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

Quantum and Classical Ergotropy From Entropies - Deffner

Download as pdf or txt
Download as pdf or txt
You are on page 1of 13

entropy

Article
Quantum and Classical Ergotropy from Relative Entropies
Akira Sone 1,2,3, * and Sebastian Deffner 4,5

1 Aliro Technologies, Inc., Boston, MA 02135, USA


2 Theoretical Division, Los Alamos National Laboratory, Los Alamos, NM 87545, USA
3 Center for Nonlinear Studies, Los Alamos National Laboratory, Los Alamos, NM 87545, USA
4 Department of Physics, University of Maryland, Baltimore County, Baltimore, MD 21250, USA;
deffner@umbc.edu
5 Instituto de Física ‘Gleb Wataghin’, Universidade Estadual de Campinas, Campinas 13083-859, Brazil
* Correspondence: akira@aliroquantum.com; Tel.: +1-646-468-5399

Abstract: The quantum ergotropy quantifies the maximal amount of work that can be extracted
from a quantum state without changing its entropy. Given that the ergotropy can be expressed as
the difference of quantum and classical relative entropies of the quantum state with respect to the
thermal state, we define the classical ergotropy, which quantifies how much work can be extracted
from distributions that are inhomogeneous on the energy surfaces. A unified approach to treat both
quantum as well as classical scenarios is provided by geometric quantum mechanics, for which we
define the geometric relative entropy. The analysis is concluded with an application of the conceptual
insight to conditional thermal states, and the correspondingly tightened maximum work theorem.

Keywords: ergotropy; geometric quantum mechanics; conditional thermal state



1. Introduction
Citation: Sone, A.; Deffner, S. According to its definition, the adjective ergotropic refers to the physiological mech-
Quantum and Classical Ergotropy anisms of a nervous system to favor an organism’s capacity to expend energy [1]. Gen-
from Relative Entropies. Entropy 2021,
eralizing this notion to physical systems, quantum ergotropy was then coined to denote
23, 1107. https://doi.org/10.3390/
the maximal amount of work that can be extracted by isentropic transformations [2]. In
e23091107
particular, the quantum ergotropy quantifies the amount of energy that is stored in active
quantum states, and which can be extracted by making the state passive [3–6]. In simple
Academic Editor: Ronnie Kosloff
terms, a passive state is diagonal in the energy basis, and its eigenstates are ordered in
descending magnitude of its eigenvalues. Gibbs states are then called completely passive [3].
Received: 2 August 2021
Accepted: 20 August 2021
The quantum ergotropy plays a prominent role in quantum thermodynamics [7]. In
Published: 25 August 2021
particular, when assessing the thermodynamic value of genuine quantum properties [8–11],
such as squeezed and nonequilibrium reservoirs [12,13], coherence [14,15], or quantum
Publisher’s Note: MDPI stays neutral
correlations [16,17], it has proven powerful. However, if the quantum system is not in
with regard to jurisdictional claims in
contact with a heat reservoir, computing the quantum ergotropy is far from trivial. This
published maps and institutional affil- is due to the fact that the ergotropy is determined by a maximum over all unitaries that
iations. can act upon the system [2]. Note that not all passive states can be reached by unitary
operations, in particular, including the completely passive state.
In this paper, given that the quantum ergotropy can be written as the difference of
quantum and classical relative entropies (the Kullback–Leibler divergence of the eigenvalue
Copyright: © 2021 by the authors.
distributions), we define a classical ergotropy, which quantifies the maximal amount of work
Licensee MDPI, Basel, Switzerland.
that can be extracted from inhomogeneities on the energy surfaces, which have been shown
This article is an open access article
to be analogous to quantum coherences [18,19].
distributed under the terms and In a second part of the analysis, we turn to a unified framework, namely geometric
conditions of the Creative Commons quantum mechanics. Exploiting this approach [20–22], we define the geometric relative
Attribution (CC BY) license (https:// entropy. With this, it becomes particularly transparent to characterize the one-time mea-
creativecommons.org/licenses/by/ surement approach to quantum work [23–27]. In this paradigm, work is determined by
4.0/). first measuring the energy of the system, and then letting it evolve under time-dependent

Entropy 2021, 23, 1107. https://doi.org/10.3390/e23091107 https://www.mdpi.com/journal/entropy


Entropy 2021, 23, 1107 2 of 13

dynamics. In contrast to the two-time measurement approach [28–46], no projective mea-


surement is taken at the end of the process. Hence, the work probability distribution is
entirely determined by the statistics conditioned on the initial energy. Here, we identify
the distinct contributions to the thermodynamic cost of projective measurements by sepa-
rating out the coherent and incoherent ergotropies, and the population mismatch in the
conditional statistics.
Hence, by expressing the quantum ergotropy as a difference of relative entropies,
we are able to (i) generalize the notion to classical scenarios, and to (ii) elucidate the
thermodynamics of projective measurements. This analysis further cements ergotropy as
one of the salient pillars of quantum thermodynamics.
The paper is organized as follow. In Section 2, we review quantum ergotropy in terms
of relative entropies and its relation to the quantum coherence in Section 3. Then, we
introduce the formulation of classical ergotropy in Section 4, and discuss the geometric
quantum mechanics approach to ergotropies in Section 5. Finally, we discuss the physical
meaning of the conditional thermal states in the second law of thermodynamics based on
its ergotropy in Section 6 before our conclusions in Section 7.

