptsubscriptΛ𝑖subscript𝜌0\xi_{i}=\Tr\left(\Lambda_{i}^{\ell}\rho_{0}\right),italic_ξ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT = roman_Tr ( roman_Λ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT roman_ℓ end_POSTSUPERSCRIPT italic_ρ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ) , (8)

Note that this term ξisubscript𝜉𝑖\xi_{i}italic_ξ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT is the same as the one appearing in the sum presented in Eq. (7). This term will provide us with the influence of the Lindbladian, which fixes the temporal evolution, onto the initial state.

III Quantum Thermal kinematics: Measures of distance and speed

The concept of thermal kinematics, well-established for classical systems recently in Ref. [54], combines arguments from stochastic thermodynamics with information geometry to analyze the thermodynamical processes [56]. For classical systems, it is possible to define a statistical distance [54], related to the classical Fisher information which quantifies the temporal variation of local flows. Therefore, for two time-varying infinitesimal processes, the line element can be defined from the Kullback-Leibler divergence (KLD) of two probability distributions, defined as

Dcl[Pcl(x,t+dt),Pcl(x,t)]=Icl(t)dt2+𝒪(dt4),subscript𝐷clsubscript𝑃cl𝑥𝑡𝑑𝑡subscript𝑃cl𝑥𝑡subscript𝐼cl𝑡𝑑superscript𝑡2𝒪𝑑superscript𝑡4D_{\text{cl}}\left[P_{\text{cl}}(x,t+dt),P_{\text{cl}}(x,t)\right]=I_{\text{cl% }}(t)dt^{2}+\mathcal{O}(dt^{4}),italic_D start_POSTSUBSCRIPT cl end_POSTSUBSCRIPT [ italic_P start_POSTSUBSCRIPT cl end_POSTSUBSCRIPT ( italic_x , italic_t + italic_d italic_t ) , italic_P start_POSTSUBSCRIPT cl end_POSTSUBSCRIPT ( italic_x , italic_t ) ] = italic_I start_POSTSUBSCRIPT cl end_POSTSUBSCRIPT ( italic_t ) italic_d italic_t start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + caligraphic_O ( italic_d italic_t start_POSTSUPERSCRIPT 4 end_POSTSUPERSCRIPT ) , (9)

where Icl(t)subscript𝐼cl𝑡I_{\text{cl}}(t)italic_I start_POSTSUBSCRIPT cl end_POSTSUBSCRIPT ( italic_t ) is the classical Fisher information (FI), and allows us to define a proper statistical distance between two states (see App. A, Eq. (57)). Note that we denote all classical quantities and variables with the subscript clcl\mathrm{cl}roman_cl. The line element is then defined from Eq. (9) as

dlcl:=Icl(t)dt.assign𝑑subscript𝑙clsubscript𝐼cl𝑡𝑑𝑡dl_{\mathrm{cl}}:=\sqrt{I_{\text{cl}}(t)}dt.italic_d italic_l start_POSTSUBSCRIPT roman_cl end_POSTSUBSCRIPT := square-root start_ARG italic_I start_POSTSUBSCRIPT cl end_POSTSUBSCRIPT ( italic_t ) end_ARG italic_d italic_t . (10)

where Icl(t)subscript𝐼cl𝑡\sqrt{I_{\text{cl}}(t)}square-root start_ARG italic_I start_POSTSUBSCRIPT cl end_POSTSUBSCRIPT ( italic_t ) end_ARG can be identified as the statistical velocity at a given time t𝑡titalic_t, namely

vcl(t):=Icl(t).assignsubscript𝑣cl𝑡subscript𝐼cl𝑡v_{\text{cl}}(t):=\sqrt{I_{\text{cl}}(t)}.italic_v start_POSTSUBSCRIPT cl end_POSTSUBSCRIPT ( italic_t ) := square-root start_ARG italic_I start_POSTSUBSCRIPT cl end_POSTSUBSCRIPT ( italic_t ) end_ARG . (11)

To study thermal kinematics in the quantum regime, we may use two different measures. The first one will be the fidelity between two states (analog to the KLD in the classical case), defined as

F(ρ1,ρ2):=Trρ1ρ2ρ1.assign𝐹subscript𝜌1subscript𝜌2tracesubscript𝜌1subscript𝜌2subscript𝜌1F(\rho_{1},\rho_{2}):=\Tr\sqrt{\sqrt{\rho_{1}}\rho_{2}\sqrt{\rho_{1}}}.italic_F ( italic_ρ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT , italic_ρ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ) := roman_Tr square-root start_ARG square-root start_ARG italic_ρ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT end_ARG italic_ρ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT square-root start_ARG italic_ρ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT end_ARG end_ARG . (12)

It measures how close two quantum states are in terms of their density matrix. It is symmetric and invariant under unitary operations. Despite it does not define a distance [108], the fidelity allows us to define a proper metric, the so-called Bures distance

[DB(ρ,σ)]2:=2(1F(ρ,σ)).assignsuperscriptdelimited-[]subscript𝐷𝐵𝜌𝜎221𝐹𝜌𝜎\left[D_{B}(\rho,\sigma)\right]^{2}:=2(1-F(\rho,\sigma)).[ italic_D start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT ( italic_ρ , italic_σ ) ] start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT := 2 ( 1 - italic_F ( italic_ρ , italic_σ ) ) . (13)

Analogous to the classical case, in our context of thermal relaxation, an infinitesimal statistical line element may be defined as follows [109]

[DB(ρ(t),ρ(t+dt))]2=14Q[ρ(t)]dt2+𝒪(dt4),superscriptdelimited-[]subscript𝐷𝐵𝜌𝑡𝜌𝑡𝑑𝑡214subscript𝑄delimited-[]𝜌𝑡𝑑superscript𝑡2𝒪𝑑superscript𝑡4\left[D_{B}\left(\rho(t),\rho(t+dt)\right)\right]^{2}=\frac{1}{4}\mathcal{I}_{% Q}[\rho(t)]dt^{2}+\mathcal{O}(dt^{4}),[ italic_D start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT ( italic_ρ ( italic_t ) , italic_ρ ( italic_t + italic_d italic_t ) ) ] start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT = divide start_ARG 1 end_ARG start_ARG 4 end_ARG caligraphic_I start_POSTSUBSCRIPT italic_Q end_POSTSUBSCRIPT [ italic_ρ ( italic_t ) ] italic_d italic_t start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + caligraphic_O ( italic_d italic_t start_POSTSUPERSCRIPT 4 end_POSTSUPERSCRIPT ) , (14)

being Qsubscript𝑄\mathcal{I}_{Q}caligraphic_I start_POSTSUBSCRIPT italic_Q end_POSTSUBSCRIPT the quantum Fisher information (QFI), with respect to the parameter time, defined as

Q[ρ(t)]:=Tr[Lt2ρ(t)],assignsubscript𝑄delimited-[]𝜌𝑡Trdelimited-[]superscriptsubscript𝐿𝑡2𝜌𝑡\mathcal{I}_{Q}[\rho(t)]:=\text{Tr}\left[L_{t}^{2}\rho(t)\right],caligraphic_I start_POSTSUBSCRIPT italic_Q end_POSTSUBSCRIPT [ italic_ρ ( italic_t ) ] := Tr [ italic_L start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_ρ ( italic_t ) ] , (15)

where Ltsubscript𝐿𝑡L_{t}italic_L start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT is the logarithmic time-derivative operator defined by ρ˙(t):=(Ltρ(t)+ρ(t)Lt)/2assign˙𝜌𝑡subscript𝐿𝑡𝜌𝑡𝜌𝑡subscript𝐿𝑡2\dot{\rho}(t):=\left(L_{t}\rho(t)+\rho(t)L_{t}\right)/2over˙ start_ARG italic_ρ end_ARG ( italic_t ) := ( italic_L start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT italic_ρ ( italic_t ) + italic_ρ ( italic_t ) italic_L start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT ) / 2, see App. A. From Eq. (14), we can directly define the line element as

dl:=14Q[ρ(t)]dt,assign𝑑𝑙14subscript𝑄delimited-[]𝜌𝑡𝑑𝑡dl:=\sqrt{\frac{1}{4}\mathcal{I}_{Q}[\rho(t)]}dt,italic_d italic_l := square-root start_ARG divide start_ARG 1 end_ARG start_ARG 4 end_ARG caligraphic_I start_POSTSUBSCRIPT italic_Q end_POSTSUBSCRIPT [ italic_ρ ( italic_t ) ] end_ARG italic_d italic_t , (16)

and thus

v(t):=14t[ρ(t)]assign𝑣𝑡14subscript𝑡delimited-[]𝜌𝑡v(t):=\sqrt{\frac{1}{4}\mathcal{I}_{t}[\rho(t)]}italic_v ( italic_t ) := square-root start_ARG divide start_ARG 1 end_ARG start_ARG 4 end_ARG caligraphic_I start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT [ italic_ρ ( italic_t ) ] end_ARG (17)

represents the quantum instantaneous statistical velocity of the system in the quantum case. The statistical length of a path taken between time tisubscript𝑡𝑖t_{i}italic_t start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT and tfsubscript𝑡𝑓t_{f}italic_t start_POSTSUBSCRIPT italic_f end_POSTSUBSCRIPT is computed as

(ti,tf)=titf14t[ρ(t)]𝑑t.subscript𝑡𝑖subscript𝑡𝑓superscriptsubscriptsubscript𝑡𝑖subscript𝑡𝑓14subscript𝑡delimited-[]𝜌𝑡differential-d𝑡\ell(t_{i},t_{f})=\int_{t_{i}}^{t_{f}}\sqrt{\frac{1}{4}\mathcal{I}_{t}\left[% \rho(t)\right]}dt.roman_ℓ ( italic_t start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT , italic_t start_POSTSUBSCRIPT italic_f end_POSTSUBSCRIPT ) = ∫ start_POSTSUBSCRIPT italic_t start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_t start_POSTSUBSCRIPT italic_f end_POSTSUBSCRIPT end_POSTSUPERSCRIPT square-root start_ARG divide start_ARG 1 end_ARG start_ARG 4 end_ARG caligraphic_I start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT [ italic_ρ ( italic_t ) ] end_ARG italic_d italic_t . (18)

As reaching the steady state during a dissipative process takes infinite time, to establish a kinematic basis for quantifying thermal relaxation kinematics, we define the quantum degree of completion as

φ(s):=(ti,ts)(ti,tf),assign𝜑𝑠subscript𝑡𝑖subscript𝑡𝑠subscript𝑡𝑖subscript𝑡𝑓\varphi(s):=\dfrac{\ell(t_{i},t_{s})}{\ell(t_{i},t_{f})},italic_φ ( italic_s ) := divide start_ARG roman_ℓ ( italic_t start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT , italic_t start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT ) end_ARG start_ARG roman_ℓ ( italic_t start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT , italic_t start_POSTSUBSCRIPT italic_f end_POSTSUBSCRIPT ) end_ARG , (19)

being a monotonically increasing function bounded between 0 and 1.

IV Heating and cooling protocols

To puzzle out the properties of cooling and heating far from equilibrium in quantum systems subject to instantaneous quenches, we define two possible experiments.

IV.1 Protocol between three temperatures

The first feasible protocol is to compare the free evolution with respect to an intermediate temperature. Hence, we define three temperatures TC<TW<THsubscript𝑇𝐶subscript𝑇𝑊subscript𝑇𝐻T_{C}<T_{W}<T_{H}italic_T start_POSTSUBSCRIPT italic_C end_POSTSUBSCRIPT < italic_T start_POSTSUBSCRIPT italic_W end_POSTSUBSCRIPT < italic_T start_POSTSUBSCRIPT italic_H end_POSTSUBSCRIPT, the subscripts corresponding to cold (C𝐶Citalic_C), warm (W𝑊Witalic_W) and hot (H𝐻Hitalic_H) respectively, as displayed in Fig. 1(a). Associated to this temperatures there are three Gibbs states, ρβith=exp[βiH]/𝒵subscriptsuperscript𝜌thsubscript𝛽𝑖subscript𝛽𝑖𝐻𝒵\rho^{\text{th}}_{\beta_{i}}=\exp[-\beta_{i}H]/\mathcal{Z}italic_ρ start_POSTSUPERSCRIPT th end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_β start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT end_POSTSUBSCRIPT = roman_exp [ - italic_β start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT italic_H ] / caligraphic_Z, with H𝐻Hitalic_H being the Hamiltonian of the system, βi=1/kBTisubscript𝛽𝑖1subscript𝑘𝐵subscript𝑇𝑖\beta_{i}=1/k_{B}T_{i}italic_β start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT = 1 / italic_k start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT italic_T start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT the inverse temperature, and 𝒵=Tr{exp[βiH]}𝒵tracesubscript𝛽𝑖𝐻\mathcal{Z}=\Tr\left\{\exp[-\beta_{i}H]\right\}caligraphic_Z = roman_Tr { roman_exp [ - italic_β start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT italic_H ] } the partition function for i{C,W,H}𝑖𝐶𝑊𝐻i\in\{C,W,H\}italic_i ∈ { italic_C , italic_W , italic_H }.

In this protocol, we can analyze the behavior of our system by the use of the fidelity, Eq. (12). As both trajectories, cooling and heating up, have the same target point, the warm thermal density matrix, we can use the fidelity between our temporal state and the target one as a measure of distance. To fix the initial conditions, we consider thermal states with equal fidelity values with respect to TWsubscript𝑇𝑊T_{W}italic_T start_POSTSUBSCRIPT italic_W end_POSTSUBSCRIPT for both TCsubscript𝑇𝐶T_{C}italic_T start_POSTSUBSCRIPT italic_C end_POSTSUBSCRIPT and THsubscript𝑇𝐻T_{H}italic_T start_POSTSUBSCRIPT italic_H end_POSTSUBSCRIPT, meaning that

F(ρβCth,ρβWth)=F(ρβHth,ρβWth).𝐹subscriptsuperscript𝜌thsubscript𝛽𝐶subscriptsuperscript𝜌thsubscript𝛽𝑊𝐹subscriptsuperscript𝜌thsubscript𝛽𝐻subscriptsuperscript𝜌thsubscript𝛽𝑊F(\rho^{\text{th}}_{\beta_{C}},\rho^{\text{th}}_{\beta_{W}})=F(\rho^{\text{th}% }_{\beta_{H}},\rho^{\text{th}}_{\beta_{W}}).italic_F ( italic_ρ start_POSTSUPERSCRIPT th end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_β start_POSTSUBSCRIPT italic_C end_POSTSUBSCRIPT end_POSTSUBSCRIPT , italic_ρ start_POSTSUPERSCRIPT th end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_β start_POSTSUBSCRIPT italic_W end_POSTSUBSCRIPT end_POSTSUBSCRIPT ) = italic_F ( italic_ρ start_POSTSUPERSCRIPT th end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_β start_POSTSUBSCRIPT italic_H end_POSTSUBSCRIPT end_POSTSUBSCRIPT , italic_ρ start_POSTSUPERSCRIPT th end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_β start_POSTSUBSCRIPT italic_W end_POSTSUBSCRIPT end_POSTSUBSCRIPT ) . (20)

The process starts with the hot and cold temperature and introduce a sudden quench to the warm temperature, monitoring the fidelity evolution. The three temperatures protocol allows us to distinguish which process is faster heating or cooling.

