Semi-classical work and quantum work identities in Weyl representation

O. Brodier1, K. Mallick2 and A. M. Ozorio de Almeida3 1 Institut Denis Poisson, Campus de Grandmont, Universit´ede Tours, 37200 TOURS 2 Institut de Physique Th´eorique,,Universit´eParis-Saclay, CEA and CNRS, 91191 Gif-sur-Yvette, France 3 Centro Brasileiro de Pesquisas F´ısicas, Rua Dr. Xavier Sigaud, 150, 22290-180 Rio de Janeiro, BRASIL (Dated: September 25, 2020) We derive a semi-classical nonequilibrium work identity by applying the Wigner-Weyl quantization scheme to the Jarzynski identity for a classical Hamiltonian. This allows us to extend the concept of work to the leading order in ~. We propose a geometric interpretation of this semi-classical Jarzynski relation in terms of trajectories in a complex phase space and illustrate it with the exactly solvable case of the quantum harmonic oscillator.

PACS numbers: 03.65.Sq,03.65.Yz

According to the second principle of , macroscopic phenomena tend to evolve towards states cor- responding to a maximum number of underlying microstates, i.e. states that maximize entropy. Combined with the first principle, this leads to a more operational statement for isothermal processes: the minimal amount of work to modify a system from a state A to a state B is given by

W ≥ F (B) − F (A), (1)

where F = U − TS is the free energy. has shown that the interpretation of macroscopic thermodynamics is statistical, by endowing microscopic states with a probability measure. Hence, the second principle should be understood as an average of a random process; one should write, in fact,

hW i ≥ F (B) − F (A), (2)

where hW i is the average of path-dependent work along the realizations of a given macroscopic process (or protocol) leading from state A to state B. In 1996, C. Jarzynski [1, 2] discovered that there exists a non-equilibrium work relation, underlying the long-established work inequality :

he−βW i = e−β(F (B)−F (A)). (3)

A remarkable consequence, from the thermodynamics point of view, is that, for (3) to be true, some individual realizations of the transform must ‘violate’ the second principle, that is, the system can reach once in a while a final state whose free energy variation is actually greater than the received work. The Jarzynski identity and its generalization by G. Crooks [3] have triggered an immense amount of work during the last two decades (see e.g. [4–6] for reviews) and has been verified experimentally [7–9]. The original approach of Jarzynski was based on classical trajectories and on the classical definition of work. It was therefore a challenge to generalize it to a quantum system. For closed quantum systems, this difficulty was overcome by the so-called two measurements process [10–16, 21], where work is defined as the difference of energy between the end and the beginning of the evolution. This scheme was also studied using the Weyl formalism [17, 18] and path integrals [19]. As explained by Talkner et al. [20], this definition of work does not correspond to a quantum observable because it can not be represented by a Hermitian operator. This explains why alternative proposals based on some quantum work operator did not obey the Jarzynski identity [22–24]. For an open system, the two measurements arXiv:1912.04969v2 [quant-ph] 24 Sep 2020 scheme could be applied by considering the system together with its environment as a global, closed, system (see e.g. [25–27]). A different strategy to study open systems is to use a quantum map that acts on the density matrix of the system [28–30]. Under suitable assumptions, this map leads to a quantum Markov evolution, described by a Lindblad equation [31]. In this dynamical framework, a quantum analog of the Jarzynski relation can be proved by defining a work operator through a generalization of the Feynman-Kac formula to quantum Markov semi-groups (see [32] and references therein). Further studies and proposals for experimental checks of the quantum Jarzynski identity have unveiled the interplay between measurement, quantum trajectories and stochastic thermodynamics [33–46]. In [47], C. Jarzynski, H. T. Quan and S. Raav studied the semi-classical limit of the two measurements process to study the correspondence between the quantum and the classical definitions of work, and between the corresponding work distributions (see also [48, 49]). The aim of the present work is to revert the logic and to derive a quantum Jarzynski identity by using the Weyl representation of quantum mechanics. The advantage of this approach is to restore a classical phase space, allowing us, in the semi-classical regime, to define a pseudo-work along pseudo- trajectories, whose classical limit coincides with the definition traditional work along the classical trajectories. Our 2 semi-classical definition of the work does not require the system to be closed and can be associated with a continuously measuring environment, such as modeled by Lindblad type equations, for which nonequilibrium work identities are valid [32]. The outline of this work is the following. In section I, we review the basic properties of the Wigner-Weyl quantization scheme that will be used afterwards. In section II, we consider a quantum system in thermal equilibrium at inverse temperature β governed by a time-independent Hamiltonian: starting from the Wigner transform of the density matrix, we define a ’pseudo-Hamiltonian’ which, in the semi-classical limit, can be viewed as the average of the true classical Hamiltonian over a trajectory in complex time, of duration ∆τ = −i~β. In section III, we study a system with a time-dependent Hamiltonian: we define a pseudo-work, interpret it as the time-integral of the power generated by the pseudo-Hamiltonian over a complex trajectory and show that this pseudo-work satisfies semi-classical Jarzynski identity. This relation is illustrated by an explicit calculation for the harmonic oscillator in section IV. The last section is devoted to concluding remarks.

I. A BRIEF REVIEW OF WEYL QUANTIZATION

In this section, we recall some basic properties of the Wigner-Weyl quantization scheme [50, 51] that we shall use in the present work. Elementary presentations can be found in [52–54] and more advanced discussions in [55, 56]. The Weyl transform allows us to construct an operator from a phase space function f(p, q) . The idea is simply to take the Fourier transform of f and then to take a modified inverse Fourier transform, where the variables p and q are replaced by the operators pb and qb. Literally, one has   1 ZZ i ZZ i (kppb+kq qb) − (kpp˜+kq q˜) fb = 2 e ~ f(˜p, q˜)e ~ dp˜ dq˜ dkp dkq. (4) (2π~) The integral gets simpler in the position representation, and, by applying Baker Campbell Hausdorff and the closure relation,

i i 1 i i 1 (kpp+kq q) kq q kpp kq q e ~ b b = e ~ 2 b e ~ b e ~ 2 b

Z i 1 i i 1 kq q kpp kq q = e ~ 2 b |pie ~ hp| e ~ 2 b dp, (5) one ends up with Z  0 00  0 00 1 i p(q0−q00) q + q hq |fb|q i = e ~ f p, dp. (6) 2π~ 2 h i This relation can be easily inverted, and, from any Hermitian operator Ab, one can define its ”Weyl symbol” Ab (p, q) W by Z h i − i pQ Q Q Ab (p, q) = e ~ hq + |Ab|q − i dQ, (7) W 2 2 which is a function of classical phase space. The Weyl symbol of a Hermitian operator is real, as can be seen easily by taking the complex conjugate of (7). This integral is sometimes called the Wigner transform. When the operator is a density operator, one adds a prefactor for normalization Z h i 1 − i pQ Q Q W (p, q) = Ab (p, q) = e ~ hq + |ρb|q − i dQ, (8) W 2π~ 2 2 Since the transform is one-to-one, the Weyl representation is strictly equivalent to regular quantum mechanics. For operators made of a single variable, it actually respects the ”correspondence principle”. For instance one has

n n [pb ]W (p, q) = p n n [qb ]W (p, q) = q . (9) On the other hand, the Weyl symbol of a product of non-commuting operators is generally not the product of the Weyl symbols of the operators. One has in fact

h i 1 Z 2i [(p1−p)(q2−q)−(p2−p)(q1−q)] AbBb (p, q) = e ~ A(p1, q1)B(p2, q2) dp1 dq1 dp2 dq2. (10) W π2~2 3

For instance one has   h i ~ ∂ pBb (p, q) = p + B(p, q) b 2i ∂q     h i ~ ∂ ~ ∂ qpBb (p, q) = q − p + B(p, q). (11) bb 2i ∂p 2i ∂q

More generally, one has [53]   h i ~ ∂ ~ ∂ AbBb (p, q) = A p + , q − B(p, q). (12) 2i ∂q 2i ∂p Although this product rule implies that products of operators are represented by complicated expressions, this simplifies in the case of symmetrized products of operators. For instance one has

[pq] (p, q) = pq + ~ bb W 2i [qp] (p, q) = pq − ~ , (13) bb W 2i and consequently, pq + qp bb bb (p, q) = pq. (14) 2 W This problem of ordering products disappears as soon as one takes the trace of a product of operators. Indeed, for every couple of operators Ab and Bb and their corresponding Weyl symbols A(p, q) and B(p, q), one has the following fundamental identity: ZZ TrAbBb = A(p, q)B(p, q) dp dq. (15)

We emphasize that the Weyl symbol of a general function of a combination of non-commuting operators is generally not the function of the corresponding Weyl symbol: h  i h i  exp Ab (p, q) 6= exp Ab (p, q) , (16) W W

(unless the operator A depends on a single variable, that is, Ab = f(pb) or Ab = f(qb)). In particular, this implies that one can not obtain a Jarzynski equality in the Weyl representation by simply quantizing the classical Jarzynski proof, which is based on the properties of the exponential function. One of the motivations of the present work is to overcome this difficulty (see in section III).