2. Quantum Ergotropy
We begin by deriving a simple expression for the quantum ergotropy, which does not
explicitly depend on the optimization over unitary maps. To this end, consider a quantum
system with Hamiltonian H and quantum state ρ. Then, the ergotropy is defined as [2]
h n oi
E (ρ) ≡ tr{ρ H } − min tr UρU † H , (1)
U ∈U

where U is the unitary group.


Our goal is now to express Equation (1) as a difference of relative entropies. To this
end, we write the quantum state ρ in its “ordered” eigenbasis
Z
ρ= ∑ pi | pi ih pi | with p i ≥ p i +1 . (2)
i

Let σ be a second quantum state, which we write


Z
σ= ∑ si |si ihsi | with s i ≥ s i +1 . (3)
i

In principle, ρ and σ can be vastly different quantum states. To better compare ρ and
σ, it is then interesting to identify the unitary operation that takes ρ as close as possible to
σ. Hence, considering the quantum relative entropy
n o
S(UρU † ||σ) ≡ tr{ρ ln (ρ)} − tr UρU † ln (σ ) , (4)

it is known that the minimization of the quantum relative entropy over all the unitary
operations is the classical relative entropy [47] (see Appendix A for the proof)
 
pi
h i Z
min S(UρU ||σ ) = ∑ pi ln

≡ D (ρ||σ) . (5)
U ∈U si
i

To this end, we choose σ as the Gibbs state

exp (− βH )
ρeq ≡ with Z ≡ tr{exp (− βH )} . (6)
Z
Entropy 2021, 23, 1107 3 of 13

For the sake of simplicity, we further assume that the eigenenergies are ordered in
ascending magnitude, Ei ≤ Ei+1 . As an alternative expression, the quantum ergotropy can
be expressed as the difference of relative entropies [14,48] (see Appendix B for the proof)

β E (ρ) = S(ρ||ρeq ) − D (ρ||ρeq ) . (7)

Note that the quantum ergotropy does not depend on the specific value of the tem-
perature, but rather Equation (7) holds for any β. In conclusion, the quantum ergotropy is
written as the difference of the quantum and classical relative entropies of the quantum
state ρ with respect to ρeq . Note that Equation (7) is entirely determined by ρ and ρeq , and
independent of any optimization.

3. Ergotropy from Quantum Coherence


It was recently recognized [14,15,17] that the quantum ergotropy (1) can be separated
into two fundamentally different contributions

E ( ρ ) = Ei ( ρ ) + E c ( ρ ) . (8)

The incoherent ergotropy Ei (ρ) denotes the maximal work that can be extracted from ρ
without changing its coherence, which is defined as [14]

Ei (ρ) ≡ tr{(ρ − τ ) H } . (9)

Here, we call τ the coherence-invariant state of ρ, which is defined as [14]


n o
tr{τH } = min tr UρU † H , (10)
U ∈U (i)

where U (i) is the set of unitary operations without changing the coherence of ρ. Refer to
ref. [14] for more details about U (i) .
The coherent ergotropy Ec (ρ) is the work that is exclusively stored in the coherences.
This can be quantified by the relative entropy of coherence [49]

C(ρ) = H(L(ρ)) − H(ρ) , (11)

where H(ρ) ≡ −tr{ρ ln (ρ)} is the von Neumann entropy of ρ, and L is the purely dephas-
ing map, i.e., the map that removes all coherences but leaves the diagonal elements in the
energy basis invariant. From the expression of the coherent ergotropy derived in ref. [14]
and Equation (5), the coherent ergotropy can be rewritten in terms of classical relative
entropy
β Ec (ρ) = C(ρ) + S(L(τ )||ρeq ) − D (ρ||ρeq ) . (12)
Hence, we conclude that there are three distinct contributions to the coherent ergotropy.
Namely, work can be extracted not only from the coherences directly, but also from the
population mismatch between the completely decohered state and the corresponding
thermal state. However, the total extractable work is lowered by the fact that generally ρ is
not diagonal in energy; hence, the classical relative entropy is different from the quantum
relative entropy of the completely decohered state.