We first focus on what we call forward protocol where the relaxation occurs to the warm temperature, TWsubscript𝑇𝑊T_{W}italic_T start_POSTSUBSCRIPT italic_W end_POSTSUBSCRIPT between the state at hot temperature THsubscript𝑇𝐻T_{H}italic_T start_POSTSUBSCRIPT italic_H end_POSTSUBSCRIPT, and cold temperature, TCsubscript𝑇𝐶T_{C}italic_T start_POSTSUBSCRIPT italic_C end_POSTSUBSCRIPT. What may be even more surprising is that heating also turns out to be faster along the reversed, the so called backward protocol. That is, we prepare the system to be in equilibrium at the warm temperature TWsubscript𝑇𝑊T_{W}italic_T start_POSTSUBSCRIPT italic_W end_POSTSUBSCRIPT and track back the relaxations at TCsubscript𝑇𝐶T_{C}italic_T start_POSTSUBSCRIPT italic_C end_POSTSUBSCRIPT and THsubscript𝑇𝐻T_{H}italic_T start_POSTSUBSCRIPT italic_H end_POSTSUBSCRIPT, respectively. This observation is remarkable as it shows that heating is inherently faster than cooling at TE conditions.

IV.2 Protocol between two temperatures

We can also proceed using a more simple protocol, namely, cooling and heating between two temperatures TC<THsubscript𝑇𝐶subscript𝑇𝐻T_{C}<T_{H}italic_T start_POSTSUBSCRIPT italic_C end_POSTSUBSCRIPT < italic_T start_POSTSUBSCRIPT italic_H end_POSTSUBSCRIPT respectively, as displayed in Fig. 1(b). Measuring with this second protocol the asymmetry between cooling and heating far from equilibrium appears more resounding and stronger. But in this protocol the absence of a reference density matrix prevents us to use the fidelity as a distance measure. We need to use a true metric distance, namely, the quantum Fisher information, Eq. (15), and the so called thermal kinematics [54].

In this scenario, starting from one of the temperatures, after a sudden quench, we let the system evolve freely to the other one. This phenomenon allows us to observe heating, i.e relaxation at THsubscript𝑇𝐻T_{H}italic_T start_POSTSUBSCRIPT italic_H end_POSTSUBSCRIPT in a temperature quench from an equilibrium prepared at TCsubscript𝑇𝐶T_{C}italic_T start_POSTSUBSCRIPT italic_C end_POSTSUBSCRIPT; and the reverse cooling, i.e. relaxation at TCsubscript𝑇𝐶T_{C}italic_T start_POSTSUBSCRIPT italic_C end_POSTSUBSCRIPT in a temperature quench from the equilibrium at THsubscript𝑇𝐻T_{H}italic_T start_POSTSUBSCRIPT italic_H end_POSTSUBSCRIPT. In order to compare the two processes in a proper way we will use the quantum degree of completion given by Eq. (19) and the quantum instantaneous statistical velocity, Eq. (17).

In the following, we test these protocols in three different quantum systems of increasing complexity. The first one is the simplest case, as it only consists of a two-level system coupled to a thermal bath at a given temperature. In this case, all the relevant quantities will be obtained analytically, since the solution for the Lindblad master equation is available exactly. This model will serve as a motivating case to perform an in-depth analysis of the two main models presented in the manuscript: the quantum harmonic oscillator and a quantum Brownian particle. For the harmonic oscillator, since the Hilbert space characterizing the system has infinite dimensions, all the computations performed are potentially more complicated. For this reason, only some of the results are obtained analytically. Finally, the results presented for a quantum Brownian particle are obtained numerically.

V A simple case: Thermal qubit

Let us start with a preliminary analysis of the simplest system of interest: a two-level system, also called a qubit, weakly coupled to a thermal bath. Despite its simplicity, this is a paradigmatic example as the coupling of few level systems to thermal baths has been mastered in the last decades in different condensed matter platforms e.g., semiconductor quantum dots [66, 67] or superconducting qubits [63, 64, 65]. They are important pieces in the development of modern quantum thermodynamic engines [110, 13, 14, 15]. This simple case provides us with analytical understanding of the problem. It is important to remark that all the final conclusions drawn for the more involved examples of the harmonic oscillator and quantum Brownian motion will be in accordance with the ones obtained from this simple analysis.

Consider a two-level system weakly coupled to a thermal bath at inverse temperature β𝛽\betaitalic_β. Transitions between the ground (n=0𝑛0n=0italic_n = 0) and the excited (n=1𝑛1n=1italic_n = 1) states, split by an energy ωPlanck-constant-over-2-pi𝜔\hbar\omegaroman_ℏ italic_ω, occur with rates W10=γn¯(ω,T)subscript𝑊10𝛾¯𝑛𝜔𝑇W_{10}=\gamma\bar{n}(\omega,T)italic_W start_POSTSUBSCRIPT 10 end_POSTSUBSCRIPT = italic_γ over¯ start_ARG italic_n end_ARG ( italic_ω , italic_T ) and W01=γ[1+n¯(ω,T)]subscript𝑊01𝛾delimited-[]1¯𝑛𝜔𝑇W_{01}=\gamma[1+\bar{n}(\omega,T)]italic_W start_POSTSUBSCRIPT 01 end_POSTSUBSCRIPT = italic_γ [ 1 + over¯ start_ARG italic_n end_ARG ( italic_ω , italic_T ) ] induced by the bath [96], with the coupling rate γ𝛾\gammaitalic_γ and an average number of photons with frequency ω𝜔\omegaitalic_ω in a bath at temperature T𝑇Titalic_T, n(ω,T)𝑛𝜔𝑇n(\omega,T)italic_n ( italic_ω , italic_T ), given by the Bose-Einstein distribution

n¯(ω,T)=[exp(ω/kBT)1]1.¯𝑛𝜔𝑇superscriptdelimited-[]Planck-constant-over-2-pi𝜔subscript𝑘𝐵𝑇11\displaystyle\bar{n}(\omega,T)=\left[\exp(\hbar\omega/k_{B}T)-1\right]^{-1}.over¯ start_ARG italic_n end_ARG ( italic_ω , italic_T ) = [ roman_exp ( start_ARG roman_ℏ italic_ω / italic_k start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT italic_T end_ARG ) - 1 ] start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT . (21)

When thermalized, the state of the system can be written as a vector formed by the diagonal elements of the density matrix giving the occupation of the two states, ρ=(ρ00ρ11)T𝜌superscriptsubscript𝜌00subscript𝜌11𝑇\rho=(\rho_{00}\ \rho_{11})^{T}italic_ρ = ( italic_ρ start_POSTSUBSCRIPT 00 end_POSTSUBSCRIPT italic_ρ start_POSTSUBSCRIPT 11 end_POSTSUBSCRIPT ) start_POSTSUPERSCRIPT italic_T end_POSTSUPERSCRIPT in the Fock-Liouville representation. Then, the Lindblad equation, Eq. (2), is a simple rate equation

ρ˙(t)=(γn¯(ω,T)γ[n¯(ω,T)+1]γn¯(ω,T)γ[n¯(ω,T)+1])ρ(t).˙𝜌𝑡𝛾¯𝑛𝜔𝑇𝛾delimited-[]¯𝑛𝜔𝑇1𝛾¯𝑛𝜔𝑇𝛾delimited-[]¯𝑛𝜔𝑇1𝜌𝑡\displaystyle\displaystyle{\dot{\rho}(t)=\left(\begin{array}[]{cc}-\gamma\,% \bar{n}(\omega,T)&\gamma\,[\bar{n}(\omega,T)+1]\\ \gamma\,\bar{n}(\omega,T)&\gamma\,[\bar{n}(\omega,T)+1]\end{array}\right)\rho(% t).}over˙ start_ARG italic_ρ end_ARG ( italic_t ) = ( start_ARRAY start_ROW start_CELL - italic_γ over¯ start_ARG italic_n end_ARG ( italic_ω , italic_T ) end_CELL start_CELL italic_γ [ over¯ start_ARG italic_n end_ARG ( italic_ω , italic_T ) + 1 ] end_CELL end_ROW start_ROW start_CELL italic_γ over¯ start_ARG italic_n end_ARG ( italic_ω , italic_T ) end_CELL start_CELL italic_γ [ over¯ start_ARG italic_n end_ARG ( italic_ω , italic_T ) + 1 ] end_CELL end_ROW end_ARRAY ) italic_ρ ( italic_t ) . (24)

Note that, in the absence of coherence in the initial state, the Hamiltonian term of the Lindblad equation (2) does not contribute and the dynamics is purely dissipative.

We are interested in the relaxation from an initial thermal state at temperature T0=T+ΔT=1/kBβ0subscript𝑇0𝑇Δ𝑇1subscript𝑘Bsubscript𝛽0T_{0}=T+\Delta T=1/k_{\text{B}}\beta_{0}italic_T start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = italic_T + roman_Δ italic_T = 1 / italic_k start_POSTSUBSCRIPT B end_POSTSUBSCRIPT italic_β start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT. The time evolution of the density matrix can be obtained solving Eq. (24)

ρ(t)𝜌𝑡\displaystyle\rho(t)italic_ρ ( italic_t ) =\displaystyle== ρβth+eΓt(eωβeωβ0)(1+eωβ)(1+eωβ0)(11),superscriptsubscript𝜌𝛽thsuperscript𝑒Γ𝑡superscript𝑒Planck-constant-over-2-pi𝜔𝛽superscript𝑒Planck-constant-over-2-pi𝜔subscript𝛽01superscript𝑒Planck-constant-over-2-pi𝜔𝛽1superscript𝑒Planck-constant-over-2-pi𝜔subscript𝛽011\displaystyle\rho_{\beta}^{\rm th}+\frac{e^{-\Gamma t}(e^{\hbar\omega\beta}-e^% {\hbar\omega\beta_{0}})}{(1+e^{\hbar\omega\beta})(1+e^{\hbar\omega\beta_{0}})}% \left(\begin{array}[]{c}-1\\ 1\end{array}\right),italic_ρ start_POSTSUBSCRIPT italic_β end_POSTSUBSCRIPT start_POSTSUPERSCRIPT roman_th end_POSTSUPERSCRIPT + divide start_ARG italic_e start_POSTSUPERSCRIPT - roman_Γ italic_t end_POSTSUPERSCRIPT ( italic_e start_POSTSUPERSCRIPT roman_ℏ italic_ω italic_β end_POSTSUPERSCRIPT - italic_e start_POSTSUPERSCRIPT roman_ℏ italic_ω italic_β start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_POSTSUPERSCRIPT ) end_ARG start_ARG ( 1 + italic_e start_POSTSUPERSCRIPT roman_ℏ italic_ω italic_β end_POSTSUPERSCRIPT ) ( 1 + italic_e start_POSTSUPERSCRIPT roman_ℏ italic_ω italic_β start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_POSTSUPERSCRIPT ) end_ARG ( start_ARRAY start_ROW start_CELL - 1 end_CELL end_ROW start_ROW start_CELL 1 end_CELL end_ROW end_ARRAY ) , (27)

with the total rate Γ=γ(1+2n¯(ω,T))=γcoth(ωβ/2)Γ𝛾12¯𝑛𝜔𝑇𝛾hyperbolic-cotangentPlanck-constant-over-2-pi𝜔𝛽2\Gamma=\gamma(1+2\bar{n}(\omega,T))=\gamma\coth(\hbar\omega\beta/2)roman_Γ = italic_γ ( 1 + 2 over¯ start_ARG italic_n end_ARG ( italic_ω , italic_T ) ) = italic_γ roman_coth ( start_ARG roman_ℏ italic_ω italic_β / 2 end_ARG ), that is is proportional to the thermal fluctuations of the bath. The fact that there is a single relaxation channel in this case emphasizes the role of the bath fluctuations (larger for higher temperatures). It is hence clear that the evolution will be faster when the system relaxes to a hotter steady state.

Refer to caption
Figure 2: Thermalization kinematics for a qubit. (a) Dependence of the decaying mode corresponding to the non-zero eigenvalue. (b) Overlap of a state thermalized at a temperature T+ΔT𝑇Δ𝑇T+\Delta Titalic_T + roman_Δ italic_T with the stationary state at a temperature T𝑇Titalic_T, for different values of ΔT/TΔ𝑇𝑇\Delta T/Troman_Δ italic_T / italic_T. The black line in (b) corresponds to the asymptotic behaviour at large ΔTΔ𝑇\Delta Troman_Δ italic_T.
Refer to caption
Figure 3: Three-temperatures protocol for a qubit. (a) Bures distance as a function of ΔTΔ𝑇\Delta Troman_Δ italic_T. The blue and red dots mark the equidistant temperatures considered for the cooling and heating protocols, respectively. These are used to compute (b) the Fidelity and (c) the velocity as function of time, and (d) the level of completion for a given time renormalized by tfin=2/γsubscript𝑡fin2𝛾t_{\rm fin}=2/\gammaitalic_t start_POSTSUBSCRIPT roman_fin end_POSTSUBSCRIPT = 2 / italic_γ. In all panels ω=1Planck-constant-over-2-pi𝜔1\hbar\omega=1roman_ℏ italic_ω = 1, TW=ω/2kBsubscript𝑇𝑊Planck-constant-over-2-pi𝜔2subscript𝑘BT_{W}=\hbar\omega/2k_{\text{B}}italic_T start_POSTSUBSCRIPT italic_W end_POSTSUBSCRIPT = roman_ℏ italic_ω / 2 italic_k start_POSTSUBSCRIPT B end_POSTSUBSCRIPT, and ΔT=THTW=ω/kBΔ𝑇subscript𝑇𝐻subscript𝑇𝑊Planck-constant-over-2-pi𝜔subscript𝑘B\Delta T=T_{H}-T_{W}=\hbar\omega/k_{\text{B}}roman_Δ italic_T = italic_T start_POSTSUBSCRIPT italic_H end_POSTSUBSCRIPT - italic_T start_POSTSUBSCRIPT italic_W end_POSTSUBSCRIPT = roman_ℏ italic_ω / italic_k start_POSTSUBSCRIPT B end_POSTSUBSCRIPT, with TC0.70ωsubscript𝑇𝐶0.70Planck-constant-over-2-pi𝜔T_{C}\approx 0.70\hbar\omegaitalic_T start_POSTSUBSCRIPT italic_C end_POSTSUBSCRIPT ≈ 0.70 roman_ℏ italic_ω chosen to be equidistant from TWsubscript𝑇𝑊T_{W}italic_T start_POSTSUBSCRIPT italic_W end_POSTSUBSCRIPT as depicted in (a). For the time evolution t0=0subscript𝑡00t_{0}=0italic_t start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = 0 in all cases. The dashed line in (a) is a quadratic expansion of the Bures distance around ΔT=0Δ𝑇0\Delta T=0roman_Δ italic_T = 0, according to Eq. (56).