II. SEMI-CLASSICAL APPROXIMATION OF A THERMAL STATE

In this section, we construct in the classical phase space a ‘pseudo-Hamiltonian’ Γ(p, q), defined from the Weyl symbol of the thermal state generated by the quantum Hamiltonian Hb. The function Γ(p, q) is defined in the following way: h i e−βHb (p, q) ≡ e−βΓ(p,q). (17) W we emphasize again that, because of (16), we have Γ(p, q) 6= H(p, q), where H(p, q) is the classical Hamiltonian. The exact formula Γ is rather complicated but we can derive an approximate expression G(p, q) ' Γ(p, q) in the semi-classical limit, by interpreting the thermal state as a Schr¨odingerpropagator during an imaginary time [57, 58]

−βHb − i ∆τHb e = e ~ , (18) where

∆τ = −i~β (19) 4 is interpreted as an imaginary time. For a real ∆τ, this propagator can be well approximated by the Van Vleck propa- gator [60], which plugged into (8), gives the semi-classical Wigner propagator calculated by M. V. Berry (see equation (21) of [59]). Therefore, using Berry’s result, the semi-classical Wigner thermal state is simply the continuation of this propagator for imaginary ∆τ. In this section, we rederive the expression of the semi-classical Wigner thermal state by the stationary phase method. This will allow us to introduce notations and techniques that will be useful in the rest of this work. Starting from equation (38.30) of [60], we write

1/2 i 1 ∂p i π − (tf−ti)Hb X f Sj (qi,ti;qf,tf)−imj hqf|e ~ |qii ' Ksc(qi, ti; qf, tf) ≡ e ~ 2 , (20) (2iπ )1/2 ∂q j ~ i where Sj is a solution of a Hamilton-Jacobi equation or its time reverse,     ∂Sj ∂Sj ∂Sj ∂Sj + H , qf = 0 − H − , qi = 0, (21) ∂tf ∂qf ∂ti ∂qi

th and coincides with the classical action calculated along one of the j classical trajectories (pj(t), qj(t)) generated by H(p, q), the classical counterpart of Hb, such that   qj(ti) = qi ∂tpj(t) = −∂qH (pj(t), qj(t)) (22) qj(tf) = qf ∂tqj(t) = ∂pH (pj(t), qj(t)) ,

There are a priori several classical trajectories, labelled by j, and the propagator (20) sums over all the trajectories having the right initial and final conditions. The integers mj in equation (20) are the Maslov indices, which count the number of times the jth trajectory crosses a caustic, that is, the set of points where ∂pf is singular [60, 61]. These ∂qi caustics are reached only after a certain amount of time; before that time, there is a unique trajectory for two given boundary conditions, here (qi, ti) and (qf, tf). Thus, we have

Z tf S(qi, ti; qf, tf) = [p(t)∂tq(t) − H (p(t), q(t))] dt ti Z tf = p(t)∂tq(t) dt − (tf − ti)H (p(ti), q(ti)) , (23) ti the last line being true only for a time independent Hamiltonian. We also have, from the fact that S is solution of (21), and see chapter 46 of [62] for the whole story,

 ∂S  ∂S  = −p  = H(p , q )  ∂q i  ∂t i i i i (24)  ∂S  ∂S  = pf  = −H(pf, qf) . ∂qf ∂tf We now take the analytical continuation of this propagator and obtain

−βHb hqf|e |qii ' Ksc(qi, τi; qf, τf) 2 1/2 1 ∂ S i S(qi,τi;qf,τf) ' e ~ , (25) 1/2 (2iπ~) ∂qf∂qi with τi ∈ iR and τf = τi + ∆τ ∈ iR purely imaginary times. We have supposed that all Maslov indices mj vanish: this requires the imaginary time to be small enough to avoid the first caustics of the complex trajectory. Like in regular WKB method, the caustics corresponds to the singularities of the second derivative of the Jacobian of S, and the Maslov indices increase by one every time the trajectory crosses a caustic. It is a consequence of the necessary change of represensation, from (p, q) to the Fourier conjugate variables (ξp, ξq), in order to have a non singular expression of the propagator in the neighbourhood of the caustic. This procedure is more delicate in the complex domain than in the real one, as one needs to chose the ”correct branch” connecting the trajectory from both sides of the caustic. This is the Stokes phenomenon. To obtain a correct ordering of the ~ corrections, the value of ∆τ = τf −τi must remain constant as ~ → 0 [70]. This corresponds to a semiclassical regime at low temperature ( large β ). Although the limit itself, ~ = 0, cannot really 5 be interpreted as a classical regime, since it must occur at 0 temperature which is the realm of quantum regime, we can expect that finite values of ~ and finite values of temperature can be in this intermediate regime. We remark that the initial imaginary ”time” τi ∈ iR is, a priori, a free parameter, since the Hamiltonian is not a function of β. Let us now have a closer look at the action S: we consider the imaginary time classical trajectory (p(τ), q(τ)) generated by H(p, q), such that   q(τi) = qi ∂τ p(τ) = −∂qH (p(τ), q(τ)) (26) q(τf) = qf ∂τ q(τ) = ∂pH (p(τ), q(τ)) , where τ is an imaginary parameter with τ ∈ [τi, τf]. This trajectory is generically unique and the action S (which is now imaginary) is calculated by integrating along this trajectory (26),

Z τf S(qi, τi; qf, τf) = [p(τ)∂τ q(τ) − H (p(τ), q(τ))] dτ τi Z τf = p(τ)∂τ q(τ) dτ − ∆τH (p(τi), q(τi)) . (27) τi We now evaluate (17) by plugging the semi-classical expression (25) in the Wigner transform (8):

h i 1 Z i Q Q −βHb − pQ −βHbt e (p, q) = e ~ hq + |e |q − i dQ W 2π~ 2 2 2 1/2 1 Z 1 ∂ S i [Stot(p,q,Q)] ' e ~ dQ . (28) 1/2 2π~ (2iπ~) ∂qf∂qi For every Q, the total action Q Q S (p, q, Q) = −pQ + S(q − , τ ; q + ; τ ), (29) tot 2 i 2 f defines, implicitly, a classical trajectory (pi, qi) → (pf, qf) such that Q q = q (τ ) = q − i i 2 Q q = q (τ ) = q + . (30) f f 2 In the semi-classical limit, ~ → 0, keeping ∆τ fixed, the stationary phase method can be used; the main contribution in the integral (28) is given by the stationary point Q?, that solves

∂ 1 ∂S  Q? Q?  1 ∂S  Q? Q?  Stot(p, q, Q)|Q? = − q − , τi; q + ; τf + q − , τi; q + ; τf − p = 0. (31) ∂Q 2 ∂qi 2 2 2 ∂qf 2 2 Thus, according to (24), we obtain p + p p = i f . (32) 2 Using equations (30) and (32), we then conclude that among the family of trajectories (p(τ), q(τ)) spanned by Q, we must select the stationary trajectory (p?(τ), q?(τ)) such that p? + p? i f = p (33) 2 q? + q? i f = q, (34) 2 ? ? ? ? ? ? ? ? where (pi , qi ) = (p (τi), q (τi)) and (pf , qf ) = (p (τf), q (τf)). To understand the structure of the solution, let us 0 0 0 0 define the imaginary time flow (p , q ) 7→ (p(τ), q(τ)) = Cτ (p , q ) with (p(τ), q(τ)) solution of equation (26) and (p(0), q(0)) = (p0, q0). Formally, one can write 1 p(τ) = p(0) + τp˙(0) + τ 2p¨(0) + ... 2 1 q(τ) = q(0) + τq˙(0) + τ 2q¨(0) + ... (35) 2 6

? ? (pi , qi )

? ? ? Im(p, q) Re(p) (pf − pi ,Q )

(p, q)

? ? (pc, qc )

? ? (pf , qf )

O Re(q) real plane

? ? ? ? ? FIG. 1: The complex trajectory crosses the real plane at (pc , qc ); its chord (pf − pi ,Q ) is imaginary and the middle of this ? ? chord is (p, q). Also, (p, q) is the image of (pc , qc ) through M ∆τ . 2

We first note that, for real (p(0), q(0)), thenp ˙(0),p ¨(0) . . . are also real, as they can be obtained from (22) and expressed in terms of derivatives of the type ∂pnqm H(p(0), q(0)) where H is real. On the other hand, τ is imaginary. Therefore one then has, for real (p0, q0),

0 0 0 0 Cτ (p , q ) = C−τ (p , q ). (36)

Let us then build the map

0 0 0 0 C ∆τ (p , q ) + C ∆τ (p , q ) 0 0 0 0 2 − 2 (p , q ) 7→ M ∆τ (p , q ) = . (37) 2 2

0 0 From (36), M ∆τ is a real map from the real phase space {(p , q )} to itself. Then, we define 2

−1 ? ?   (p , q ) = M ∆τ (p, q). (38) c c 2

Obviously,

? ? ? ? (p , q ) = C ∆τ (p , q ) i i − 2 c c ? ? ? ? (p , q ) = C ∆τ (p , q ) (39) f f 2 c c then fulfill conditions (33) and (34). This construction makes it clear that

? ? pi = pf ? ? qi = qf , (40)

