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

11institutetext: Center for Energy Research and Technology, North Carolina A&\&T State University, Greensboro, North Carolina 27411, USA

Jarzynski equality and the second law of thermodynamics beyond the weak-coupling limit: The quantum Brownian oscillator

Ilki Kim e-mail: hannibal.ikim@gmail.com
(August 12, 2025)
Abstract

We consider a time-dependent quantum linear oscillator coupled to a bath at an arbitrary strength. We then introduce a generalized Jarzynski equality (GJE) which includes the terms reflecting the system-bath coupling. This enables us to study systematically the coupling effect on the linear oscillator in a non-equilibrium process. This is also associated with the second law of thermodynamics beyond the weak-coupling limit. We next take into consideration the GJE in the classical limit. By this generalization we show that the Jarzynski equality in its original form can be associated with the second law, in both quantal and classical domains, only in the vanishingly small coupling regime.

pacs:
05.40.JcBrownian motion and 05.70.-aThermodynamics

1 Introduction

Since it was introduced, the Jarzynski equality (JE) has been attracting a great deal of interest due to its remarkable attribute. It explicitly states that if a given system, initially prepared in thermal equilibrium, is driven far from this equilibrium by an external perturbation, then this non-equilibrium process satisfies JAR97

eβW(tf)=𝑑WP(W)eβW(tf)=eβΔF,\left\langle e^{-\beta\,W(t_{\text{\tiny f}})}\right\rangle\,=\,\int dW\cdot P(W)\cdot e^{-\beta\,W(t_{\text{\tiny f}})}\,=\,e^{-\beta\,\Delta F}\,, (1)

where the symbol W(tf)W(t_{\text{\scriptsize f}}) denotes the microscopic work performed on the system in a single run for a variation of the external parameter {λ(t)| 0ttf}\{\lambda(t)\,|\,0\leq t\leq t_{\text{\scriptsize f}}\} according to the pre-determined protocol, and the symbol P(W)P(W) is the probability distribution of the work value WW; the symbol β=1/(kBT)\beta=1/(k_{\mbox{\tiny B}}T) denotes the inverse temperature of the environment, and ΔF\Delta F is the Helmholtz free energy change of the system between the initial and final states, equivalent to the average reversible work Wrev\langle W\rangle_{\text{\scriptsize rev}} in the corresponding isothermal process CAL85 . As such, the average \langle\cdots\rangle is evaluated over a large number of runs. Then, applying the Jensen inequality to the JE (1), we can easily obtain WΔF\langle W\rangle\geq\Delta F as an expression of the second law of the thermodynamics JAR08 ; BOK10 . Further, as a generalization of JE, Crooks introduced the fluctuation theorem given by CRO98 ; CRO99

Pf(+W)Pr(W)=exp{β(WΔF)},\frac{P_{\text{\scriptsize f}}(+W)}{P_{\text{\scriptsize r}}(-W)}\;=\;\exp\{\beta\,(W-\Delta F)\}\,, (2)

where the symbols Pf(W)P_{\text{\scriptsize f}}(W) and Pr(W)P_{\text{\scriptsize r}}(W) are the probability densities of performing the work WW when its protocol runs in the forward and reverse directions, respectively.

The JE has been verified in experiments with small-scale systems, where the fluctuations of work values are sufficiently observable JAR02 , e.g., in determination of the free energy change between the unfolded and folded conformations of a single RNA molecule LIP02 ; BUS05 . However, its strict validity is still in dispute, especially in association with the second law beyond the weak coupling between system and bath (see, e.g., an instructive critique in COH04 as well as Jarzynski’s reply in JAR04 ).

The JE has also been discussed in the quantum domain, where no notion of trajectory in the phase space is available and so instead the spectrum information has to be used for determination of the work probability distribution P(W)P(W). Most of those attempts were made within isolated or weakly-coupled systems TAS00 ; MUK03 ; ROE04 ; ESP06 ; TAL07 ; MAH07 ; ENG07 ; CRO08 ; AND08 ; DEF08 ; TAK09 ; MAH10 ; MAH11 ; NGO12 . However, the finite coupling strength between system and bath in small-scale devices gives rise to some quantum subtleties which can no longer be neglected for studying the thermodynamic properties of such devices SHE02 ; MAH04 ; CAP05 .

In CAM09 , the fluctuation theorem (2), immediately reproducing the JE, was discussed for arbitrary open quantum systems with no restriction on the coupling strength. Its key idea was such that if the total system (i.e., open quantum system plus bath) is initially in a thermal canonical state, but otherwise isolated, and an external perturbation λ(t)\lambda(t) acts solely on the open system, then the work performed on the total system may be interpreted as the work performed on the open system alone. Then the final result was explicitly obtained [cf. (17)-(17a)] by mimicking the classical case such that the work on the system-of-interest Hs(x,λ)H_{s}(x,\lambda) beyond the weak-coupling limit is directly related to the free energy change of a potential of mean force JAR04

s(x,λ)\displaystyle{\mathcal{H}}_{s}^{*}(x,\lambda) =\displaystyle= Hs(x,λ)β1×\displaystyle H_{s}(x,\lambda)\,-\,\beta^{-1}\times
ln(𝑑yexp[β{Hb(y)+hint(x,y)}]𝑑yexp{βHb(y)}),\displaystyle\ln\left(\frac{\int dy\,\exp\left[-\beta\,\{H_{b}(y)\,+\,h_{\text{\tiny int}}(x,y)\}\right]}{\int dy\,\exp\left\{-\beta\,H_{b}(y)\right\}}\right)\,,

where the symbols Hb(y)H_{b}(y) and hint(x,y)h_{\text{\tiny int}}(x,y) denote the bath and interaction Hamiltonians, respectively; note here that on the right-hand side the λ\lambda-dependency exists only in Hs(x,λ)H_{s}(x,\lambda) [cf. (16)]. This quantum result was subsequently applied to a solvable model (e.g., TAL09 ).

However, as will be discussed below, it is not clear if this quantum fluctuation theorem is associated directly with the system-of-interest H^s(t)\hat{H}_{s}(t) alone, beyond the weak-coupling limit, being the quantum description of Hs(x,λ)H_{s}(x,\lambda) in (1); in fact, an attempt to relate the quantum description of s(x,λ){\mathcal{H}}_{s}^{*}(x,\lambda) to the second law of thermodynamics within the coupled system H^s(t)\hat{H}_{s}(t) would even lead to a violation of this law. In this paper we intend to resolve this issue, within a time-dependent quantum Brownian oscillator as a mathematically manageable scheme, by introducing explicitly a generalized Jarzynski equality (GJE) directly associated with the open system H^s(t)\hat{H}_{s}(t) with no restriction on the coupling strength, and then discussing the resultant second law with no violation. The general layout of this paper is the following. In Sect. 2 we briefly review the results known from the references and useful for our discussions. In Sect. 3 we discuss the second law beyond the weak-coupling regime, which shows that the JE in its known form can be associated with this law only in the vanishingly small coupling regime. In Sect. 4 we introduce the GJE consistent with the second law, valid at an arbitrary coupling strength. Finally we give the concluding remarks of this paper in Sect. 5.

2 Quantum Brownian oscillator in a non-equilibrium thermal process

The quantum Brownian oscillator in consideration is given by the Caldeira-Leggett model Hamiltonian ING98 ; WEI08

H^(t)=H^s(t)+H^b+H^sb,\hat{H}(t)\;=\;\hat{H}_{s}(t)\,+\,\hat{H}_{b}\,+\,\hat{H}_{sb}\,, (4)

where a system linear oscillator, a bath, and a system-bath interaction are given by

H^s(t)\displaystyle\hat{H}_{s}(t)\; =p^22M+My2(t)2q^2\displaystyle=\;\frac{\hat{p}^{2}}{2M}\,+\,\frac{M\,y^{2}(t)}{2}\,\hat{q}^{2} (4a)
H^b\displaystyle\hat{H}_{b}\; =j=1N(p^j22mj+mjωj22x^j2)\displaystyle=\;\sum_{j=1}^{N}\left(\frac{\hat{p}_{j}^{2}}{2m_{j}}\,+\,\frac{m_{j}\,\omega_{j}^{2}}{2}\,\hat{x}_{j}^{2}\right) (4b)
H^sb\displaystyle\hat{H}_{sb}\; =q^j=1Ncjx^j+q^2j=1Ncj22mjω¯j2,\displaystyle=\;-\hat{q}\sum_{j=1}^{N}c_{j}\,\hat{x}_{j}\,+\,\hat{q}^{2}\sum_{j=1}^{N}\frac{c_{j}^{2}}{2m_{j}\,\bar{\omega}_{j}^{2}}\,, (4c)

respectively. Here the angular frequency y(t)>0y(t)>0 varies in the time interval [0,tf][0,t_{\text{\scriptsize f}}] according to an arbitrary but pre-determined protocol (for the sake of convenience, let y0=y(0)y_{0}=y(0) in what follows), and the constants cjc_{j} denote the coupling strengths. The total system H^(0)\hat{H}(0) initially prepared is in a canonical thermal equilibrium state ρ^β=eβH^(0)/Zβ(y0)\hat{\rho}_{\beta}=e^{-\beta\hat{H}(0)}/Z_{\beta}(y_{0}) with the initial partition function Zβ(y0)Z_{\beta}(y_{0}). Then the initial internal energy of the coupled oscillator is given by Tr{H^s(0)ρ^β}=Trs{H^s(0)R^s(0)}\mbox{Tr}\{\hat{H}_{s}(0)\,\hat{\rho}_{\beta}\}=\mbox{Tr}_{s}\{\hat{H}_{s}(0)\,\hat{R}_{s}(0)\}, where the initial state of the oscillator R^s(0)=Trb(ρ^β)\hat{R}_{s}(0)=\mbox{Tr}_{b}(\hat{\rho}_{\beta}). Here the symbol Trb\mbox{Tr}_{b} denotes the partial trace for the bath degrees of freedom only; it is explicitly given by GRA88 ; WEI08

q|R^s(0)|q\displaystyle\langle q|\hat{R}_{s}(0)|q^{\prime}\rangle =\displaystyle= 12πq^2β×\displaystyle\frac{1}{\sqrt{2\pi\langle\hat{q}^{2}\rangle_{\beta}}}\,\times
exp{(q+q)28q^2βp^2β(qq)222},\displaystyle\exp\left\{-\frac{(q+q^{\prime})^{2}}{8\,\langle\hat{q}^{2}\rangle_{\beta}}-\frac{\langle\hat{p}^{2}\rangle_{\beta}\cdot(q-q^{\prime})^{2}}{2\hbar^{2}}\right\}\,,

expressed in terms of the equilibrium fluctuations, q^2β=Tr(q^2ρ^β)\langle\hat{q}^{2}\rangle_{\beta}=\mbox{Tr}(\hat{q}^{2}\hat{\rho}_{\beta}) and p^2β=Tr(p^2ρ^β)=M2q^˙2β\langle\hat{p}^{2}\rangle_{\beta}=\mbox{Tr}(\hat{p}^{2}\hat{\rho}_{\beta})=M^{2}\langle\dot{\hat{q}}^{2}\rangle_{\beta}. Beyond the weak-coupling limit, this reduced density matrix is not any longer in form of a canonical thermal state eβH^s(0)\propto e^{-\beta\hat{H}_{s}(0)}, immediately leading to the fact that the coupled oscillator H^s(0)\hat{H}_{s}(0) is not with its well-defined local equilibrium temperature KIM10 .

For the below discussions, we will need the fluctuation-dissipation theorem in the initial (equilibrium) state FOR88

12q^(t1)q^(t2)+q^(t2)q^(t1)β=π×\displaystyle\frac{1}{2}\,\left\langle\hat{q}(t_{1})\,\hat{q}(t_{2})\,+\,\hat{q}(t_{2})\,\hat{q}(t_{1})\right\rangle_{\beta}\;=\;\frac{\hbar}{\pi}\,\times
0𝑑ωcoth(βω2)cos{ω(t2t1)}Im{χ~(ω+i 0+)}.\displaystyle\int_{0}^{\infty}d\omega\,\coth\left(\frac{\beta\hbar\omega}{2}\right)\;\cos\{\omega(t_{2}-t_{1})\}\;\mbox{Im}\{\tilde{\chi}(\omega+i\,0^{+})\}\,.

Here the dynamic susceptibility is given by

χ~(ω)=1M1y02ω2iωγ~(ω),\tilde{\chi}(\omega)\;=\;\frac{1}{M}\,\frac{1}{y_{0}^{2}-\omega^{2}-i\omega\,\tilde{\gamma}(\omega)}\,, (7)

where the symbol γ~(ω)\tilde{\gamma}(\omega) denotes the Fourier-Laplace transformed damping kernel. This can be rewritten as LEV88

χ~(ω)=1Mj=1N(ω2ωj2)k=0N(ω2ω¯k2)\displaystyle\tilde{\chi}(\omega)\;=\;-\frac{1}{M}\,\frac{\displaystyle\prod_{j=1}^{N}\,(\omega^{2}-\omega_{j}^{2})}{\displaystyle\prod_{k=0}^{N}\,(\omega^{2}-\bar{\omega}_{k}^{2})} (7a)

in terms of the normal-mode frequencies {ω¯k}\{\bar{\omega}_{k}\} of the total system H^(0)\hat{H}(0). It is straightforward to verify that Im{χ~(ω+i0+)}π/(2My0)δ(ωy0)\mbox{Im}\{\tilde{\chi}(\omega+i0^{+})\}\to\pi/(2My_{0})\cdot\delta(\omega-y_{0}) for an uncoupled (or isolated) oscillator (i.e., all system-bath coupling strengths cj0c_{j}\equiv 0).