It is however convenient to look further into the details of the dynamics, as introduced in Sec. II, since it allows for an analytical treatment. We start by obtaining the eigenvalues of \mathcal{L}caligraphic_L. In this simple case, the spectrum is reduced to only two values: λ1=0subscript𝜆10\lambda_{1}=0italic_λ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT = 0 related to the trivial stationary state, and λ2=Γsubscript𝜆2Γ\lambda_{2}=-\Gammaitalic_λ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT = - roman_Γ, which takes into account the decay mode and depends on the bath parameters, contained in n¯(ω,T)¯𝑛𝜔𝑇\bar{n}(\omega,T)over¯ start_ARG italic_n end_ARG ( italic_ω , italic_T ) and in the coupling γ𝛾\gammaitalic_γ. The temperature dependence of the decaying mode corresponding to λ2subscript𝜆2\lambda_{2}italic_λ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT is plotted in Fig. 2(a). Their corresponding (right) eigenvectors, see Eq. (5), are given by

Λ1r=1eβω+1(eβω1),superscriptsubscriptΛ1𝑟1superscript𝑒𝛽Planck-constant-over-2-pi𝜔1superscript𝑒𝛽Planck-constant-over-2-pi𝜔1\displaystyle\displaystyle{\Lambda_{1}^{r}=\frac{1}{e^{\beta\hbar\omega}+1}% \left(\begin{array}[]{c}e^{\beta\hbar\omega}\\ 1\end{array}\right),}roman_Λ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_r end_POSTSUPERSCRIPT = divide start_ARG 1 end_ARG start_ARG italic_e start_POSTSUPERSCRIPT italic_β roman_ℏ italic_ω end_POSTSUPERSCRIPT + 1 end_ARG ( start_ARRAY start_ROW start_CELL italic_e start_POSTSUPERSCRIPT italic_β roman_ℏ italic_ω end_POSTSUPERSCRIPT end_CELL end_ROW start_ROW start_CELL 1 end_CELL end_ROW end_ARRAY ) , (30)

corresponding to the stationary state of the system, characterized by λ1=0subscript𝜆10\lambda_{1}=0italic_λ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT = 0, and

Λ2r=(11),superscriptsubscriptΛ2𝑟11\displaystyle\displaystyle{\Lambda_{2}^{r}=\left(\begin{array}[]{c}1\\ -1\end{array}\right),}roman_Λ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_r end_POSTSUPERSCRIPT = ( start_ARRAY start_ROW start_CELL 1 end_CELL end_ROW start_ROW start_CELL - 1 end_CELL end_ROW end_ARRAY ) , (33)

which is related to the decaying mode, λ2=Γsubscript𝜆2Γ\lambda_{2}=-\Gammaitalic_λ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT = - roman_Γ. These results agree with the exact evolution obtained in Eq. (27).

To analyze the dynamics and the influence of the initial state, we need to compute the overlap between Λ2superscriptsubscriptΛ2\Lambda_{2}^{\ell}roman_Λ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT roman_ℓ end_POSTSUPERSCRIPT and the initial state, see Eq. (8). As we are starting from thermal equilibrium, the initial state coincides with the stationary state at temperature T0=T+ΔTsubscript𝑇0𝑇Δ𝑇T_{0}=T+\Delta Titalic_T start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = italic_T + roman_Δ italic_T, meaning that ρ0=Λ1r(T+ΔT)subscript𝜌0superscriptsubscriptΛ1𝑟𝑇Δ𝑇\rho_{0}=\Lambda_{1}^{r}(T+\Delta T)italic_ρ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = roman_Λ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_r end_POSTSUPERSCRIPT ( italic_T + roman_Δ italic_T ). Then, using

Λ2=11+eωβ(1eωβ),superscriptsubscriptΛ211superscript𝑒Planck-constant-over-2-pi𝜔𝛽1superscript𝑒Planck-constant-over-2-pi𝜔𝛽\Lambda_{2}^{\ell}=\frac{1}{1+e^{\hbar\omega\beta}}(1\ -e^{\hbar\omega\beta}),roman_Λ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT roman_ℓ end_POSTSUPERSCRIPT = divide start_ARG 1 end_ARG start_ARG 1 + italic_e start_POSTSUPERSCRIPT roman_ℏ italic_ω italic_β end_POSTSUPERSCRIPT end_ARG ( 1 - italic_e start_POSTSUPERSCRIPT roman_ℏ italic_ω italic_β end_POSTSUPERSCRIPT ) , (34)

the overlap is

ξ=Tr(Λ2ρ0(T))=eβ0ωeβω(eβ0ω+1)(eβω+1).𝜉tracesuperscriptsubscriptΛ2subscript𝜌0𝑇superscript𝑒subscript𝛽0Planck-constant-over-2-pi𝜔superscript𝑒𝛽Planck-constant-over-2-pi𝜔superscript𝑒subscript𝛽0Planck-constant-over-2-pi𝜔1superscript𝑒𝛽Planck-constant-over-2-pi𝜔1\xi=\Tr(\Lambda_{2}^{\ell}\rho_{0}(T))=\frac{e^{\beta_{0}\hbar\omega}-e^{\beta% \hbar\omega}}{\left(e^{\beta_{0}\hbar\omega}+1\right)\left(e^{\beta\hbar\omega% }+1\right)}.italic_ξ = roman_Tr ( start_ARG roman_Λ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT roman_ℓ end_POSTSUPERSCRIPT italic_ρ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ( italic_T ) end_ARG ) = divide start_ARG italic_e start_POSTSUPERSCRIPT italic_β start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT roman_ℏ italic_ω end_POSTSUPERSCRIPT - italic_e start_POSTSUPERSCRIPT italic_β roman_ℏ italic_ω end_POSTSUPERSCRIPT end_ARG start_ARG ( italic_e start_POSTSUPERSCRIPT italic_β start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT roman_ℏ italic_ω end_POSTSUPERSCRIPT + 1 ) ( italic_e start_POSTSUPERSCRIPT italic_β roman_ℏ italic_ω end_POSTSUPERSCRIPT + 1 ) end_ARG . (35)

Note that the overlap has the same modulus under the exchange of temperatures T𝑇Titalic_T and T0subscript𝑇0T_{0}italic_T start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT. Hence, in a two-temperatures protocol, the overlap is the same in both ways (cooling and heating): |ξ(TH,TC)|=|ξ(TC,TH)|𝜉subscript𝑇𝐻subscript𝑇𝐶𝜉subscript𝑇𝐶subscript𝑇𝐻|\xi(T_{H},T_{C})|=|\xi(T_{C},T_{H})|| italic_ξ ( italic_T start_POSTSUBSCRIPT italic_H end_POSTSUBSCRIPT , italic_T start_POSTSUBSCRIPT italic_C end_POSTSUBSCRIPT ) | = | italic_ξ ( italic_T start_POSTSUBSCRIPT italic_C end_POSTSUBSCRIPT , italic_T start_POSTSUBSCRIPT italic_H end_POSTSUBSCRIPT ) |. As a consequence, any asymmetry in this protocol is to be attributed only to the monotonous increase of the coupling rate shown in Fig. 2(a): γ𝛾\gammaitalic_γ is larger when relaxing to a hot bath, so |λ2H|>|λ2C|superscriptsubscript𝜆2𝐻superscriptsubscript𝜆2𝐶|\lambda_{2}^{H}|>|\lambda_{2}^{C}|| italic_λ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_H end_POSTSUPERSCRIPT | > | italic_λ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_C end_POSTSUPERSCRIPT |. As shown in Fig. 2(b), the overlap increases monotonically with ΔTΔ𝑇\Delta Troman_Δ italic_T, i.e., far from equilibrium states are more strongly overlapped. Remarkably there is a suppression in |ξ|𝜉|\xi|| italic_ξ | for low temperature states that agree with Nernst’s unattainability principle [11]: the gap in the overlap at low T𝑇Titalic_T and low ΔTΔ𝑇\Delta Troman_Δ italic_T prevents a cold system to be further cooled down, cf. the blue curve in Fig. 2(b). Note also that this Nernst gap is accompanied by a low-temperature plateau of the decaying rate, with |λ2|γsubscript𝜆2𝛾|\lambda_{2}|\to\gamma| italic_λ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT | → italic_γ, see Fig. 2(a). In the opposite case, the overlap of an infinite and a zero temperature states is maximal: |ξ|1/5𝜉15|\xi|\to 1/5| italic_ξ | → 1 / 5, with the bound |ξ(T+ΔT,T)||ξasym|𝜉𝑇Δ𝑇𝑇subscript𝜉asym|\xi(T+\Delta T,T)|\leq|\xi_{\rm asym}|| italic_ξ ( italic_T + roman_Δ italic_T , italic_T ) | ≤ | italic_ξ start_POSTSUBSCRIPT roman_asym end_POSTSUBSCRIPT | for the asymptotic value

|ξasym|12tanh(ω2kBT)subscript𝜉asym12Planck-constant-over-2-pi𝜔2subscript𝑘𝐵𝑇|\xi_{\rm asym}|\to\frac{1}{2}\tanh\left(\frac{\hbar\omega}{2k_{B}T}\right)| italic_ξ start_POSTSUBSCRIPT roman_asym end_POSTSUBSCRIPT | → divide start_ARG 1 end_ARG start_ARG 2 end_ARG roman_tanh ( divide start_ARG roman_ℏ italic_ω end_ARG start_ARG 2 italic_k start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT italic_T end_ARG ) (36)

when ΔTTmuch-greater-thanΔ𝑇𝑇\Delta T\gg Troman_Δ italic_T ≫ italic_T, see Fig. 2(b).

To compute the fidelity as a measure of distance between two thermal states, let the system be described by a Gibbs state at an inverse temperature β𝛽\betaitalic_β and frequency ω𝜔\omegaitalic_ω in the Fock-Liouville space, ρβthsubscriptsuperscript𝜌th𝛽\rho^{\text{th}}_{\beta}italic_ρ start_POSTSUPERSCRIPT th end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_β end_POSTSUBSCRIPT given by Eq. (30). Consider two thermal states at different inverse temperatures β1subscript𝛽1\beta_{1}italic_β start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT and β2subscript𝛽2\beta_{2}italic_β start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT. In this case, the fidelity is simply

F(ρβ1th,ρβ2th)=1+eω(β1+β2)/2[(1+eβ1ω)(1+eβ2ω)]1/2.𝐹superscriptsubscript𝜌subscript𝛽1thsuperscriptsubscript𝜌subscript𝛽2th1superscript𝑒Planck-constant-over-2-pi𝜔subscript𝛽1subscript𝛽22superscriptdelimited-[]1superscript𝑒subscript𝛽1Planck-constant-over-2-pi𝜔1superscript𝑒subscript𝛽2Planck-constant-over-2-pi𝜔12F(\rho_{\beta_{1}}^{\rm th},\rho_{\beta_{2}}^{\rm th})=\frac{1+e^{\hbar\omega(% \beta_{1}+\beta_{2})/2}}{[(1+e^{\beta_{1}\hbar\omega})(1+e^{\beta_{2}\hbar% \omega})]^{1/2}}.italic_F ( italic_ρ start_POSTSUBSCRIPT italic_β start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT end_POSTSUBSCRIPT start_POSTSUPERSCRIPT roman_th end_POSTSUPERSCRIPT , italic_ρ start_POSTSUBSCRIPT italic_β start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT end_POSTSUBSCRIPT start_POSTSUPERSCRIPT roman_th end_POSTSUPERSCRIPT ) = divide start_ARG 1 + italic_e start_POSTSUPERSCRIPT roman_ℏ italic_ω ( italic_β start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT + italic_β start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ) / 2 end_POSTSUPERSCRIPT end_ARG start_ARG [ ( 1 + italic_e start_POSTSUPERSCRIPT italic_β start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT roman_ℏ italic_ω end_POSTSUPERSCRIPT ) ( 1 + italic_e start_POSTSUPERSCRIPT italic_β start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT roman_ℏ italic_ω end_POSTSUPERSCRIPT ) ] start_POSTSUPERSCRIPT 1 / 2 end_POSTSUPERSCRIPT end_ARG . (37)

With this expression, we calculate the Bures distance DB2=2[1F(ρβ1th,ρβ2th)]subscriptsuperscript𝐷2𝐵2delimited-[]1𝐹superscriptsubscript𝜌subscript𝛽1thsuperscriptsubscript𝜌subscript𝛽2thD^{2}_{B}=2[1-F(\rho_{\beta_{1}}^{\rm th},\rho_{\beta_{2}}^{\rm th})]italic_D start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT = 2 [ 1 - italic_F ( italic_ρ start_POSTSUBSCRIPT italic_β start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT end_POSTSUBSCRIPT start_POSTSUPERSCRIPT roman_th end_POSTSUPERSCRIPT , italic_ρ start_POSTSUBSCRIPT italic_β start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT end_POSTSUBSCRIPT start_POSTSUPERSCRIPT roman_th end_POSTSUPERSCRIPT ) ], see Eq. (13), which we plot in Fig. 3(a). In Sec. VI.3 a similar analysis of the Bures distance and its dependence with the temperature difference is made for the harmonic oscillator and quantum Brownian particle cases.