Q? which implies, from equation (30) with q being real, that 2 is in fact the imaginary part of qf, with

Q? = −Q?. (41)

To summarize, the arc (p?(τ), q?(τ)) is symmetric with regard to the real phase space plane, and it intersects this ? ? ? τi+τf ? τi+τf ? ? ? real phase space plane at (pc , qc ) = (p ( 2 ), q ( 2 )). Moreover, the chord (pf − pi ,Q ) of this arc is purely imaginary, and the middle of this chord is (p, q). The picture is shown on figure 1. Finally, we retrieve for the Weyl symbol, defined in (28), an expression equivalent to the one in [59], but in imaginary time, that is h i e−βHb (p, q) ≡ e−βΓ(p,q) ' N (p, q)e−βG(p,q) (42) W ~→0 7 with 1  Q? Q?  G(p, q) = − −pQ? + S(q − , τ ; q + , τ ) ∆τ 2 i 2 f Z τf  ? ? 1 ? ? ? = H (pc , qc ) − p (τ)∂τ q (τ) dτ − pQ ∆τ τi 1 = H (p?, q?) − A(p, q) (43) c c |∆τ| and the prefactor

1 1 ∂2S? 1/2 r π N (p, q) = 1/2 i 00 ? 2π (2iπ ) ∂qf∂qi − S (Q ) ~ ~ ~ tot 2 ? 1/2 v 1 1 ∂ S i π u 8π~ = e 4 u 1/2 2 ? 2 ? 2 ? 2π (2iπ ) ∂qf∂qi t ∂ S ∂ S ∂ S ~ ~ 2 + 2 − 2 ∂qf ∂qi ∂qf∂qi v u 2 ∂2S? 1 u ∂qf∂qi = u . (44) 2π t ∂2S? 1  ∂2S? ∂2S?  ~ − 2 + 2 ∂qf∂qi 2 ∂qf ∂qi

? R τf ? ? ? Here, S represents the action S evaluated at the saddle-point. The term A(p, q) = i p (τ)∂τ q (τ) dτ − ipQ is τi a real number that can be interpreted as the area between the complex arc (p?(τ), q?(τ)) and its chord, as shown in figure 1. As a consequence, G(p, q) is also a real function. On the other hand, Γ(p, q) is also real since the Weyl representation of any Hermitian operator is real, as can easily be checked from (7). The prefactor (44) is the product of two terms: one arises from the Van Vleck propagator and the other is generated by the stationary phase method (i. e. an imaginary Gaussian integration). In the following, we shall keep the leading order in ~ only and the prefactor N (p, q) will be omitted.

III. A JARZYNSKI IDENTITY IN THE WEYL REPRESENTATION

In the previous section, using a semi-classical approach, we have derived an expression for the function G(p, q) such that e−βG(p,q) represents the quantum thermal state generated by the (quantum) Hamiltonian Hˆ . We shall now introduce an explicit time dependence in the Hamiltonian Ht, define a pseudo-work, calculate its semi-classical expression ∂tGt and derive a formal Jarzynski identity. The heart of the matter resides in the geometric interpretation of the semi-classical trajectories.

A. Time-dependent Hamiltonian

We use the semi-classical scheme constructed in section II, but with a time dependent Hamiltonian Ht(p, q). In this context, it is important to be aware that the ’imaginary time’ τ of the trajectory (26), which is related to the temperature 1/β, has nothing to do with the physical time t. In particular, the physical time t must remain frozen during the imaginary time propagation in (26). In other words, the imaginary time trajectory (pt(τ), qt(τ)) obeys   q(τi) = qi ∂τ p(τ) = −∂qHt (p(τ), q(τ)) (45) q(τf) = qf ∂τ q(τ) = ∂pHt (p(τ), q(τ)) , with a fixed value of t. This also means that S? and Q? and (p?(τ), q?(τ)) are then functions of t.

Remark: We emphasize that the trajectory (45) is not the analytical continuation of the real trajectory generated by Ht(p, q). Indeed, such a continuation (¯p(τ), q¯(τ)) would obey a slightly different equation   q¯(τi) = qi ∂τ q¯(τ) = −∂q¯Hτ (¯p(τ), q¯(τ)) (46) q¯(τf) = qf ∂τ q¯(τ) = ∂p¯Hτ (¯p(τ), q¯(τ)) . 8

Here, the Hamiltonian is changing along the trajectory. For any given value of t, we calculate Gt(p, q), the Van Vleck approximation of the pseudo-Hamiltonian h i  log e−βHbt (p, q) , by using equation (43). W We now define the pseudo-work as ∂tGt(p, q). The expression of this time derivative is actually simpler than the expression of Gt(p, q) itself, as we shall now show. We first consider the time-dependent stationary phase trajectory, ? ? (pt (τ), qt (τ)). Using equation (29), we write  Q? Q?  − ∆τG (p, q) = S (p, q, Q?, t) = −pQ? + S? q − , τ ; q + , τ , (47) t tot t 2 i 2 f where Q? is a function of time t. Taking derivative with respect to time, we obtain after using (31): ∂ ∂ ∂Q? −∆τ∂ G (p, q) = S (p, q, Q?, t) + S (p, q, Q?, t) t t ∂t tot ∂Q? tot ∂t ∂ = S (p, q, Q?, t) ∂t tot ∂  Q? Q?  = S? q − , τ ; q + , τ ∂t t 2 i 2 f ?  Q? Q?  ?  Q? Q?  St+dt q − 2 , τi; q + 2 , τf − St q − 2 , τi; q + 2 , τf = lim . (48) dt→0 dt ? ? We made explicit the latter time derivative so that the reader can remind that St+dt and St actually live on two different stationary phase trajectories, with an implicit t dependence. However, the trajectory for time t + dt can be seen as a fluctuation (p(τ) + δp(τ), q(τ) + δq(τ)) around the trajectory (p(τ), q(τ)) for time t, therefore, 1 Z τf −∆τ∂tGt(p, q) = lim [(p(τ) + δp(τ)) ∂τ (q(τ) + δq(τ)) − Ht+dt (p(τ) + δp(τ), q(τ) + δq(τ)) dt→0 dt τi − p(τ)∂τ q(τ) + Ht (p(τ), q(τ))] dτ , (49) and, because of the stationarity of the action around (p(τ), q(τ)) with the same initial and final positions and times, the first order terms in (δp(τ), δq(τ)) cancels out, and only remains the derivative with respect to the explicit time dependence of Ht, therefore

Z τf 1 ? ? ∂tGt(p, q) = ∂tHt (pt (τ), qt (τ)) dτ. (50) ∆τ τi

To summarize, we have shown that in the semi-classical limit e−βHbt can be approximately represented by a function −βGt(p,q) e in the Weyl space and that the associated work ∂tGt(p, q) is given by a simple expression as an average of ∂tHt(p, q) over a complex trajectory. We note that

lim Gt(p, q) 6= Ht(p, q), (51) ~→0 because, when ∆τ is fixed, the ~ → 0 limit has to be taken together with β → +∞: this is not a classical limit. On the other hand, we do have

lim Gt(p, q) = Ht(p, q), (52) ∆τ→0 In order to have all the ingredients to build a quantum identity which resembles formally to a Jarzynski identity, we need trajectories in the Weyl space (along which the pseudo-work is integrated). We now explain how to construct these trajectories with the help of techniques developed in [63].

B. Trajectories in Weyl space

A key ingredient of the Jarzynski identity [1, 2] is the power ∂tHt(p(t), q(t)) along a classical trajectory (p(t), q(t)), whose integral over t gives the work. The exponential of the Jarzynski work relates the initial distribution Π0(p, q) = −βH0(p,q) e to the distribution at the final time Πt(p(t), q(t)) defined as

−βHt(p(t),q(t)) Πt(p(t), q(t)) ≡ Πt(p(t), q(t)|p, q) = e 9 where (p(t), q(t)) is the image under the Hamiltonian flow of the initial point (p(0), q(0)) = (p, q) in the phase space. −βHbt Γt(p,q) Now, if Πb t = e is the, not normalized, quantum thermal density operator, and if the Weyl symbol of Πb t is e , Γt(p,q) from equation (17), we would like to translate Πt(p(t), q(t)) as a kind of propagation of e in phase space. We can first remark that Γt(p, q), considered as a Hamiltonian, can generate trajectories (pΓ(t), qΓ(t)) in phase space. Then, eΓt(pΓ(t),qΓ(t)) could serve as a backbone for a Jarzynski equality in the Weyl representation, by formally reproducing the initial Jarzynski proof as if Γt(p, q) was a classical Hamiltonian ( see appendix A ). However this formal dynamics in phase space has no connection with the quantum evolution, which seems desirable if we want to obtain a quantum Jarzynski identity with physical meaning. For this reason, in order to find a satisfactory translation of Πt(p(t), q(t)) in the Weyl representation, we will focus on the fact that it is simply the Liouville propagation of the classical distribution Πt(p, q), that is d (Π (p(t), q(t))) = ∂ Π (p(t), q(t)) + ∂ Π (p(t), q(t))∂ p(t) + ∂ Π (p(t), q(t))∂ q(t) dt t t t p t t q t t = ∂tΠt(p(t), q(t)) − ∂pΠt(p(t), q(t))∂qHt(p(t), q(t)) + ∂qΠt(p(t), q(t))∂pHt(p(t), q(t))