From (2), it follows that FOR85

q^2β\displaystyle\hskip-19.91684pt\langle\hat{q}^{2}\rangle_{\beta} =\displaystyle= π0𝑑ωcoth(βω2)Im{χ~(ω+i 0+)}\displaystyle\frac{\hbar}{\pi}\,\int_{0}^{\infty}d\omega\,\coth\left(\frac{\beta\hbar\omega}{2}\right)\,\mbox{Im}\{\tilde{\chi}(\omega+i\,0^{+})\} (8a)
q^˙2β\displaystyle\hskip-19.91684pt\langle\dot{\hat{q}}^{2}\rangle_{\beta} =\displaystyle= π0𝑑ωω2coth(βω2)Im{χ~(ω+i 0+)},\displaystyle\frac{\hbar}{\pi}\,\int_{0}^{\infty}d\omega\,\omega^{2}\,\coth\left(\frac{\beta\hbar\omega}{2}\right)\,\mbox{Im}\{\tilde{\chi}(\omega+i\,0^{+})\}\,, (8b)

as well-known, both of which give the initial internal energy of the coupled oscillator

Us(0):=H^s(0)β=M2π×\displaystyle U_{s}(0)\;:=\;\langle\hat{H}_{s}(0)\rangle_{\beta}\;=\;\frac{M\hbar}{2\pi}\,\times (9)
0𝑑ω(y02+ω2)coth(βω2)Im{χ~(ω+i 0+)}.\displaystyle\int_{0}^{\infty}d\omega\,(y_{0}^{2}+\omega^{2})\,\coth\left(\frac{\beta\hbar\omega}{2}\right)\cdot\mbox{Im}\{\tilde{\chi}(\omega+i\,0^{+})\}\,.

In the absence of the system-bath coupling, this reduces to the well-known expression of internal energy CAL85

e(β,y0)=y0(12+n^β)=y02coth(βy02),\displaystyle e(\beta,y_{0})\;=\;\hbar y_{0}\left(\frac{1}{2}\,+\,\langle\hat{n}\rangle_{\beta}\right)\;=\;\frac{\hbar y_{0}}{2}\,\coth\left(\frac{\beta\hbar y_{0}}{2}\right)\,, (9a)

where the average quantum number n^β=1/(eβy01)\langle\hat{n}\rangle_{\beta}=1/(e^{\beta\hbar y_{0}}-1). With β0\beta\hbar\to 0, its classical value also appears as ecl(β)=1/βe_{\text{\scriptsize cl}}(\beta)=1/\beta.

In comparison, there is a widely used alternative approach to the thermodynamic energy of the coupled oscillator in equilibrium, given by 𝒰s(0):=(/β)ln𝒵β(y0){\mathcal{U}}_{s}^{*}(0):=-(\partial/\partial\beta)\,\ln{\mathcal{Z}}_{\beta}^{*}(y_{0}) FOR05 ; HAE05 ; HAE06 ; FOR06 ; KIM06 ; KIM07 ; the symbol 𝒵β(y0){\mathcal{Z}}_{\beta}^{*}(y_{0}) denotes the reduced partition function associated with the Hamiltonian of mean force CAM09 ; JAR04

^s(t)|t=0:=1βln[Trb{eβH^(t)}Trb(eβH^b)]|t=0,\left.\hat{{\mathcal{H}}}_{s}^{*}(t)\right|_{t=0}\;:=\;-\frac{1}{\beta}\,\left.\ln\left[\frac{\mbox{Tr}_{b}\{e^{-\beta\hat{H}(t)}\}}{\mbox{Tr}_{b}(e^{-\beta\hat{H}_{b}})}\right]\right|_{t=0}\,, (10)

and thus 𝒵β(y0)=Trs{eβ^s(0)}=Zβ(y0)/(Zb)β{\mathcal{Z}}_{\beta}^{*}(y_{0})=\mbox{Tr}_{s}\{e^{-\beta\hat{{\mathcal{H}}}_{s}(0)}\}=Z_{\beta}(y_{0})/(Z_{b})_{\beta} with the partition function (Zb)β(Z_{b})_{\beta} associated with the isolated bath H^b\hat{H}_{b}. With the vanishing coupling strengths (cj=0c_{j}=0), this partition function precisely reduces, as required, to its standard form of zβ(y0)={csch(βy0/2)}/2z_{\beta}(y_{0})=\{\mbox{csch}(\beta\hbar y_{0}/2)\}/2 for an isolated oscillator. Then it can be shown that HAE06

𝒰s(0)=1β{1+n=12y02+νnγ^(νn)νn2γ^(νn)νn2+νnγ^(νn)+y02}{\mathcal{U}}_{s}^{*}(0)\;=\;\frac{1}{\beta}\,\left\{1+\sum_{n=1}^{\infty}\frac{2\,y_{0}^{2}+\nu_{n}\,\hat{\gamma}(\nu_{n})-\nu_{n}^{2}\,\hat{\gamma}^{\prime}(\nu_{n})}{\nu_{n}^{2}+\nu_{n}\,\hat{\gamma}(\nu_{n})+y_{0}^{2}}\right\} (11a)
with νn=2πn/β\nu_{n}=2\pi n/\beta\hbar and γ^(z)=γ~(iz)\hat{\gamma}(z)=\tilde{\gamma}(iz), whereas
Us(0)=1β{1+n=12y02+νnγ^(νn)νn2+νnγ^(νn)+y02}.U_{s}(0)\;=\;\frac{1}{\beta}\,\left\{1+\sum_{n=1}^{\infty}\frac{2\,y_{0}^{2}+\nu_{n}\,\hat{\gamma}(\nu_{n})}{\nu_{n}^{2}+\nu_{n}\,\hat{\gamma}(\nu_{n})+y_{0}^{2}}\right\}\,. (11b)

It is seen that 𝒰s(0)Us(0){\mathcal{U}}_{s}^{*}(0)\neq U_{s}(0) unless the damping model is Ohmic. In fact, all thermodynamic quantities resulting from the partition function 𝒵β(y0){\mathcal{Z}}_{\beta}^{*}(y_{0}) cannot exactly describe the well-defined thermodynamics of the reduced system H^s(0)\hat{H}_{s}(0) in its state (2) beyond the weak-coupling limit KIM10 ; see also ING12 for interesting discussions of the different behaviors between ^s(0)\hat{{\mathcal{H}}}_{s}^{*}(0) and H^s(0)\hat{H}_{s}(0) in terms of the specific heat, within a free damped quantum particle given by H^(t)\hat{H}(t) with y(t)0y(t)\equiv 0 in (4).

It is also instructive to consider a quasi-static (or reversible) process briefly, which the system of interest undergoes change infinitely slowly throughout. Then the coupled oscillator remains in equilibrium exactly in form of Eq. (2) in every single step such that

q|R^eq{y(t)}|q=q|R^s(0)|q|y0y(t).\langle q|\hat{R}_{\text{\scriptsize eq}}\{y(t)\}|q^{\prime}\rangle\;=\;\left.\langle q|\hat{R}_{s}(0)|q^{\prime}\rangle\right|_{y_{0}\to y(t)}\,. (12)

Here the time tt is understood merely as a parameter specifying the frequency value y(t)y(t). Accordingly, the thermodynamic energy 𝒰s(t){\mathcal{U}}_{s}^{*}(t) turns out to be in form of (11a) with y(t)y0y(t)\leftarrow y_{0}. For later discussions, we rewrite it as FOR85 ; KIM07

𝒰s(t)=1π0𝑑ωe(β,ω)Im{ddωlnχ~t(ω+i0+)},{\mathcal{U}}_{s}^{*}(t)\,=\,\frac{1}{\pi}\int_{0}^{\infty}d\omega\,e(\beta,\omega)\cdot\mbox{Im}\left\{\frac{d}{d\omega}\,\ln\tilde{\chi}_{t}(\omega+i0^{+})\right\}\,, (13)

where the second factor of the integrand

Im{ddωlnχ~t(ω+i0+)}\displaystyle\;\mbox{Im}\left\{\frac{d}{d\omega}\,\ln\tilde{\chi}_{t}(\omega+i0^{+})\right\}
=\displaystyle= π{k=0Nδ(ωω¯k,t)j=1Nδ(ωωj)}.\displaystyle\;\pi\left\{\sum_{k=0}^{N}\delta(\omega-\bar{\omega}_{k,t})\,-\,\sum_{j=1}^{N}\delta(\omega-\omega_{j})\right\}\,. (13a)

Here the susceptibility χ~t(ω)\tilde{\chi}_{t}(\omega) is given by (7) with y(t)y0y(t)\leftarrow y_{0}, as well as {ω¯k,t}={ω¯k}y0y(t)\{\bar{\omega}_{k,t}\}=\{\bar{\omega}_{k}\}_{y_{0}\to y(t)}. In the absence of the system-bath coupling, Eq. (13a) reduces to πδ{ωy(t)}\pi\,\delta\{\omega-y(t)\}, as required. Likewise, the free energy defined as s(t):=(1/β)ln𝒵β{y(t)}{\mathcal{F}}_{s}^{*}(t):=-(1/\beta)\,\ln{\mathcal{Z}}_{\beta}^{*}\{y(t)\}, being the total system free energy minus the bare bath free energy, can also be expressed as FOR85

s(t)=1π0𝑑ωf(β,ω)Im{ddωlnχ~t(ω+i0+)},{\mathcal{F}}_{s}^{*}(t)\,=\,\frac{1}{\pi}\int_{0}^{\infty}d\omega\,f(\beta,\omega)\cdot\mbox{Im}\left\{\frac{d}{d\omega}\,\ln\tilde{\chi}_{t}(\omega+i0^{+})\right\}\,, (14)

where the free energy of an isolated oscillator

f(β,ω)=ω2+1βln(1eβω),\displaystyle f(\beta,\omega)\;=\;\frac{\hbar\omega}{2}\,+\,\frac{1}{\beta}\,\ln\left(1-e^{-\beta\hbar\omega}\right)\,, (14a)

with fcl(β,ω)={ln(βω)}/βf_{\text{\scriptsize cl}}(\beta,\omega)=\{\ln(\beta\hbar\omega)\}/\beta in the classical limit.

Further, to see explicitly the different behaviors of (10) from its classical value (1), we now apply to the exponentiated negative total Hamiltonian H^(t)\hat{H}(t) in (10) the Zassenhaus formula with X^:=H^s(t)\hat{X}:=\hat{H}_{s}(t) and Y^:=H^b+H^sb\hat{Y}:=\hat{H}_{b}+\hat{H}_{sb} WIL67 ,

es(X^+Y^)=esX^esY^es2C^2es3C^3e^{s\,(\hat{X}+\hat{Y})}\;=\;e^{s\,\hat{X}}\cdot e^{s\,\hat{Y}}\cdot e^{s^{2}\,\hat{C}_{2}}\cdot e^{s^{3}\,\hat{C}_{3}}\,\cdots (15)

where C^2=(1/2)[X^,Y^]\hat{C}_{2}=-(1/2)\,[\hat{X},\hat{Y}], and C^3=(2/3)[Y^,C^2](1/3)[X^,C^2]\hat{C}_{3}=-(2/3)\,[\hat{Y},\hat{C}_{2}]-(1/3)\,[\hat{X},\hat{C}_{2}]; this enables Eq. (10) to be rewritten as

^s(t)=H^s(t)1βln{eβY^eβ2C^2eβ3C^3(t)Trb(eβH^b)}.\hat{\mathcal{H}}_{s}^{*}(t)\;=\;\hat{H}_{s}(t)\,-\,\frac{1}{\beta}\,\ln\left\{\frac{e^{-\beta\,\hat{Y}}\,e^{\beta^{2}\,\hat{C}_{2}}\,e^{-\beta^{3}\,\hat{C}_{3}(t)}\,\cdots}{\mbox{Tr}_{b}(e^{-\beta\,\hat{H}_{b}})}\right\}\,. (16)

Here the operator C^3=C^3(t)\hat{C}_{3}=\hat{C}_{3}(t), and so the second term on the right-hand side is time-dependent, which is not the case in its classical value (1), to be noted for our discussions below.

