<<

arXiv:1803.04658v3 [quant-ph] 17 Dec 2018 lopoieasml itr o h rgno i agaduies inflation. universe and bang big of origin the may for dynamics picture inflationary simple The a provide thermodynamics. also quantum sep- and border realm, classical the quantum provides the deep criticality arating the dynamical corresponding in The allocated proved. thermodynamics, naturally of is law third temperature, the nonequilibrium which negative character- from a transition with phase associated quantum dynamics inflationary dynamical by of ized memory occurrence non-Markovian the by (3) induced is and thermo- which dynamics; regime, classical coupling of strong breakdown the the in dynamics (2) regime, hypothesis; coupling system-reservoir equilibrium weak introducing ther- the without classical in of dynamics emergence quantum the from the reservoir, modynamics (1) heat unambiguously a cavity unravels with the interacting theory to system theory nonequilibrium atomic the two-level apply we the When and depends states. system that quantum regime of dynamics nonequilibrium detailed into the temperature on of concept with contact the in to generalize systems, We particle theory single a for even develop systems, applicable small we is arbitrary that paper, to systems quantum this of In thermodynamics nonequilibrium that exact mechanics. the dynamics study quantum the of through principle emerges thermodynamics the equilibrium follows how clear less is it ∗ † [email protected] [email protected] unu hroyaiso igepril systems particle single of thermodynamics Quantum lsia hroyaisi ul ihtecneto qiiru tts However, states. equilibrium of concept the with built is thermodynamics Classical 1 eateto hsc,Ntoa hn ugUiest,Tai University, Kung Cheng National , of Department d aiu Ali, Manirul Md. 1, ∗ e-igHuang, Wei-Ming Abstract 1 1 n e-i Zhang Wei-Min and a 00,Taiwan 70101, nan 1, † reservoir. a exact Introduction

Recent development in the field of open quantum systems initiates a renewed interest on the issue of taking place under nonequilibrium situations [1–8]. Meanwhile, a conceptual understanding of thermodynamic laws in the quantum domain has also been central to research for quantum thermodynamics [9–16]. The questions of how classical thermodynamics emerges from quantum dynamics, how do quantum systems dynamically equilibrate and thermalize, and whether thermalization is always reachable in quantum systems, have attracted much attention in recent years. However, the theory of quantum thermodynamics that has conceptually no ambigurity and can be widely acceptable has yet been reached from various approximation methods, due to lack of the exact solution of nonequilibrium dynamics for open quantum systems. In this paper, we will answer these questions by rigorously and exactly solving the nonequilibrium dynamics based on the exact we developed recently for a large class of open quantum systems [17–21].

Recall that classical thermodynamics is built with the concept of equilibrium states that are completely characterized by specifying the relation between the internal energy E and a set of other extensive parameters: the entropy S, the volume V , and the particle num-

ber Ni of different components i = 1, 2, ··· , etc. Explicitly, classical thermodynamics is

described by the thermodynamic fundamental equation, E = E(S,V,N1, N2, ··· ), under the extremum principle of either maximizing the entropy or minimizing the energy [22].

The thermodynamic temperature T , the pressure P and the chemical potential µ1,µ2, ··· are defined by the first derivative of energy with respect to entropy, the volume and the particle number: T = ∂E , P = ∂E , µ = ∂E , ··· such that ∂S V,N1,N2,··· ∂V S,N1,N2,··· 1 ∂N1 S,V,N2,··· dE = TdS + PdV + µ1dN1 + ··· . Microscopically, statistical averages of various thermo- dynamic quantities in equilibrium thermodynamics are calculated over a probability distri- bution function of canonical or grant canonical ensemble in the thermodynamic limit (the system with a large number of particles coupled very weakly to a reservoir at a relatively high termature).

Contrary to the above classical thermodynamics, here we consider the thermodynamics of single quantum systems coupled strongly or weakly to the reservoir at arbitrary temper- ature, including the zero-temperature (the deep quantum realm). Then the internal energy E should depend only on one extensive parameter, the entropy S of the system, and all

2 = + +

ࡴ ࡴࡿ ࡴࡾ ࡴࡵ

System

Reservoir k

ࢀ૙

FIG. 1: Schematic illustration of an open quantum system. An open quantum

system described by the Hamiltonian HS is in contact with a heat reservoir with

Hamiltonian HR, and HI describes the interaction between the system and the reservoir. The reservoir is taken initially in a thermal equilibrium state with an initial temperature

T0.

other extensive parameters V, N1, ··· are absent for such simple open quantum systems. Correspondingly, the fundamental equation of thermodynamics for such a simple quantum system takes, in principle, the form E = E(S). Conventionally, there is no classical concept of single particle temperature, whereas the single quantum particle coupled with a reser- voir can have ensemble of various microstates validating the existence of entropy. Thus, the temperature can always be introduced by the definition, T ≡ ∂E/∂S, from the above fundamental equation. In this paper, we begin with the exact nonequilibrium quantum dynamics where the system is coupled to a reservoir. We investigate the emerging thermodynamics through the real-time dynamics of various properly defined thermodynamic quantities, i.e., energy, entropy, temperature, and free energy, much before the system approaches to the steady state. More specifically, we establish the quantum thermodynamics of the system fully within the framework of .

Results

1. Formulation of quantum thermodynamics. Consider a standard open quantum system coupled to an arbitrary reservoir that is described by the Hamiltonian,

H = HS + HR + HI , (1)

which consists of the Hamiltonians of the system and the reservoir, denoted respectively

3 by HS and HR, and the interaction between them, HI. The general dynamics of the open quantum system can be determined by the exact master equation of the reduced density

matrix ρS(t), which is defined by tracing over all the reservoir degrees of freedom from the

total ρtot(t) of the system plus reservoir: ρS(t) = TrR[ρtot(t)]. The total

density matrix ρtot(t) obeys the von Neumann equation [23]: i~ρ˙tot(t) = [H, ρtot(t)]. Thus, the formal solution of the von Neumann equation gives

† ρS (t) = TrR[U(t, t0)ρtot(t0)U (t, t0)], (2)

1 where the time-evolution operator U(t, t0) = exp{ i~ H(t − t0)}, and ρtot(t0) is the initial

state of the total system (the system plus the reservoir). The reduced density matrix ρS(t) provides all the dynamics of the system under the time evolution and obey the exact master equation for a large class of open quantum systems given in Refs. [17–21]. From this exact formalism, we can use the von Neumann definition of entropy to define the nonequilibrium entropy of the system, which provides in general the lower bound of the entropy of the system and it is also the entanglement entropy between the system and the reservoir,

S(t)= −k TrS [ρS(t) log ρS (t)] . (3)