= ∂tΠt(p(t), q(t)) − {Ht, Πt} (p(t), q(t)). (53)

† This Liouville propagation will then be translated into the quantum unitary propagation Ubt Πb tUbt of Πb t, and we shall use the fact that any quantum evolution in the Weyl representation can be described, in the semiclassical limit, as a generalized Liouville propagation, see [63]. We can find indeed a trajectory (ˇp(t), qˇ(t)) in phase space, such that

h † i h i Ubt AbUbt (p, q) ' Ab (ˇp(t), qˇ(t)) (54) W W where (ˇp(t), qˇ(t)) tends to the classical trajectory (p(t), q(t)) when ~ → 0. To resume, the solution Πt(p(t), q(t)) of Γt(ˇp(t),qˇ(t)) † (53) will be mapped to e , which is the approximate Weyl representation of Ubt Πb tUbt, solution of

d  †  †   † 1 h i Ubt Πb tUbt = Ubt ∂tΠb t Ubt − Ubt Hbt, Πb t Ubt. (55) dt i~

This path (ˇp(t), qˇ(t)) is defined as in [63], the difference being that eΓt(ˇp(t),qˇ(t)) is interpreted as an imaginary time propagator. The original construction of [63] is thus shifted into the complex domain, and(ˇp(t), qˇ(t)) appears as the middle curve of two complex (nonreal) classical trajectories. † −βHbT We begin with the Weyl symbol of Ub(t0,T ) e Ub(t0,T ). Once again we use the matrix element (20) into the Weyl transform (7), in real time for the real propagation, and in imaginary time for the temperature propagation,

h i Z i Q Q † −βHbT − pQ † −βHbT Ub(t0,T ) e Ub(t0,T ) (p, q) = e ~ hq + |Ub(t0,T ) e Ub(t0,T )|q − i dQ W 2 2

ZZZ i Q Q − pQ † −βHbT = e ~ hq + |Ub(t0,T ) |q1ihq1|e |q2ihq2|Ub(t0,T )|q − i dQdq1dq2 2 2   ZZZ i Q Q − pQ −βHbT = e ~ hq1|Ub(t0,T )|q + i hq1|e |q2ihq2|Ub(t0,T )|q − i dQdq1dq2 2 2

ZZZ i {Stot(Q,q1,q2,T )} ' e ~ dQdq1dq2, (56) where z is the complex conjugate of z, and the total action Stot is now  Q   Q  S (p, q, Q, q , q , t ,T ) = −pQ − S q + , t ; q ,T + S (q , τ ; q ; τ ) + S q − , t ; q ,T , (57) tot 1 2 0 − 2 0 1 T 2 i 1 f + 2 0 2 with τf = τi + ∆τ ∈ iR. Hence the total action is made of the real time action S+ counted positively, the complex time action ST and the real time action S− counted negatively. We warn the reader again that the complex time action is not the analytical continuations of the real time ones, as already mentioned in section III A. Q R T The action S+(q − , t0; q2,T ) = [p+(t)∂tq+(t) − Ht(p+(t), q+(t))] dt is calculated along a regular classical 2 t0 trajectory (p+(t), q+(t)), generated by Ht(p, q) with running t, that is,  Q   q (t ) = q − ∂tp+(t) = −∂qHt (p+(t), q+(t)) + 0 2 (58) ∂tq+(t) = ∂pHt (p+(t), q+(t)) ;  q+(T ) = q2 10

R τf the action ST (q2, τi; q1, τf) = [pT (τ)∂τ qT (τ) − HT (pT (τ), qT (τ))] dτ. is calculated along the classical trajectory τi τ 7→ (pT (τ), qT (τ)) generated by HT (p, q), in which the time dependence T has been frozen. More explicitly, we have   qT (τi) = q2 ∂τ pT (τ) = −∂qHT (pT (τ), qT (τ)) (59) qT (τf) = q1 ∂τ qT (τ) = ∂pHT (pT (τ), qT (τ)) ,

Q R T where τ is imaginary. Finally, S−(q1, t0; q + ,T ) = [p−(t)∂tq−(t) − Ht(p−(t), q−(t))] dt, is calculated along a 2 t0 regular classical trajectory (p−(t), q−(t)), generated by Ht(p, q) with running t, that is,  Q   q (t ) = q + ∂tp−(t) = −∂qHt (p−(t), q−(t)) − 0 2 (60) ∂tq−(t) = ∂pHt (p−(t), q−(t)) .  q−(T ) = q1

 Q  Notice that −S− q + 2 , t0; q1,T can be interpreted both as the opposite of an action with forward trajec-  Q  tory, describing the matrix element hq1|Ub(t0,T )|q + 2 i ; and as a positive action along a backward trajectory,     Q ˜ Q Q † −S− q + 2 , t0; q1,T = S− q1, t0; q + 2 ,T , describing the matrix element hq + 2 |Ub(t0,T ) |q1i. We found that the former way makes calculations easier to follow. We evaluate now expression (56) with stationary phase method, that is, we replace the integral over Q, q1, q2 by a ? ? ? single value of the integrand at the generically unique triple (Q , q1 , q2 ) such that ∂ 1  Q?  1  Q?  S (p, q, Q?, q?, q?, t ,T ) = − ∂ S q + , t ; q?,T − ∂ S q − , t ; q?,T − p = 0 (61) ∂Q tot 1 2 0 2 qi − 2 0 1 2 qi + 2 0 2  ?  ∂ ? ? ? ? ? Q ? Stot(p, q, Q , q1 , q2 , t0,T ) = ∂qi ST (q2 , τi; q1 ; τf) + ∂qf S+ q − , t0; q2 ,T = 0 (62) ∂q2 2  ?  ∂ ? ? ? Q ? ? ? Stot(p, q, Q , q1 , q2 , t0,T ) = −∂qf S− q + , t0; q1 ,T + ∂qf ST (q2 , τi; q1 ; τf) = 0 (63) ∂q1 2 This defines three trajectories,

? ?   Q? ?  ? Q? ? ? • p+(t), q+(t) , associated with S+ q − 2 , t0; q2 ,T , solution of (58) with q+(t0) = q − 2 and q+(T ) = q2 ,

? ? ? ? ? ? ? ? • (pT (τ), qT (τ)), associated with ST (q2 , τi; q1 ; τf), solution of (59) with qT (τi) = q2 and qT (τf) = q1 ,

? ?   Q? ?  ? ? ? Q? • p−(t), q−(t) , associated with S− q + 2 , t0; q1 ,T , solution of (60) with q−(T ) = q1 and q−(t0) = q + 2 , such that, according to equation (61), and using (24), we have

p? (t ) + p? (t ) − 0 + 0 = p 2 ? ? q? (t ) + q? (t ) q + Q + q − Q and − 0 + 0 = 2 2 = q . (64) 2 2 Then, from equation (62) with (24), we obtain

? ? p+(T ) = pT (τi) ? ? ? q+(T ) = qT (τi) = q2 . (65) And equation (63) with (24) imply

? ? p−(T ) = pT (τf) ? ? ? q−(T ) = qT (τf) = q1 . (66) To construct this set of three trajectories, one proceed as in the time independent case. We define the complex time [t] propagator Cτ such that

[t] 0 0 0 0 [t] 0 0  [t] 0 0  C0 (p , q ) = (p , q ) ∂τ Cτ (p , q ) = J ∇Ht Cτ (p , q ) , (67) 11

 0 −1  with J = , and the real time classical propagator R 0 , such that 1 0 t ,t

0 0 0 0 0 0 0 0 Rt0,t0 (p , q ) = (p , q ) ∂tRt0,t(p , q ) = J ∇Ht (Rt0,t(p , q )) . (68)

We already saw in equation (36) of section II that, for real (p0, q0), the positive imaginary time propagation is the complex conjugate of the negative imaginary time propagation, that is,

[t] 0 0 [t] 0 0 C−τ (p , q ) = Cτ (p , q ). (69) On the other hand, for any complex (p00, q00), one has

00 00 00 00 Rt0,t(p , q ) = Rt0,t(p , q ), (70) since the Hamiltonian Ht and times t and t0 are real, that is, the real time propagation conserves the complex conjugate relation between two phase space points. As a consequence, the map M ∆τ , defined by t0,T, 2

−1  [t] 0 0  −1  [t] 0 0  (Rt0,T ) C ∆τ (p , q ) + (Rt0,T ) C ∆τ (p , q ) 0 0 0 0 2 − 2 (p , q ) 7→ Mt ,T, ∆τ (p , q ) = , (71) 0 2 2 is a real map, that is, it maps the real phase space (p0, q0) ∈ R2 to itself. It is then easy to define the inverse image ? ? (p , q ) of (p, q) through map M ∆τ , that is c c t0,T, 2

−1 ? ?   (p , q ) = M ∆τ (p, q), (72) c c t0,T, 2 and, from this point, to define the whole stationary phase trajectory. One has indeed,

? ? [t] ? ? (pT (τi), qT (τi)) = C ∆τ (pc , qc ) − 2 ? ? [t] ? ? (pT (τf), qT (τf)) = C ∆τ (pc , qc ) 2 ? ? −1 ? ? (p+(t), q+(t)) = (Rt,T ) (pT (τi), qT (τi)) ? ? −1 ? ? (p−(t), q−(t)) = (Rt,T ) (pT (τf), qT (τf)). (73) The picture is represented on figure 2. ? ? ? ? ?  ? ?  Notice that, Q , q1 , q2 and, according to the above construction, p+(t), q+(t) and p−(t), q−(t) , are all implicit functions of times t0 and T . In other words, if one changes time t0 or T , the whole construction of the set of three trajectories is different. But, if one changes t0 or T in a infinitesimal way, then the trajectories will only shift infinitesimally. This will simplify further calculations because of the stationarity of the action.