An extension of the Crooks fluctuation theorem (2) to arbitrary open quantum systems was introduced in CAM09 , which is valid regardless of the system-bath coupling strength; this reads as

Pf(+W)Pr(W)=eβ(WΔs),\frac{P_{\text{\scriptsize f}}(+W)}{P_{\text{\scriptsize r}}(-W)}\;=\;e^{\beta\,(W-\Delta{\mathcal{F}}_{s}^{*})}\,, (17)

where Δs(tf)=s(tf)s(0)\Delta{\mathcal{F}}_{s}^{*}(t_{\text{\scriptsize f}})={\mathcal{F}}_{s}^{*}(t_{\text{\scriptsize f}})-{\mathcal{F}}_{s}^{*}(0), explicitly given in (14) for a coupled oscillator. This fluctuation theorem leads to the Jarzynski equality

eβWPf(W)=eβΔs,\displaystyle\langle e^{-\beta\,W}\rangle_{\scriptscriptstyle{P}_{\text{\tiny f}}(W)}\;=\;e^{-\beta\,\Delta{\mathcal{F}}_{s}^{*}}\,, (17a)

and it follows that the average work performed by the external perturbation satisfies WΔs\langle W\rangle\geq\Delta{\mathcal{F}}_{s}^{*}. However, it is normally non-trivial to determine the work distribution Pf(W)P_{\text{\scriptsize f}}(W) explicitly, which can be obtained only from the spectrum information of the (isolated) total system H^(t)\hat{H}(t). Further, from the time-dependent behavior of (16) it is not apparent whether the free energy change Δs\Delta{\mathcal{F}}_{s}^{*} is precisely tantamount to the minimum work on the system-of-interest H^s(t)\hat{H}_{s}(t) only, especially beyond the weak-coupling limit.

For later purposes, we also point out that an explicit expression of the work distribution is known for an isolated oscillator as TAL07 ; DEF08

P{W(tf,y0)}=m,n\displaystyle P\{W(t_{\text{\scriptsize f}},y_{0})\}\,=\,\sum_{m,n} δ(W[Em{y(tf)}En(y0)])×\displaystyle\delta\left(W-[E_{m}\{y(t_{\text{\scriptsize f}})\}-E_{n}(y_{0})]\right)\times (18)
Pm,n{y(tf)}Pn(y0)\displaystyle P_{m,n}\{y(t_{\text{\scriptsize f}})\}\cdot P_{n}(y_{0})

with the microscopic work W(tf,y0)W(t_{\text{\scriptsize f}},y_{0}) in a single event of the variation {y0y(tf)}\{y_{0}\to y(t_{\text{\scriptsize f}})\}. Here En{y(t)}=y(t)(n+1/2)E_{n}\{y(t)\}=\hbar y(t)\,(n+1/2) is the instantaneous energy eigenvalue, and the symbol Pn(y0)=eβEn(y0)/zβ(y0)P_{n}(y_{0})=e^{-\beta\,E_{n}(y_{0})}/z_{\beta}(y_{0}) denotes the probability of the occupation number in the initial preparation at temperature TT, as well as Pm,n{y(tf)}=|ψm(tf)|U^(tf)|ψn(0)|2P_{m,n}\{y(t_{\text{\scriptsize f}})\}\;=\;|\langle\psi_{m}(t_{\text{\scriptsize f}})|\hat{U}(t_{\text{\scriptsize f}})|\psi_{n}(0)\rangle|^{2} the transition probability between initial and final states nn and mm, expressed in terms of the instantaneous eigenstates |ψn(t)|\psi_{n}(t)\rangle and the time-evolution U^(tf)\hat{U}(t_{\text{\scriptsize f}}) of the time-dependent harmonic oscillator. Then the JE is explicitly given by eβWP=eβΔf{y(tf),y0}\langle e^{-\beta\,W}\rangle_{\scriptscriptstyle{P}}=e^{-\beta\,\Delta f\{y(t_{\text{\tiny f}}),y_{0}\}}, where the free energy change Δf{y(tf),y0}=f{β,y(tf)}f(β,y0)\Delta f\{y(t_{\text{\scriptsize f}}),y_{0}\}=f\{\beta,y(t_{\text{\scriptsize f}})\}-f(\beta,y_{0}); and the average work turns out to be WP=𝑑WWP(W)=(/2){Qy(tf)y0}coth(βy0/2)e{β,y(tf)}e(β,y0)\langle W\rangle_{\scriptscriptstyle{P}}=\int_{-\infty}^{\infty}dW\,W\,P(W)=(\hbar/2)\,\{Q^{*}\,y(t_{\text{\scriptsize f}})-y_{0}\}\,\coth(\beta\hbar y_{0}/2)\neq e\{\beta,y(t_{\text{\scriptsize f}})\}-e(\beta,y_{0}) DEF08 , where Q{y0,y(tf)}=Q^{*}\{y_{0},y(t_{\text{\scriptsize f}})\}=

12y0y(tf)[y02{y2(tf)X2+X˙2}+{y2(tf)Y2+Y˙2}]\frac{1}{2\,y_{0}\,y(t_{\text{\scriptsize f}})}\left[y_{0}^{2}\,\{y^{2}(t_{\text{\scriptsize f}})\cdot X^{2}+\dot{X}^{2}\}+\{y^{2}(t_{\text{\scriptsize f}})\cdot Y^{2}+\dot{Y}^{2}\}\right] (19)

with both XX and YY expressed in terms of the Airy functions Ai and Bi. We stress that this average work evaluated with the distribution P(W)P(W) is not a quantum-mechanical expectation value of an observable TAL07 .

3 Discussion of the second law within the Drude-Ullersma model

Now we intend to discuss the second law within an oscillator coupled to a bath at an arbitrary strength. We do this in the Drude-Ullersma model for the damping kernel in (7) with γ~d(ω+i0+)=γoωd/(ωdiω)\tilde{\gamma}_{d}(\omega+i0^{+})=\gamma_{\mbox{\tiny o}}\,\omega_{d}/(\omega_{d}-i\omega), where a damping parameter γo\gamma_{\mbox{\tiny o}}, representing the system-bath coupling strength, and a cut-off frequency ωd\omega_{d} ULL66 . It is convenient to adopt, in place of {y(t),ωd,γo}\{y(t),\omega_{d},\gamma_{\mbox{\tiny o}}\}, the parameters {y¯t,Ωt,γ¯t}\{\bar{y}_{t},\Omega_{t},\bar{\gamma}_{t}\} through the relations FOR06

y2(t)=y¯t2ΩtΩt+γ¯t;ωd=Ωt+γ¯t\displaystyle\displaystyle y^{2}(t)\;=\;\bar{y}_{t}^{2}\;\frac{\Omega_{t}}{\Omega_{t}\,+\,\bar{\gamma}_{t}}\;\;;\;\;\omega_{d}\;=\;\Omega_{t}\,+\,\bar{\gamma}_{t}
γo=γ¯tΩt(Ωt+γ¯t)+y¯t2(Ωt+γ¯t)2.\displaystyle\displaystyle\gamma_{\mbox{\tiny o}}\;=\;\bar{\gamma}_{t}\;\frac{\Omega_{t}\,(\Omega_{t}\,+\,\bar{\gamma}_{t})\,+\,\bar{y}_{t}^{2}}{(\Omega_{t}\,+\,\bar{\gamma}_{t})^{2}}\,. (20)

Then it can be shown that

Ωt+zt,1+zt,2=ωd;Ωtzt,1zt,2=y2(t)ωd\displaystyle\Omega_{t}+z_{t,1}+z_{t,2}\,=\,\omega_{d}\;\;;\;\;\Omega_{t}\,z_{t,1}\,z_{t,2}\,=\,y^{2}(t)\cdot\omega_{d} (21)
Ωtzt,1+zt,1zt,2+zt,2Ωt=y¯t2+Ωtγ¯t=y2(t)+ωdγo,\displaystyle\Omega_{t}\,z_{t,1}+z_{t,1}\,z_{t,2}+z_{t,2}\,\Omega_{t}\,=\,\bar{y}_{t}^{2}+\Omega_{t}\,\bar{\gamma}_{t}\,=\,y^{2}(t)+\omega_{d}\,\gamma_{\mbox{\tiny o}}\,,

where (zt,1,zt,2)=(γ¯t/2+i𝐰¯t,γ¯t/2i𝐰¯t)(z_{t,1},z_{t,2})=(\bar{\gamma}_{t}/2+i\bar{{\mathbf{w}}}_{t},\bar{\gamma}_{t}/2-i\bar{{\mathbf{w}}}_{t}) with 𝐰¯t={y¯t2(γ¯t/2)2}1/2\bar{{\mathbf{w}}}_{t}=\{\bar{y}_{t}^{2}-(\bar{\gamma}_{t}/2)^{2}\}^{1/2}. Assuming that γo0\gamma_{\mbox{\tiny o}}\neq 0, the susceptibility (7) reduces to FOR06 ; KIM06 ; KIM07

χ~t(d)(ω+i0+)=1Mω+i(Ωt+zt,1+zt,2)(ω+iΩt)(ω+izt,1)(ω+izt,2).\tilde{\chi}_{t}^{(d)}(\omega+i0^{+})\;=\;-\frac{1}{M}\,\frac{\omega\,+\,i\,(\Omega_{t}\,+\,z_{t,1}\,+\,z_{t,2})}{(\omega\,+\,i\Omega_{t})(\omega\,+\,iz_{t,1})(\omega\,+\,iz_{t,2})}\,. (22)

In comparison, it turns out that in an isolated case (γo=0\gamma_{\mbox{\tiny o}}=0), we have γ¯t=0\bar{\gamma}_{t}=0 and so y(t)=y¯t=𝐰¯ty(t)=\bar{y}_{t}=\bar{{\mathbf{w}}}_{t} and ωd=Ωt\omega_{d}=\Omega_{t}, as well as zt,1=iy(t)z_{t,1}=iy(t) and zt,2=iy(t)z_{t,2}=-iy(t), therefore Eq. (22) reduces to (M{y2(t)ω2})1(M\{y^{2}(t)-\omega^{2}\})^{-1}, simply real-valued; in this case, the imaginary number, πi/{2My(t)}δ{ωy(t)}\pi i/\{2My(t)\}\cdot\delta\{\omega-y(t)\} must be added to this real number for the actual susceptibility.

Now we substitute (22) into (13), which gives

𝒰s(t)=0𝑑ωe(β,ω)𝒫t(ω),{\mathcal{U}}_{s}^{*}(t)\;=\;\int_{0}^{\infty}d\omega\;e(\beta,\omega)\cdot{\mathcal{P}}_{t}^{*}(\omega)\,, (23)

where the distribution

𝒫t(ω)=\displaystyle{\mathcal{P}}_{t}^{*}(\omega)\;= 1πIm{ddωlnχ~t(ω+i0+)}\displaystyle\;\frac{1}{\pi}\;\mbox{Im}\left\{\frac{d}{d\omega}\,\ln\tilde{\chi}_{t}(\omega+i0^{+})\right\}
=\displaystyle= 1π[{l=13ωl¯(t)ω2+ωl¯2(t)}ωdω2+ωd2]\displaystyle\;\frac{1}{\pi}\left[\left\{\sum_{l=1}^{3}\frac{\underline{\omega_{l}}(t)}{\omega^{2}+\underline{\omega_{l}}^{2}(t)}\right\}\,-\,\frac{\omega_{d}}{\omega^{2}+\omega_{d}^{2}}\right] (23a)

with (ω1¯,ω2¯,ω3¯):=(Ωt,zt,1,zt,2)(\underline{\omega_{1}},\underline{\omega_{2}},\underline{\omega_{3}}):=(\Omega_{t},z_{t,1},z_{t,2}), as well as with the normalization 0𝑑ω𝒫t(ω)=1\int_{0}^{\infty}d\omega\,{\mathcal{P}}_{t}^{*}(\omega)=1 for γo0\gamma_{\mbox{\tiny o}}\neq 0 ILK13 . With the aid of (21), Eq. (23a) can be rewritten as a compact expression

𝒫t(ω)=ωd2γogt(ω)/π(ω2+ωd2)gt(ω),{\mathcal{P}}_{t}^{*}(\omega)\;=\;\frac{\omega_{d}^{2}\,\gamma_{\mbox{\tiny o}}\cdot g_{t}^{*}(\omega)/\pi}{(\omega^{2}+\omega_{d}^{2})\cdot g_{t}(\omega)}\,, (24)

where both factors

gt(ω)\displaystyle g_{t}(\omega)\; =(ω2+Ωt2){γ¯t2ω2+(ω2y¯t2)2}\displaystyle=\;(\omega^{2}+\Omega_{t}^{2})\,\{\bar{\gamma}_{t}^{2}\,\omega^{2}+(\omega^{2}-\bar{y}_{t}^{2})^{2}\} (24a)
gt(ω)\displaystyle g_{t}^{*}(\omega)\; = 3ω4+{ωd2ωdγoy2(t)}ω2+{ωdy(t)}2.\displaystyle=\;3\,\omega^{4}\,+\,\{\omega_{d}^{2}-\omega_{d}\,\gamma_{\mbox{\tiny o}}-y^{2}(t)\}\,\omega^{2}\,+\,\{\omega_{d}\cdot y(t)\}^{2}\,. (24b)

It is observed that the polynomial gt(ω)g_{t}(\omega) is non-negative, while gt(ω)g_{t}^{*}(\omega) can be negative and so can 𝒫t(ω){\mathcal{P}}_{t}^{*}(\omega); the behaviors of 𝒫t(ω){\mathcal{P}}_{t}^{*}(\omega) are plotted in Fig. 1. Then it is easy to see that in the classical case, Eq. (23) reduces to 𝒰s,cl(t)=ecl(β){\mathcal{U}}_{s,\text{\scriptsize cl}}^{*}(t)=e_{\text{\scriptsize cl}}(\beta) regardless of the coupling strength γo\gamma_{\mbox{\tiny o}}. Likewise, we also have the free energy expressed as

s(t)=0𝑑ωf(β,ω)𝒫t(ω).{\mathcal{F}}_{s}^{*}(t)\;=\;\int_{0}^{\infty}d\omega\;f(\beta,\omega)\cdot{\mathcal{P}}_{t}^{*}(\omega)\,. (25)

In comparison, Eq. (23a) identically vanishes in an isolated case; however, for the aforementioned reason, we have 𝒫t(ω)=δ{ωy(t)}{\mathcal{P}}_{t}^{*}(\omega)=\delta\{\omega-y(t)\} indeed in this case. Therefore, we may say that the smoothness of the distribution (24) reflects the system-bath coupling.