Here we have introduced the Boltzmann constant k in the above definition, in order to make a clear connection to the thermal entropy in the later discussion. Whether or not this nonequilibrium entropy can be properly defined as the quantum thermodynamic en- tropy lies on the fact whether it can correctly reproduce the equilibrium entropy in classical thermodynamics when the system reaches the equilibrium state. On the other hand, the nonequilibrium internal energy of the system at time t is quantum mechanically given by

E(t) = TrS [ρS(t)HS] , (4) which is also fully determined by ρS(t). Thus, equations (3-4) establishes the fundamental re- lation of nonequilibrium quantum thermodynamics between the energy and entropy through

the quantum density matrix ρS(t) of open quantum systems under a rigorous nonequilibrium description. Based on the above fundamental relations for nonequilibrium quantum thermodynam- ics, we generalize the concept of temperature far-beyond equilibrium. We introduce a temperature-like quantity T (t) which is defined as the derivative of the nonequilibrium

4 internal energy with respect to the nonequilibrium entropy of the system as follows:

∂E(t) T (t) ≡ (5) ∂S(t) We call this quantity as the nonequilibrium temperature (or dynamical temperature) of the system because it evolves through various quantum states much before it approaches to the steady state or thermal equilibrium. Conditionally, the nonequilibrium temperature defined by Eq. (5) is physically meaningful if it is coincident to the thermal equilibrium temperature in the steady-state limit when the system approaches the thermal equilibrium state in the weak-coupling regime, as we will show later in the paper. Consequently, using Legendre transformation of E(t) with respect to S(t) that replaces the entropy by the nonequilibrium temperature as the independent variable, we obtained the nonequilibrium Helmholtz free energy,

F (t)= E(t) − T (t)S(t), (6)

which is a nonequilibrium quantum thermodynamic potential resulting from the time evo- lution of the system. Based on such nonequilibrium quantum thermodynamic potential F (t), one can define the nonequilibrium quantum thermodynamic work, which is the time- dependent dissipative work done by the system, as the decrease of the nonequilibrium Helmholtz free energy of the system: dW (t)= −dF (t). Thus, the time-dependent (nonequi- librium) heat transferred into the system dQ(t) can be obtained as a consequence of the energy conservation,

dQ(t)= dE(t) − dF (t), (7)

where dE(t) is the change of the internal energy in the nonequilibrium process. With this new formulation of nonequilibrium quantum thermodynamics, we show next (i) how the classical thermodynamics emerges from quantum dynamics of single particle systems in the weak system-reservoir coupling regime, without introducing any hypothesis on equilibrium state; (ii) how the classical thermodynamics breaks down in the strong coupling regime due to non-Markovian memory dynamics; and (iii) a dynamical quantum phase transition associated with the occurrence of negative temperature when the reservoir is

initially in vacuum state or its initial thermal energy kT0 is less than the system initial

energy E(t0). The dynamical criticality in this nonequilibrium quantum phase transition

5 provides a inflationary dynamics that may reveal a simple physical picture to the origin of universe inflation [24]. We also show how the third law of thermodynamics, as a true quantum effect, is obtained in this theory of quantum thermodynamics.

2. Application to a cavity system. We first consider the system as a single-mode † photonic cavity with HS = ~ωsa a, in contact with a heat bath (it is also applicable to fermionic system in a similar way [18, 19]). Here ωs is the frequency of the photon field in the cavity. The bath is modeled as a collection of infinite photonic modes with a continuous frequency spectrum, and the system-bath interaction is described by the Fano-Anderson model that has wide applications in atomic physics and condensed matter physics [25–27]. Thus, the total Hamiltonian of the system plus the bath is given by

~ † ~ † ~ † ∗ † H = ωsa a + ωkbkbk + (Vka bk + Vk bka). (8) Xk Xk The reduced density matrix of the system obey the following exact master equation [17, 20, 28]: d ρ (t)= −iω′ (t, t ) a†a, ρ(t) + γ(t, t ) 2aρ (t)a† −a†aρ (t)−ρ (t)a†a dt S s 0 0 S S S   †  † † † +γ(t, t0) a ρS (t)a + aρS(t)a − a aρS(t)−ρS(t)aa , (9)   in which the renormalized systeme frequency, the dissipation (relaxation) and fluctuation ′ (noise) coefficients are determined exactly by ωs(t, t0) = −Im[u ˙(t, t0)/u(t, t0)], γ(t, t0) =

−Re[u ˙(t, t0)/u(t, t0)] and γ(t, t0) =v ˙(t, t) − 2v(t, t)Re[u ˙(t, t0)/u(t, t0)]. These exact coeffi- cients are determined by the two basic Schwinger-Keldysh nonequilibrium Green function e † ′ † ′ u(t, t0)= h[a(t), a (t0)]i, and v(t, t ) ∼ha (t )a(t)i (apart by an initial-state dependent term) that satisfy the following integro-differential equations [17, 20]: d t u(t, t0)+ iωsu(t, t0)+ dτg(t, τ)u(τ, t0)=0. (10a) dt t0 t t Z ∗ v(t, t)= dτ1 dτ2 u(t, τ1)g(τ1, τ2)u (t, τ2), (10b) Zt0 Zt0 ∞ −iω(t−τ) ∞ −iω(τ1−τ2) The integral kernels g(t, τ)= 0 dωJ(ω)e ande g(τ1, τ2)= 0 dωJ(ω)n(ω, T0)e 1 are determined by the spectralR density J(ω) of the reservoir,R and n(ω, T0) = ~ω/kT0 is e e −1 the initial particle number distribution of the reservoir. The Green’s function v(t, t) which characterizes the nonequilibrium thermal fluctuations gives the nonequilibrium fluctuation- dissipation theorem [17]. The exact analytic solution of the nonequilibrium propagating

6 ∞ Green function u(t, t0) has been recently given [17] as u(t, t0)= −∞ dωD(ω) exp{−iω(t−t0)} with D(ω) = Dl(ω)+ Dc(ω), where Dl(ω) is the contributionR of a dissipationless localized bound state in the Fano model, and Dc(ω) is the continuous part of the spectra (corre- sponding to the level broadening induced by the reservoir) that characterizes the relax- ation of the system. Without loss of generality, we consider an Ohmic spectral density

J(ω)= ηω exp (−ω/ωc), where η is the coupling strength between the system and the ther- mal reservoir, and ωc is the frequency cutoff of the reservoir spectra [29]. In this case, the system has a localized bound state when the coupling strength η exceeds the critical value

ηc = ωs/ωc. With the above specification, all quantum thermodynamical quantities can be calculated exactly. Explicitly, let the system be prepared in an energy eigenstate with the initial density matrix ρ(t0)= |E0ihE0|, where E0 = n0~ωs and n0 is an integer. The reservoir is initially in thermal equilibrium state with an initial temperature T0. After the initial time t0, both the system and the reservoir evolve into nonequilibrium states. The system state at arbitrary later time t can be obtained rigorously by solving the exact master equation of Eq. (9). The result is [30]:

∞ n0 ρS(t)= Wn (t)|EnihEn|, (11a) n=0 X [v(t, t)]n W n0 (t)= [1 − A (t, t )]n0 n [1 + v(t, t)]n+1 0

min{n0,n} k n0 n 1 A(t, t ) × 0 , (11b) k ! k!"v(t, t) 1 − A(t, t0)# Xk=0 2 |u(t,t0)| where A (t, t0) = 1+v(t,t) . From these results, all the nonequilibrium thermodynamic quan- tities, namely the time-dependent energy, entropy, temperature and free energy, etc., can

n0 be calculated straightforward. Explicitly, the function Wn (t) is the probability of finding the quantum system with energy En = n~ωs at time t. The internal energy Eq. (4) of the

n0 system at time t is given by E(t)= n Wn (t)En. The von Neumann entropy of the density ∞ n0 n0 operator ρ(t), i.e., Eq. (3), is simplyP reduced to S(t) = −k n=0 Wn (t) log Wn (t). These results establish explicitly the fundamental relation betweenP the nonequilibrium energy and entropy of the system in the last section, from which one can determine the nonequilib- rium temperature T (t) defined by Eq. (5), and all other related nonequilibrium quantum thermodynamical quantities. The results are given as follows.

7 (i). Emergence of thermodynamics and thermo-statistics from quantum dynamics in the clas-

sical condition kT0 > E0. It is indeed straightforward to see how the system is dynamically

thermalized in the weak system-reservoir coupling regime (η ≪ ηc). Initially the system is

at zero temperature (because it is in the pure state |E0i) with zero entropy, and the reservoir

is in thermal equilibrium at a relatively high temperature T0 with kT0 > E0 = n0~ωs. Then the system and the reservoir undergo a nonequilibrium evolution due to their contact with each other. The nonequilibrium dynamics is fully determined by the nonequilibrium Green

functions u(t, t0) and v(t, t) of Eq. (10). One can indeed show explicitly that in the long

time steady-state limit t → ts, we have u(ts, t0) → 0, and the average photon number in the cavity obey the Bose-Einstein statistical distribution,

† 1 ha (ts)a(ts)= v(ts, ts) → n(ωs, T0)= ~ . (12) e ωs/kT0 − 1

n0 The probability Wn (t → ts) of occupying the n-th energy state is then solely determined by the Bose-Einstein statistical distribution, n(ωs, T0), and the system state ultimately ap- proaches unambiguously to the thermal equilibrium with the reservoir [6, 30]

∞ n H [n(ωs, T0)] 1 − S ρ(t → t )= |E ihE | = e kT0 (13) s [1 + n(ω , T )]n+1 n n Z n=0 s 0 X where Z = TrS[exp(−HS/kT0)], and T0 is the initial temperature of the reservoir. This pro- vides a rigorous proof of the dynamical thermalization within quantum mechanical frame- work. One can then calculate the time-dependent internal energy and the entropy. In Fig. 2, we show the dynamics of nonequilibrium internal energy E(t), entropy S(t), temperature T (t) and free energy F (t) in the weak system-reservoir coupling regime for several different initial states of the system. As we see, both the energy and entropy of the system increases with time due to its contact with the thermal reservoir through a nonequilibrium process (see Fig. 2a and 2b). The nonequilibrium temperature T (t) of the system increases gradually

(Fig. 2c) and finally approaches to the temperature that is identical to the temperature T0 of the reservoir, namely the system reaches the equilibrium with the reservoir. The results in Fig. 2 show precisely how thermalization is dynamically approached in the weak system-reservoir coupling for a single quantum system, independent of its initial states. As it is expected, the entropy always increases in nonequilibrium processes [see Fig. 2b], demonstrating the second law of thermodynamics. The second law places constraints on

8 15 4 a 3 b 10 3 2 5 1 0 0 5 1 0 0 150 300 450 0 150 300 450 15 10 c d n = 1 10 0 -15 n = 5 0 5 n = 10 0 0 -40 0 150 300 450 0 150 300 450

FIG. 2: Nonequilibrium quantum thermodynamics of a photonic cavity. The nonequilibrium dynamics of the quantum system in terms of various thermodynamic quantities in the weak system-reservoir coupling regime with different initial states.

Different curves represent different initial states: |E0i = |En0 i with n0 =1, 5 and 10. The

initial temperature of the thermal reservoir is taken at kT0 = 15~ωs, the cut-off frequency

ωc =5ωs, and the system-reservoir coupling constant η =0.01ηc.

state transformations [31], according to which the system evolving nonequilibriumly from the

initial state ρ(t0) makes the free energy F (t) goes down (i.e., delivering work by the system), as shown in Fig. 2d. Meantime, the increasing internal energy with the decreasing free energy (work done by the system) always makes the heat transfer into the system (nonequilibrium extension of the first law of thermodynamics), resulting in the increase of the nonequilibrium temperature T (t), as shown in Fig. 2c. In the stead-state limit, the free energy approaches to a minimum value with maximum entropy in the equilibrium state. This solution in the weak system-reservoir coupling regime produces precisely the whole theory of classical thermodynamics without introducing any hypothesis on equilibrium state of the system from the beginning. It is also very different from the previous investigations with Born-Markovian approximation, where the reservoir is required to remain in the equilibrium state in all the time. Indeed, reproducing the classical thermodynamics in the classical regime from the exact nonequilibrium quantum thermodynamics with arbitrary initial states of the system is the necessary condition for unambiguously formulating quantum thermodynamics.

9 33 4.5

= 0.1 b c = 0.3 20 2.5 c = 0.4 c = 1.3 c = 1.5 a c 5 0 0 25 50 75 100 0 25 50 75 100 5 30 d -30 20 -80 10 c 0 -130 0 25 50 75 100 0 25 50 75 100

FIG. 3: Nonequilibrium quantum thermodynamics of a photonic cavity in the strong-coupling regime. The nonequilibrium dynamics of the quantum system in terms of various thermodynamic quantities in the strong system-reservoir coupling regime.

Different curves represent different values of the coupling strengths η =0.1ηc (blue),

η =0.3ηc (green), η =0.4ηc (red), η =1.3ηc (black), η =1.5ηc (magenta). The other

parameters are ωc =5ωs, kT0 = 20~ωs, and the system is initially in the state |E0i = |E5i.

(ii). Breakdown of thermodynamics and thermo-statistics in strong system-reservoir coupling regime. In fact, the above thermalization crucially relies on the assumption of the very weak interaction between the system and the reservoir. In the literature, there is a specific thermalization condition on the system-reservoir coupling [32]: the strength of the coupling

η induces the effect that has to be negligible in comparison to initial thermal energy kT0.