? ? −1 ? ? (p+(t0), q+(t0)) = (Rt0,T ) (pT (τf, qT (τf))

? ? [T ] ? ? (pT (τf), qT (τf)) = C∆τ (pc, qc ) 2

(p, q)

? ? (pc, qc )

? ? −1 ? ? (p−(t0), q−(t0)) = (Rt0,T ) (pT (τi), qT (τi)) ? ? [T ] ? ? (pT (τi), qT (τi)) = C ∆τ (pc, qc ) − 2

? ? ? ? FIG. 2: The application M ∆τ maps (pc , qc ) to M ∆τ (pc , qc ) = (p, q). t0,T, 2 t0,T, 2 12

? ?  ? ?  Remark: although p−(T ), q−(T ) is the propagation of p+(T ), q+(T ) with complex time ∆τ and Hamiltonian HT , ? ?  ? ?  p+(t0), q+(t0) is not equal to the propagation of p−(t0), q−(t0) with complex time −∆τ and Hamiltonian Ht0 , because (59) does not commute with (58), as explained in the paragraph before (46). If they did commute then the whole action would be the integral of an exact 1 form, so it would only depend on initial and final conditions ? ?  ? ?  p+(t0), q+(t0) and p−(t0), q−(t0) , that is, it would be possible to deform the triple trajectory to a simple arc ? ?  ? ?  joining p+(t0), q+(t0) to p−(t0), q−(t0) . See the illustration on figure 3. For additional remarks on this complex structure, see appendix B.

? ? (p+(t0), q+(t0))

? ? (p+(T ), q+(T ))

−∆τ (p, q) ∆τ

? ? (p−(t0), q−(t0)) ? ? (p−(T ), q−(T ))

? ? ? ? ? ? FIG. 3: Whereas (p−(T ), q−(T )) is the propagation of (p+(T ), q+(T )) with complex time ∆τ,(p+(t0), q+(t0)) is not equal to ? ? the propagation of (p−(t0), q−(t0)) with complex time −∆τ.

One has finally h i i ? ? ? ? ? ? ? ? ? ? ? † −βHbT {−p(q (T )−q (T ))+S (q (t0),t0;q (T ),T )+S (q (T ),τi;q (T ),τf)−S (q (t0),t0;q (T ),T )} Ub(t0,T ) e Ub(t0,T ) (p, q) ' e ~ + − + + + T + − − − − . W (74)

(t0) Then we define GT (p, q) by

(t0) ? ? ? − ∆τGT (p, q) = Stot(p, q, Q , q1 , q2 , t0,T ). (75)

(t0) The superscript (t0) in G (p, q) is meant to distinguish the argument of the exponential in h i T h i † −βHbT −βHbT Ub(t0,T ) e Ub(t0,T ) (p, q) from the argument of the exponential in e (p, q), which would sim- W W ply be GT (p, q). (t0) The time evolution of GT (p, q) can be evaluated in the same way as Gt(p, q) in the previous subsection. One can indeed write ? ? ? d ∂ ∂ ∂Q ∂ ∂q1 ∂ ∂q2 Stot = Stot + Stot + Stot + Stot dT ∂T ∂Q ∂T ∂q1 ∂T ∂q2 ∂T ∂ = S ∂T tot ∂ ∂ ∂ = S? + S? + S? , (76) ∂T + ∂T − ∂T T because the Q, q1 and q2 derivatives of Stot are equal to 0 from stationary phase conditions (61), (62) and (63). Then, from (24), one has

∂ S? q? (t ), t ; q? (T ),T  = −H p? (T ), q? (T ) ∂T + + 0 0 + T + + ∂ S? q? (t ), t ; q? (T ),T  = −H p? (T ), q? (T ) (77) ∂T − − 0 0 − T − − 13 whereas, with the same arguments of stationarity used in (49) to prove (50), one has

Z τi ∂ ? ? ? ? ? ST (qT (τi), τi; qT (τf), τf) = − ∂T HT (pT (τ), qT (τ)) dτ. (78) ∂T τi ? ?  Further, the energy is conserved along the imaginary time trajectory (59), therefore HT p+(T ), q+(T ) = ? ?  HT p−(T ), q−(T ) , keeping in mind (65) and (66). Therefore only remains

Z τf d  (t0)  1 ? ? GT (p, q) = ∂T HT (pT (τ), qT (τ)) dτ, (79) dT ∆τ τi ? ? We note that it is the same expression as (50), except that the complex arc (pT (τ), qT (τ)) has been propagated, so (p, q) is no longer the center of its chord. Let us then define (ˇp(T ), qˇ(T )) as the center of the chord to the imaginary ? ? arc {(pT (τ), qT (τ)) , τ ∈ [τi, τf]}, that is, remembering (65) and (66), p? (τ ) + p? (τ ) p? (T ) + p? (T ) pˇ(T ) = T i T f = + − 2 2 q? (τ ) + q? (τ ) q? (T ) + q? (T ) qˇ(T ) = T i T f = + − . (80) 2 2 (ˇp(T ), qˇ(T )) is then our ”pseudo classical trajectory”, it is like the real shadow of the couple of complex Hamiltonian ? ? ? ? trajectories (p+, q+) and (p−, q−), as is illustrated on figure 4.

? ? (p+(t0), q+(t0))

Im(p, q) Re(p) ? ? ? ? (p+(T ), q+(T )) = (pT (τi), qT (τi)) (p, q) = (pˇ(t0), qˇ(t0))

(p?, q?) ? ?  c c (ˇp(T ), qˇ(T )) p−(t0), q−(t0)

? ? ? ?  (pT (τf), qT (τf)) = p−(T ), q−(T ) O Re(q) real plane

d (t0) FIG. 4: Whereas ∂tGt(p, q) is an average over a complex arc whose middle of the chord is (p, q), like in figure 1, dT [GT (p, q)] is averaged over a complex arc whose middle is the propagation of (p, q) along a ”pseudo classical trajectory”.

Hence, we have d   G(t0)(p, q) = ∂ G (ˇp(T ), qˇ(T )). (81) dT T T T Equation (81) actually looks like a backward Liouville propagation of the function defined by (50). The difference is that, in the Liouville case, (p(T ), q(T )) is the classical propagation of (p, q), whereas here, (ˇp(T ), qˇ(T )) is the average of two classical propagations. We have lim (ˇp(T ), qˇ(T )) = (p(T ), q(T )) . (82) ∆τ→0

C. A semi-classical Jarzynski identity

We define the semi-classical work along a trajectory (ˇp(t), qˇ(t)) as Z T

Wt0,T (p, q) = ∂tGt(ˇp(t), qˇ(t)) dt. (83) t0 14

According to (81) one has

Z T d (t0) Wt0,T (p, q) = [Gt (p, q)] dt t0 dt

(t0) = GT (p, q) − Gt0 (p, q) (84) that is, the pseudo-work is the difference between the final and initial pseudo-energies. This allows us to follow the steps of the original Jarzynski proof in [1]. Let Gt(p, q) be the pseudo-Hamiltonian associated with Ht(p, q), and ZZ −βGt(p,q) −βFsc(t) Zsc(t) = e dpdq ≡ e .

The average of the exponential of the pseudo-work W, in the sense defined by the expression (15), is given by ZZ −βWt ,T −βW(p,q) he 0 isc = ρT (p, q) e dp dq

ZZ −βGt0 (p,q) R T e −β ∂tGt(ˇp(t),qˇ(t)) dt = e t0 dp dq Zsc(t0) −βG (p,q)   ZZ e t0 −β G(t0)(p,q)−G (p,q) = e T t0 dp dq Zsc(t0) ZZ 1 −βG(t0)(p,q) = e T dp dq Zsc(0) Z (T ) = sc Zsc(t0) = e−β∆Fsc . (85) with

ZZ (t0) ZZ (t0) h i h i −βG (p,q) −βΓ (p,q) † −βHbT −βHbT Zsc(T ) = e T dp dq ' e T dp dq = Tr Ub(t0,T ) e Ub(t0,T ) = Tr e ~→0 ZZ ZZ h i −βGt (p,q) −βΓt (p,q) −βHbt Zsc(t0) = e 0 dp dq ' e 0 dp dq = Tr e 0 . (86) ~→0 Finally, we have obtained a Jarzynski identity in the semi-classical limit

ZZ −βGt (p,q) e 0 −βW (p,q) Zsc(T ) e t0,T dp dq = (87) Zsc(t0) Zsc(t0)

R T where the pseudo-work Wt ,T (p, q) = ∂tGt(ˇp(t), qˇ(t)) dt is evaluated along the pseudo-trajectory (ˇp(t), qˇ(t)) which 0 t0 −βG (p,q) starts at (p, q), with a probability given by the thermal state e t0 . The average of the exponential of this Z(t0) pseudo-work then gives the ratio of the quantum partition functions in the semi-classical limit.