As the density matrix time evolution is given by Eq. (27), we can find an analytical expression for the time evolution of the fidelity of a system initially at a state ρβ0thsuperscriptsubscript𝜌subscript𝛽0th\rho_{\beta_{0}}^{\rm th}italic_ρ start_POSTSUBSCRIPT italic_β start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_POSTSUBSCRIPT start_POSTSUPERSCRIPT roman_th end_POSTSUPERSCRIPT with respect to the stationary state as it is put in contact with a bath at inverse temperature β𝛽\betaitalic_β, namely

F[ρ(t),ρβth]=eωβ[eωβ0+Aββ0(t)]+1Aββ0(t)(1+eωβ)(1+eωβ0),𝐹𝜌𝑡superscriptsubscript𝜌𝛽thsuperscript𝑒Planck-constant-over-2-pi𝜔𝛽delimited-[]superscript𝑒Planck-constant-over-2-pi𝜔subscript𝛽0subscript𝐴𝛽subscript𝛽0𝑡1subscript𝐴𝛽subscript𝛽0𝑡1superscript𝑒Planck-constant-over-2-pi𝜔𝛽1superscript𝑒Planck-constant-over-2-pi𝜔subscript𝛽0F[\rho(t),\rho_{\beta}^{\rm th}]=\frac{\sqrt{e^{\hbar\omega\beta}[e^{\hbar% \omega\beta_{0}}+A_{\beta\beta_{0}}(t)]}+\sqrt{1-A_{\beta\beta_{0}}(t)}}{(1+e^% {\hbar\omega\beta})(1+e^{\hbar\omega\beta_{0}})},italic_F [ italic_ρ ( italic_t ) , italic_ρ start_POSTSUBSCRIPT italic_β end_POSTSUBSCRIPT start_POSTSUPERSCRIPT roman_th end_POSTSUPERSCRIPT ] = divide start_ARG square-root start_ARG italic_e start_POSTSUPERSCRIPT roman_ℏ italic_ω italic_β end_POSTSUPERSCRIPT [ italic_e start_POSTSUPERSCRIPT roman_ℏ italic_ω italic_β start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_POSTSUPERSCRIPT + italic_A start_POSTSUBSCRIPT italic_β italic_β start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_POSTSUBSCRIPT ( italic_t ) ] end_ARG + square-root start_ARG 1 - italic_A start_POSTSUBSCRIPT italic_β italic_β start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_POSTSUBSCRIPT ( italic_t ) end_ARG end_ARG start_ARG ( 1 + italic_e start_POSTSUPERSCRIPT roman_ℏ italic_ω italic_β end_POSTSUPERSCRIPT ) ( 1 + italic_e start_POSTSUPERSCRIPT roman_ℏ italic_ω italic_β start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_POSTSUPERSCRIPT ) end_ARG , (38)

where the time dependence is encapsulated in the term Aββ0(t)(1eΓt)(eβωeβ0ω)/(1+eβω)subscript𝐴𝛽subscript𝛽0𝑡1superscript𝑒Γ𝑡superscript𝑒𝛽Planck-constant-over-2-pi𝜔superscript𝑒subscript𝛽0Planck-constant-over-2-pi𝜔1superscript𝑒𝛽Planck-constant-over-2-pi𝜔A_{\beta\beta_{0}}(t)\equiv{(1-e^{-\Gamma t}})(e^{\beta\hbar\omega}-e^{\beta_{% 0}\hbar\omega})/(1+e^{\beta\hbar\omega})italic_A start_POSTSUBSCRIPT italic_β italic_β start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_POSTSUBSCRIPT ( italic_t ) ≡ ( 1 - italic_e start_POSTSUPERSCRIPT - roman_Γ italic_t end_POSTSUPERSCRIPT ) ( italic_e start_POSTSUPERSCRIPT italic_β roman_ℏ italic_ω end_POSTSUPERSCRIPT - italic_e start_POSTSUPERSCRIPT italic_β start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT roman_ℏ italic_ω end_POSTSUPERSCRIPT ) / ( 1 + italic_e start_POSTSUPERSCRIPT italic_β roman_ℏ italic_ω end_POSTSUPERSCRIPT ). Note that at time t=0𝑡0t=0italic_t = 0, A(0)=0𝐴00A(0)=0italic_A ( 0 ) = 0, thereby recovering the fidelity given by Eq. (37). The fidelity as a function of time is plotted in Fig. 3(b) for the heating up and cooling down processes. This result confirming that the heating protocol is faster than the cooling one. We have verified that this effect is much stronger for low temperatures.

Even in this simple case the dynamics are less trivial than what one may suspect from the above discussion. This is revealed by using the quantum Fisher information to compute the thermal kinematic distance and speed, see Eqs. (16) and (17). In this case, being the density matrix diagonal, and tρ00(t)=tρ11(t)subscript𝑡subscript𝜌00𝑡subscript𝑡subscript𝜌11𝑡\partial_{t}\rho_{00}(t)=\partial_{t}\rho_{11}(t)∂ start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT italic_ρ start_POSTSUBSCRIPT 00 end_POSTSUBSCRIPT ( italic_t ) = ∂ start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT italic_ρ start_POSTSUBSCRIPT 11 end_POSTSUBSCRIPT ( italic_t ), the QFI reads Q[ρ(t)]=[tρ00(t)]2/ρ00(t)ρ11(t)subscript𝑄delimited-[]𝜌𝑡superscriptdelimited-[]subscript𝑡subscript𝜌00𝑡2subscript𝜌00𝑡subscript𝜌11𝑡{\cal I}_{Q}[\rho(t)]=[\partial_{t}\rho_{00}(t)]^{2}/\rho_{00}(t)\rho_{11}(t)caligraphic_I start_POSTSUBSCRIPT italic_Q end_POSTSUBSCRIPT [ italic_ρ ( italic_t ) ] = [ ∂ start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT italic_ρ start_POSTSUBSCRIPT 00 end_POSTSUBSCRIPT ( italic_t ) ] start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT / italic_ρ start_POSTSUBSCRIPT 00 end_POSTSUBSCRIPT ( italic_t ) italic_ρ start_POSTSUBSCRIPT 11 end_POSTSUBSCRIPT ( italic_t ), where we have also used ρ00(t)+ρ11(t)=1subscript𝜌00𝑡subscript𝜌11𝑡1\rho_{00}(t)+\rho_{11}(t)=1italic_ρ start_POSTSUBSCRIPT 00 end_POSTSUBSCRIPT ( italic_t ) + italic_ρ start_POSTSUBSCRIPT 11 end_POSTSUBSCRIPT ( italic_t ) = 1, leading to

Q[ρ(t)]=Γ2[eωβκ(t)1][κ(t)+1],subscript𝑄delimited-[]𝜌𝑡superscriptΓ2delimited-[]superscript𝑒Planck-constant-over-2-pi𝜔𝛽𝜅𝑡1delimited-[]𝜅𝑡1{\mathcal{I}}_{Q}[\rho(t)]=\frac{\Gamma^{2}}{[e^{\hbar\omega\beta}\kappa(t)-1]% [\kappa(t)+1]},caligraphic_I start_POSTSUBSCRIPT italic_Q end_POSTSUBSCRIPT [ italic_ρ ( italic_t ) ] = divide start_ARG roman_Γ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG [ italic_e start_POSTSUPERSCRIPT roman_ℏ italic_ω italic_β end_POSTSUPERSCRIPT italic_κ ( italic_t ) - 1 ] [ italic_κ ( italic_t ) + 1 ] end_ARG , (39)

where the time-dependence is encapsulated in the term

κ(t)1+eωβ0eωβeωβ0eΓt,𝜅𝑡1superscript𝑒Planck-constant-over-2-pi𝜔subscript𝛽0superscript𝑒Planck-constant-over-2-pi𝜔𝛽superscript𝑒Planck-constant-over-2-pi𝜔subscript𝛽0superscript𝑒Γ𝑡\kappa(t)\equiv\frac{1+e^{\hbar\omega\beta_{0}}}{e^{\hbar\omega\beta}-e^{\hbar% \omega\beta_{0}}}e^{\Gamma t},italic_κ ( italic_t ) ≡ divide start_ARG 1 + italic_e start_POSTSUPERSCRIPT roman_ℏ italic_ω italic_β start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_POSTSUPERSCRIPT end_ARG start_ARG italic_e start_POSTSUPERSCRIPT roman_ℏ italic_ω italic_β end_POSTSUPERSCRIPT - italic_e start_POSTSUPERSCRIPT roman_ℏ italic_ω italic_β start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_POSTSUPERSCRIPT end_ARG italic_e start_POSTSUPERSCRIPT roman_Γ italic_t end_POSTSUPERSCRIPT , (40)

and the instantaneous statistical velocity is

v(t)=Γ/2eωβκ(t)1κ(t)+1.𝑣𝑡Γ2superscript𝑒Planck-constant-over-2-pi𝜔𝛽𝜅𝑡1𝜅𝑡1v(t)=\frac{\Gamma/2}{\sqrt{e^{\hbar\omega\beta}\kappa(t)-1}\sqrt{\kappa(t)+1}}.italic_v ( italic_t ) = divide start_ARG roman_Γ / 2 end_ARG start_ARG square-root start_ARG italic_e start_POSTSUPERSCRIPT roman_ℏ italic_ω italic_β end_POSTSUPERSCRIPT italic_κ ( italic_t ) - 1 end_ARG square-root start_ARG italic_κ ( italic_t ) + 1 end_ARG end_ARG . (41)

This expression is displayed in Fig. 3(c). The velocity of the process is initially larger for the heating mechanism, confirming our hypothesis, but eventually it becomes slower than the cooling one. It can be concluded that as the system gets closer to thermalization the dynamics get slower. To analyze the full process, we also compute the statistical length

(t0,t)=1Γ{arctan[|(eωβ1)κ(t)22[eωβκ(t)1][κ(t)+1]|]arctan[|(eωβ1)κ(t0)22[eωβκ(t0)1][κ(t0)+1]|]}.\displaystyle\begin{aligned} \ell(t_{0},t)&=\frac{1}{\Gamma}\left\{\arctan% \left[\left|\frac{(e^{\hbar\omega\beta}{-}1)\kappa(t)-2}{2\sqrt{[e^{\hbar% \omega\beta}\kappa(t){-}1][\kappa(t){+}1]}}\right|\right]\right.\\ &-\left.\arctan\left[\left|\frac{(e^{\hbar\omega\beta}-1)\kappa(t_{0})-2}{2% \sqrt{[e^{\hbar\omega\beta}\kappa(t_{0}){-}1][\kappa(t_{0}){+}1]}}\right|% \right]\right\}.\end{aligned}start_ROW start_CELL roman_ℓ ( italic_t start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT , italic_t ) end_CELL start_CELL = divide start_ARG 1 end_ARG start_ARG roman_Γ end_ARG { roman_arctan [ | divide start_ARG ( italic_e start_POSTSUPERSCRIPT roman_ℏ italic_ω italic_β end_POSTSUPERSCRIPT - 1 ) italic_κ ( italic_t ) - 2 end_ARG start_ARG 2 square-root start_ARG [ italic_e start_POSTSUPERSCRIPT roman_ℏ italic_ω italic_β end_POSTSUPERSCRIPT italic_κ ( italic_t ) - 1 ] [ italic_κ ( italic_t ) + 1 ] end_ARG end_ARG | ] end_CELL end_ROW start_ROW start_CELL end_CELL start_CELL - roman_arctan [ | divide start_ARG ( italic_e start_POSTSUPERSCRIPT roman_ℏ italic_ω italic_β end_POSTSUPERSCRIPT - 1 ) italic_κ ( italic_t start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ) - 2 end_ARG start_ARG 2 square-root start_ARG [ italic_e start_POSTSUPERSCRIPT roman_ℏ italic_ω italic_β end_POSTSUPERSCRIPT italic_κ ( italic_t start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ) - 1 ] [ italic_κ ( italic_t start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ) + 1 ] end_ARG end_ARG | ] } . end_CELL end_ROW (42)

with which we plot the ratio φ(t)=(0,t)/(0,tfin)𝜑𝑡0𝑡0subscript𝑡fin\varphi(t)=\ell(0,t)/\ell(0,t_{\rm fin})italic_φ ( italic_t ) = roman_ℓ ( 0 , italic_t ) / roman_ℓ ( 0 , italic_t start_POSTSUBSCRIPT roman_fin end_POSTSUBSCRIPT ) in Fig. 3(d). Due to the initial higher velocity for the heating process, the degree of completion is larger at all times for the heating protocol.

Refer to caption
Figure 4: Simulation of the protocols for the harmonic oscillator. (a) Bures distance as a function to the temperature difference with respect to the equilibrium state at TWsubscript𝑇𝑊T_{W}italic_T start_POSTSUBSCRIPT italic_W end_POSTSUBSCRIPT. The blue and red dots mark the initial temperatures TCsubscript𝑇𝐶T_{C}italic_T start_POSTSUBSCRIPT italic_C end_POSTSUBSCRIPT and THsubscript𝑇𝐻T_{H}italic_T start_POSTSUBSCRIPT italic_H end_POSTSUBSCRIPT for the cooling and heating protocols, respectively. (b) Fidelity for the state of the system with respect to the thermal state at TWsubscript𝑇𝑊T_{W}italic_T start_POSTSUBSCRIPT italic_W end_POSTSUBSCRIPT. The process corresponds to the three-temperature protocol, evolving from TCsubscript𝑇𝐶T_{C}italic_T start_POSTSUBSCRIPT italic_C end_POSTSUBSCRIPT to TWsubscript𝑇𝑊T_{W}italic_T start_POSTSUBSCRIPT italic_W end_POSTSUBSCRIPT (orange line) and THsubscript𝑇𝐻T_{H}italic_T start_POSTSUBSCRIPT italic_H end_POSTSUBSCRIPT to TWsubscript𝑇𝑊T_{W}italic_T start_POSTSUBSCRIPT italic_W end_POSTSUBSCRIPT (blue line). (c) Instantaneous statistical velocity for the harmonic oscillator computed for the heating and cooling protocol in the two-temperature scheme. (d) Degree of completion for the harmonic oscillator in the two-temperature scheme. In all plots, the temperature ranges are such that n(ω,TC)=1𝑛𝜔subscript𝑇𝐶1n(\omega,T_{C})=1italic_n ( italic_ω , italic_T start_POSTSUBSCRIPT italic_C end_POSTSUBSCRIPT ) = 1 and n(ω,TH)=10𝑛𝜔subscript𝑇𝐻10n(\omega,T_{H})=10italic_n ( italic_ω , italic_T start_POSTSUBSCRIPT italic_H end_POSTSUBSCRIPT ) = 10, the coupling parameter is γ=0.1𝛾0.1\gamma=0.1italic_γ = 0.1, ω=1Planck-constant-over-2-pi𝜔1\hbar\omega=1roman_ℏ italic_ω = 1, and tfin=50subscript𝑡fin50t_{\textrm{fin}}=50italic_t start_POSTSUBSCRIPT fin end_POSTSUBSCRIPT = 50.

VI Increasing complexity

VI.1 Quantum harmonic oscillator

After having introduced the main concepts presented in the manuscript with a clear and analytically solvable case, we will perform a similar analysis for more complicated and richer systems. The quantum harmonic oscillator allows us to derive some analytical expressions for the behavior of the system but, on the other hand, numerical methods are required to compute quantities of interest such as the quantum speed through the quantum Fisher information.