We find that thermalization occurs when η ≪ ηc, where ηc is the critical value of the system-reservoir coupling for the occurrence of localized bound state [17]. In particular, our numerical calculation shows that the ideal thermalization condition occurs when η . 0.01ηc. On the other hand, the role of strong coupling on equilibrium quantum thermodynamics was only discussed in the context of damped harmonic oscillators for quantum Brownian motion [33, 34]. Some efforts have been made to formulate quantum thermodynamics for a strongly coupled system [35–37] under nonequilibrium conditions subject to an external time-dependent perturbation. However, the real-time dynamics of thermodynamic quantities were not explored from a full quantum dynamics description in the strong system-reservoir

10 coupling regime. We find that the real-time quantum thermodynamics of the strong system-reservoir cou- pling is indeed quite distinct. As it is shown in Fig. 3, when the system-reservoir cou-

pling strengths are not so weak (in the intermediate range: 0.01ηc < η 6 ηc), the system will quickly approaches to different thermal equilibrium states (called thermal-like states in Ref. [6]) for different coupling strengths with different steady-state temperatures that are larger than the initial reservoir temperature T0 (see the blue, green and red curves in Fig. 3c that correspond to different system-reservoir coupling strengths). For this intermediate coupling strength region, different equilibrium states are reached by different steady-state minima of the free energy F (t). These results already show some deviation away from the classical thermodynamics prediction, namely after a long enough time, the system must equilibrate with the reservoir that is characterized by the reservoir temperature T0. In fact, due to the relatively stronger couplings between the system and the reservoir, both the system and the reservoir undergo a nonequilibrium evolution, and then approach to a new common equilibrium state characterized by the new equilibrium temperature T (t → ts). In

Fig. 3c, it shows that ts ≃ 25/ωs, and T (ts) changes as η varies. But the principle of ther- modynamics, namely the existence of equilibrium state and the associated thermodynamic laws in the steady state, is still valid. Furthermore, when the system-reservoir coupling strength gets very strong (goes beyond the critical value η > ηc), the system indeed does not approach to thermal equilibrium [6]. In this strong coupling regime, the average energy E(t) initially increases, and then it starts decreasing with time, indicating a backflow of energy (see Fig. 3a) from the system into the reservoir, and then it oscillates. A similar dynamical behavior of entropy S(t) is seen (Fig. 3b) under this strong coupling, namely, although the entropy increases in the short time, it is decreased and oscillates in later time. Thus, strong system-reservoir couplings cause significant departure from the classical thermalization. We also do not see (Fig. 3c) a monotonous increase of dynamical temperature T (t) that can eventually approach to a steady-state value. In the strong system-reservoir coupling regime, the system indeed cannot reach to a thermal equilibrium, insteadly it approaches to a more complex nonequilibrium (oscillatory) steady state [6]. Under this situation, the nonequilibrium free energy F (t) also shows nontrivial oscillations with local minima, instead of having a global minimum, as shown in Fig. 3d.

11 The physical origin for the breakdown of classical thermodynamics in the strong system- reservoir coupling regime comes from the important effect of nonequilibrium dynamics, i.e. the non-Markovian back-action memory effect, as we have discovered earlier [6, 17]. In this situation, the system can memorize forever some of its initial state information, i.e. initial state information cannot be completely wiped out. Therefore, the basic hypothesis of , namely after a long enough time, the system will evolve into an equilibrium state with the reservoir that is independent of its initial state, breaks down and the system cannot be thermalized. The oscillation behavior in Fig. 3 is a manifestation of non-Markovian memory effect originating from the localized bound state of the system, a general non-Markovian dynamic feature in open quantum systems [17]. The localized bound states of the system produces dissipationless dynamics that do not allow the system to be thermalized with its environment, which was noticed earlier by Anderson [25, 27] and is justified recently by one of us [6]. The localized bound state can appear when the

system-environment coupling strength exceeds some critical value ηc that is determined by

the structure of spectral density J(ω). For the Ohmic spectral density, ηc = ωs/ωc. This phenomenon also reveals the fact that the transition from thermalization to localization can occur not only in many-body systems [40], but also dynamically in a single particle quantum system in contact with a reservoir [41].

(iii). Dynamical quantum phase transition and quantum criticality with quantum-classical border. Now we discuss a very interesting situation occurring in the weak-coupling quantum

regime. The quantum regime is defined as the reservoir is initially in vacuum state (T0 = 0)

or the initial thermal energy of the reservoir is less then the system initial energy: kT0 < E0,

where the system is initially in the energy eigenstate |E0i. In this situation, the energy dissipation dominates the nonequilibrium process so that the system energy always decreases in time (see Fig. 4a), which is very different from the situation in Fig. 2a where the reservoir is in a relatively high temperature classical regime (kT0 > E0) and the energy flows in an opposite way. On the other hand, as shown in Fig. 4b the system entropy (beginning with zero) will increase to a maximum value at an intermediate time, then the entropy turns to decrease and eventually approaches to zero (the blue curve in Fig. 4b for kT0 = 0 case). This phenomenon can happen because the system is initially in a pure state (zero entropy), then the system takes relaxation and goes eventually into the vacuum state (back to a

12 pure state) when all energy of the (small) system is eventually dissipated into the (large) reservoir. Thus, at the beginning and in the end of the nonequilibrium process, the entropy of the system is zero. But during the nonequilibrium evolution, the entropy must go up (increase) in time and reach the maximum at some intermediate time (corresponding to the maximizing mixed state of the system), and then go down (decrease) to zero again in the rest of relaxation process. On the other hand, because the system energy always decreases, the nonequilibrium temperature which starts from zero, must decrease to negative infinity as the entropy approaches to the maximum. It then jumps from the negative infinite temperature to the positive infinite temperature at the maximum entropy point, and then goes down as the entropy decreases, and eventually reaches to zero temperature at the steady state, as shown by the blue curve in Fig. 4c. This dramatic change of the dynamical temperature shows that for the nonequilibrium processes of open systems in the quantum regime, the second law of thermodynamics cannot be always maintained (because of the appearance of a negative temperature and the decrease of the entropy). But at the end, the entropy and the temperature both approach to zero,

5 a 2 b kT = 0 2.5 0 kT = E 1 0 3 0 0 0 400 800 0 400 800 100 100 c 100 d 0 0 -100 -20 67 68 69 80 800 0 0 10

0 80 800 -100 -100 0 50 100 150 0 50 100 150

FIG. 4: Inflationary dynamics and dynamical quantum phase transition. The dynamics of (a) energy E(t), (b) entropy S(t), (c) dynamical temperature T (t), and (d) free energy F (t) are shown for the reservoir initially in vacuum state (blue curves) and for the reservoir with initial thermal energy kT0 =3~ω0 (red curves). The system is initially in the state |En0 i = |E5i, and a weak system-reservoir coupling strength η =0.01ηc (blue) is taken.