IV. A SOLVABLE EXAMPLE: THE HARMONIC OSCILLATOR

It is useful to illustrate formal non-equilibrium identities on some specific systems. In the case of the quantum work relations, exactly solvable models are rare and mostly limited to the harmonic oscillator, non-interacting quantum gases, or two-level systems [64–69]. Here, we shall apply the formulas of the previous sections to the harmonic oscillator, whose classical Hamiltonian is defined by 1 1 H(p, q) = p2 + mω2q2 (88) 2m 2 and the corresponding quantum Hamiltonian is

2 ~ 2 1 2 2 Hb = − ∂ + mω q . (89) 2m q 2 15

A. Exact canonical distribution in Weyl representation

In order to calculate the canonical distribution in Weyl representation, we decompose the density operator of the thermal state in the Fock states |ni, and then use the Weyl representation of a projector on a Fock state, |nihn|, which is known to be [53, 56]:

n   (−1) − 2 H(p,q) 4 Wn(p, q) = e ~ω Ln H(p, q) . (90) π~ ~ω Then, we have X e−βHb = e−β~ω(n+1/2)|nihn|. (91) n It is also useful to introduce X eβHb = eβ~ω(n+1/2)|nihn|. (92) n

By virtue of the linearity of the Weyl transform, the Weyl symbol of e−βHb is given by

h −βHb i X −β ω(n+1/2) e (p, q) = e ~ Wn(p, q) W n n   X −β ω(n+1/2) (−1) − 2 H(p,q) 4 = e ~ e ~ω L H(p, q) , (93) π n ω n ~ ~ where we recognize the generating function of the Legendre polynomials,

− YX X e 1−X XnL (Y ) = , (94) n 1 − X n from which we deduce that     h −βHb i 1 2 β~ω e (p, q) =   exp − tanh H(p, q) . (95) W β~ω ~ω 2 2π~ cosh 2

B. Canonical distribution from the stationary phase approximation

Dealing with a harmonic oscillator, we expect the stationary phase method to be exact. The imaginary time ? ? trajectory can simply be obtained. Starting from the real intersection of the complex arc with the real plane, (pc , qc ), it is given by

? ? ? p (τ) = pc cos ωτ − mωqc sin ωτ p? q?(τ) = q? cos ωτ + c sin ωτ, (96) c mω ∆τ ∆τ Hence, from (27), where the above choice of origin implies that τi = − 2 and τf = 2 , we obtain (p?)2 1  sinh ω β S = −i c − mω2(q?)2 ~ 2m 2 c ω imω cosh ( βω)((q?)2 + (q?)2) − 2q?q? = ~ f i f i (97) 2 sinh (~βω) On the other hand, from the relation

p?( ∆τ ) + p?(− ∆τ )  ∆τ   ∆τ  pQ? = 2 2 q? − q? − , (98) 2 2 2 16 we deduce  ? 2  i ? ? 1 (pc ) 1 2 ? 2 (S − pQ ) = − sinh (ω~β) + mω (qc ) . (99) ~ ~ω 2m 2 Noting that

p?(− ∆τ ) + p?( ∆τ ) ω β p = 2 2 = p? cosh ~ 2 c 2 q?(− ∆τ ) + q?( ∆τ ) ω β q = 2 2 = q? cosh ~ , (100) 2 c 2 we conclude, from (43) and (44), that

    −βG(p,q) 1 2 β~ω e =   exp − tanh H(p, q) (101) ω~β ~ω 2 2π~ cosh 2 and we recover expression (95).

C. Quantum power and pseudo-work

The Jarzynski power of the Hamiltonian (88) with time dependent ω is given by

2 ∂tHt = mω˙ tωtq . (102)

Applying equation (50), we obtain

∆τ Z 2 1 ? 2 ∂tGt(p, q) = mωtω˙ t (q (τ)) dτ ∆τ ∆τ − 2 ~β Z − 2  ? 2  1 ? 2 0 2 (pc ) 0 2 2i ? ? 0 0 0 = − mωtω˙ t (qc ) cosh (ωtt ) − 2 2 sinh (ωtt ) + pc qc cosh (ωtt ) sinh (ωtt ) dt β ~β m ω mωt ~ 2 t    ? 2   1 ? 2 sinh (ωt~β) ~β (pc ) sinh (ωt~β) ~β = mωtω˙ t (qc ) + − 2 2 − , (103) ~β 2ωt 2 m ωt 2ω 2 which, using equation (100), leads to

   2   ω˙ t 2 1 2 2 sinh (β~ωt) p sinh (β~ωt) ∂tGt(p, q) = mωt q + 1 + 1 − . (104) ωt 1 + cosh β~ωt 2 (β~ωt) 2m β~ωt

One can verify that

 2 ? 2 lim ∂tGt(p, q) = mωtω˙ t lim qc = mωtω˙ tq = ∂tHt(p, q) (105) β~→0 β~→0 Remark: we note that the pseudo-trajectory (ˇp(t), qˇ(t)) actually coincides with the classical trajectory (p(t), q(t)) because of the linearity of the dynamics of the harmonic oscillator with respect to the initial conditions,

 p(t)   a(t) b(t)   p(0)  = . (106) q(t) c(t) d(t) q(0)

V. CONCLUSION

We have derived a formal Jarzynski identity in the Weyl representation, whose classical limit is the celebrated nonequilibrium work identity for the corresponding classical Hamiltonian. This formal identity involves the average of 17 the exponential of a pseudo-work ∂tGt along pseudo-trajectories which start from initial phase space points distributed according to a thermal distribution. This average is shown to be equal to the ratio of the partition functions of the semi-classical thermal distribution which is approximately equal to the ratio of the quantum partition functions in the semi-classical regime – as long as the semi-classical Van Vleck propagator can be considered to be valid approximation. In other words, we have obtained a semi-classical estimate of the quantum free energy from a ~ correction of the Jarzynski identity which involve only classical quantities. This relation has been verified for the quantum harmonic oscillator. The logical path of the present work can be summarized by the following diagram, which explains that we have combined the Weyl quantization scheme with a semi-classical (i.e. stationary phase) calculation and draws a parallel between the classical derivation of Jarzynski and its mirror image in the Weyl quantum phase space:

Work t Jarzynski −βH0(p,q) −βHt(p,q) R 0 0 0 RR e −βWt(p,q) Zcl(t) e −−−−→ Wt(p, q) = ∂tHt0 (p(t ), q(t )) dt −−−−−−→ e dpdq = 0 Zcl(0) Zcl(0)   yQuant. e−βHbt   yWeyl (107) e−βΓt(p,q)   ysemiclassical

Work t Jarzynski −βG0(p,q) −βGt(p,q) R 0 0 0 RR e −βWt(p,q) Zsc(t) e −−−−→ Wt(p, q) = ∂tGt0 (ˇp(t ), qˇ(t )) dt −−−−−−→ e dpdq = 0 Zsc(0) Zsc(0) A natural extension of this study would be to use the formal identity in the Weyl space to calculate quantum corrections to the classical Jarzynski formula – order by order with respect to ~ – and to find some geometric interpretation of these corrections. It would also be interesting to compare the pseudo-work defined here with the work operator defined for Lindblad equations [32], in the markovian framework of an open system constantly monitored by its environment.

Appendix A: Other trajectories

The approach we followed led us to define a pseudo-Hamiltonian Γt(p, q), whose semiclassical approximation is Gt(p, q). It might have seemed natural to use this pseudo-Hamiltonian as a proper classical-like Hamiltonian, and generate real classical trajectories (pr(t), qr(t)) directly from Hamilton equations ∂Γ p˙ (t) = − t (p (t), q (t)) r ∂q r r ∂Γ q˙ (t) = t (p (t), q (t)). (A1) r ∂p r r From the structure of these equations, we would then automatically have d ∂ Γ (p (t), q (t)) = Γ (p (t), q (t)), (A2) dt t r r ∂t t r r which is enough to derive a Jarzynski identity, by following exactly the classical proof for a Hamiltonian closed system as presented in [1]. It would give simply, by using the same argument of symplecticity of trajectories,

−βΓ (p (t ),q (t )) Z e 0 r 0 r 0 R T ∂ 1 Z −β ∂t Γt(pr (t),qr (t)) dt −βΓ0(pr (T ),qr (T )) e t0 dpr(t0)dqr(t0) = e dpr(t0)dqr(t0) Z(t0) Z(t0) 1 Z −βΓ0(pr (T ),qr (T )) = e dpr(T )dqr(T ) Z(t0) Z(T ) = , (A3) Z(t0) where the second line uses the fact that the Jacobian of (pr(t0), qr(t0)) 7→ (pr(T ), qr(T )) is equal to 1, because of symplecticity. 18