A harmonic oscillator is described by the following Hamiltonian

H𝐻\displaystyle Hitalic_H =\displaystyle== ωaa,Planck-constant-over-2-pi𝜔superscript𝑎𝑎\displaystyle\hbar\omega a^{\dagger}a,roman_ℏ italic_ω italic_a start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT italic_a , (43)

where Planck-constant-over-2-pi\hbarroman_ℏ is the Planck’s constant, ω𝜔\omegaitalic_ω is the oscillator frequency and a,a𝑎superscript𝑎a,a^{\dagger}italic_a , italic_a start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT are the annihilation and creation bosonic operators, respectively. The interaction with a thermal bath is described by the jump operators

L+=γn¯(ω,T)a,L=γ[n¯(ω,T)+1]a,subscript𝐿absent𝛾¯𝑛𝜔𝑇superscript𝑎subscript𝐿absent𝛾delimited-[]¯𝑛𝜔𝑇1𝑎\displaystyle\begin{aligned} L_{+}&=\sqrt{\gamma\,\bar{n}(\omega,T)}\;a^{% \dagger},\\ L_{-}&=\sqrt{\gamma\,[\bar{n}(\omega,T)+1]}\;a,\end{aligned}start_ROW start_CELL italic_L start_POSTSUBSCRIPT + end_POSTSUBSCRIPT end_CELL start_CELL = square-root start_ARG italic_γ over¯ start_ARG italic_n end_ARG ( italic_ω , italic_T ) end_ARG italic_a start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT , end_CELL end_ROW start_ROW start_CELL italic_L start_POSTSUBSCRIPT - end_POSTSUBSCRIPT end_CELL start_CELL = square-root start_ARG italic_γ [ over¯ start_ARG italic_n end_ARG ( italic_ω , italic_T ) + 1 ] end_ARG italic_a , end_CELL end_ROW (44)

where γ𝛾\gammaitalic_γ is the coupling strength with the bath, and n(ω,T)𝑛𝜔𝑇n(\omega,T)italic_n ( italic_ω , italic_T ) is the average number of excitations in the bath at a given temperature T𝑇Titalic_T as in the two-level system case, see Eq. (21). The state of a thermal harmonic oscillator is simply determined by its average number of excitations, n(ω,T)𝑛𝜔𝑇n(\omega,T)italic_n ( italic_ω , italic_T ), as

ρβth=n=0[n¯((ω,T)]n[1+n¯((ω,T)]n+1|nn|=n=0enωβ(1eωβ)|nn|,\displaystyle\begin{aligned} \rho^{\text{th}}_{\beta}&=\sum_{n=0}^{\infty}% \dfrac{\left[\bar{n}((\omega,T)\right]^{n}}{[1+\bar{n}((\omega,T)]^{n+1}}\ket{% n}\bra{n}\\ &=\sum_{n=0}^{\infty}e^{-n\omega\beta}\left(1-e^{-\omega\beta}\right)\ket{n}% \bra{n},\end{aligned}start_ROW start_CELL italic_ρ start_POSTSUPERSCRIPT th end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_β end_POSTSUBSCRIPT end_CELL start_CELL = ∑ start_POSTSUBSCRIPT italic_n = 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∞ end_POSTSUPERSCRIPT divide start_ARG [ over¯ start_ARG italic_n end_ARG ( ( italic_ω , italic_T ) ] start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT end_ARG start_ARG [ 1 + over¯ start_ARG italic_n end_ARG ( ( italic_ω , italic_T ) ] start_POSTSUPERSCRIPT italic_n + 1 end_POSTSUPERSCRIPT end_ARG | start_ARG italic_n end_ARG ⟩ ⟨ start_ARG italic_n end_ARG | end_CELL end_ROW start_ROW start_CELL end_CELL start_CELL = ∑ start_POSTSUBSCRIPT italic_n = 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∞ end_POSTSUPERSCRIPT italic_e start_POSTSUPERSCRIPT - italic_n italic_ω italic_β end_POSTSUPERSCRIPT ( 1 - italic_e start_POSTSUPERSCRIPT - italic_ω italic_β end_POSTSUPERSCRIPT ) | start_ARG italic_n end_ARG ⟩ ⟨ start_ARG italic_n end_ARG | , end_CELL end_ROW (45)

being |nket𝑛\ket{n}| start_ARG italic_n end_ARG ⟩ the pure state of a system with n𝑛nitalic_n photons. Note that as n¯(ω,T)¯𝑛𝜔𝑇\bar{n}(\omega,T)over¯ start_ARG italic_n end_ARG ( italic_ω , italic_T ) depends on the temperature and the frequency in terms, in the thermal state this dependence is implicitly included. Initially, we shall consider the system to be in such thermal state.

If a system is in a Gaussian state, including a thermal state, and the interaction with the bath is also Gaussian, its state would be entirely characterized by the evolution of its occupation numbers adelimited-⟨⟩𝑎\left<a\right>⟨ italic_a ⟩ and aadelimited-⟨⟩superscript𝑎𝑎\left\langle a^{\dagger}a\right\rangle⟨ italic_a start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT italic_a ⟩. This fact reduces the problem to the computation of the evolution of the expected values instead of the whole density matrix, leading to a single ordinary differential equation. The dynamics of adelimited-⟨⟩𝑎\langle a\rangle⟨ italic_a ⟩ and aadelimited-⟨⟩superscript𝑎𝑎\left\langle a^{\dagger}a\right\rangle⟨ italic_a start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT italic_a ⟩ are described by the following expressions [111]

dadt𝑑delimited-⟨⟩𝑎𝑑𝑡\displaystyle\frac{d\langle a\rangle}{dt}divide start_ARG italic_d ⟨ italic_a ⟩ end_ARG start_ARG italic_d italic_t end_ARG =\displaystyle== i(ω+Γ2)a,𝑖𝜔Γ2delimited-⟨⟩𝑎\displaystyle-i\left(\omega+\dfrac{\Gamma}{2}\right)\langle a\rangle,- italic_i ( italic_ω + divide start_ARG roman_Γ end_ARG start_ARG 2 end_ARG ) ⟨ italic_a ⟩ ,
daadt𝑑delimited-⟨⟩superscript𝑎𝑎𝑑𝑡\displaystyle\frac{d\left\langle a^{\dagger}a\right\rangle}{dt}divide start_ARG italic_d ⟨ italic_a start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT italic_a ⟩ end_ARG start_ARG italic_d italic_t end_ARG =\displaystyle== Γaa+Γn¯(ω,T).Γdelimited-⟨⟩superscript𝑎𝑎Γ¯𝑛𝜔𝑇\displaystyle-\Gamma\left\langle a^{\dagger}a\right\rangle+\Gamma\bar{n}(% \omega,T).- roman_Γ ⟨ italic_a start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT italic_a ⟩ + roman_Γ over¯ start_ARG italic_n end_ARG ( italic_ω , italic_T ) . (46)

[being n¯(ω,T)¯𝑛𝜔𝑇\bar{n}(\omega,T)over¯ start_ARG italic_n end_ARG ( italic_ω , italic_T ) the average number of excitations of the bath in resonance with the oscillator. Focusing on the temporal evolution of the average number of the system excitations, the solution to this differential equation can be obtained, leading us to the variation in the average number of excitations

aat=subscriptdelimited-⟨⟩superscript𝑎𝑎𝑡absent\displaystyle\left<a^{\dagger}a\right>_{t}=⟨ italic_a start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT italic_a ⟩ start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT = aa0eΓt+0tΓn(ω,T)eΓ(ts)𝑑s,subscriptdelimited-⟨⟩superscript𝑎𝑎0superscript𝑒Γ𝑡superscriptsubscript0𝑡Γ𝑛𝜔𝑇superscript𝑒Γ𝑡𝑠differential-d𝑠\displaystyle\left<a^{\dagger}a\right>_{0}e^{-\Gamma t}+\int_{0}^{t}\Gamma n(% \omega,T)e^{-\Gamma(t-s)}ds,⟨ italic_a start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT italic_a ⟩ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT italic_e start_POSTSUPERSCRIPT - roman_Γ italic_t end_POSTSUPERSCRIPT + ∫ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_t end_POSTSUPERSCRIPT roman_Γ italic_n ( italic_ω , italic_T ) italic_e start_POSTSUPERSCRIPT - roman_Γ ( italic_t - italic_s ) end_POSTSUPERSCRIPT italic_d italic_s , (47)

Note that the sub-index t𝑡titalic_t represents the time dependence and 00 the initial value for the average number of excitations. We consider that the system and the bath are in contact at t=0𝑡0t=0italic_t = 0, without loss of generality. The protocol is modeled by a quench, i.e. a step function, so the n(ω,T)𝑛𝜔𝑇n(\omega,T)italic_n ( italic_ω , italic_T ) term in the integral is constant. Therefore, the evolution reduces to

aat=aa0eΓt+n(ω,T)(1eΓt).subscriptdelimited-⟨⟩superscript𝑎𝑎𝑡subscriptdelimited-⟨⟩superscript𝑎𝑎0superscript𝑒Γ𝑡𝑛𝜔𝑇1superscript𝑒Γ𝑡\left<a^{\dagger}a\right>_{t}=\left<a^{\dagger}a\right>_{0}e^{-\Gamma t}+n(% \omega,T)\left(1-e^{-\Gamma t}\right).⟨ italic_a start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT italic_a ⟩ start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT = ⟨ italic_a start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT italic_a ⟩ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT italic_e start_POSTSUPERSCRIPT - roman_Γ italic_t end_POSTSUPERSCRIPT + italic_n ( italic_ω , italic_T ) ( 1 - italic_e start_POSTSUPERSCRIPT - roman_Γ italic_t end_POSTSUPERSCRIPT ) . (48)

This expression for aatsubscriptdelimited-⟨⟩superscript𝑎𝑎𝑡\left<a^{\dagger}a\right>_{t}⟨ italic_a start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT italic_a ⟩ start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT directly provides us with the temporal evolution of the average number of excitations of the harmonic oscillator. To compare the state of the system with the final thermal state, whose average number of excitations is the one of the bath, we analytically calculate the fidelity for the system at a time t𝑡titalic_t, starting at time t=0𝑡0t=0italic_t = 0 with an occupation number aa0subscriptdelimited-⟨⟩superscript𝑎𝑎0\left<a^{\dagger}a\right>_{0}⟨ italic_a start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT italic_a ⟩ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT compared to a final thermal state at temperature T𝑇Titalic_T, defined by n¯(ω,T)¯𝑛𝜔𝑇\bar{n}(\omega,T)over¯ start_ARG italic_n end_ARG ( italic_ω , italic_T ). Then, the Fidelity reads

F(ρSth(t),ρF)=Trn=0[(aatn¯(ω,T))n(1+aat)1+n(1+n¯(ω,T))1+n]1/2|nn|𝐹subscriptsuperscript𝜌th𝑆𝑡subscript𝜌Ftracesuperscriptsubscript𝑛0superscriptdelimited-[]superscriptsubscriptdelimited-⟨⟩superscript𝑎𝑎𝑡¯𝑛𝜔𝑇𝑛superscript1subscriptdelimited-⟨⟩superscript𝑎𝑎𝑡1𝑛superscript1¯𝑛𝜔𝑇1𝑛12ket𝑛bra𝑛\displaystyle F(\rho^{\text{th}}_{S}(t),\rho_{\text{F}})=\Tr\sum_{n=0}^{\infty% }\left[\dfrac{\left(\left<a^{\dagger}a\right>_{t}\bar{n}(\omega,T)\right)^{n}}% {(1+\left<a^{\dagger}a\right>_{t})^{1+n}(1+\bar{n}(\omega,T))^{1+n}}\right]^{1% /2}\ket{n}\bra{n}italic_F ( italic_ρ start_POSTSUPERSCRIPT th end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_S end_POSTSUBSCRIPT ( italic_t ) , italic_ρ start_POSTSUBSCRIPT F end_POSTSUBSCRIPT ) = roman_Tr ∑ start_POSTSUBSCRIPT italic_n = 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∞ end_POSTSUPERSCRIPT [ divide start_ARG ( ⟨ italic_a start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT italic_a ⟩ start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT over¯ start_ARG italic_n end_ARG ( italic_ω , italic_T ) ) start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT end_ARG start_ARG ( 1 + ⟨ italic_a start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT italic_a ⟩ start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT ) start_POSTSUPERSCRIPT 1 + italic_n end_POSTSUPERSCRIPT ( 1 + over¯ start_ARG italic_n end_ARG ( italic_ω , italic_T ) ) start_POSTSUPERSCRIPT 1 + italic_n end_POSTSUPERSCRIPT end_ARG ] start_POSTSUPERSCRIPT 1 / 2 end_POSTSUPERSCRIPT | start_ARG italic_n end_ARG ⟩ ⟨ start_ARG italic_n end_ARG | =1[(1+aat)(1+n¯(ω,T))]1/211rabsent1superscriptdelimited-[]1subscriptdelimited-⟨⟩superscript𝑎𝑎𝑡1¯𝑛𝜔𝑇1211𝑟\displaystyle=\dfrac{1}{[(1+\left<a^{\dagger}a\right>_{t})(1+\bar{n}(\omega,T)% )]^{1/2}}\dfrac{1}{1-r}= divide start_ARG 1 end_ARG start_ARG [ ( 1 + ⟨ italic_a start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT italic_a ⟩ start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT ) ( 1 + over¯ start_ARG italic_n end_ARG ( italic_ω , italic_T ) ) ] start_POSTSUPERSCRIPT 1 / 2 end_POSTSUPERSCRIPT end_ARG divide start_ARG 1 end_ARG start_ARG 1 - italic_r end_ARG (49)

where r={aatn¯(ω,T)/[(1+aat)(1+n¯(ω,T))]}1/2𝑟superscriptsubscriptdelimited-⟨⟩superscript𝑎𝑎𝑡¯𝑛𝜔𝑇delimited-[]1subscriptdelimited-⟨⟩superscript𝑎𝑎𝑡1¯𝑛𝜔𝑇12r=\left\{\left<a^{\dagger}a\right>_{t}\bar{n}(\omega,T)/[(1+\left<a^{\dagger}a% \right>_{t})(1+\bar{n}(\omega,T))]\right\}^{1/2}italic_r = { ⟨ italic_a start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT italic_a ⟩ start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT over¯ start_ARG italic_n end_ARG ( italic_ω , italic_T ) / [ ( 1 + ⟨ italic_a start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT italic_a ⟩ start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT ) ( 1 + over¯ start_ARG italic_n end_ARG ( italic_ω , italic_T ) ) ] } start_POSTSUPERSCRIPT 1 / 2 end_POSTSUPERSCRIPT. In Fig. 4 both the Bures distance as a function of the temperature, panel 4(a), and the Fidelity as a function of time, panel 4(b), are displayed. The results confirm behavior obtained for the two-level system. One main difference is that to obtain the same distance, in the Bures sense, we need a higher temperature difference in the harmonic oscillator case that for the two-level system. This is due to the infinite size of the Hilbert space of the harmonic oscillator, in comparison to a two-dimensional Hilbert space.

To analyze the thermal kinematics of the system we have computed numerically the quantum Fisher Information of the two-temperatures protocol. The results are displayed in Figure 4, panels 4(c) and 4(d). It is clear that, even if the harmonic oscillator is a different and more complicated system its thermal behavior is similar to the one for the two-level system. This supports our hypothesis that the thermal asymmetry is a general trend of open quantum systems. In the next section we check this behavior with an even more complex system as the Quantum Brownian particle.