13 as a manifestation of the third law of thermodynamics in equilibrium state. Note that although the third law of thermodynamics is hypothesized in classical thermodynamics, it cannot be demonstrated within classical thermodynamics, namely the zero temperature is not reachable as the entropy cannot approach to zero in the classical regime. This is because zero temperature locates in the very deep quantum mechanical realm (corresponding to the situation of the system staying in the ). Therefore as a criteria, a correct theory of quantum thermodynamics must capture the generalized third law of thermodynamics, namely the entropy becomes zero as the temperature approach to zero in quantum regime. Here we show explicitly how the third law of thermodynamics at equilibrium state is obtained dynamically from quantum mechanics.

On the other hand, the negative temperature in open quantum systems we find here, although it sounds to be very strange, only appears in the nonequilibrium region. We find that this negative nonequilibrium temperature phenomenon always occurs when the system

is initially in a pure state with the condition that the system initial energy E0 is large than the initial thermal energy kT0 of the reservoir. More importantly, we find that the occurrence of the negative nonequilibrium temperature is accompanied with a nontrivial behavior of the entropy changing from increasing to decreasing, as shown in Fig. 4b-c. Here we would like to argue that this unusual dynamics may provide a simple physical picture to the origin of the universe inflation [24]. Suppose that in the very beginning, our universe is a single system staying in a pure [49] and nothing else, so that the reservoir is in vacuum. Then the universe should evolve through a nonequilibrium process similar to that given in Fig. 4. The sudden jump from the negative infinite temperature to the positive infinite temperature at the transition point (corresponding to the entropy varies from increase to decrease in time) shown in Fig. 4c allows the system to suddenly generate an infinite amount of heat energy dQ = d[T S] → ∞ [i.e. an inflationary energy as a result of the maximizing information (entropy) of the system]. At the same time, the free energy of the system is always increased during the nonequilibrium evolution, but only at the transition point it is suddenly dropped. This sudden change corresponds to deliver an inflationary work dW →∞ (i.e., an inflationary decrease of free energy of the system, dF → −∞) to the reservoir, see Fig. 4d. This may provide naturally a simple physical process for the origin of big bang and the universe inflation [24, 49] which is a long-standing problem that has not been solved so far in . The continuous decrease of the entropy in the rest of the

14 FIG. 5: Dynamical quantum phase transition in quantum thermodynamics. Contour color-code plots of (a) the dynamical energy E(t) and (b) the nonequilibrium

temperature T (t) are presented as varying initial thermal energy kT0 of the reservoir. The other parameters are taken as η =0.01 ηc, ωc =5ωs, and the system is considered to be in an initial state |En0 i = |E5i. A dynamical quantum phase transition occurs when initial thermal energy kT0 of the reservoir is very close to the initial system energy En0 . nonequilibrium evolution may explain how the universe evolves to flatness. Interestingly, an inflationary dynamics in Bose-Einstein condensates (BEC) crossing a quantum critical point is recently observed [42]. How to make the above inflationary dynamics into a more realistic cosmology model remains for further investigation. Now if we change the initial state of the reservoir from vacuum to a thermal state with a low thermal energy, kT0 < E0, the time-dependent entropy and the temperature of the system show the same behavior, except for the steady state of the system which becomes now a thermal state with temperature T0, equilibrating to the reservoir, as shown by the red curves in Fig. 4. Thus, when we change the initial thermal energy (kT0) of the reservoir from a lower value to a higher value crossing the system initial energy E0 (or alternatively change the system initial energy from E0 < kT0 to E0 > kT0 which is experimentally more feasible), we observe (see Fig. 5) a nontrivial dynamical quantum phase transition when the thermal energy kT0 of the reservoir is very close to the initial system energy E0. This quantum criticality provides indeed a dynamical border separating the classical thermodynamics from quantum thermodynamics, namely classical thermodynamics is only valid for a rather high temperature regime (kT0 > E0), while in the quantum regime: kT0 < E0, only quantum thermodynamics can capture thermodynamics phenomenon, manifested by the occurrence of negative temperature and the decrease of entropy in the nonequilibrium process, and it

15 is finally also justified by the third law of thermodynamics in the steady-state limit.

3. Application to a two-level atomic system. To demonstrate the universality of the above finding in the simple cavity system, we investigate next the real-time thermodynamics of another widely studied open quantum system, namely a two-level atomic system (or a general spin-1/2 particle) interacting with a bosonic reservoir, to see how thermodynamics emerges from quantum dynamics of spontaneous decay process. The total Hamiltonian of the system plus reservoir is given by

~ ~ † ~ ∗ † H = ωsσ+σ− + ωkbkbk + gkσ+bk + gkσ−bk , (14) Xk Xk   where σ+ = |1ih0| and σ− = |0ih1| are the raising and lowering operators of the Pauli

matrix with ground state |0i, excited state |1i, and transition frequency ωs. The reservoir Hamiltonian is described by a collection of infinite bosonic modes with the bosonic operators † bk and bk, and gk is the coupling strength between the system and kth mode of the reservoir. We consider that the system is prepared initially in the excited state |1i and the reservoir in a vacuum state. The time evolution of the reduced density matrix of the two-level system is given by the exact master equation [50, 51] d ρ(t)= −iω′ (t, t )[σ σ , ρ(t)] + γ(t, t ) [2σ ρ(t)σ − σ σ ρ(t) − ρ(t)σ σ ] , (15) dt s 0 + − 0 − + + − + − ′ where the renormalized transition frequency ωs(t, t0) = −Im (u ˙(t, t0)/u(t, t0)) and the dis-

sipation coefficient γ(t, t0) = −Re (u ˙(t, t0)/u(t, t0)), which is the same as that given in the

general exact master equation of Eq. (9). The Green function u(t, t0) satisfies also the same integro-differential equation as Eq. (10a) d t u(t, t )+ iω u(t, t )+ dτf(t, τ)u(τ, t )=0, (16) dt 0 s 0 0 Zt0 where the two-time reservoir correlation function f(t, τ) is related to the spectral density J(ω) of the reservoir as f(t, τ)= dωJ(ω)e−iω(t−τ). Here we consider a Lorentzian spectral density, R 2 1 γ0λ J(ω)= 2 2 , (17) 2π (ω0 − ω) + λ

for which the Green function u(t, t0) of Eq. (16) can easily be solved with the exact solution