To observe explicitly how the system-bath coupling strength affects both energies 𝒰s(t){\mathcal{U}}_{s}^{*}(t) and s(t){\mathcal{F}}_{s}^{*}(t), we apply to (23a) the identity obtained from the interplay between generalized functions and the theory of moments KAN04 such that

𝒫t(ω)=n=0(1)nμnδ(n){ωy(t)}n!.{\mathcal{P}}_{t}^{*}(\omega)\;=\;\sum_{n=0}^{\infty}\frac{(-1)^{n}\cdot\mu_{n}^{*}\cdot\delta^{(n)}\{\omega-y(t)\}}{n!}\,. (26)

Here the symbol δ(n){}\delta^{(n)}\{\cdots\} denotes the nnth derivative, and the nnth moment μn=𝑑ω{ωy(t)}n𝒫t(ω)Θ(ω)\mu_{n}^{*}=\int_{-\infty}^{\infty}d\omega\,\{\omega-y(t)\}^{n}\,{\mathcal{P}}_{t}^{*}(\omega)\,\Theta(\omega) with the Heaviside step function Θ(ω)\Theta(\omega). Then, with the aid of (21), the formal decomposition (26) can explicitly be evaluated as

𝒫t(ω)\displaystyle{\mathcal{P}}_{t}^{*}(\omega) =\displaystyle= δ{ωy(t)}μ1(t)δ(1){ωy(t)}+\displaystyle\delta\{\omega-y(t)\}\,-\,\mu_{1}^{*}(t)\cdot\delta^{(1)}\{\omega-y(t)\}\,+
μ2(t)2δ(2){ωy(t)}+,\displaystyle\frac{\mu_{2}^{*}(t)}{2}\cdot\delta^{(2)}\{\omega-y(t)\}\,+\,\cdots\,,

where the moments are given by μ0(t)=1\mu_{0}^{*}(t)=1 and

μ1(t)=\displaystyle\mu_{1}^{*}(t)= y(t)+\displaystyle\;-y(t)\;+ (3a)
1π[ωdln{ωdy(t)}l=13ωl¯(t)ln{ωl¯(t)y(t)}];\displaystyle\;\frac{1}{\pi}\left[\omega_{d}\,\ln\left\{\frac{\omega_{d}}{y(t)}\right\}\,-\,\sum_{l=1}^{3}\underline{\omega_{l}}(t)\cdot\ln\left\{\frac{\underline{\omega_{l}}(t)}{y(t)}\right\}\right]\;;
μ2(t)=\displaystyle\mu_{2}^{*}(t)= ωdγo2y(t)μ1(t);μ3(t)=.\displaystyle\;\omega_{d}\,\gamma_{\mbox{\tiny o}}-2\,y(t)\cdot\mu_{1}^{*}(t)\;;\;\mu_{3}^{*}(t)\,=\,\cdots\,. (3b)

It is easy to verify that in the weak-coupling limit γo0\gamma_{\mbox{\tiny o}}\to 0, all moments μn0\mu_{n}^{*}\to 0 where n1n\geq 1. The substitution of (3) into (23) then allows us to have

𝒰s(t)=e{β,y(t)}+n=1μn(t){ωne(β,ω)}ωy(t),{\mathcal{U}}_{s}^{*}(t)\,=\,e\{\beta,y(t)\}\,+\,\sum_{n=1}^{\infty}\mu_{n}^{*}(t)\cdot\{\partial_{\omega}^{n}\,e(\beta,\omega)\}_{\omega\to y(t)}\,, (28)

where the summation on the right-hand side reflects all system-bath coupling. An expression with the same structure immediately follows for the free energy s(t){\mathcal{F}}_{s}^{*}(t), too.

It is a noteworthy fact that as a simple case, we have f(,ω)=ω/2f(\infty,\omega)=\hbar\omega/2 at zero temperature, and so Eq. (25) can be rewritten as a reduced expression 0𝑑ωf(,ω)δ{ωy0(t)}\int_{0}^{\infty}d\omega\,f(\infty,\omega)\cdot\delta\{\omega-y_{0}^{*}(t)\} with y0(t):=k=0ω¯k,tj=1ωjy_{0}^{*}(t):=\sum_{k=0}\bar{\omega}_{k,t}-\sum_{j=1}\omega_{j} [cf. (13a)]. Therefore, it looks like a free energy of the Hamiltonian ^s(t)\hat{{\mathcal{H}}}_{s}^{*}(t) which is in an isolated pure state; there is no heat flow between system and bath at zero temperature. On the other hand, the system-of-interest H^s(t)\hat{H}_{s}(t) is in a mixed state due to the system-bath coupling even at zero temperature [cf. (2) and (12)]. This also suggests to us that the free energy s(t){\mathcal{F}}_{s}^{*}(t) cannot exactly be associated with the reduced system H^s(t)\hat{H}_{s}(t) alone.

Similarly to (23), we next introduce another distribution which is useful for describing the the internal energy Us(t)U_{s}(t) beyond the weak-coupling. By using (9) with y0y(t)y_{0}\to y(t), we can easily get

Us(t)=0𝑑ωe(β,ω)Pt(ω),U_{s}(t)\;=\;\int_{0}^{\infty}d\omega\;e(\beta,\omega)\cdot P_{t}(\omega)\,, (29)

where the distribution

Pt(ω)=Mπω2+y2(t)ωIm{χ~t(ω+i0+)},\displaystyle P_{t}(\omega)\;=\;\frac{M}{\pi}\,\frac{\omega^{2}+y^{2}(t)}{\omega}\cdot\mbox{Im}\{\tilde{\chi}_{t}(\omega+i0^{+})\}\,, (29a)

reducing to δ{ωy(t)}\delta\{\omega-y(t)\} in the identically vanishing coupling. Within the Drude-Ullersma model, we have KIM07

Im{χ~t(ω+i0+)}=1Ml=13λt(l)ωω2+ωl¯2(t),\mbox{Im}\{\tilde{\chi}_{t}(\omega+i0^{+})\}\;=\;-\frac{1}{M}\,\sum_{l=1}^{3}\,\lambda_{t}^{(l)}\,\frac{\omega}{\omega^{2}+\underline{\omega_{l}}^{2}(t)}\,, (30)

directly obtained from (22). Here the coefficients

λt(1)=zt,1+zt,2(Ωtzt,1)(zt,2Ωt)\displaystyle\lambda_{t}^{(1)}\;=\;\frac{z_{t,1}\,+\,z_{t,2}}{(\Omega_{t}\,-\,z_{t,1})(z_{t,2}\,-\,\Omega_{t})}
λt(2)=Ωt+zt,2(zt,1Ωt)(zt,2zt,1)\displaystyle\lambda_{t}^{(2)}\;=\;\frac{\Omega_{t}\,+\,z_{t,2}}{(z_{t,1}\,-\,\Omega_{t})(z_{t,2}\,-\,z_{t,1})} (30a)
λt(3)=Ωt+zt,1(zt,2Ωt)(zt,1zt,2)\displaystyle\lambda_{t}^{(3)}\;=\;\frac{\Omega_{t}\,+\,z_{t,1}}{(z_{t,2}\,-\,\Omega_{t})(z_{t,1}\,-\,z_{t,2})}

satisfy the relations

l=13λt(l)ωl(t)¯=1y2(t);l=13λt(l)= 0\displaystyle\sum_{l=1}^{3}\,\frac{\lambda_{t}^{(l)}}{\underline{\omega_{l}(t)}}\,=\,-\frac{1}{y^{2}(t)}\;;\;\sum_{l=1}^{3}\,\lambda_{t}^{(l)}\,=\,0 (30b)
l=33λt(l)ωl¯(t)= 1;l=13λt(l)ωl¯3(t)=y2(t)γoωd\displaystyle\sum_{l=3}^{3}\,\lambda_{t}^{(l)}\cdot\underline{\omega_{l}}(t)\,=\,1\;;\;\sum_{l=1}^{3}\,\lambda_{t}^{(l)}\cdot\underline{\omega_{l}}^{3}(t)\,=\,-y^{2}(t)-\gamma_{\mbox{\tiny o}}\,\omega_{d}
l=13λt(l)ωl¯2(t)= 0;l=13λt(l)ωl¯4(t)=γoωd2,\displaystyle\sum_{l=1}^{3}\,\lambda_{t}^{(l)}\cdot\underline{\omega_{l}}^{2}(t)\,=\,0\;;\;\sum_{l=1}^{3}\,\lambda_{t}^{(l)}\cdot\underline{\omega_{l}}^{4}(t)\,=\,-\gamma_{\mbox{\tiny o}}\,\omega_{d}^{2}\,,

which will be useful below. In the weak-coupling limit γo0\gamma_{\mbox{\tiny o}}\to 0, we easily get

(λt(1),λt(2),λt(3))(0,12iy(t),12iy(t)).\displaystyle\left(\lambda_{t}^{(1)},\lambda_{t}^{(2)},\lambda_{t}^{(3)}\right)\to\left(0,\frac{1}{2i\,y(t)},\frac{-1}{2i\,y(t)}\right)\,. (30c)

Eqs. (8a) and (8b) can then be expressed explicitly as KIM07

q^2β(t)\displaystyle\langle\hat{q}^{2}\rangle_{\beta}(t) =\displaystyle= 1Mβy2(t)+πMl=13λt(l)ψ(1+βωl¯(t)2π)\displaystyle\frac{1}{M\,\beta\,y^{2}(t)}+\frac{\hbar}{\pi M}\sum_{l=1}^{3}\lambda_{t}^{(l)}\cdot\psi\left(1+\frac{\beta\hbar\underline{\omega_{l}}(t)}{2\pi}\right)
q^˙2β(t)\displaystyle\langle\dot{\hat{q}}^{2}\rangle_{\beta}(t) =\displaystyle= 1MβπM×\displaystyle\frac{1}{M\,\beta}-\frac{\hbar}{\pi M}\,\times (31)
l=13λt(l)ωl¯2(t)ψ(1+βωl¯(t)2π)\displaystyle\sum_{l=1}^{3}\lambda_{t}^{(l)}\,\underline{\omega_{l}}^{2}(t)\cdot\psi\left(1+\frac{\beta\hbar\underline{\omega_{l}}(t)}{2\pi}\right)

in terms of the digamma function ψ()\psi(\cdots), respectively, thus immediately giving the internal energy Us(t)U_{s}(t) in its closed form. Here we also used the relation ψ(1+z)=ψ(z)+1/z\psi(1+z)=\psi(z)+1/z ABS74 . From this, we can easily verify that in the classical limit 0\hbar\to 0, the internal energy Us(t)ecl(β)U_{s}(t)\to e_{\text{\scriptsize cl}}(\beta) regardless of the coupling strength γo\gamma_{\mbox{\tiny o}}.

Now we substitute (30) with (30b) into (29a), which will yield

Pt(ω)\displaystyle P_{t}(\omega) =\displaystyle= {ω2+y2(t)}πl=13λt(l)ω2+ωl¯2(t)\displaystyle-\frac{\{\omega^{2}+y^{2}(t)\}}{\pi}\sum_{l=1}^{3}\,\frac{\lambda_{t}^{(l)}}{\omega^{2}+\underline{\omega_{l}}^{2}(t)} (32)
=\displaystyle= (ωd2γo/π){ω2+y2(t)}/gt(ω).\displaystyle(\omega_{d}^{2}\,\gamma_{\mbox{\tiny o}}/\pi)\,\{\omega^{2}+y^{2}(t)\}/g_{t}(\omega)\,.