Refer to caption
Figure 5: Simulation of the protocols for the Quantum Brownian particle. (a) Bures distance dependence with the temperature difference with respect to the equilibrium state, as in the harmonic oscillator case. (b) Fidelity for the state with respect to the thermal state at TWsubscript𝑇𝑊T_{W}italic_T start_POSTSUBSCRIPT italic_W end_POSTSUBSCRIPT. The process corresponds to the three-temperature protocol, evolving from TCsubscript𝑇𝐶T_{C}italic_T start_POSTSUBSCRIPT italic_C end_POSTSUBSCRIPT to TWsubscript𝑇𝑊T_{W}italic_T start_POSTSUBSCRIPT italic_W end_POSTSUBSCRIPT (orange line) and THsubscript𝑇𝐻T_{H}italic_T start_POSTSUBSCRIPT italic_H end_POSTSUBSCRIPT to TWsubscript𝑇𝑊T_{W}italic_T start_POSTSUBSCRIPT italic_W end_POSTSUBSCRIPT (blue line). (c) Instantaneous statistical velocity for the quantum Brownian particle computed for the heating and cooling protocol in the two-temperature scheme. (d) Degree of completion for the quantum Brownian particle in the two-temperature scheme. In the case of the quantum Brownian particle, the values of the parameters are tfin=5103,m=1,Ω=103,Λ=1,ζ=0.1formulae-sequencesubscript𝑡fin5superscript103formulae-sequence𝑚1formulae-sequenceΩsuperscript103formulae-sequenceΛ1𝜁0.1t_{\textrm{fin}}=5\cdot 10^{3},m=1,\Omega=10^{-3},\Lambda=1,\zeta=0.1italic_t start_POSTSUBSCRIPT fin end_POSTSUBSCRIPT = 5 ⋅ 10 start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT , italic_m = 1 , roman_Ω = 10 start_POSTSUPERSCRIPT - 3 end_POSTSUPERSCRIPT , roman_Λ = 1 , italic_ζ = 0.1.
Refer to caption
Figure 6: Near-temperature simulations for the three-temperature protocol for the harmonic oscillator. The Fidelity is calculated between the time evolution of the state of the system, ρ(t)𝜌𝑡\rho(t)italic_ρ ( italic_t ), and the thermal state at n(ω,TW)𝑛𝜔subscript𝑇𝑊n(\omega,T_{W})italic_n ( italic_ω , italic_T start_POSTSUBSCRIPT italic_W end_POSTSUBSCRIPT ). In all cases ω=1𝜔1\omega=1italic_ω = 1, the average excitation number for the cold bath is n(ω,TC)=1𝑛𝜔subscript𝑇𝐶1n(\omega,T_{C})=1italic_n ( italic_ω , italic_T start_POSTSUBSCRIPT italic_C end_POSTSUBSCRIPT ) = 1 in all cases, and the hot temperature is for panels (a), (d) n(ω,TH)=1.1𝑛𝜔subscript𝑇𝐻1.1n(\omega,T_{H})=1.1italic_n ( italic_ω , italic_T start_POSTSUBSCRIPT italic_H end_POSTSUBSCRIPT ) = 1.1; for panels (b), (e) n(ω,TH)=2𝑛𝜔subscript𝑇𝐻2n(\omega,T_{H})=2italic_n ( italic_ω , italic_T start_POSTSUBSCRIPT italic_H end_POSTSUBSCRIPT ) = 2; and for panels (c), (f) n(ω,TH)=5𝑛𝜔subscript𝑇𝐻5n(\omega,T_{H})=5italic_n ( italic_ω , italic_T start_POSTSUBSCRIPT italic_H end_POSTSUBSCRIPT ) = 5. In the linear regime, for close temperatures TCsubscript𝑇𝐶T_{C}italic_T start_POSTSUBSCRIPT italic_C end_POSTSUBSCRIPT and THsubscript𝑇𝐻T_{H}italic_T start_POSTSUBSCRIPT italic_H end_POSTSUBSCRIPT, both curves collapse and the processes remain symmetric. However, as the temperature difference increases, the lines of heating up and cooling down evolve differently creating the asymmetry between the protocols.

VI.2 Quantum Brownian particle in a trap

The third model introduced to analyze the protocols is a quantum Brownian particle trapped in a harmonic trap, following the results and experiments already performed in the classical case [54]. This is the most sophisticated case that is treated in the article, where all the relevant quantities need to be computed numerically.

A quantum Brownian particle interacting with a bosonic bath is described by the following Hamiltonian [112, 113, 89, 90]

H𝐻\displaystyle Hitalic_H =\displaystyle== HS+HB+HIsubscript𝐻𝑆subscript𝐻𝐵subscript𝐻𝐼\displaystyle H_{S}+H_{B}+H_{I}italic_H start_POSTSUBSCRIPT italic_S end_POSTSUBSCRIPT + italic_H start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT + italic_H start_POSTSUBSCRIPT italic_I end_POSTSUBSCRIPT (50)
=\displaystyle== pS22mS+i=1nκi22mBiωBixS2+ϕ(xS)superscriptsubscript𝑝𝑆22subscript𝑚𝑆superscriptsubscript𝑖1𝑛subscriptsuperscript𝜅2𝑖2subscript𝑚subscript𝐵𝑖subscript𝜔subscript𝐵𝑖superscriptsubscript𝑥𝑆2italic-ϕsubscript𝑥𝑆\displaystyle\frac{p_{S}^{2}}{2m_{S}}+\sum_{i=1}^{n}\frac{\kappa^{2}_{i}}{2m_{% B_{i}}\omega_{B_{i}}}x_{S}^{2}+\phi(x_{S})divide start_ARG italic_p start_POSTSUBSCRIPT italic_S end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG 2 italic_m start_POSTSUBSCRIPT italic_S end_POSTSUBSCRIPT end_ARG + ∑ start_POSTSUBSCRIPT italic_i = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT divide start_ARG italic_κ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT end_ARG start_ARG 2 italic_m start_POSTSUBSCRIPT italic_B start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT end_POSTSUBSCRIPT italic_ω start_POSTSUBSCRIPT italic_B start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT end_POSTSUBSCRIPT end_ARG italic_x start_POSTSUBSCRIPT italic_S end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + italic_ϕ ( italic_x start_POSTSUBSCRIPT italic_S end_POSTSUBSCRIPT )
+\displaystyle++ i=1nωBiaB,iaB,ii=1nκixB,ixS,superscriptsubscript𝑖1𝑛Planck-constant-over-2-pisubscript𝜔subscript𝐵𝑖subscriptsuperscript𝑎𝐵𝑖subscript𝑎𝐵𝑖superscriptsubscript𝑖1𝑛subscript𝜅𝑖subscript𝑥𝐵𝑖subscript𝑥𝑆\displaystyle\sum_{i=1}^{n}\hbar\omega_{B_{i}}a^{\dagger}_{B,i}a_{B,i}-\sum_{i% =1}^{n}\kappa_{i}x_{B,i}x_{S},∑ start_POSTSUBSCRIPT italic_i = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT roman_ℏ italic_ω start_POSTSUBSCRIPT italic_B start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT end_POSTSUBSCRIPT italic_a start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_B , italic_i end_POSTSUBSCRIPT italic_a start_POSTSUBSCRIPT italic_B , italic_i end_POSTSUBSCRIPT - ∑ start_POSTSUBSCRIPT italic_i = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT italic_κ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT italic_x start_POSTSUBSCRIPT italic_B , italic_i end_POSTSUBSCRIPT italic_x start_POSTSUBSCRIPT italic_S end_POSTSUBSCRIPT ,

where the indexes S,B𝑆𝐵S,Bitalic_S , italic_B hold for the system and bath operators respectively. Here mSsubscript𝑚𝑆m_{S}italic_m start_POSTSUBSCRIPT italic_S end_POSTSUBSCRIPT is the mass of the Brownian particle, xSsubscript𝑥𝑆x_{S}italic_x start_POSTSUBSCRIPT italic_S end_POSTSUBSCRIPT its position, pSsubscript𝑝𝑆p_{S}italic_p start_POSTSUBSCRIPT italic_S end_POSTSUBSCRIPT its momentum and ϕ(x)italic-ϕ𝑥\phi(x)italic_ϕ ( italic_x ) is a trapping potential. Similarly, mBisubscript𝑚subscript𝐵𝑖m_{B_{i}}italic_m start_POSTSUBSCRIPT italic_B start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT end_POSTSUBSCRIPT, ωBisubscript𝜔subscript𝐵𝑖\omega_{B_{i}}italic_ω start_POSTSUBSCRIPT italic_B start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT end_POSTSUBSCRIPT and xBisubscript𝑥subscript𝐵𝑖x_{B_{i}}italic_x start_POSTSUBSCRIPT italic_B start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT end_POSTSUBSCRIPT are the mass, frequency, and position of the i𝑖iitalic_i-th bath particle, for all i=1,,n𝑖1𝑛i=1,\ldots,nitalic_i = 1 , … , italic_n. The factors κisubscript𝜅𝑖\kappa_{i}italic_κ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT represent the coupling between the system and the i𝑖iitalic_i-th bath mode.

The trapping potential is customarily taken as a harmonic term so that

ϕ(x)=m2ω~2x2,italic-ϕ𝑥𝑚2superscript~𝜔2superscript𝑥2\phi(x)=\dfrac{m}{2}\tilde{\omega}^{2}x^{2},italic_ϕ ( italic_x ) = divide start_ARG italic_m end_ARG start_ARG 2 end_ARG over~ start_ARG italic_ω end_ARG start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_x start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT , (51)

for a given trap frequency ω~~𝜔\tilde{\omega}over~ start_ARG italic_ω end_ARG.

A treatment for the problem can be performed by a Lindblad-equation-like transformation of the equations of motion of the system. The global evolution of the system and bath may be described by a unitary operator, and the state of the system at a given time t𝑡titalic_t is described by

ρ˙S(t)=TrB{U(t)(ρS(0)ρB)U(t)}[ρS(t)],subscript˙𝜌𝑆𝑡subscripttrace𝐵𝑈𝑡tensor-productsubscript𝜌𝑆0subscript𝜌𝐵superscript𝑈𝑡delimited-[]subscript𝜌𝑆𝑡\dot{\rho}_{S}(t)=\Tr_{B}\left\{U(t)\left(\rho_{S}(0){\otimes}\rho_{B}\right)U% ^{\dagger}(t)\right\}\equiv\mathcal{L}[\rho_{S}(t)],over˙ start_ARG italic_ρ end_ARG start_POSTSUBSCRIPT italic_S end_POSTSUBSCRIPT ( italic_t ) = roman_Tr start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT { italic_U ( italic_t ) ( italic_ρ start_POSTSUBSCRIPT italic_S end_POSTSUBSCRIPT ( 0 ) ⊗ italic_ρ start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT ) italic_U start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT ( italic_t ) } ≡ caligraphic_L [ italic_ρ start_POSTSUBSCRIPT italic_S end_POSTSUBSCRIPT ( italic_t ) ] , (52)

being \mathcal{L}caligraphic_L the Liouvillian superoperator of the coherent dynamics. Given the fact that the interaction between the system and the bath is linear and assuming it to be also weak, we can consider a single Lindblad operator 𝒜𝒜\mathcal{A}caligraphic_A such that [113]

𝒜(T)𝒜𝑇\displaystyle\mathcal{A}(T)caligraphic_A ( italic_T ) =\displaystyle== αx+βp,𝛼𝑥𝛽𝑝\displaystyle\alpha x+\beta p,italic_α italic_x + italic_β italic_p , (53)

for some parameters α,β𝛼𝛽\alpha,\beta\in\mathbb{C}italic_α , italic_β ∈ blackboard_C. In order to match the coefficients represented in Eq. (53) with a general Born-Markov treatment of the problem within the Caldeira-Leggett limit, these α𝛼\alphaitalic_α and β𝛽\betaitalic_β must be given by

α=(2mζkBT)1/2,𝛼superscript2𝑚𝜁subscript𝑘𝐵𝑇12Planck-constant-over-2-pi\alpha=\dfrac{\left(2m\zeta k_{B}T\right)^{1/2}}{\hbar},italic_α = divide start_ARG ( 2 italic_m italic_ζ italic_k start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT italic_T ) start_POSTSUPERSCRIPT 1 / 2 end_POSTSUPERSCRIPT end_ARG start_ARG roman_ℏ end_ARG , (54)

and

Re(β)=ζkBT2αΛ,Im(β)=ζ2α.formulae-sequenceRe𝛽𝜁subscript𝑘𝐵𝑇superscriptPlanck-constant-over-2-pi2𝛼ΛIm𝛽𝜁2Planck-constant-over-2-pi𝛼\operatorname{Re}(\beta)=-\dfrac{\zeta k_{B}T}{\hbar^{2}\alpha\Lambda},\quad% \operatorname{Im}(\beta)=\dfrac{\zeta}{2\hbar\alpha}.roman_Re ( italic_β ) = - divide start_ARG italic_ζ italic_k start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT italic_T end_ARG start_ARG roman_ℏ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_α roman_Λ end_ARG , roman_Im ( italic_β ) = divide start_ARG italic_ζ end_ARG start_ARG 2 roman_ℏ italic_α end_ARG . (55)

In these relations, ΛΛ\Lambdaroman_Λ is the so-called Lorentz-Drude cutoff appearing in baths with Ohmic spectral density; ζ𝜁\zetaitalic_ζ is a damping constant, whose inverse is related to the relaxation scales; T𝑇Titalic_T is the temperature of the bath and m𝑚mitalic_m the mass of the oscillators. The Caldeira-Leggett limit is satisfied for large temperature and cut-off limits. Under this regime, one recovers the Caldeira-Leggett equation for general diffusion processes in a quantum framework for a quantum Brownian particle [89, 90]. Note that the temperature of the baths is related to the average number of excitations via the Bose-Einstein relation as in the previous models.

For this model both the fidelity and quantum Fisher information may be calculated numerically, as an analytical description is intractable. In Fig. 6 we observe a similar analysis to the ones performed for both the two-level system and the harmonic oscillator. In this case, the temperature range that we need to consider is even larger, due to the complexity of the bath. All the results are similar, confirming the general character of our results. One interesting feature is that during the heating up process in the three-temperatures protocol [c.f. Fig. 6(b)] the fidelity reaches a value close to one in a finite time, and then bounces down. This interesting behavior suggest that the system suffers from hysteresis, an interesting feature specially due to the Markovian character of the dynamics.