λ Γ(t − t0) λ Γ(t − t0) u(t, t =0)= e(−iωs− 2 )(t−t0) cosh + sinh , (18) 0 2 Γ 2   16 2 where Γ = λ − 2γ0λ, which has no localized bound state because the spectral density covers the wholep range of the frequency domain. The spectral width of the reservoir λ is −1 connected to the reservoir correlation time τB by the relation τB ≈ λ . The relaxation time scale τR over which the state of the system changes is related to the coupling strength γ0 by −1 the relation τR ≈ γ0 . The time evolved density matrix ρ(t) depicts the dissipative relaxation process of the system at far-from-equilibrium. Once we have the density matrix ρ(t), the time-dependent nonequilibrium thermodynamic quantities are determined by the relations (3-7). In Fig. 6, we show the dynamics of nonequilibrium thermodynamic quantities, namely, internal energy E(t), entropy S(t), temperature T (t) and the free energy F (t). We consider

two different regime of the system-environment parameters: (i) λ > 2γ0, for which the reservoir correlation time is small compared to the relaxation time (τB < τR) of the system

and the behavior of u(t, t0) shows a Markovian exponential decay (ii) λ< 2γ0, in this regime

the reservoir correlation time τB is large or comparable to the relaxation time scale τR where non-Markovian effects come into play [52]. We investigate the real-time thermodynamics of the system in the above two physically distinct regimes (i) and (ii). In the weak coupling

case with τR = 5τB, the energy E(t) shows a monotonous decay due to dissipation, as shown in Fig. 6a (blue curve). On the other hand, Figure 6b (blue curve) shows that the entropy S(t) increases due to the contact of the system to the reservoir, and attains a maximum value at an intermediate time, henceforth the entropy starts decreasing and eventually approaches to zero. This happens because the system is initially in a pure state (zero entropy), then the entropy increases during the nonequilibrium evolution and reach the maximum corresponding to maximum mixing state of the system, finally the entropy goes to zero again when the system relaxes into the vacuum state (pure state) in order to equilibrate with the reservoir. The decrease in energy with increasing entropy results in a negative nonequilibrium temperature of the system, as shown in Fig. 6c. The nonequilibrium temperature T (t) decreased to negative infinity as the entropy approaches to the maximum, it then jumps from negative infinite temperature to positive infinite temperature at the maximum entropy point, eventually the system reaches to zero temperature at the steady state as shown in Fig. 6c, the exactly same as what we find in the cavity system in the weak-coupling quantum regime.

The physical origin of negative nonequilibrium temperature for this two-level atomic system is also the same as that in cavity system: The free energy increases instead of

17 1.0 0.8 a b (i) λ > 2γ0 (τR = 5 τB) 0.8 0.6 (ii) λ < 2γ0 (τR = 0 .2τB) ) 0.6 ) t t ( ( 0.4 S E 0.4 0.2 0.2 0.0 0.0 0 10 20 30 40 50 0 10 20 30 40 50 40 c 40 d 20 20 ) ) t t ( 0 ( 0 T F −k 20 −20 −40 −40 0 2 4 6 8 10 0 2 4 6 8 10 γ0t γ0t

FIG. 6: Nonequilibrium quantum thermodynamics of a two-level atomic system. We show the dynamics of nonequilibrium thermodynamic quantities at two physically distinct regime (i) λ> 2γ0 and (ii) λ< 2γ0. We plot (a) average energy E(t) (b) entropy S(t) (c) nonequiblibrium temperature T (t) and (d) free energy F (t), where the

blue curves represent weak coupling regime (i) with τR =5τB, and the red curves indicate

the strong coupling regime (ii) with τR =0.2τB. The system is initially taken in the excited state, and the reservoir is initially in vacuum state. decreasing (see Fig. 6d), an amount of work dW is done on the system by the reservoir resulting to an amount of heat dQ dissipated from the system into the reservoir. Hence, for this two-level system, the second law of thermodynamics also cannot be maintained when the system evolves from a pure state to a mixed state and then back to a pure state in the whole spontaneous decay processes. Nevertheless, the entropy and the temperature both dynamically approach to zero in the steady-state limit, as a manifestation of the third law of thermodynamics in equilibrium state in the quantum regime. In the strong coupling case

with τR =0.2τB (see the red curves of Fig. 6), we show the nonequilibrium thermodynamic

dynamics in the regime λ< 2γ0. In this regime, the average energy E(t) initially decreases, then it shows a small oscillation, demonstrating a backflow of energy (red curve of Fig. 6a) from the system into the reservoir, as a non-Markovian memory effect. We see a similar dynamical behavior in the entropy S(t) under this strong coupling, namely, the entropy increases in the short time, then decreases and oscillates in later time (red curve of Fig. 6b).

18 The nonequilibrium temperature and free energy in the strong coupling regime (red curves of Fig. 6c-d) also show similar behaviors except a small shift in the time-scale. The dynamical transition associated with the inflationary dynamics is the same as we found in the single cavity system. The concept of negative temperatures has been discussed in the literature [43, 44] which demand the system to be in the equilibrium and also to have a finite spectrum of energy states, bounded both from above and below. Such negative equilibrium temperature has been realized in localized spin systems [45–47], where the discrete finite spectrum naturally provides both lower and upper energy bounds. A similar situation is also found recently where a cloud of potassium atoms is tuned to negative temperatures via quantum phase transition [48]. However, it is important to mention here that the physical origin or condition under which the negative temperature phenomenon occurs here in our cases is quite distinct from the previous studies [43–48]. In previous studies [43–48], it was emphasized that

in thermal equilibrium, the probability for a particle to occupy a state with energy En is proportional to the Boltzmann factor exp(−En/kT ). For negative temperatures, the

Boltzmann factor increases exponentially with increasing En and the high-energy states are then occupied more than the low-energy states. Contrary to that, the emergence of negative temperature here we see is dynamical, and it has the clear physical origin in the nonequilibrium dynamics evolution that the entropy increases and then decreases as the system starts from a pure state, evolves into a maximum mixed state and then returns back to a pure state (or a low mixed state) dominated by the spontaneous emissions when the reservoir is initially in vacuum state (or an initial state with a very low thermal energy). Thus, our finding is indeed universal for other open quantum systems.

Conclusions and outlook

Classical thermodynamics is built with the concept of equilibrium states, where dynamics is absent and all thermodynamic quantities are time-independent. In this paper, we devel- oped an exact nonequilibrium theory for quantum thermodynamics validating for arbitrary quantum systems in contact with reservoirs at arbitrary time. We generalize the concept of temperature to nonequilibrium that depends on the details of quantum states of the system and their dynamical evolution. We show that the exact nonequilibrium theory unravels the intimate connection between laws of thermodynamics and their quantum origin. We ex-