Therefore, the distribution Pt(ω)P_{t}(\omega) is non-negative. It is easy to verify the normalization 0𝑑ωPt(ω)=1\int_{0}^{\infty}d\omega P_{t}(\omega)=1 for γo0\gamma_{\mbox{\tiny o}}\neq 0. The behaviors of Pt(ω)P_{t}(\omega) are displayed in Fig. 2. In comparison, Eq. (32) identically vanishes in an isolated case; however, likewise with 𝒫t(ω){\mathcal{P}}_{t}^{*}(\omega), we have Pt(ω)=δ{ωy(t)}P_{t}(\omega)=\delta\{\omega-y(t)\} indeed in this case. We can also obtain, as the counterpart to (3),

Pt(ω)\displaystyle P_{t}(\omega) =\displaystyle= δ{ωy(t)}μ1(t)δ(1){ωy(t)}+\displaystyle\delta\{\omega-y(t)\}\,-\,\mu_{1}(t)\cdot\delta^{(1)}\{\omega-y(t)\}\,+ (33)
μ2(t)2δ(2){ωy(t)}+,\displaystyle\frac{\mu_{2}(t)}{2}\cdot\delta^{(2)}\{\omega-y(t)\}\,+\,\cdots\,,

where the moments are given by μ0(t)=1\mu_{0}(t)=1 and

μ1(t)=\displaystyle\mu_{1}(t)= y(t)+\displaystyle\;-y(t)\;+ (33a)
1πl=13λt(l){y2(t)ωl¯2(t)}ln{ωl¯(t)y(t)}\displaystyle\;\frac{1}{\pi}\,\sum_{l=1}^{3}\,\lambda_{t}^{(l)}\,\left\{y^{2}(t)-\underline{\omega_{l}}^{2}(t)\right\}\cdot\ln\left\{\frac{\underline{\omega_{l}}(t)}{y(t)}\right\}
μ2(t)=\displaystyle\mu_{2}(t)= ωdγo/22y(t)μ1(t);μ3(t)=.\displaystyle\;\omega_{d}\,\gamma_{\mbox{\tiny o}}/2-2\,y(t)\cdot\mu_{1}(t)\;;\;\mu_{3}(t)\,=\,\cdots\,. (33b)

In the weak-coupling limit γo0\gamma_{\mbox{\tiny o}}\to 0, all moments μn0\mu_{n}\to 0 where n1n\geq 1. The substitution of (33) into (29) can immediately give rise to the sum rule for Us(t)U_{s}(t), which is the counterpart to (28). In addition, we stress that the probability density Pt(ω)P_{t}(\omega) is not a quantum-mechanical quantity.

Next we intend to express the distribution 𝒫t(ω){\mathcal{P}}_{t}^{*}(\omega) in terms of Pt(ω)P_{t}(\omega), which will enable us to relate the thermodynamic energy 𝒰s(t){\mathcal{U}}_{s}^{*}(t) directly to thermodynamic quantities of the reduced system H^s(t)\hat{H}_{s}(t). We first compare (24) and (32), which easily leads to 𝒫t(ω)=Pt(ω)+𝒫~t(ω){\mathcal{P}}_{t}^{*}(\omega)=P_{t}(\omega)+\tilde{\mathcal{P}}_{t}^{*}(\omega) with

𝒫~t(ω)\displaystyle\tilde{\mathcal{P}}_{t}^{*}(\omega) :=\displaystyle:= 2/π(ω2+ωd2)l=13λt(l)ωl¯2(t)ω2+ωl¯2(t)×\displaystyle\frac{-2/\pi}{(\omega^{2}+\omega_{d}^{2})}\sum_{l=1}^{3}\frac{\lambda_{t}^{(l)}\cdot\underline{\omega_{l}}^{2}(t)}{\omega^{2}+\underline{\omega_{l}}^{2}(t)}\,\times (34)
{y2(t)+ωl¯2(t)+ωdγo/2}.\displaystyle\{y^{2}(t)+\underline{\omega_{l}}^{2}(t)+\omega_{d}\,\gamma_{\mbox{\tiny o}}/2\}\,.

It is also straightforward to verify that 0𝒫~t(ω)𝑑ω=0\int_{0}^{\infty}\tilde{\mathcal{P}}_{t}^{*}(\omega)\,d\omega=0. In the Ohmic limit, Eq. (34) vanishes. Then we can get

𝒫t(ω)=[1+{ω2y2(t)}2ωdγo{ω2+y2(t)}]Pt(ω)ωd/πω2+ωd2.{\mathcal{P}}_{t}^{*}(\omega)\;=\;\left[1+\frac{\{\omega^{2}-y^{2}(t)\}^{2}}{\omega_{d}\,\gamma_{\mbox{\tiny o}}\,\{\omega^{2}+y^{2}(t)\}}\right]\,P_{t}(\omega)\,-\,\frac{\omega_{d}/\pi}{\omega^{2}+\omega_{d}^{2}}\,. (35)

By substituting this into (23) and then applying Eqs. (2) and (29), we can finally arrive at the expression

𝒰s(t)=Us(t)+𝒰~s(t),{\mathcal{U}}_{s}^{*}(t)\;=\;U_{s}(t)\,+\,\tilde{\mathcal{U}}_{s}^{*}(t)\,, (36)

where the coupling-induced term (cf. Appendix A)

𝒰~s(t)=0𝑑ωe(β,ω)𝒫~t(ω)\displaystyle\tilde{\mathcal{U}}_{s}^{*}(t)\,=\,\int_{0}^{\infty}d\omega\;e(\beta,\omega)\cdot\tilde{\mathcal{P}}_{t}^{*}(\omega) (36a)
=\displaystyle=\; ωd2πψ(1+βωd2π)+2πωdγo×\displaystyle\frac{\hbar\omega_{d}}{2\pi}\cdot\psi\left(1+\frac{\beta\hbar\omega_{d}}{2\pi}\right)\,+\,\frac{\hbar}{2\pi\,\omega_{d}\,\gamma_{\mbox{\tiny o}}}\,\times (36b)
l=13λt(l){y2(t)+ωl¯2(t)}2ψ(1+βωl¯(t)2π).\displaystyle\sum_{l=1}^{3}\lambda_{t}^{(l)}\,\{y^{2}(t)+\underline{\omega_{l}}^{2}(t)\}^{2}\cdot\psi\left(1+\frac{\beta\hbar\underline{\omega_{l}}(t)}{2\pi}\right)\,.

It can be shown that this vanishes indeed with γo0\gamma_{\mbox{\tiny o}}\to 0. This also vanishes in the classical limit 0\hbar\to 0. Eq. (36b) is displayed in Fig. 3.

Similarly, the free energy can also be expressed as

s(t)=F¯s(t)+~s(t),{\mathcal{F}}_{s}^{*}(t)\;=\;\bar{F}_{s}(t)\,+\,\tilde{\mathcal{F}}_{s}^{*}(t)\,, (37)

where a generalized “free energy” F¯s(t)=0𝑑ωf(β,ω)Pt(ω)\bar{F}_{s}(t)=\int_{0}^{\infty}d\omega\,f(\beta,\omega)\cdot P_{t}(\omega), and

~s(t)=0𝑑ωf(β,ω)𝒫~t(ω)\displaystyle\tilde{\mathcal{F}}_{s}^{*}(t)\,=\,\int_{0}^{\infty}d\omega\,f(\beta,\omega)\cdot\tilde{\mathcal{P}}_{t}^{*}(\omega) (37a)
=\displaystyle=\; 1βln{Γ(1+βωd2π)}+1βωdγo×\displaystyle\frac{1}{\beta}\cdot\ln\left\{\Gamma\left(1+\frac{\beta\hbar\omega_{d}}{2\pi}\right)\right\}\,+\,\frac{1}{\beta\,\omega_{d}\,\gamma_{\mbox{\tiny o}}}\,\times (37b)
l=13λt(l)ωl¯(t){y2(t)+ωl¯2(t)}2ln{Γ(1+βωl¯(t)2π)},\displaystyle\sum_{l=1}^{3}\frac{\lambda_{t}^{(l)}}{\underline{\omega_{l}}(t)}\,\{y^{2}(t)+\underline{\omega_{l}}^{2}(t)\}^{2}\cdot\ln\left\{\Gamma\left(1+\frac{\beta\hbar\underline{\omega_{l}}(t)}{2\pi}\right)\right\}\,,

expressed in terms of the gamma function Γ()\Gamma(\cdots). It can be shown that Eq. (37b) vanishes with γo0\gamma_{\mbox{\tiny o}}\to 0, as well as with 0\hbar\to 0. Fig. 4 displays different behaviors of (37b).

Now we consider the free energy change Δs(tf)=0𝑑ωf(β,ω){𝒫tf(ω)𝒫0(ω)}\Delta{\mathcal{F}}_{s}^{*}(t_{\text{\scriptsize f}})=\int_{0}^{\infty}d\omega\,f(\beta,\omega)\,\{{\mathcal{P}}_{t_{\text{\tiny f}}}^{*}(\omega)-{\mathcal{P}}_{0}^{*}(\omega)\}. This turns out to be identical to the quantum-mechanical average value

𝒲rev{β,y(tf)}=y0y(tf)𝑑yH^s{y(t)}yβ,{\mathcal{W}}_{\text{\scriptsize rev}}\{\beta,y(t_{\text{\scriptsize f}})\}\;=\;\int_{y_{0}}^{y(t_{\text{\tiny f}})}dy\,\left\langle\frac{\partial\hat{H}_{s}\{y(t)\}}{\partial y}\right\rangle_{\beta}\,, (38)

evaluated along an isothermal process, i.e., in the infinitesimally slow variation of frequency (cf. Appendix B). However, this “work” 𝒲rev{β,y(tf)}{\mathcal{W}}_{\text{\scriptsize rev}}\{\beta,y(t_{\text{\scriptsize f}})\} may conceptually not be interpreted as a minimum average work performed on the reduced system H^s(t)\hat{H}_{s}(t), due to the fact that the actual average work should be defined as an average value of a classical stochastic variable with transition probabilities derived from quantum mechanics, rather than as an expectation value of some “work” operator TAL07 ; accordingly, the minimum average work comes out when the individual work of each run is performed only for the reversible process.

Next let us discuss the second law of thermodynamics within the system-of-interest H^s(t)\hat{H}_{s}(t) in terms of ΔUs(tf)=0𝑑ωe(β,ω){Ptf(ω)P0(ω)}\Delta U_{s}(t_{\text{\scriptsize f}})=\int_{0}^{\infty}d\omega\,e(\beta,\omega)\,\{P_{t_{\text{\tiny f}}}(\omega)-P_{0}(\omega)\} and Δs(tf)\Delta{\mathcal{F}}_{s}^{*}(t_{\text{\scriptsize f}}). To do so, we reinterpret an entire process {y(t)| 0ttf}\{y(t)\,|\,0\leq t\leq t_{\text{\scriptsize f}}\} as a sum of the following three sub-processes, based on the fact that all thermodynamic state functions are path-independent; in the sub-process (I), we completely decouple the system oscillator H^s(0)\hat{H}_{s}(0) from a bath by letting the non-vanishing coupling strengths cj0c_{j}\to 0 in an isothermal fashion. The internal energy change through this first sub-process is given by ΔUs,1=e(β,y0)Us(0)\Delta U_{s,1}=e(\beta,y_{0})-U_{s}(0) while the free energy change needed for decoupling the system from the bath, Δs,1=f(β,y0)s(0)\Delta{\mathcal{F}}_{s,1}^{*}=f(\beta,y_{0})-{\mathcal{F}}_{s}^{*}(0) FOR06 ; KIM06 ; KIM07 . In the next sub-process (II) we vary the frequency y(t)y(t) of the resultant isolated oscillator according to the pre-determined protocol, followed by its coupling weakly to the bath which makes the oscillator come back to the thermal equilibrium at temperature TT. The relevant internal energy change and free energy change are ΔUs,2(tf)=e{β,y(tf)}e(β,y0)\Delta U_{s,2}(t_{\text{\scriptsize f}})=e\{\beta,y(t_{\text{\scriptsize f}})\}-e(\beta,y_{0}) and Δs,2(tf)=f{β,y(tf)}f(β,y0)\Delta{\mathcal{F}}_{s,2}^{*}(t_{\text{\scriptsize f}})=f\{\beta,y(t_{\text{\scriptsize f}})\}-f(\beta,y_{0}), respectively. In this sub-process, the JE (1) holds true. In the last sub-process (III) we increase the coupling strengths cjc_{j} up to their original values in an isothermal fashion; the internal energy change ΔUs,3=Us(tf)e{β,y(tf)}\Delta U_{s,3}=U_{s}(t_{\text{\scriptsize f}})-e\{\beta,y(t_{\text{\scriptsize f}})\}, and the free energy change needed for coupling the system to the bath, Δs,3=s(tf)f{β,y(tf)}\Delta{\mathcal{F}}_{s,3}^{*}={\mathcal{F}}_{s}^{*}(t_{\text{\scriptsize f}})-f\{\beta,y(t_{\text{\scriptsize f}})\}. As a result, the total internal energy change and free energy change through sub-processes (I)-(III) equal ΔUs(tf)\Delta U_{s}(t_{\text{\scriptsize f}}) and Δs(tf)\Delta{\mathcal{F}}_{s}^{*}(t_{\text{\scriptsize f}}), respectively.

We remind that the inequality ΔUs,3<Δs,3\Delta U_{s,3}<\Delta{\mathcal{F}}_{s,3}^{*} is valid (so is ΔUs,1>Δs,1\Delta U_{s,1}>\Delta{\mathcal{F}}_{s,1}^{*}), particularly in the low-temperature regime KIM07 ; FOR06 ; KIM06 . Therefore, this inequality says that if we consider a next decoupling process after (III) and then identified the free change Δs,3\Delta{\mathcal{F}}_{s,3}^{*} with the (maximum) useful work spontaneously releasable from the coupled system H^s(tf)\hat{H}_{s}(t_{\text{\scriptsize f}}), then the second law within the system H^s(tf)\hat{H}_{s}(t_{\text{\scriptsize f}}) could be violated. This tells us that the free energy change Δs(tf)\Delta{\mathcal{F}}_{s}^{*}(t_{\text{\scriptsize f}}) cannot be the (minimum) actual work performed on the system H^s(t)\hat{H}_{s}(t) beyond the weak-coupling limit, and so the JE (17a) is not allowed to associate itself directly with the second law within the coupled oscillator.