We didn’t retain this option because, although the proof is formally correct, we could not find any quantum signification to these trajectories in terms of the Weyl representation of a quantum evolution. On the other hand, our pseudo-trajectory (ˇp(t), qˇ(t)) naturally describes the Weyl representation of the quantum evolution of an operator, in the semiclassical limit. We have indeed,

h † i Ub(t0,T ) AbUb(t0,T ) (p, q) = A(ˇp(T ), qˇ(T )). (A4) W

p2 1 2 2 The example of the harmonic oscillator Ht(p, q) = 2m + 2 mωt q is quite suggestive here. The pseudo-Hamiltonian 1 β~ω Γt(p, q) is just the proper Hamiltonian multiplied by λ = tanh , and therefore the classical-like trajectory ~ω 2 generated by Γt(p, q) is a classical trajectory of the Harmonic oscillator, but where the time has been scaled by λ. This obviously comes from the Hamilton equations, ∂H p˙ (t) = −λ t (p (t), q (t)) r ∂q r r ∂H q˙ (t) = λ t (p (t), q (t)), (A5) r ∂p r r which can be written ∂p ∂H r = − t (p (λt), q (λt)) ∂(λt) ∂q r r ∂q ∂H r = t (p (λt), q (λt)). (A6) ∂(λt) ∂p r r On the other hand, the pseudo-trajectory (ˇp(t), qˇ(t)) is exactly the classical trajectory (p(t), q(t)) with the correct time t, ∂H p˙(t) = − t (p(t), q(t)) ∂q ∂H q˙(t) = t (p(t), q(t)), (A7) ∂p which suggests that this choice is indeed more natural.

Appendix B: Complex symplectic structure

The complex structure of the trajectories is made of two types of trajectories, the imaginary ’formal time’ trajectories ? ? ? ?  ? ?  (pT (τ), qT (τ)), and the real time ones, p+(t), q+(t) and p−(t), q−(t) . They do not commute, therefore the mixed complex phase space trajectories that we consider cannot be considered as the analytical continuation of classical phase space (p(t), q(t)). Obviously, freezing the time real t of the Hamiltonian Ht(p, q) and propagating through Hamiltonian dynamics with complex formal time, like for instance (59) if one restricts to purely imaginary formal time, will generate trajectories whose reunion is the manifold Ht(p, q) = E, where E is a complex constant. For instance, in the scheme of FIG. 4, the ? ? ? ? ? ? energy Ec = Ht0 (pT (τ), qT (τ)), along the arc which goes from (pT (τi), qT (τi)) to (pT (τf), qT (τf)), is a real constant, ? ? because the arc crosses the real plane (Re (p), Im (q)) at (pc , qc ), and the Hamiltonian is real. ? ?  On the other hand, the energy is not conserved any more through the real time propagations p+(t), q+(t) and ? ?  ? ?  ? ?  p−(t), q−(t) , driven by (58). Still, if we define E+(t) = Ht p+(t), q+(t) and E−(t) = Ht p−(t), q−(t) , then we have E+(t) = E−(t), that is, both energies are complex conjugate. Also, we have E+(t0) = E−(t0) = Ec ∈ R, that is, the energy is initially real at t = t0, then they draw two complex conjugate branches until t = T . The midpoint of the final tips of these two complex conjugate branches is the real phase space point (p, q) where the Weyl function is evaluated (see FIG. 4). ? ?  Let us follow for instance p+(t), q+(t) , and then let us freeze the time t at a later value t = t1 > t0. Then, ? ?  propagating from p+(t1), q+(t1) with complex formal time would generate a new set of trajectories whose reunion is  2 ? ?  the manifold (p, q) ∈ C such that Ht1 (p, q) = Ht1 p+(t1), q+(t1) = E+(t1) , where E+(t1) ∈ C. Since Ht1 (p, q) 6=

Ht0 (p, q), nothing prevents this new manifold to intersect the initial Ec manifold. To resume, for a fixed time t, the whole two dimensional complex phase space (p, q) ∈ C2 can be orga- nized as a foliation Ft of one complex dimensional manifolds. Each one of this complex manifold is stable 19 through the Hamiltonian dynamics generated by Ht with frozen time dependence t, and it can be identified with  2 (p, q) ∈ C such that Ht(p, q) = E , with some E ∈ C, so we may call it Lt(E). The reunion of all the Lt(E) for all 2 values of E gives back C . In particular, every imaginary formal time dynamics (59) lives on such a manifold Lt(E). However, changing time t changes the whole foliation. As a consequence, a real time trajectory (p(t), q(t)), defined by (58), which crosses a leave Lt(E) at time t, will then cross an other leave Lt+dt(E + dE) at time t + dt, with dE defined by d dE = H (p(t), q(t)) dt dt t ∂ = H (p(t), q(t)) dt, (B1) ∂t t

dE so dt is a function of the trajectory (p(t), q(t)). Although Lt(E + dE) does not intersect Lt(E), Lt+dt(E + dE) can a priori intersect Lt(E) at points (pa, qa) solutions of ( H (p , q ) = E + dE t+dt a a (B2) Ht (pa, qa) = E, that is ( Ht (pa, qa) = E ∂ dE ∂ (B3) ∂t Ht (pa, qa) = dt = ∂t Ht(p(t), q(t)).

For instance, if Ht(x) = H0(x) + F t · x, where x = (p, q), then these equations become ( H (x ) = E t a (B4) F˙ t · xa = F˙ t · x(t), which implies that xa is at the intersection between the complex line x(t) + aJ F˙ t, a ∈ C and the surface of constant energy E. In other words, xa = x(t) + aJ F˙ t, where a ∈ C is such that   Ht x(t) + aJ F˙ t = E. (B5)

As soon as Ht is quadratic or more, there can be other solutions than a = 0. A natural symplectic structure can be attached to each manifold Lt(E) by using Hamilton’s equations (59), gen- eralized to arbitrary complex τ, as Schwartz identities. On the other hand, the real time dynamics, driven by (58), can be supplied with an other symplectic structure, adding (E, t) as an extra couple of canonical variables to (p, q), and this structure can also be extended to complex (E, t). However, how to connect both symplectic structures is not clear to us by now.

Acknowledgments

The authors would like to thank D. Ullmo, S. Tomsovic and R. Sopracase for useful discussions and comments. We also thank the referees for very stimulating remarks.

[1] C. Jarzynski, Nonequilibrium equality for free energy differences, Phys. Rev. Lett., 78, 2690 (1997). [2] C. Jarzynski, Equilibrium free-energy differences from nonequilibrium measurements: A master-equation approach, Phys. Rev. E 56, 5018 (1997). [3] G. E. Crooks, Entropy production fluctuation theorem and the nonequilibrium work relation for free energy differences, Phys. Rev. E, 60, 2721 (1999). [4] C. Jarzynski, Nonequilibrium work relations: foundations and applications, Eur. Phys. J. B 64, 331 (2008). [5] C. Jarzynski, Equalities and Inequalities: Irreversibility and the Second Law of Thermodynamics at the Nanoscale, Ann. Rev. Cond. Mat. Phys, 2, 329 (2011). [6] U. Seifert, Stochastic thermodynamics, fluctuation theorems and molecular machines, Rep. Prog. Phys. 75, 126001 (2012). 20