19 plicitly and exactly solved the dynamical evolution of the quantum system interacting with a heat reservoir. We demonstrated the emerging thermodynamics of the system through real-time dynamics of various thermo-dynamic quantities, namely the time-dependent en- ergy, entropy, temperature, free energy, heat and work in the weak system-reservoir coupling regime under the situation much before the system approaches to thermal equilibrium, and showed how the system dynamically approaches to a thermal equilibrium state asymptot- ically in the steady-state limit. In the strong system-reservoir coupling regime, a revival dynamics in the nonequilibrium thermodynamic quantities is observed, indicating back- reactions between the system and the environment as a non-Markovian memory effect, so that equilibrium thermodynamics is unreachable, as noticed early by Anderson [25] and justified recently by one of us [6]. Our most remarkable findings in this paper appear in the nonequilibrium thermodynam-

ics with the reservoir thermal energy kT0 being less than the system initial energy, which is defined as the quantum regime. Under this situation, we observe a dramatic change in the thermodynamic behavior of the system that the system dissipates its energy into the reservoir in time, while the entropy of the system first increases with the decrease of sys- tem energy so that the system is characterized with a negative nonequilibrium temperature. After the entropy reaches a maximum value at an intermediate time (corresponding to the maximized mixed state of the system), the entropy decreases in the rest of nonequilibrium evolution. Correspondingly, the nonequilibrium temperature jumps from the negative infi- nite to the positive infinite values at the maximum entropy point, then decreases to the value equilibrating to the reservoir. This dramatic change of thermodynamic behavior induces an inflationary heat energy in the system that suddenly delivers into the reservoir as an infla- tionary work done by the system, which may provide a simple picture for the origin of big bang and universe inflation. The ending steady state manifests the third law of thermody- namics in the deep quantum realm. We also observed a nontrivial dynamical quantum phase transition when the thermal energy of the reservoir is very close to the system initial energy. The corresponding dynamical criticality provides clearly a border separating quantum and classical thermodynamics. This dynamical quantum phase transition is manifested through nonequilibrium quantum states that cannot be captured by equilibrium thermodynamics. These new findings for quantum thermodynamics could be examined through BEC ex- periments or cavity QED experiments. More specifically, one can experimentally explore

20 the photon dynamics or boson dynamics via quantum non-demolition measurement [53, 54] to justify these findings. For example, by preparing the cavity in a Fock state, and sending sequences of circular Rydberg atoms through the photonic cavity in different environment, which carry the cavity state information without destroying the cavity photon state. The cavity photon state can be measured using an experimental setup similar to that given in Ref. [53, 55]. Another way to observe the nontrivial quantum thermodynamics we obtained is to measure the energy loss and its state evolution of a two-level atom in cavity system. It may be also possible to use the BEC systems [42] to find these new features. On the other hand, it is also straightforward to apply this quantum thermodynamics theory to electronic systems in nanostructures where the exact master equation with quantum transport has been developed [17, 56]. Furthermore, our nonequilibrium theory for quantum thermodynamics can provide a platform to probe the universal validity of second law of thermodynamics in many other systems [57–60]. We will also investigate various fluctuation relations [37, 61–63] in the quantum regime within this framework.

[1] Trotzky, S. et al. Probing the relaxation towards equilibrium in an isolated strongly correlated one-dimensional Bose gas. Nature Phys. 8, 325 (2012). [2] Gring, M. et al. Relaxation and prethermalization in an isolated quantum system. Science 337, 1318 (2012). [3] Langen, T., Geiger, R., Kuhnert, M., Rauer, B. & Schmiedmayer, J. Local emergence of thermal correlations in an isolated quantum many-body system. Nature Phys. 9, 640 (2013). [4] Kosloff, R. Quantum thermodynamics: a dynamical viewpoint. Entropy 15, 2100 (2013). [5] Eisert, J., Friesdorf, M. & Gogolin, C. Quantum many-body systems out of equilibrium. Nature Phys. 11, 124 (2015). [6] Xiong, H. N., Lo, P. Y., Zhang, W. M., Feng, D. H. & Nori, F. Non-Markovian complexity in the quantum-to-classical transition. Sci. Rep. 5, 13353 (2015). [7] Millen, J. & Xuereb, A. Perspective on quantum thermodynamics. New J. Phys. 18, 011002 (2016). [8] Anders, J. & Esposito, M. Focus on quantum thermodynamics. New J. Phys. 19, 010201 (2017).

21 [9] Gemmer, J., Michel, M. & Mahler, G. Quantum Thermodynamics: Emergence of Thermody- namic Behavior Within Composite Quantum Systems (Springer, Berlin, 2009). [10] Campisi, M., Hnggi, P. and Talkner, P. Colloquium: Quantum Fluctuation Relations: Foun- dations and Applications, Rev. Mod. Phys. 83, 771 (2011). [11] Jarzynski, C. Equalities and Inequalities: Irreversibility and the Second Law of Thermody- namics at the Nanoscale, Annu. Rev. Condens. Matter Phys. 2, 329 (2011). [12] Scully, M. O., Chapin, K. R., Dorfman, K. E., Kim, M. B., & Svidzinsky, A. Quantum heat engine power can be increased by noise-induced , PNAS 108, 15097 (2011). [13] Saira, O.-P., Yoon, Y., Tanttu, Mttnen, T., M., Averin, D. V. & Pekola, J. P. Test of the Jarzynski and Crooks Fluctuation Relations in an Electronic System, Phys. Rev. Lett. 109, 180601 (2012). [14] Esposito, M. Ochoa, M. A. & Galperin, M. Quantum thermodynamics: A nonequilibrium Green’s function approach. Phys. Rev. Lett. 114, 080602 (2015). [15] Pekola, J. P. Towards Quantum Thermodynamics in Electronic Circuits, Nat. Phys. 11, 118 (2015). [16] Roßnagel, J., Dawkins, S. T., Tolazzi, K. N., Abah, O., Lutz, E., Schmidt-Kaler, F. & Singer, K. A single-atom heat engine. Science 352, 325 (2016). [17] Zhang, W. M., Lo, P. Y., Xiong, H. N., Tu, M. W. Y. & Nori, F. General non-Markovian dynamics of open quantum systems. Phys. Rev. Lett. 109, 170402 (2012). [18] Tu, M. W. Y. & Zhang, W. M. Non-Markovian decoherence theory for a double-dot charge qubit. Phys. Rev. B 78, 235311 (2008). [19] Jin, J., Tu, M. W. Y., Zhang, W. M. & Yan, J. Y. Non-equilibrium quantum theory for nanodevices based on the Feynman-Vernon influence functional, New. J. Phys. 12, 083013 (2010). [20] Lei, C. U. & Zhang, W. M. A quantum photonic dissipative transport theory. Ann. Phys. 327, 1408 (2012). [21] Zhang, W. M. Exact master equation and general non-Markovian dynamics in open quantum systems, arXiv:1807.01965 (2018). [22] Callen, H. B. Thermodynamics and an Introduction to Thermostatistics (John Wiley & Sons, New York, 1985). [23] von Neumann, J. Mathematical foundations of quantum mechanics. (Princeton university