VI.3 Analysis of the results

In the three-temperature protocol, the fidelity of the state at a given time t𝑡titalic_t has been compared to the thermal state at the intermediate temperature, TWsubscript𝑇𝑊T_{W}italic_T start_POSTSUBSCRIPT italic_W end_POSTSUBSCRIPT, so that it increases to one, when thermalization takes place. As it is indicated in IV.1, the thermal state at warm temperature, ρβWthsubscriptsuperscript𝜌thsubscript𝛽𝑊\rho^{\text{th}}_{\beta_{W}}italic_ρ start_POSTSUPERSCRIPT th end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_β start_POSTSUBSCRIPT italic_W end_POSTSUBSCRIPT end_POSTSUBSCRIPT, is chosen to be equidistant to the cold, ρβCthsubscriptsuperscript𝜌thsubscript𝛽𝐶\rho^{\text{th}}_{\beta_{C}}italic_ρ start_POSTSUPERSCRIPT th end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_β start_POSTSUBSCRIPT italic_C end_POSTSUBSCRIPT end_POSTSUBSCRIPT, and hot states, ρβHthsubscriptsuperscript𝜌thsubscript𝛽𝐻\rho^{\text{th}}_{\beta_{H}}italic_ρ start_POSTSUPERSCRIPT th end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_β start_POSTSUBSCRIPT italic_H end_POSTSUBSCRIPT end_POSTSUBSCRIPT, so that the Bures distance is the same in both cases. The distance is depicted in Figs. 3(a), 4(a), and 5(a) as a function of the temperature difference ΔTΔ𝑇\Delta Troman_Δ italic_T with respect to TWsubscript𝑇𝑊T_{W}italic_T start_POSTSUBSCRIPT italic_W end_POSTSUBSCRIPT , meaning that ΔTC=TCTWΔsubscript𝑇𝐶subscript𝑇𝐶subscript𝑇𝑊\Delta T_{C}=T_{C}-T_{W}roman_Δ italic_T start_POSTSUBSCRIPT italic_C end_POSTSUBSCRIPT = italic_T start_POSTSUBSCRIPT italic_C end_POSTSUBSCRIPT - italic_T start_POSTSUBSCRIPT italic_W end_POSTSUBSCRIPT for the initial cold state (heating-up protocol), and ΔTH=THTWΔsubscript𝑇𝐻subscript𝑇𝐻subscript𝑇𝑊\Delta T_{H}=T_{H}-T_{W}roman_Δ italic_T start_POSTSUBSCRIPT italic_H end_POSTSUBSCRIPT = italic_T start_POSTSUBSCRIPT italic_H end_POSTSUBSCRIPT - italic_T start_POSTSUBSCRIPT italic_W end_POSTSUBSCRIPT for the initial hot state (cooling-down protocol). The temperature of both initial points is represented by a hollow red circle for the case of heating and by a hollow blue circle for the cooling. Both points evolve to the equilibrium thermal state, clearly represented by a minimum. It is worth noticing that the asymmetry in the different protocols is clearly appreciated here. Even if they are starting from the same initial point, the length of the path followed to the equilibrium state is clearly larger in the cooling protocol.

The asymmetry in the three-temperature protocols is analyzed by the use of the fidelity between the initial states and the target one, as a function of time. This is displayed in Figs. 3(b), 4(b), and 5(b). Remarkably, the necessary times to thermalize differ by several orders of magnitude due to the increasing complexity of each configuration. However, the three of them show a similar behavior. This fact is also relevant for the velocity analysis.

Regarding the two-temperature protocol, we shall make use of the instantaneous velocity quantity, Eq. (17), and degree of completion, Eq. (19). Figures 3(c), 4(c) and 5(c) represent this quantity for the qubit, the harmonic oscillator, and the Brownian particle respectively. As expected, the velocity in both cases for the heating process (TCTH)subscript𝑇𝐶subscript𝑇𝐻(T_{C}\rightarrow T_{H})( italic_T start_POSTSUBSCRIPT italic_C end_POSTSUBSCRIPT → italic_T start_POSTSUBSCRIPT italic_H end_POSTSUBSCRIPT ) is larger than the cooling counterpart until (THTC)subscript𝑇𝐻subscript𝑇𝐶(T_{H}\rightarrow T_{C})( italic_T start_POSTSUBSCRIPT italic_H end_POSTSUBSCRIPT → italic_T start_POSTSUBSCRIPT italic_C end_POSTSUBSCRIPT ) it reaches a null value, i.e. the system thermalizes at the final temperature. As we anticipated in the previous paragraph, the thermalization times differ by several orders of magnitude. This is also represented by the scale in the velocity axis in both of the aforementioned plots. Finally, the computation of the degree of completion is directly derived considering the values of the instantaneous velocity, Eqs. (18) and (19). Figures 3(d), 4(d) and 5(d) show the temporal evolution of the degree of completion and, as it is expected, the functions are similar regardless of the sort of system, and the heating process takes less amount of time to be completed than the cooling. The main implication in the function is related to the quicker saturation to the unity of the function in the former case.

Although all methods based on thermal kinematics theory are useful for analyzing the dynamics towards the equilibrium state, they do not provide theoretical insight into its origin. The Bures distance with respect to the equilibrium temperature in the three-temperature case (see Figs. 3(a), 4(a), and 5(a)) offers intuition about the protocols. However, it is solely a measure of the distance and does not offer any fundamental reasoning; it is merely a consequence of a more fundamental phenomenon.

Refer to caption
Figure 7: Eigenvalues of the Liouvillian operator \mathcal{L}caligraphic_L for the harmonic oscillator (a), and a quantum Brownian particle (c) for n¯(ω,TH)=10¯𝑛𝜔subscript𝑇𝐻10\bar{n}(\omega,T_{H})=10over¯ start_ARG italic_n end_ARG ( italic_ω , italic_T start_POSTSUBSCRIPT italic_H end_POSTSUBSCRIPT ) = 10 (red points) and n¯(ω,TC)=1¯𝑛𝜔subscript𝑇𝐶1\bar{n}(\omega,T_{C})=1over¯ start_ARG italic_n end_ARG ( italic_ω , italic_T start_POSTSUBSCRIPT italic_C end_POSTSUBSCRIPT ) = 1 (blue points). Note the difference of scales in the x𝑥xitalic_x-axis. Panels (b) and (d) represent the first eigenvalues of the respective panels (a) and (c) with a size proportional to the overlap with the thermal state at the opposite temperature for the harmonic oscillator, see Eq. (8). In both cases, the truncated dimension of the Hilbert space is N=150𝑁150N=150italic_N = 150.

VI.4 Linear response regime

We shall conclude the section providing a brief comment on the near-equilibrium, i.e. linear regime, for thermal evolution close to the equilibrium temperature in the three-level protocol.

The linear response theory, developed mainly by Kubo [5, 6], is the cornerstone to analyze the near-equilibrium behavior in non-equilibrium classical thermodynamics. The fluctuation-dissipation theorem states that the fluctuation properties of a system in TE determine its linear response to an external perturbation [6]. In the quantum counterpart, this theorem has been derived for closed quantum systems and recently for open quantum systems [114, 115]. This extension allows us to apply the existing results from isolated equilibrium systems to open systems, with Lindbladian dynamics [114, 115]. Within this regime, one expects to recover the same as in classical thermodynamics results, where the asymmetry between heating and protocols is absent. That means, for a small temperature differences in both protocols, we expect to observe how the asymmetry diminish, since we are approaching the equilibrium state.

For the qubit case this phenomenon is clearly appreciated in the analytical derivation of the Fidelity comparing two states, Eq. (37). For small ΔTΔ𝑇\Delta Troman_Δ italic_T, the Fidelity is quadratic

F(ρβ0th,ρβth)=1eβω8(1+eβω)2(ωΔTT2)2+𝒪(ΔTT)3,𝐹superscriptsubscript𝜌subscript𝛽0thsuperscriptsubscript𝜌𝛽th1superscript𝑒𝛽Planck-constant-over-2-pi𝜔8superscript1superscript𝑒𝛽Planck-constant-over-2-pi𝜔2superscriptPlanck-constant-over-2-pi𝜔Δ𝑇superscript𝑇22𝒪superscriptΔ𝑇𝑇3F(\rho_{\beta_{0}}^{\rm th},\rho_{\beta}^{\rm th})=1-\frac{e^{\beta\hbar\omega% }}{8\left(1+e^{\beta\hbar\omega}\right)^{2}}\left(\frac{\hbar\omega\Delta T}{T% ^{2}}\right)^{2}+{\cal O}\left(\frac{\Delta T}{T}\right)^{3},italic_F ( italic_ρ start_POSTSUBSCRIPT italic_β start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_POSTSUBSCRIPT start_POSTSUPERSCRIPT roman_th end_POSTSUPERSCRIPT , italic_ρ start_POSTSUBSCRIPT italic_β end_POSTSUBSCRIPT start_POSTSUPERSCRIPT roman_th end_POSTSUPERSCRIPT ) = 1 - divide start_ARG italic_e start_POSTSUPERSCRIPT italic_β roman_ℏ italic_ω end_POSTSUPERSCRIPT end_ARG start_ARG 8 ( 1 + italic_e start_POSTSUPERSCRIPT italic_β roman_ℏ italic_ω end_POSTSUPERSCRIPT ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG ( divide start_ARG roman_ℏ italic_ω roman_Δ italic_T end_ARG start_ARG italic_T start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + caligraphic_O ( divide start_ARG roman_Δ italic_T end_ARG start_ARG italic_T end_ARG ) start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT , (56)

i.e. no asymmetry is expected for states close to the equilibrium.

Regarding the simulations for the harmonic oscillator, the results for close temperatures is depicted in Fig. 6(a) and 6(d). As the temperature difference increases, the asymmetry starts to appear, making this discrepancy in both protocols more acute the larger is this gap. Figures 6(b)-(c) and 6(e)-(f) show this behavior. Figures 6(a)-(c) represent the fidelity with respect to the thermal state at warm temperature TWsubscript𝑇𝑊T_{W}italic_T start_POSTSUBSCRIPT italic_W end_POSTSUBSCRIPT, i.e. in the three-temperature protocol. Similarly, Figs. 6(d)-(f) showcase the asymmetry in the two-temperature scenario, displayed in the velocity needed to reach the opposite state. It is clear that the asymmetry arises as one deviates from equilibrium when the temperature difference increases.

VII Spectral analysis

Performing a spectral analysis of the Liouvillian, we can gain intuition about the relaxation time of our models for the heating and cooling protocols, as the dynamics of the system is given by Eq. (7). In Fig. 7 the eigenvalues of both the harmonic oscillator and the Brownian particle are displayed. Due to the infinite size of the Hilbert space of the systems, we have used a truncated Fock basis of dimension N𝑁Nitalic_N, large enough to display the general behavior. As discussed in Sec. II, the spectrum of the Lindbladian is composed of eigenvalues whose real part is negative, apart from the null eigenvalue which determines the stationary state. This decomposition does not depend on the initial state we consider but on the parameters defined in the Lindbladian, i.e. the Hamiltonian and jump operators, as well as the constants and variables defined therein. The fact that an open quantum system, when initialized at a given state, say ρ0subscript𝜌0\rho_{0}italic_ρ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT, evolves differently than from another initial state, say ρ~0subscript~𝜌0\tilde{\rho}_{0}over~ start_ARG italic_ρ end_ARG start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT, for the same Lindbladian depends on the overlap between the considered initial state and the left eigenvectors, see Eq. (8).

The spectra for the heating and cooling cases and the different systems are displayed in Fig. 7. For the Lindbladians with higher temperature there is a spreading of the eigenvalues towards the negative real axis, see Fig. 7(a) for the harmonic oscillator and Fig. 7(c) for the quantum Brownian particle. This means that, in the heating-up processes, there are many more fast-decaying modes than in the cooling-down counterparts, indicating that heating will be faster. This behavior is in agreement with Fig. 2(a) for the thermal qubit. In Figs. 7(b) and 7(d), the slowest decay modes, which act as bottlenecks to the dynamics, are plotted with the symbol size being proportional to the overlap with the initial state of the cooling down and heating protocols. The initial state in each case is chosen to be equal to a thermal state at the same temperature as the opposite process. This means that if the cooling-down/heating-up process is causing the system to evolve to n¯(ω,TC)/n¯(ω,TH)¯𝑛𝜔subscript𝑇𝐶¯𝑛𝜔subscript𝑇𝐻\bar{n}(\omega,T_{C})/\bar{n}(\omega,T_{H})over¯ start_ARG italic_n end_ARG ( italic_ω , italic_T start_POSTSUBSCRIPT italic_C end_POSTSUBSCRIPT ) / over¯ start_ARG italic_n end_ARG ( italic_ω , italic_T start_POSTSUBSCRIPT italic_H end_POSTSUBSCRIPT ), the initial state will be ρβHth/ρβCthsubscriptsuperscript𝜌thsubscript𝛽𝐻subscriptsuperscript𝜌thsubscript𝛽𝐶\rho^{\text{th}}_{\beta_{H}}/\rho^{\text{th}}_{\beta_{C}}italic_ρ start_POSTSUPERSCRIPT th end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_β start_POSTSUBSCRIPT italic_H end_POSTSUBSCRIPT end_POSTSUBSCRIPT / italic_ρ start_POSTSUPERSCRIPT th end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_β start_POSTSUBSCRIPT italic_C end_POSTSUBSCRIPT end_POSTSUBSCRIPT.

It is clear that in the cooling protocol, there is a higher overlap with slower decay modes in both cases. Moreover, the number of slower modes in the cooling case (represented by the blue stars) is larger than the number of modes in the heating case (red dots), apart from being closer between them and to the null eigenvalue. This spectral analysis provides us with a justification for the asymmetry in all the processes, referring all of them to mere observations of the decaying modes appearing in the spectra of the Liouvillians. For instance, the asymmetry appears naturally when computing the overlap of the state of the considered system with the generator of the dynamics, i.e. the Liouvillian: the ovelap is larger for the heating up protocol (blue stars) than for the cooling down (red dots). This explanation allows us to justify and clarify all the results obtained throughout the article. We recall however that this asymmetry in the overlap was not present in the thermal qubit case.

VIII Conclusions

In this paper, we have studied and unveiled an intriguing effect of non-equilibrium open quantum systems: the asymmetry of the time-evolution of heating up and cooling down trajectories. By introducing quantum information measures such as the fidelity, the Bures distance, and the quantum Fisher information we analyzed this phenomenon in two different protocols. The first protocol involves an intermediate temperature, equidistant between a hotter and a colder one, while the second protocol works between two absolute temperatures. The measures developed in this work are general and applicable to various other dissipative processes.

We extended the thermal kinematics to open quantum systems and applied these protocols to three different configurations of increasing complexity: a thermal qubit, a harmonic oscillator coupled to a bosonic heat bath, as well as a canonical model for the quantum Brownian motion. The qubit system provides an analytical description that can be solved exactly for all the studied magnitudes, suggesting a connection to the third law of thermodynamics; the other systems are analyzed numerically. Our results unequivocally indicate that heating up and cooling down are intrinsically different processes, with heating up always being the fastest. In the limit of small temperature differences we recover a symmetric behavior in accordance with equilibrium thermodynamics in the quantum regime.