Comments deserve here. From Eqs. (29) and (36)-(37) it is tempted to rewrite the internal energy as Us(t)=βlnZ¯β{y(t)}U_{s}(t)=-\partial_{\beta}\ln\bar{Z}_{\beta}\{y(t)\} in terms of a generalized partition function Z¯β{y(t)}:=exp{0𝑑ωlnzβ(ω)Pt(ω)}\bar{Z}_{\beta}\{y(t)\}:=\exp\{\int_{0}^{\infty}d\omega\,\ln z_{\beta}(\omega)\,P_{t}(\omega)\} beyond the weak-coupling limit, and the resultant “free energy” as F¯s(t)=β1lnZ¯β{y(t)}\bar{F}_{s}(t)=-\beta^{-1}\,\ln\bar{Z}_{\beta}\{y(t)\}, with F¯s(tf)s(tf)\bar{F}_{s}(t_{\text{\scriptsize f}})\neq{\mathcal{F}}_{s}^{*}(t_{\text{\scriptsize f}}). In the classical case, on the other hand, Eq. (37a) vanishes, and so F¯s,cl(tf)=s,cl(tf)\bar{F}_{s,{\text{\scriptsize cl}}}(t_{\text{\scriptsize f}})={\mathcal{F}}_{s,{\text{\scriptsize cl}}}^{*}(t_{\text{\scriptsize f}}) indeed, explicitly given by

F¯s,cl(t)=12βl=13λt(l){ωl¯2(t)y2(t)}ln{ωl¯(t)}ωl¯(t).\bar{F}_{s,{\text{\scriptsize cl}}}(t)\;=\;\frac{1}{2\beta}\sum_{l=1}^{3}\frac{\lambda_{t}^{(l)}\,\{\underline{\omega_{l}}^{2}(t)-y^{2}(t)\}\cdot\ln\{\underline{\omega_{l}}(t)\}}{\underline{\omega_{l}}(t)}\,. (39)

It is also notable that Eq. (39) is independent of \hbar, whereas this is not true for fcl(β,ω)={ln(βω)}/βf_{\text{\scriptsize cl}}(\beta,\omega)=\{\ln(\beta\hbar\omega)\}/\beta. This means that the classical free energy can be conceptually rescued only when the system-bath coupling is reflected (cf. of course, Δfcl(β,ω)=(ln{y(tf)/y0})/β\Delta f_{\text{\scriptsize cl}}(\beta,\omega)=(\,\ln\{y(t_{\text{\scriptsize f}})/y_{0}\})/\beta with {ω|y0y(tf)}\{\omega\,|\,y_{0}\to y(t_{\text{\scriptsize f}})\}, independent of \hbar). This free energy F¯s(t)\bar{F}_{s}(t) will be employed below for our discussion of a generalized Jarzynski equality.

4 Quantum Jarzynski equality beyond the weak-coupling limit

We will introduce a generalized Jarzynski equality consistent with the second law within the oscillator coupled to a bath at an arbitrary strength. This will need an appropriate definition of the work performed on the system. Therefore we first consider a reversible process in which it is straightforward to evaluate the work. Then the generalized free energy change in the variation of frequency can be expressed as

ΔF¯s(tf)=0𝑑ωΔf(ω,y0){Ptf(ω)P0(ω)},\Delta\bar{F}_{s}(t_{\text{\tiny f}})\;=\;\int_{0}^{\infty}d\omega\,\Delta f(\omega,y_{0})\cdot\{P_{t_{\text{\tiny f}}}(\omega)-P_{0}(\omega)\}\,, (40)

where Δf(ω,y0)=f(β,ω)f(β,y0)\Delta f(\omega,y_{0})=f(\beta,\omega)-f(\beta,y_{0}). From this, a generalized Jarzynski equality (GJE) beyond the weak-coupling limit is introduced as

ΔF¯s(tf)=1β0dω{lneβW(β,ω)P}{Ptf(ω)P0(ω)}\Delta\bar{F}_{s}(t_{\text{\tiny f}})=-\frac{1}{\beta}\int_{0}^{\infty}d\omega\,\{\ln\langle e^{-\beta\,W(\beta,\omega)}\rangle_{\scriptscriptstyle{P}}\}\,\{P_{t_{\text{\tiny f}}}(\omega)-P_{0}(\omega)\} (41)

where the average P\langle\cdots\rangle_{\scriptscriptstyle{P}} is carried out with the work distribution P(W)P(W) in (18) for an isolated oscillator in the frequency variation {y0ω(tf)}\{y_{0}\to\omega(t_{\text{\tiny f}})\}, as well as the probability density Pt(ω)P_{t}(\omega) reflects the actual coupling between H^s(t)\hat{H}_{s}(t) and H^b\hat{H}_{b}.

In an isolated case the GJE easily reduces to the known form (1). As shown in (41), we now need to deal with a sum of the Jarzynski equalities, each being valid for an isolated system, with the initial and final “weights” P0P_{0} and PtfP_{t_{\text{\tiny f}}} obtained directly from the susceptibility χt(ω)\chi_{t}(\omega) (cf. Appendix B). Technically this means that we first turn off the coupling cj’s0c_{j}\mbox{'s}\to 0, which makes the initial reduced density matrix (2) reduce to the canonical thermal state eβH^s(0)/zβ(y0)e^{-\beta\hat{H}_{s}(0)}/z_{\beta}(y_{0}); next we carry out the JE processes independently for many different frequencies ω\omega’s with the two weights. From this, we can extract the free energy change ΔF¯s(tf)\Delta\bar{F}_{s}(t_{\text{\tiny f}}), without a measurement of any other “work” directly on the coupled system. It is instructive to remind that the JE (17a) can also be rewritten as

Δs(tf)=1β0𝑑ω{lneβW(β,ω)}{𝒫tf(ω)𝒫0(ω)}\Delta{\mathcal{F}}_{s}^{*}(t_{\text{\tiny f}})=-\frac{1}{\beta}\int_{0}^{\infty}d\omega\,\{\ln\langle e^{-\beta\,W(\beta,\omega)}\rangle\}\,\{{\mathcal{P}}^{*}_{t_{\text{\tiny f}}}(\omega)-{\mathcal{P}}^{*}_{0}(\omega)\} (42)

in its form, where the distributions 𝒫t(ω){\mathcal{P}}_{t}^{*}(\omega)’s come from the susceptibility as well [cf. (23a) and (25)], but not guaranteed to be non-negative.

Next, in order to observe explicitly the deviation of the GJE from the JE in its known form, we substitute the sum rule (33) into (41), which yields

ΔF¯s(tf)\displaystyle\Delta\bar{F}_{s}(t_{\text{\tiny f}}) =\displaystyle= 1β[lneβW{β,y(tf)}P+\displaystyle-\frac{1}{\beta}\left[\ln\left\langle e^{-\beta\,W\{\beta,y(t_{\text{\tiny f}})\}}\right\rangle_{\scriptscriptstyle{P}}\,+\right.
n=1μn(t)n!(y)nlneβW{β,y(t)}P|t=0t=tf].\displaystyle\left.\sum_{n=1}^{\infty}\frac{\mu_{n}(t)}{n!}\,\left(\frac{\partial}{\partial y}\right)^{n}\left.\ln\left\langle e^{-\beta\,W\{\beta,y(t)\}}\right\rangle_{\scriptscriptstyle{P}}\right|_{t=0}^{t=t_{\text{\tiny f}}}\right]\,.

Therefore we can now look into the sufficiently weak, but not necessarily vanishingly small, coupling regime, simply by adding the low-order coupling-induced terms (33a) and (33b). On the other hand, beyond the weak coupling limit we cannot simply neglect higher-order moments (cf. Fig. 5 as well as Figs. 1, 2). As a result, we may say that the JE (1) is exactly valid only in the vanishingly small coupling limit.

Now we briefly discuss (4) in the classical limit β0\beta\hbar\to 0. Then the term with n=1n=1 easily reduces to

μ1(t)yf{β,y(t)}|t=0t=tf1β{μ1(tf)y(tf)μ1(0)y0},\mu_{1}(t)\cdot\partial_{y}f\{\beta,y(t)\}\Big{|}_{t=0}^{t=t_{\text{\tiny f}}}\,\to\,\frac{1}{\beta}\,\left\{\frac{\mu_{1}(t_{\text{\scriptsize f}})}{y(t_{\text{\scriptsize f}})}\,-\,\frac{\mu_{1}(0)}{y_{0}}\right\}\,, (44)

non-vanishing indeed. Likewise, so are all terms with n2n\geq 2. Therefore, even the classical JE (1) does not exactly hold true any longer beyond the weak-coupling limit.

It is also instructive to add remarks on the effect of system-bath coupling to the Jarzynski equality (41); in an isolated case, albeit the JE in its known form is well-known, we cannot perform an isothermal process, in which the (minimum) average work exactly amounts to the free energy change. And in an isothermal process we might think heat exchange through the system-bath coupling at every single moment in such a way that we switch off the coupling (“decoupling”) and then perform the external perturbation {y0y(tf)}\{y_{0}\to y(t_{\text{\scriptsize f}})\} in the resultant isolated case, followed by contacting with the bath again (“coupling”); so the total heat exchange over the actual isothermal process would be equivalent to the amount of final thermal relaxation leading to the end equilibrium state in the above picture a pair of decoupling and coupling is added to. However, as discussed after (38), this picture could lead to a violation of the second law, and so is not acceptable. Moreover, in order to make the JE useful, the system under consideration needs to be sufficiently small-scaled, in which the work fluctuations are observable; in this scale, however, the system-bath coupling is normally non-negligible. As a result, we may argue that the usefulness of the JE in its known form is fairly limited.

Now we discuss the relevance of the GJE to the second law of thermodynamics within the system-of-interest H^s(t)\hat{H}_{s}(t) beyond the weak-coupling limit. We first introduce the average work in an irreversible process as

W(tf)¯=0𝑑ωW{β,ω(tf)}P{Ptf(ω)P0(ω)},\hskip-5.69046pt\overline{\langle W(t_{\text{\scriptsize f}})\rangle}=\int_{0}^{\infty}d\omega\,\left\langle W\{\beta,\omega(t_{\text{\scriptsize f}})\}\right\rangle_{\scriptscriptstyle{P}}\cdot\{P_{t_{\text{\tiny f}}}(\omega)-P_{0}(\omega)\}\,, (45)

where each partial average work W(β,ω)P\langle W(\beta,\omega)\rangle_{\scriptscriptstyle{P}} is explicitly given by (/2){Q(y0,ω)ωy0}coth(βy0/2)(\hbar/2)\{Q^{*}(y_{0},\omega)\cdot\omega-y_{0}\}\coth(\beta\hbar y_{0}/2) [cf. (19)]. As such, the average work W(tf)¯\overline{\langle W(t_{\text{\scriptsize f}})\rangle}, being not a quantum-mechanical expectation value, can be determined without any measurement of the work directly on the system coupled to a bath. Applying the Jensen inequality to (41), we can then obtain an expression of the second law of thermodynamics beyond the weak-coupling

0𝑑ω{W(β,ω)PΔf(ω,y0)}{Ptf(ω)P0(ω)}0,\hskip-5.69046pt\int_{0}^{\infty}d\omega\,\left\{\langle W(\beta,\omega)\rangle_{\scriptscriptstyle{P}}-\Delta f(\omega,y_{0})\right\}\,\{P_{t_{\text{\tiny f}}}(\omega)-P_{0}(\omega)\}\geq 0\,, (46)

equivalent to W(tf)¯F¯s(tf)\overline{\langle W(t_{\text{\scriptsize f}})\rangle}\geq\bar{F}_{s}(t_{\text{\scriptsize f}}). It is now needed to ask if this equality is valid indeed; Figs. 6-7 demonstrate its validity for many different sets of parameters {y(t),ωd,γo}\{y(t),\omega_{d},\gamma_{\mbox{\tiny o}}\}. Subsequently, the first law allows us to have the heat Q(tf)¯=ΔUs(tf)W(tf)¯\overline{\langle Q(t_{\text{\scriptsize f}})\rangle}\,=\,\Delta U_{s}(t_{\text{\scriptsize f}})\,-\,\overline{\langle W(t_{\text{\scriptsize f}})\rangle}.

Lastly, a couple of short comments deserve here. First, we remind that our approach was made entirely in the local picture H^s(t)\hat{H}_{s}(t), rather than the total picture H^(t)\hat{H}(t). Accordingly, there is no room appropriate for detecting system-bath entanglement directly within our generalized Jarzynski equality, while it was discussed, on the other hand, within the Jarzynski equality (17a) in VED10 . Secondly, in practical terms the number of independent experimental runs needed for obtaining a sufficiently visible convergence of the JE (1) grows exponentially with the system size GRO05 , and so the computational cost is high enough. This cost will be even higher when we deal with a system beyond the weak-coupling limit, due to the additional averaging needed for the GJE. Further it was shown GON13 that a significantly faster convergence of the JE can be achieved via accelerated adiabatic control. Even in this scheme the computational cost is expected to increase beyond the weak-coupling limit, from our result.