22 press, 1955). [24] Guth, A. H. Infationary universe: a possible solution to the horizon and fatness problems. Phys. Rev. D 23, 347 (1981). [25] Anderson, P. W. Absence of diffusion in certain random lattices. Phys. Rev. 109, 1492 (1958). [26] Fano, U. Effects of configuration interaction on intensities and phase shifts. Phys. Rev. 124, 1866 (1961). [27] Mahan, G. D. Many-Body Physics (Kluwer Academic/Plenum Publishers, New York, 2000). [28] Xiong, H. N., Zhang, W. M., Wang, X., & Wu, M. H. Exact non-Markovian cavity dynamics strongly coupled to a reservoir, Phys. Rev. A 82 (2010) 012105. [29] Leggett, A. J. et al. Dynamics of the dissipative two-state system. Rev. Mod. Phys. 59, 1 (1987). [30] Lo, P. Y., Xiong, H. N. & Zhang, W. M. Breakdown of Bose-Einstein distribution in photonic crystals. Sci. Rep. 5, 9423 (2015). [31] Brand˜ao, F., Horodecki, M., Ng N., Oppenheim, J. & Wehner, S. The second laws of quantum thermodynamics. Proc. Natl. Acad. Sci. 112, 3275 (2015). [32] Riera, A., Gogolin, C. & Eisert, J. Thermalization in nature and on a quantum computer. Phys. Rev. Lett. 108, 080402 (2012). [33] Ford, G. W., Lewis, J. T. & O’Connell R. F. Quantum oscillator in a blackbody radiation field. Phys. Rev. Lett. 55, 2273 (1985). [34] Suba¸si, Y., Fleming, C. H., Taylor, J. M. & Hu, B. L. Equilibrium states of open quantum systems in the strong coupling regime. Phys. Rev. E 86, 061132 (2012). [35] Seifert, U. First and second law of thermodynamics at strong coupling. Phys. Rev. Lett. 116, 020601 (2016). [36] Gallego, R., Riera, A. & Eisert, J. Thermal machines beyond the weak coupling regime. New J. Phys. 16, 125009 (2014). [37] Campisi, M., Talkner, P. & H¨anggi, P. Fluctuation theorem for arbitrary open quantum systems. Phys. Rev. Lett. 102, 210401 (2009). [38] Ali, M. M., Lo, P. Y., Tu, M. W. Y. & Zhang, W. M. Non-Markovianity measure using two-time correlation functions. Phys. Rev. A 92, 062306 (2015). [39] Ali, M. M. & Zhang, W. M. Nonequilibrium transient dynamics of photon statistics. Phys. Rev. A 95, 033830 (2017).

23 [40] Nandkishore, R. & Huse, D. A. Many-body localization and thermalization in quantum sta- tistical mechanics. Ann. Rev. of Conden. Matter Phys. 6, 15 (2015). [41] Lin, Y. C., Yang, P. Y. & Zhang, W. M. Non-equilibrium quantum phase transition via entanglement decoherence dynamics. Sci. Rep. 6, 34804 (2016). [42] Feng, L., Clark, L. W., Gaj, A. & Chin, C. Coherent inflationary dynamics for Bose-Einstein condensates crossing a quantum critical point. Nature Phys. 14, 269 (2018). [43] Ramsey, N. F. Thermodynamics and statistical mechanics at negative absolute temperatures. Phys. Rev. 103, 20 (1956). [44] Klein, M. J. Negative Absolute Temperatures. Phys. Rev. 104, 589 (1956). [45] Purcell, E. M. & Pound R. V. A nuclear spin system at negative temperature. Phys. Rev. 81, 279 (1951). [46] Oja, A. S. & Lounasmaa, O. V. Nuclear magnetic ordering in simple metals at positive and negative nanokelvin temperatures. Rev. Mod. Phys. 69, 1 (1997). [47] Medley, P., Weld, D. M., Miyake, H., Pritchard, D. E. & Ketterle, W. Spin gradient demag- netization cooling of ultracold atoms. Phys. Rev. Lett. 106, 195301 (2011). [48] Braun, S. et al. Negative absolute temperature for motional degrees of freedom. Science 339, 52 (2013). [49] Hartle, J. B. & Hawking, S. W. of the Universe. Phys. Rev. D28, 2960 (1983). [50] Garraway, B. M. Nonperturbative decay of an atomic system in a cavity, Phys. Rev. A 55 (1997) 2290. [51] Vacchini, B. & Breuer, H. P. Exact master equations for the non-Markovian decay of a qubit. Phys. Rev. A 81, 042103 (2010). [52] Bellomo, B., Franco, R. L. & Compagno, G. Non-Markovian Effects on the Dynamics of Entanglement, Phys. Rev. Lett. 99, 160502 (2007) [53] Guerlin, C. et al. Progressive field-state collapse and quantum non-demolition photon count- ing. Nature 448, 889 (2007). [54] Del´eglise, S. et al. Reconstruction of non-classical cavity field states with snapshots of their decoherence. Nature 455, 510 (2008). [55] Sayrin, C. et al. Real-time quantum feedback prepares and stabilizes photon number states. Nature 477, 73 (2011). [56] Yang, P. Y. & Zhang, W. M. Master equation approach to transient quantum transport in

24 nanostructures. Front. Phys. 12, 127204 (2017). [57] Allahverdyan, A. E. & Nieuwenhuizen, T. M. Extraction of work from a single thermal bath in the quantum regime. Phys. Rev. Lett. 85, 1799 (2000). [58] Wang, G. M., Sevick, E. M., Mittag, E., Searles, D. J. & Evans, D. J. Experimental demon- stration of violations of the second law of thermodynamics for small systems and short time scales. Phys. Rev. Lett. 89, 050601 (2002). [59] Scully, M. O., Zubairy, M. S., Agarwal, G. S. & Walther, H. Extracting work from a single heat bath via vanishing quantum coherence. Science 299, 862 (2003). [60] Roßnagel, J., Abah, O., Kaler, F. S., Singer, K. & Lutz, E. Nanoscale heat engine beyond the Carnot limit. Phys. Rev. Lett. 112, 030602 (2014). [61] Jarzynski, C. Diverse phenomena, common themes. Nature Phys. 11, 105 (2015). [62] Campisi, M., H¨anggi, P. & Talkner, P. Quantum fluctuation relations: Foundations and ap- plications. Rev. Mod. Phys. 83, 771 (2011). [63] H¨anggi, P. & Talkner, P. The other QFT. Nature Phys. 11, 108 (2015).

Acknowledgments WMZ acknowledges the support from the Ministry of Science and Technology of Taiwan under Contract No. NSC-102-2112-M-006-016-MY3. MMA acknowledges the support from the Ministry of Science and Technology of Taiwan and the Physics Division of National Center for Theoretical Sciences, Taiwan.

Author Contributions MMA performed the detailed calculations and WMH performed the calculation for the two- level systems. WMZ proposed the ideas and interpreted the physics. MMA and WMZ both have contributed on discussions, physical formulation and writing the manuscript.

Competing financial interests The authors declare no competing financial interests.

Additional Information Correspondence and requests for materials should be addressed to M. M. Ali (maniquan-

25 [email protected]) and W. M. Zhang ([email protected]).

26