By studying the Liouvillian spectrum of the system, we observe that the eigenvalues spread towards the negative real line as temperature increases. This indicates that for thermal baths at higher temperatures there are more fast decaying modes, making the evolution faster. Additionally, the overlap between the initial state and the fast-decaying modes confirms that the heating up is always a faster mechanism than the heating up.

Despite their simplicity, the proposed configurations can be readily be tested experimentally is various platforms e.g., semiconductor qubits [67] or superconducting cavity quantum thermodynamic circuits [65]. As systems with higher complexity require longer times to thermalize, harmonic oscillators or quantum Brownian motors are ideal canditates to detect thermalization asymmetries.

Acknowledgements.
The authors would like to thank M. Skotiniotis, G. Menesse and G. H. Camillo for their insightful comments. The research leading to these results has received funding from the fellowship FPU20/02835 and from the Projects of I+D+i Ref. PID2020-113681GB-I00, Ref. PID2020-116567GB-C22, Ref. PID2021-128970OA-I00, Ref. PID2022-136374NB-C21, PID2022-142911NB-I00, financed by MICIN/AEI/10.13039/501100011033 and FEDER “A way to make Europe”, and Projects Ref. A-FQM-175-UGR18, Ref. P20_00173 and Ref. A-FQM-644-UGR20, C-EXP-251-UGR23 and through the “María de Maeztu” Programme for Units of Excellence in R&D CEX2023-001316-M, financed by the Spanish Ministerio de Ciencia, Innovación y Universidades and European Regional Development Fund, Junta de Andalucía-Consejería de Economía y Conocimiento 2014-2020. We are also grateful for the computing resources and related technical support provided by PROTEUS, the supercomputing center of the Institute Carlos I for Theoretical and Computational Physics in Granada, Spain. Finally, this work has also been financially supported by the Ministry for Digital Transformation and of Civil Service of the Spanish Government through the QUANTUM ENIA project call - Quantum Spain project, and by the European Union through the Recovery, Transformation and Resilience Plan - NextGenerationEU within the framework of the Digital Spain 2026 Agenda.

References

Appendix A Classical and Quantum Fisher Information

In classical parameter estimation a canonical measure is the classical Fisher Information, I(θ)𝐼𝜃I(\theta)italic_I ( italic_θ ), of a probability density p(x,θ)𝑝𝑥𝜃p(x,\theta)italic_p ( italic_x , italic_θ ), defined as

Icl(θ):=(dlogp(x,θ)dθ)2p(x,θ)𝑑x.assignsubscript𝐼cl𝜃superscriptsubscriptsuperscript𝑑𝑝𝑥𝜃𝑑𝜃2𝑝𝑥𝜃differential-d𝑥I_{\text{cl}}(\theta):=\int_{-\infty}^{\infty}\left(\dfrac{d\log p(x,\theta)}{% d\theta}\right)^{2}p(x,\theta)dx.italic_I start_POSTSUBSCRIPT cl end_POSTSUBSCRIPT ( italic_θ ) := ∫ start_POSTSUBSCRIPT - ∞ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∞ end_POSTSUPERSCRIPT ( divide start_ARG italic_d roman_log italic_p ( italic_x , italic_θ ) end_ARG start_ARG italic_d italic_θ end_ARG ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_p ( italic_x , italic_θ ) italic_d italic_x . (57)

The geometric interpretation of the Fisher information arises from defining a statistical line element, ds𝑑𝑠dsitalic_d italic_s, such that ds2:=I(θ)dθ2assign𝑑superscript𝑠2𝐼𝜃𝑑superscript𝜃2ds^{2}:=I(\theta)d\theta^{2}italic_d italic_s start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT := italic_I ( italic_θ ) italic_d italic_θ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT. Therefore, the line element ds𝑑𝑠dsitalic_d italic_s can be regarded as a dimensionless distance between probability densities p(x,θ)𝑝𝑥𝜃p(x,\theta)italic_p ( italic_x , italic_θ ) and p(x,θ+dθ)𝑝𝑥𝜃𝑑𝜃p(x,\theta+d\theta)italic_p ( italic_x , italic_θ + italic_d italic_θ ).

Consider a quantum state, ρ𝜌\rhoitalic_ρ, parametrized by an n𝑛nitalic_n-dimension vector θ=(θ1,,θn)𝜃subscript𝜃1subscript𝜃𝑛\vec{\theta}=\left(\theta_{1},\ldots,\theta_{n}\right)over→ start_ARG italic_θ end_ARG = ( italic_θ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT , … , italic_θ start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ), and denoted by ρ(θ)𝜌𝜃\rho(\vec{\theta})italic_ρ ( over→ start_ARG italic_θ end_ARG ). For an infinitesimal change in the parameters, one can relate the Bures distance to the Quantum Fisher Information Matrix (QFIM), \mathcal{I}caligraphic_I, so that

[DB(ρ(θ),ρ(θ+dθ))]2=14i,jijdxidxj+𝒪(dx4).superscriptdelimited-[]subscript𝐷𝐵𝜌𝜃𝜌𝜃𝑑𝜃214subscript𝑖𝑗subscript𝑖𝑗𝑑subscript𝑥𝑖𝑑subscript𝑥𝑗𝒪𝑑superscript𝑥4\left[D_{B}\left(\rho(\vec{\theta}),\rho(\vec{\theta}+d\vec{\theta})\right)% \right]^{2}=\dfrac{1}{4}\sum_{i,j}\mathcal{I}_{ij}dx_{i}dx_{j}+\mathcal{O}(dx^% {4}).[ italic_D start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT ( italic_ρ ( over→ start_ARG italic_θ end_ARG ) , italic_ρ ( over→ start_ARG italic_θ end_ARG + italic_d over→ start_ARG italic_θ end_ARG ) ) ] start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT = divide start_ARG 1 end_ARG start_ARG 4 end_ARG ∑ start_POSTSUBSCRIPT italic_i , italic_j end_POSTSUBSCRIPT caligraphic_I start_POSTSUBSCRIPT italic_i italic_j end_POSTSUBSCRIPT italic_d italic_x start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT italic_d italic_x start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT + caligraphic_O ( italic_d italic_x start_POSTSUPERSCRIPT 4 end_POSTSUPERSCRIPT ) . (58)

The complete derivation can be found in [109]. The elements of the QFIM are given by

ij=Tr[Lθiρ(θ)Lθj],subscript𝑖𝑗tracesubscript𝐿subscript𝜃𝑖𝜌𝜃subscript𝐿subscript𝜃𝑗\mathcal{I}_{ij}=\Tr\left[L_{\theta_{i}}\rho({\vec{\theta}})L_{\theta_{j}}% \right],caligraphic_I start_POSTSUBSCRIPT italic_i italic_j end_POSTSUBSCRIPT = roman_Tr [ italic_L start_POSTSUBSCRIPT italic_θ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT end_POSTSUBSCRIPT italic_ρ ( over→ start_ARG italic_θ end_ARG ) italic_L start_POSTSUBSCRIPT italic_θ start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT end_POSTSUBSCRIPT ] , (59)

where {Lθi}i=1nsuperscriptsubscriptsubscript𝐿subscript𝜃𝑖𝑖1𝑛\left\{L_{\theta_{i}}\right\}_{i=1}^{n}{ italic_L start_POSTSUBSCRIPT italic_θ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT end_POSTSUBSCRIPT } start_POSTSUBSCRIPT italic_i = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT are the symmetric logarithmic derivative (SLD) operators for the klimit-from𝑘k-italic_k -th parameter, implicitly defined as

ρ(θ)θk:=Lθkρ(θ)+ρ(θ)Lθk2.assign𝜌𝜃subscript𝜃𝑘subscript𝐿subscript𝜃𝑘𝜌𝜃𝜌𝜃subscript𝐿subscript𝜃𝑘2\dfrac{\partial\rho(\vec{\theta})}{\partial{\theta_{k}}}:=\dfrac{L_{\theta_{k}% }\rho(\vec{\theta})+\rho(\vec{\theta})L_{\theta_{k}}}{2}.divide start_ARG ∂ italic_ρ ( over→ start_ARG italic_θ end_ARG ) end_ARG start_ARG ∂ italic_θ start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT end_ARG := divide start_ARG italic_L start_POSTSUBSCRIPT italic_θ start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT end_POSTSUBSCRIPT italic_ρ ( over→ start_ARG italic_θ end_ARG ) + italic_ρ ( over→ start_ARG italic_θ end_ARG ) italic_L start_POSTSUBSCRIPT italic_θ start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT end_POSTSUBSCRIPT end_ARG start_ARG 2 end_ARG . (60)

We are only interested in single-parameter estimation, in this case the QFIM reads

θ=Tr[Lθ2ρ(θ)].subscript𝜃tracesuperscriptsubscript𝐿𝜃2𝜌𝜃\mathcal{I}_{\theta}=\Tr\left[L_{\theta}^{2}\rho(\theta)\right].caligraphic_I start_POSTSUBSCRIPT italic_θ end_POSTSUBSCRIPT = roman_Tr [ italic_L start_POSTSUBSCRIPT italic_θ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_ρ ( italic_θ ) ] . (61)

We are intended to obtain an operational expression for the SLD. In the eigenbasis of ρ(θ)𝜌𝜃\rho(\theta)italic_ρ ( italic_θ ), by means of the spectral theorem, the density matrix can be decomposed in terms of its eigenvalues and eigenvectors

ρ(θ)=i=1nλi(θ)|λi(θ)λi(θ)|.𝜌𝜃superscriptsubscript𝑖1𝑛subscript𝜆𝑖𝜃ketsubscript𝜆𝑖𝜃brasubscript𝜆𝑖𝜃\rho(\theta)=\sum_{i=1}^{n}\lambda_{i}(\theta)\ket{\lambda_{i}(\theta)}\bra{% \lambda_{i}(\theta)}.italic_ρ ( italic_θ ) = ∑ start_POSTSUBSCRIPT italic_i = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT italic_λ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ( italic_θ ) | start_ARG italic_λ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ( italic_θ ) end_ARG ⟩ ⟨ start_ARG italic_λ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ( italic_θ ) end_ARG | . (62)

Hence, in the eigenbasis of the state ρ(θ)𝜌𝜃\rho(\theta)italic_ρ ( italic_θ ), the SLD operator is simply given by

Lθ=2i,jλi(θ)|dρ(θ)dθ|λj(θ)λi(θ)+λj(θ)|λi(θ)λj(θ)|,subscript𝐿𝜃2subscript𝑖𝑗quantum-operator-productsubscript𝜆𝑖𝜃𝑑𝜌𝜃𝑑𝜃subscript𝜆𝑗𝜃subscript𝜆𝑖𝜃subscript𝜆𝑗𝜃ketsubscript𝜆𝑖𝜃brasubscript𝜆𝑗𝜃L_{\theta}=2\sum_{i,j}\dfrac{\left\langle\lambda_{i}(\theta)\middle|\dfrac{d% \rho(\theta)}{d\theta}\middle|\lambda_{j}(\theta)\right\rangle}{\lambda_{i}(% \theta)+\lambda_{j}(\theta)}\left|\lambda_{i}(\theta)\right\rangle\left\langle% \lambda_{j}(\theta)\right|,italic_L start_POSTSUBSCRIPT italic_θ end_POSTSUBSCRIPT = 2 ∑ start_POSTSUBSCRIPT italic_i , italic_j end_POSTSUBSCRIPT divide start_ARG ⟨ italic_λ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ( italic_θ ) | divide start_ARG italic_d italic_ρ ( italic_θ ) end_ARG start_ARG italic_d italic_θ end_ARG | italic_λ start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ( italic_θ ) ⟩ end_ARG start_ARG italic_λ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ( italic_θ ) + italic_λ start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ( italic_θ ) end_ARG | italic_λ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ( italic_θ ) ⟩ ⟨ italic_λ start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ( italic_θ ) | , (63)

where {|λk(θ)}k=1nsuperscriptsubscriptketsubscript𝜆𝑘𝜃𝑘1𝑛\left\{\ket{\lambda_{k}(\theta)}\right\}_{k=1}^{n}{ | start_ARG italic_λ start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT ( italic_θ ) end_ARG ⟩ } start_POSTSUBSCRIPT italic_k = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT is the eigenbasis of ρ(θ)𝜌𝜃\rho(\theta)italic_ρ ( italic_θ ) for λ(θ)i+λ(θ)j0,i,j=1,,nformulae-sequence𝜆subscript𝜃𝑖𝜆subscript𝜃𝑗0for-all𝑖𝑗1𝑛\lambda(\theta)_{i}+\lambda(\theta)_{j}\neq 0,\ \forall i,j=1,\ldots,nitalic_λ ( italic_θ ) start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT + italic_λ ( italic_θ ) start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ≠ 0 , ∀ italic_i , italic_j = 1 , … , italic_n. For our analysis, we use only the time as a parameter, giving

Lt=2i,jλi(t)|dρ(t)dt|λj(t)λi(t)+λj(t)|λi(t)λj(t)|,subscript𝐿𝑡2subscript𝑖𝑗quantum-operator-productsubscript𝜆𝑖𝑡𝑑𝜌𝑡𝑑𝑡subscript𝜆𝑗𝑡subscript𝜆𝑖𝑡subscript𝜆𝑗𝑡ketsubscript𝜆𝑖𝑡brasubscript𝜆𝑗𝑡L_{t}=2\sum_{i,j}\dfrac{\left\langle\lambda_{i}(t)\middle|\dfrac{d\rho(t)}{dt}% \middle|\lambda_{j}(t)\right\rangle}{\lambda_{i}(t)+\lambda_{j}(t)}\left|% \lambda_{i}(t)\right\rangle\left\langle\lambda_{j}(t)\right|,italic_L start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT = 2 ∑ start_POSTSUBSCRIPT italic_i , italic_j end_POSTSUBSCRIPT divide start_ARG ⟨ italic_λ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ( italic_t ) | divide start_ARG italic_d italic_ρ ( italic_t ) end_ARG start_ARG italic_d italic_t end_ARG | italic_λ start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ( italic_t ) ⟩ end_ARG start_ARG italic_λ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ( italic_t ) + italic_λ start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ( italic_t ) end_ARG | italic_λ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ( italic_t ) ⟩ ⟨ italic_λ start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ( italic_t ) | , (64)

and the QFI

Q=Tr[Lt2ρ(t)].subscript𝑄tracesuperscriptsubscript𝐿𝑡2𝜌𝑡\mathcal{I}_{Q}=\Tr\left[L_{t}^{2}\rho(t)\right].caligraphic_I start_POSTSUBSCRIPT italic_Q end_POSTSUBSCRIPT = roman_Tr [ italic_L start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_ρ ( italic_t ) ] . (65)