[7] J. Liphardt, S. Dumont, S. B. Smith, I. Tinoco and C. Bustamante, Equilibrium information from nonequilibrium mea- surements in an experimental test of Jarzynski’s equality, Science, 296, 1832 (2002). [8] D. Collin, F. Ritort, C. Jarzynski, S. B. Smith, I. Tinoco and C. Bustamante, Verification of the Crooks fluctuation theorem and recovery of RNA folding free energies, Nature 437 231 (2005). [9] F. Douarche, S. Ciliberto, A. Petrosyan and I. Rabbiosi , An experimental test of the Jarzynski equality in a mechanical experiment, Eur. Phys. Lett 70, 593 (2005). [10] J. Kurchan, A quantum fluctuation theorem, arXiv:cond-mat/0007360v2 (2000). [11] H. Tasaki, Jarzynski relations for quantum systems and some applications, arXiv:cond-mat/0009244v2 (2000). [12] S. Mukamel, Quantum extension of the Jarzynski relation: Analogy with stochastic dephasing, Phys. Rev. Lett. 90, 170604 (2003); M. Esposito and S. Mukamel, Fluctuation theorems for quantum master equations, Phys. Rev. E 73, 046129 (2006). [13] W. de Roeck and C. Maes, Quantum version of free-energyirreversible-work relations, Phys. Rev. E 69, 026115 (2004). [14] C. Jarzynski and D. K. W´okcik, Classical and Quantum Fluctuation Theorems for Heat Exchange, Phys. Rev. Lett. 92, 230602 (2004). [15] D. Andrieux and P. Gaspard, Quantum Work Relations and Response Theory, Phys. Rev. Lett. 100, 230404 (2008); D. Andrieux, P. Gaspard, T. Monnai and S. Tasaki, for currents in open quantum systems arXiv:0811.3687 (2008) New J. Phys. 11, 043014 (2009). [16] D. Cohen and Y. Imry, Straightforward quantum-mechanical derivation of the Crooks fluctuation theorem and the Jarzynski equality, Phys. Rev. E 86, 011111 (2012). [17] Zhaoyu Fei, H. T. Quan, and Fei Liu, Quantum corrections of work statistics in closed quantum systems, Physical Review E 98, 012132 (2018). [18] Yixiao Qian and Fei Liu, Computing characteristic functions of quantum work in phase space, Physical Review E 100, 062119 (2019). [19] Ken Funo and H. T. Quan, Path Integral Approach to Quantum Thermodynamics, Phys. Rev. Lett. 121, 040602 (2018). [20] P. Talkner, E. Lutz and P. H¨anggi, Fluctuation theorems: Work is not an observable, Phys. Rev. E 75, 050102(R) (2007). [21] M. Campisi, P. Talkner and P. H¨anggi, Thermodynamics and Fluctuation Theorems for a Strongly Coupled Open Quantum System: An Exactly Solvable Case, J. Phys. A 42, 392002 (2009); M. Campisi, P. Talkner, and P. H¨anggi, Influence of measurements on the statistics of work performed on a quantum system, Phys. Rev. E 83, 041114 (2011); M. Campisi , P. Talkner and P. H¨anggi, Quantum BochkovKuzovlev work fluctuation theorems, Phil. Trans. R. Soc. A 369, 291 (2011). [22] S. Yukawa, A quantum analogue of the Jarzynski equality, J. Phys. Soc. Jap. 69, 2367 (2000). [23] A. E. Allahverdyan and Th. M. Nieuwenhuizen, Fluctuations of work from quantum subensembles: The case against quantum work-fluctuation theorems, Phys. Rev. E 71, 066102 (2005). [24] H. J. D. Miller and J. Anders, Time-reversal symmetric work distributions for closed quantum dynamics in the histories framework, New J. Phys. 19, 062001 (2016). [25] M. Esposito, U. Harbola, and S. Mukamel, Nonequilibrium fluctuations, fluctuation theorems, and counting statistics in quantum systems, Rev. Mod. Phys. 81, 1665 (2009). [26] M. Campisi, P. H¨anggiand P. Talkner, Colloquium : Quantum Fluctuation Relations : Foundations and Applications, Rev. Mod. Phys. 83, 771 (2011). [27] J. Aberg, Fully Quantum Fluctuation Theorems, Phys. Rev. X 8, 011019 (2018). [28] G. E. Crooks, Quantum Operation Time Reversal, Phys. Rev. A 77 034101 (2008). [29] G. E. Crooks, On the Jarzynski relation for dissipative quantum dynamics, J. Stat. Mech. P10023, (2008). [30] G. Manzano, J. M. Horowitz, and J. M. R. Parrondo, Nonequilibrium potential and fluctuation theorems for quantum maps, Phys. Rev. E 92, 032129 (2015). [31] S. Haroche and J.-M. Raymond, Exploring the Quantum, Oxford Univ.Press (2006). [32] R. Chetrite and K. Mallick, Quantum Fluctuation Relations for the Lindblad Master Equation, J. Stat. Phys. 148, 480 (2012). [33] G. Huber, F. Schmidt-Kaler, S. Deffner and E. Lutz Employing Trapped Cold Ions to Verify the Quantum Jarzynski Equality, Phys. Rev. Lett. 101, 070403 (2008). [34] J. P. Garrahan and I. Lesanovsky, Thermodynamics of Quantum Jump Trajectories, Phys. Rev. Lett. 104, 160601 (2010). [35] S. Deffner and E. Lutz, Nonequilibrium Entropy Production for Open Quantum Systems, Phys. Rev. Lett. 107, 140404 (2011). [36] J. Teifel and G. Mahler, Autonomous modular quantum systems: Contextual Jarzynski relations, Phys. Rev. E 83, 041131 (2011). [37] R. Kosloff, Quantum Thermodynamics, Entropy 15, 2100 (2013). [38] A. J. Roncaglia, F. Cerisola and J. P. Paz, Work Measurement as a Generalized Quantum Measurement, Phys. Rev. Lett. 113, 250601 (2014). [39] T. B. Batalh˜ao,A. M. Souza, R. S. Sarthour, I. S. Oliveira, M. Paternostro, E. Lutz, and R. M. Serra, Irreversibility and the Arrow of Time in a Quenched Quantum System, Phys. Rev. Lett. 115, 190601 (2015). [40] P. Talkner and P. H¨anggi, Open system trajectories specify fluctuating work but not heat, Phys. Rev. E 94, 022143 (2016). [41] R. Sampaio, S. Suomela and T. Ala-Nissila, Calorimetric measurement of work for a driven harmonic oscillator, Phys. Rev. E 94, 062122 (2016). [42] C. Elouard, D. A. Herrera-Mart´ı,M. Clusel and A. Auff`eves The role of quantum measurement in stochastic thermody- namics, Nature Quantum Information 3, 9 (2017). [43] S. Vinjanampathy and J. Anders, Quantum thermodynamics, Contemp. Phys. 57, 545 (2016). [44] A. E. Rastegin, Non-equilibrium equalities with unital quantum channels, J. Stat. Mech. P06016 (2013). 21

[45] A. E. Rastegin and K. Zyczkowski, Jarzynski equality for quantum stochastic maps, Phys. Rev. E 89, 012127 (2014). [46] A. Smith et al, Verification of the quantum nonequilibrium work relation in the presence of decoherence, New J. Phys. 20, 013008 (2018). [47] C. Jarzynski, H. T. Quan and S. Rahav, The quantum-classical correspondence principle for work distributions, Phys. Rev. X 5, 031038 (2015). [48] I. Garca-Mata, A. J. Roncaglia and D. A. Wisniacki, Quantum-to-classical transition in the work distribution for chaotic systems, Physical Review E 95(5), 050102 (2017). [49] I. Garca-Mata, A. J. Roncaglia and D. A. Wisniacki, Semiclassical approach to the work distribution, EPL (Europhysics Letters) 120(3), 30002 (2017). [50] E. Wigner, On the quantum correction for thermodynamic equilibrium, Phys. Rev. 40, 749 (1932). [51] H. Weyl and H. Robertson, Theory of Groups and Quantum Mechanics, (Dover, New York, 1931). [52] W. B. Case, Wigner functions and Weyl transforms for pedestrians, Am. J. Phys. 76, 937 (2008). [53] T. L. Curtright, D. B. Fairlie, C. K. Zachos, A Concise Treatise on Quantum Mechanics in Phase Space, (World Scientific, 2014). [54] V. S. Varadarajan, Reflections on Quanta, Symmetries and Supersymmetries, (Springer, New York, 2011). [55] M. V. Berry, Semi-classical mechanics in phase space: a study of Wigners function Phil. Trans. R. Soc. A. 287, 237 (1977). [56] A.-M. Ozorio de Almeida, the Weyl representation in classical and quantum mechanics, Phys. Rep. 295 265 (1998). [57] J. W. Negele, H. Orland, Quantum Many-particle Systems, (Perseus, Advanced Books Classics, 1998). [58] P. Cartier and C. De Witt-Morette, Functional Integration Action and Symmetries, (Cambridge University Press, 2006). [59] M. V. Berry, Quantum scars of classical closed orbits in phase space, Proc. R. Soc. Lond. A 423, 219 (1989). [60] Predrag Cvitanovi´c, Roberto Artuso, Ronnie Mainieri, Gregor Tanner, Gbor Vattay, Classical and Quantum Chaos, http://chaosbook.org/chapters/ChaosBook.pdf. [61] V. P. Maslov and M. V. Fedoriuk, Semiclassical approximation in quantum mechanics (Reidel, Dordrecht, 1981). [62] V. I. Arnold, Mathematical methods in classical mechanics (Springer, Verlag, 1978). [63] P. P. de M. Rios and A.-M. Ozorio de Almeida, the Weyl representation in classical and quantum mechanics, J. Phys. A 35 307 (2002). [64] C. Jarzynski and H. T. Quan, Validity of nonequilibrium work relations for the rapidly expanding quantum piston, Phys. Rev. E 85, 031102 (2012). [65] F. Liu, Calculating work in adiabatic two-level quantum Markovian master equations: A characteristic function method, Phys. Rev. E 90, 032121 (2014). [66] J. Teifel and G. Mahler, Model studies on the quantum Jarzynski relation, Phys. Rev. E 76, 051126 (2007); H. Schr¨oder, J. Teifel, and G. Mahler, Work and work fluctuations in quantum systems, Eur. Phys. J. Special Topics 151, 181 (2007). [67] S. Deffner and E. Lutz, Nonequilibrium work distribution of a quantum harmonic oscillator, Phys. Rev. E 77, 021128 (2008). [68] P. Talkner, P. S. Burada and P. H¨anggi, Statistics of work performed on a forced quantum oscillator, Phys. Rev. E 78, 011115 (2008). [69] A. Sindona, J. Goold, N. Lo Gullo, F. Plastina, Statistics of the work distribution for a quenched Fermi gas, New J. Phys. 16, 1367 (2014). 2 1/2 1 ∂ S [70] This will ensure that the prefactor 1/2 is of higher order in ~ than the exponential of S. (2iπ~) ∂qf∂qi