5 Concluding remarks

In summary, we derived a generalized Jarzynski equality in the scheme of a time-dependent quantum Brownian oscillator within the Drude-Ullersma damping model. This equality is associated with the second law of thermodynamics (in its generalized form) within the system oscillator coupled to a bath at an arbitrary strength. This finding also enables us to look systematically into the coupling effect on the non-equilibrium thermodynamics of the local system-of-interest beyond the weak-coupling limit. As a result, the Jarzynski equality in its original form (and all other relevant fluctuation theorems) was shown to be valid only in the vanishingly small coupling limit, which fact also holds true in the classical limit of β0\beta\hbar\to 0.

We believe that our finding will provide a useful starting point for derivation of a generalized Jarzynski equality associated with the second law in more generic quantum dissipative systems. In fact, if a smooth probability density, such as Pt(ω)P_{t}(\omega) in (29), reflecting the system-bath coupling is explicitly available, this derivation becomes conceptually rather a straightforward issue, while the technical procedure for an exact evaluation of such a probability density would be a formidable task for a broad class of nonlinear systems.

Acknowledgments

The author thanks G. Mahler (Stuttgart), G.J. Iafrate (NC State), and J. Kim (KIAS) for helpful remarks. He also acknowledges financial support provided by the U.S. Army Research Office (Grant No. W911NF-13-1-0323).

Appendix A : Detailed derivation of Eq. (36b)

We first substitute the identity ING98

e(β,ω)=ω2coth(βω2)=1β(1+2n=1ω2ω2+νn2),e(\beta,\omega)\,=\,\frac{\hbar\omega}{2}\,\coth\left(\frac{\beta\hbar\omega}{2}\right)\,=\,\frac{1}{\beta}\,\left(1+2\sum_{n=1}^{\infty}\frac{\omega^{2}}{\omega^{2}+\nu_{n}^{2}}\right)\,, (47)

with νn=2πn/(β)\nu_{n}=2\pi n/(\beta\hbar), into (36a), finally leading to 𝒰~s(t)=\tilde{\mathcal{U}}_{s}^{*}(t)=

ωdγoβn=1νn2(νn+ωd)(νn+Ωt)(νn+z1,t)(νn+z2,t).\frac{-\omega_{d}\,\gamma_{\mbox{\tiny o}}}{\beta}\sum_{n=1}^{\infty}\frac{\nu_{n}^{2}}{(\nu_{n}+\omega_{d})(\nu_{n}+\Omega_{t})(\nu_{n}+z_{1,t})(\nu_{n}+z_{2,t})}\,. (48)

This can be rewritten in terms of the digamma function as

𝒰~s(t)=ωdγo2π×\displaystyle\tilde{\mathcal{U}}_{s}^{*}(t)\,=\,\frac{\hbar\omega_{d}\gamma_{\mbox{\tiny o}}}{2\pi}\,\times (49)
j=03ψ{1+βωj¯(t)/2π}ωj¯2(t){ωj¯(t)ωj+1¯(t)}{ωj¯(t)ωj+2¯(t)}{ωj¯(t)ωj+3¯(t)}\displaystyle\sum_{j=0}^{3}\frac{\psi\{1+\beta\hbar\underline{\omega_{j}}(t)/2\pi\}\cdot\underline{\omega_{j}}^{2}(t)}{\{\underline{\omega_{j}}(t)-\underline{\omega_{j+1}}(t)\}\{\underline{\omega_{j}}(t)-\underline{\omega_{j+2}}(t)\}\{\underline{\omega_{j}}(t)-\underline{\omega_{j+3}}(t)\}}

where ω0¯:=ωd\underline{\omega_{0}}:=\omega_{d}, and the subscript j¯=j(mod 4)\underline{j}=j\,\mbox{(mod 4)}. Subsequently, with the aid of (30b) and (31) we can finally arrive at the expression in (36b).

Appendix B : Evaluation of Eq. (38)

We begin with [cf. (4a)]

H^s{y(t)}yβ=My{q^2β(y)}\left\langle\frac{\partial\hat{H}_{s}\{y(t)\}}{\partial y}\right\rangle_{\beta}\,=\,M\,y\,\left\{\langle\hat{q}^{2}\rangle_{\beta}(y)\right\} (50)

at every single frequency value y(t)y(t). With the aid of (8a) and e(β,ω)=ωωf(β,ω)e(\beta,\omega)=\omega\,\partial_{\omega}f(\beta,\omega), it turns out that

𝒲s{β,y(tf)}\displaystyle{\mathcal{W}}_{s}\{\beta,y(t_{\text{\scriptsize f}})\} =\displaystyle= 2Mπ0dω(ωf)×\displaystyle\frac{2M}{\pi}\int_{0}^{\infty}d\omega\,(\partial_{\omega}f)\,\times (51)
y0y(tf)𝑑yyIm{χ~t(ω+i 0+)}.\displaystyle\int_{y_{0}}^{y(t_{\text{\tiny f}})}dy\;y\;\mbox{Im}\{\tilde{\chi}_{t}(\omega+i\,0^{+})\}\,.

Next, with the aid of (21) we can express the susceptibility (30) in terms of {y(t),ωd,γo}\{y(t),\omega_{d},\gamma_{\mbox{\tiny o}}\} as Im{χ~t(ω+i0+)}=ωd2γoω/{MΥt(ω)}\mbox{Im}\{\tilde{\chi}_{t}(\omega+i0^{+})\}=\omega_{d}^{2}\,\gamma_{\mbox{\tiny o}}\,\omega/\{M\,\Upsilon_{t}(\omega)\}, where

Υt(ω)=ω6+{ωd(ωd2γo)2y2(t)}ω4+\displaystyle\Upsilon_{t}(\omega)\,=\,\omega^{6}+\{\omega_{d}\,(\omega_{d}-2\gamma_{\mbox{\tiny o}})-2\,y^{2}(t)\}\,\omega^{4}+ (52)
{(ωdγo)22ωd(ωdγo)y2(t)+y4(t)}ω2+ωdy2(t).\displaystyle\{(\omega_{d}\,\gamma_{\mbox{\tiny o}})^{2}-2\,\omega_{d}\,(\omega_{d}-\gamma_{\mbox{\tiny o}})\,y^{2}(t)+y^{4}(t)\}\,\omega^{2}+\omega_{d}\cdot y^{2}(t)\,.

Substituting this into (51), we can finally obtain

𝒲s{β,y(tf)}\displaystyle{\mathcal{W}}_{s}\{\beta,y(t_{\text{\scriptsize f}})\} =\displaystyle= ωd2γoπ0dω(ωωf)ω2+ωd2×\displaystyle\frac{\omega_{d}^{2}\,\gamma_{\mbox{\tiny o}}}{\pi}\int_{0}^{\infty}\frac{d\omega\,(\omega\,\partial_{\omega}f)}{\omega^{2}+\omega_{d}^{2}}\,\times (53)
y02y2(tf)dzz2+bz+c,\displaystyle\int_{y_{0}^{2}}^{y^{2}(t_{\text{\tiny f}})}\frac{dz}{z^{2}+bz+c}\,,

where z:=y2(t)z:=y^{2}(t), and

b:=\displaystyle b\;:= 2ω2(ωdγoωd2ω2)ω2+ωd2\displaystyle\;\frac{2\,\omega^{2}\,(\omega_{d}\,\gamma_{\mbox{\tiny o}}-\omega_{d}^{2}-\omega^{2})}{\omega^{2}+\omega_{d}^{2}} (53a)
c:=\displaystyle c\;:= ω2{ω4+(ωd22ωdγo)ω2+(ωdγo)2}ω2+ωd2.\displaystyle\;\frac{\omega^{2}\,\{\omega^{4}+(\omega_{d}^{2}-2\,\omega_{d}\,\gamma_{\mbox{\tiny o}})\,\omega^{2}+(\omega_{d}\,\gamma_{\mbox{\tiny o}})^{2}\}}{\omega^{2}+\omega_{d}^{2}}\,.

Using the relation GRA07

dzz2+bz+c=2(b24c)1/2arctanh{2z+b(b24c)1/2}\int\frac{dz}{z^{2}+bz+c}\;=\;\frac{-2}{(b^{2}-4c)^{1/2}}\cdot\mbox{arctanh}\left\{\frac{2z+b}{(b^{2}-4c)^{1/2}}\right\} (54)

where arctanh(z)={ln(1+z)ln(1z)}/2\mbox{arctanh}(z)=\{\ln(1+z)-\ln(1-z)\}/2, we can arrive at the expression

𝒲s{β,y(tf)}=1π0dω(ωf)×\displaystyle{\mathcal{W}}_{s}\{\beta,y(t_{\text{\scriptsize f}})\}\,=\,\frac{1}{\pi}\int_{0}^{\infty}d\omega\,(\partial_{\omega}f)\,\times (55)
arctan{(ω2+ωd2)y2(t)ω2ωd(ωdγo)ω4ωd2γoω}|t=0t=tf\displaystyle\left.\mbox{arctan}\left\{\frac{(\omega^{2}+\omega_{d}^{2})\cdot y^{2}(t)-\omega^{2}\,\omega_{d}\,(\omega_{d}-\gamma_{\mbox{\tiny o}})-\omega^{4}}{\omega_{d}^{2}\,\gamma_{\mbox{\tiny o}}\,\omega}\right\}\right|_{t=0}^{t=t_{\text{\tiny f}}}

(note that γo0\gamma_{\mbox{\tiny o}}\not\equiv 0). By integration by parts, this can be transformed into

𝒲s{β,y(tf)}=0𝑑ωf(β,ω){𝒫¯tf(ω)𝒫¯0(ω)},{\mathcal{W}}_{s}\{\beta,y(t_{\text{\scriptsize f}})\}\;=\;\int_{0}^{\infty}d\omega\,f(\beta,\omega)\cdot\{\bar{\mathcal{P}}_{t_{\text{\tiny f}}}^{*}(\omega)-\bar{\mathcal{P}}_{0}^{*}(\omega)\}\,, (56)

where the distribution 𝒫¯t(ω):=\bar{\mathcal{P}}_{t}(\omega):=

1πddωarctan{ω4+ωd(ωdγo)ω2(ω2+ωd2)y2(t)ωd2γoω}\displaystyle\frac{1}{\pi}\,\frac{d}{d\omega}\,\mbox{arctan}\left\{\frac{\omega^{4}+\omega_{d}\,(\omega_{d}-\gamma_{\mbox{\tiny o}})\,\omega^{2}-(\omega^{2}+\omega_{d}^{2})\cdot y^{2}(t)}{\omega_{d}^{2}\,\gamma_{\mbox{\tiny o}}\,\omega}\right\} (56a)
=(ωd2γo/π)gt(ω)/{gt(ω)(ω2+ωd2)}=𝒫t(ω).\displaystyle=\;(\omega_{d}^{2}\,\gamma_{\mbox{\tiny o}}/\pi)\cdot g_{t}^{*}(\omega)/\{g_{t}(\omega)\cdot(\omega^{2}+\omega_{d}^{2})\}\;=\;{\mathcal{P}}_{t}^{*}(\omega)\,. (56b)

Here we also used both (d/dx)arctan(x)=1/(1+x2)(d/dx)\,\mbox{arctan}(x)=1/(1+x^{2}) and the fact that Eq. (24a) can be rewritten in terms of {y(t),ωd,γo}\{y(t),\omega_{d},\gamma_{\mbox{\tiny o}}\} as

gt(ω)\displaystyle\hskip-5.69046ptg_{t}(\omega) =\displaystyle= ω6+{ωd22ωdγo2y2(t)}ω4+{ωdy2(t)}2\displaystyle\omega^{6}\,+\,\{\omega_{d}^{2}-2\,\omega_{d}\,\gamma_{\mbox{\tiny o}}-2\,y^{2}(t)\}\,\omega^{4}\,+\,\{\omega_{d}\cdot y^{2}(t)\}^{2} (57)
+[{ωdγo+y2(t)}22{ωdy(t)}2]ω2.\displaystyle+\,\left[\{\omega_{d}\,\gamma_{\mbox{\tiny o}}+y^{2}(t)\}^{2}-2\,\{\omega_{d}\cdot y(t)\}^{2}\right]\,\omega^{2}\,.

References

  • (1) C. Jarzynski, Phys. Rev. Lett. 78, 2690 (1997).
  • (2) H. B. Callen, Thermodynamics and an introduction to thermostatics, 2nd ed. (John Wiley, New York, 1985).
  • (3) C. Jarzynski, Eur. Phys. J. B 64, 331 (2008).
  • (4) E. Boksenbojm, B. Wynants, and C. Jarzynski, Physica A 389, 4406 (2010).
  • (5) G. E. Crooks, J. Stat. Phys. 90, 1481 (1998).
  • (6) G. E. Crooks, Phys. Rev. E 60, 2721 (1999).
  • (7) C. Jarzynski, What is the microscopic response of a system driven far from equilibrium?, in Lecture Notes in Physics 597 (Springer, Berlin, 2002), pp 63-82.
  • (8) J. Liphardt, S. Dumont, S. B. Smith, I. Tinoco Jr, and C. Bustamante, Science 296, 1832 (2002).
  • (9) C. Bustamante, J. Liphardt, and F. Ritort, Phys. Today 58 (7), 43 (2005).
  • (10) E. G. D. Cohen and D. Mauzerall, J. Stat. Mech. (2004) P07006.
  • (11) C. Jarzynski, J. Stat. Mech. (2004) P09005.
  • (12) H. Tasaki, arXiv:cond-mat/0009244 (2000).
  • (13) S. Mukamel, Phys. Rev. Lett. 90, 170604 (2003).
  • (14) W. De Roeck and C. Maes, Phys. Rev. E 69, 026115 (2004).
  • (15) M. Esposito and S. Mukamel, Phys. Rev. E 73, 046129 (2006).
  • (16) P. Talkner, E. Lutz, and P. Hänggi, Phys. Rev. E 75, 050102(R) (2007).
  • (17) J. Teifel and G. Mahler, Phys. Rev. E 76, 051126 (2007).
  • (18) A. Engel and R. Nolte, Eur. Phys. Lett. 79, 10003, (2007).
  • (19) G. E. Crooks, J. Stat. Mech. (2008) P10023.
  • (20) D. Andrieux and P. Gaspard, Phys. Rev. Lett. 100, 230404 (2008).
  • (21) S. Deffner and E. Lutz, Phys. Rev. E 77, 021128 (2008).
  • (22) P. Talkner, M. Campisi, and P. Hänggi, J. Stat. Mech. (2009) P02025.
  • (23) J. Teifel and G. Mahler, Eur. Phys. J. B. 75, 275 (2010).
  • (24) J. Teifel and G. Mahler, Phys. Rev. E 83, 041131 (2011).
  • (25) V. A. Ngo and S. Haas, Phys. Rev. E 86, 031127 (2012).
  • (26) D. P. Sheehan, Quantum limits to the second law, AIP Conf. Proc. 643 (AIP, Melville, NY, 2002).
  • (27) J. Gemmer, M. Michel, and G. Mahler, Quantum Thermodynamics (Springer, Berlin, 2004).
  • (28) C. Vladislav and D. P. Sheehan, Challenges to the second law of thermodynamics: theory and experiment (Springer, New York, 2005).
  • (29) M. Campisi, P. Talkner, and P. Hänggi, Phys. Rev. Lett. 102, 210401 (2009).
  • (30) M. Campisi, P. Talkner, and P. Hänggi, J. Phys. A 42, 392002 (2009).
  • (31) G.-L. Ingold, Dissipative quantum systems, in Quantum transport and dissipation (Wiley-VCH, Weinheim, 1998), pp 213-248.
  • (32) U. Weiss, Quantum dissipative systems, 3rd ed. (World Scientific, Singapore, 2008).
  • (33) H. Grabert, P. Schramm, and G.-L. Ingold, Phys. Rep. 168, 115 (1988).
  • (34) I. Kim and G. Mahler, Phys. Rev. E 81, 011101 (2010).
  • (35) G. W. Ford, J. T. Lewis, and R. F. O’Connell, Ann. Phys. (N.Y.) 185, 270 (1988).
  • (36) A. M. Levine, M. Shapiro, and E. Pollak, J. Chem. Phys. 88, 1959 (1988).
  • (37) G. W. Ford, J. T. Lewis, and R. F. O’Connell, Phys. Rev. Lett. 55, 2273 (1985).
  • (38) G. W. Ford and R. F. O’Connell, Physica E 29, 82 (2005).
  • (39) P. Hänggi and G.-L. Ingold, Chaos 15, 026105 (2005).
  • (40) P. Hänggi and G.-L. Ingold, Acta Phys. Pol. B 37, 1537 (2006).
  • (41) G. W. Ford and R. F. O’Connell, Phys. Rev. Lett. 96, 020402 (2006).
  • (42) I. Kim and G. Mahler, Eur. Phys. J. B 54, 405 (2006).
  • (43) I. Kim and G. Mahler, Eur. Phys. J. B 60, 401 (2007).
  • (44) G.-L. Ingold, Eur. Phys. J. B 85, 30 (2012) and references therein.
  • (45) R. M. Wilcox, J. Math. Phys. 8, 962 (1967).
  • (46) P. Ullersma, Physica 32, 27, 56, 74, 90 (1966).
  • (47) To verify the normalization conditions for 𝒫t(ω){\mathcal{P}}_{t}^{*}(\omega) and Pt(ω)P_{t}(\omega), we apply 𝑑ωa/(ω2+a2)=arctan(ω/a)\int d\omega\cdot a/(\omega^{2}+a^{2})=\mbox{arctan}(\omega/a) to Eqs. (23a) and (32), respectively.
  • (48) R. P. Kanwal, Generalized Functions: Theory and Applications, 3rd ed. (Birkuäuser, Boston, 2004).
  • (49) M. Abramowitz and I. Stegun, Handbook of Mathematical Functions with Formulas, Graphs, and Mathematical Tables (Dover, New York, 1974).
  • (50) J. Hide and V. Vedral, Phys. Rev. A 81, 062303 (2010).
  • (51) R. C. Lua and A. Y. Grosberg, J. Phys. Chem. B 109, 6805 (2005).
  • (52) J. Deng, Q. Wang, Z. Liu, P. Hänggi, and J. Gong, Phys. Rev. E 88, 062122 (2013).
  • (53) I.S. Gradshteyn and I.M. Ryzhik, Table of Integrals, Series, and Products, 7th ed. (Academic Press, San Diego, 2007).
Refer to caption
Figure 1:

Fig. 1: (Color online) The distribution 𝒫=𝒫0(ω){\mathcal{P}}={\mathcal{P}}_{0}^{*}(\omega) given in (24). Here we set y0=7y_{0}=7. (I) Solid lines with ωd=3\omega_{d}=3, from top to bottom in the maximum values, 1st) green: γo=9\gamma_{\mbox{\tiny o}}=9; 2nd) black: γo=30\gamma_{\mbox{\tiny o}}=30 (can be negative). (II) Dash lines with ωd=10\omega_{d}=10, likewise, 1st) red: γo=2\gamma_{\mbox{\tiny o}}=2; 2nd) blue: γo=9\gamma_{\mbox{\tiny o}}=9, in comparison with 𝒫0{\mathcal{P}}\equiv 0 (khaki dashdot); cf. 𝒫δ(ω7){\mathcal{P}}\to\delta(\omega-7) with γo0\gamma_{\mbox{\tiny o}}\to 0.

Refer to caption
Figure 2:

Fig. 2: (Color online) The distribution P=P0(ω)P=P_{0}(\omega) given in (32). The same as in Fig. 1.

Refer to caption
Figure 3:

Fig. 3: (Color online) The dimensionless energy y=𝒰~s(0)/Egy=\tilde{\mathcal{U}}_{s}^{*}(0)/E_{g} given in (36b) with Eg=y0/2E_{g}=\hbar y_{0}/2 versus dimensionless temperature kBT/y¯0k_{\mbox{\tiny B}}T/\hbar\bar{y}_{0}. Here we set M==kB=y¯0=1M=\hbar=k_{\mbox{\tiny B}}=\bar{y}_{0}=1. From top to bottom, 1st) green solid: Ω0=3\Omega_{0}=3 and γ¯0=3\bar{\gamma}_{0}=3 (overdamped); 2nd) blue dash: Ω0=1\Omega_{0}=1 and γ¯0=3\bar{\gamma}_{0}=3 (overdamped); 3rd) red solid: Ω0=3\Omega_{0}=3 and γ¯0=1\bar{\gamma}_{0}=1 (underdamped); 4th) black dashdot: Ω0=1\Omega_{0}=1 and γ¯0=1\bar{\gamma}_{0}=1 (underdamped); here “overdamped” means y¯0<γ¯0/2\bar{y}_{0}<\bar{\gamma}_{0}/2 whereas “underdamped” y¯0γ¯0/2\bar{y}_{0}\geq\bar{\gamma}_{0}/2, after (21). With TT\to\infty, y0y\to 0.

Refer to caption
Figure 4:

Fig. 4: (Color online) The dimensionless free energy y=~s(0)/Egy=\tilde{\mathcal{F}}_{s}^{*}(0)/E_{g} given in (37b) with Eg=y0/2E_{g}=\hbar y_{0}/2 versus dimensionless temperature kBT/y¯0k_{\mbox{\tiny B}}T/\hbar\bar{y}_{0}. The same as in Fig. 3.

Refer to caption
Figure 5:

Fig. 5: (Color online) The relative error y=(ΔF¯s(tf)/Δf{y(tf),y0})1y=(\Delta\bar{F}_{s}(t_{\text{\tiny f}})/\Delta f\{y(t_{\text{\scriptsize f}}),y_{0}\})-1 given in (4) versus dimensionless temperature kBT/y0k_{\mbox{\tiny B}}T/\hbar y_{0}, with Δf{y(tf),y0}=f{β,y(tf)}f(β,y0)\Delta f\{y(t_{\text{\scriptsize f}}),y_{0}\}=f\{\beta,y(t_{\text{\scriptsize f}})\}-f(\beta,y_{0}) being the leading term on the right-hand side of (4). Here we set =kB=y0=1\hbar=k_{\mbox{\tiny B}}=y_{0}=1 and y(tf)=5y(t_{\text{\scriptsize f}})=5. From top to bottom at T=7T=7, 1st) green solid: ωd=2\omega_{d}=2 and γo=5\gamma_{\mbox{\tiny o}}=5 (strong-coupling limit); 2nd) blue dash: ωd=7\omega_{d}=7 and γo=5\gamma_{\mbox{\tiny o}}=5 (Ohmic and strong-coupling limit); 3rd) red solid: ωd=2\omega_{d}=2 and γo=1\gamma_{\mbox{\tiny o}}=1 (weak-coupling limit); 4th) black dashdot: ωd=7\omega_{d}=7 and γo=1\gamma_{\mbox{\tiny o}}=1 (Ohmic and weak-coupling limit); 5th) khaki solid: ωd=7\omega_{d}=7 and γo=0.01\gamma_{\mbox{\tiny o}}=0.01 (Ohmic and vanishingly small coupling limit).

Refer to caption
Figure 6:

Fig. 6: (Color online) The dimensionless quantity y={W(β,ω)PΔf}{Ptf(ω)P0(ω)}/(y0)y=\{\left\langle W(\beta,\omega)\right\rangle_{\scriptscriptstyle{P}}-\Delta f\}\,\{P_{t_{\text{\tiny f}}}(\omega)-P_{0}(\omega)\}/(\hbar y_{0}) given in (46) versus dimensionless frequency ω/y0\omega/y_{0}. Here we set =kB=y0=1\hbar=k_{\mbox{\tiny B}}=y_{0}=1 and y(tf)=5y(t_{\text{\scriptsize f}})=5, and ωd=7\omega_{d}=7, as well as T=1T=1 (low-temperature regime). Let z:=0𝑑ωy(ω)z:=\int_{0}^{\infty}d\omega\cdot y(\omega). (I) Solid lines with the duration tf=5t_{\text{\scriptsize f}}=5 (slow change), 1st) orange, with a peak at ω=y(tf)\omega=y(t_{\text{\scriptsize f}}): γo=0.01\gamma_{\mbox{\tiny o}}=0.01 (vanishingly small coupling) and z=0.3822z=0.3822; 2nd) green, with maximum value shifted a little to the right: γo=1\gamma_{\mbox{\tiny o}}=1 and z=2.3590z=2.3590; 3rd) grey, with maximum value shifted further to the right: γo=5\gamma_{\mbox{\tiny o}}=5 (strong coupling) and z=1.9298z=1.9298. (II) Dash lines with tf=1t_{\text{\scriptsize f}}=1 (fast change), in the same way as in (I), 1st) black: z=0.6526z=0.6526; 2nd) blue: z=4.2329z=4.2329; 3rd) red: z=3.6977z=3.6977. As demonstrated, (1) the smaller tft_{\text{\scriptsize f}}, the larger yy-value, i.e., 1/tf1/t_{\text{\scriptsize f}} is an irreversibility measure of the process; (2) the yy-value can be negative-valued due to its factor Ptf(ω)P0(ω)P_{t_{\text{\tiny f}}}(\omega)-P_{0}(\omega), however, the integral zz is non-negative.

Refer to caption
Figure 7:

Fig. 7: (Color online) The same as in Fig. 6, but evaluated at T=5T=5 (high-temperature regime). (I) Solid lines with tf=5t_{\text{\scriptsize f}}=5 (slow change), 1st) orange: z=2.2343z=2.2343; 2nd) green: z=13.9243z=13.9243; 3rd) grey: z=11.5305z=11.5305. (II) Dash lines with tf=1t_{\text{\scriptsize f}}=1 (fast change), 1st) black: z=3.4881z=3.4881; 2nd) blue: z=22.6124z=22.6124; 3rd) red: z=19.7273z=19.7273.