4. Classical Ergotropy from Inhomogeneity


Remarkably, the above discussion of the quantum treatment can be generalized to
purely classical scenarios. It was recently recognized that distributions that are inhomoge-
neous on the energy surfaces can be considered the classical equivalent of quantum states
with coherences [18,19]. Therefore, we proceed by defining the classical ergotropy, which
quantifies the maximal work that can be extracted from inhomogeneous distributions
under Hamiltonian dynamics, i.e., under the classical equivalent of unitary maps.
Entropy 2021, 23, 1107 4 of 13

We start with the classical distribution, p(Γ), which measures how likely it is to find
a system at a point in phase space Γ. Now consider a situation in which Γ is sampled
microcanonically from an (initial) energy surface A; we then let p A (Γ) evolve under
Liouville’s equation. We are interested in assessing how close to equilibrium the system is
driven. To this end, consider the joint distribution of finding Γ0 on energy surface B, given
that Γ was sampled from energy surface A

p B| A (Γ, Γ0 ) = p(Γ0 |Γ) p A (Γ) , (13)

where p(Γ0 |Γ) is the classical transition probability distribution, which satisfies
Z Z
dΓ p(Γ0 |Γ) = dΓ0 p(Γ0 |Γ) = 1 , (14)

which follows from Liouville’s theorem and normalization. In the following, we formulate
the classical ergotropy by focusing on the joint distribution p B| A (Γ, Γ0 ) in general classical
systems.
In complete analogy to the quantum case, we now consider the relative entropy of
p B| A (Γ, Γ0 ) with respect to the thermal distribution on energy surface B

exp (− βEB (Γ0 ))


Z
eq
p B (Γ0 ) = dΓ0 exp − βEB (Γ0 ) .

with Z≡ (15)
Z
We can write

p B| A (Γ, Γ0 )
Z Z
!
eq 0 0
D ( p B| A || p B ) = dΓ dΓ p B| A (Γ, Γ ) ln eq . (16)
p B (Γ0 )

Equation (16) is a divergence-like quantity, which becomes non-negative only for the
thermodynamic scenario (See Appendix C). Note that the normalization of the transition
eq
probabilities (14) is essential to guarantee that the classical distributions, p B| A and p B , have
the same support.
As before, we then seek a “transformed” joint distribution QB| A for which the relative
 
eq
entropy D QB| A || p B becomes minimal. This QB| A can be written as
Z
QB| A (Γ00 , Γ) ≡ dΓ0 q(Γ00 |Γ0 ) p(Γ0 |Γ) p A (Γ) , (17)
 
eq
and we need to minimize D QB| A || p B as a function of the transition probability distribu-
tion q(Γ00 |Γ0 ). We start by recognizing that the convolution of two transition probability
distributions is also a transition probability distribution
Z
ξ (Γ00 |Γ) ≡ dΓ0 q(Γ00 |Γ0 ) p(Γ0 |Γ) . (18)
 
eq
Then, we have QB| A (Γ00 , Γ) = ξ (Γ00 |Γ) p A (Γ), thus we need to minimize D QB| A || p B
as a function ξ. In a complete analogy to the quantum case, we can choose ξ (Γ0 |Γ) =
δ(Γ0 − Γ) and obtain the following result (see Appendix D for the proof)
h  i  
eq eq
min D QB| A || p B = D p A || p B . (19)
ξ

Accordingly, we define the classical ergotropy as


   
eq eq
β Eclass ( p B| A ) ≡ D p B| A || p B − D p A || p B , (20)
Entropy 2021, 23, 1107 5 of 13

which quantifies the maximal amount of work that can be extracted from the joint dis-
tribution p B| A under Liouvillian maps. Remarkably, both the quantum (7) as well as
the classical (20) ergotropy comprise the classical relative entropy with respect to a ther-
mal state.
Equation (20) can also be re-written to resemble more closely the established expres-
sion of the quantum ergotropy (1). We have
Z
Eclass ( p B| A ) = dΓ0 ϕ B (Γ0 ) EB (Γ0 ) (21)

where we introduced Z
ϕ B (Γ0 ) = dΓ p B| A (Γ0 , Γ) − p A (Γ0 ) . (22)

In this form, it becomes apparent that the classical ergotropy quantifies the maximal
amount of work stored in the inhomogeneities. Notice that ϕ B is not an explicit function
of the Hamiltonian of the system, which was shown to be a classical equivalent of the
quantum coherences [18,19]. This is the analogy of how the quantum ergotropy quantifies
the maximal work extractable from quantum coherences.

5. Ergotropy in Geometric Quantum Mechanics


Thus far, we have seen that in quantum as well as in classical systems, work can
be extracted by “reshaping” the states in phase space without changing their entropy.
Remarkably, in either case, the ergotropy is given by a difference of relative entropies
(see Equations (7) and (20)). The natural question arises as to whether the quantum-to-
classical limit can be taken systematically, or rather the seemingly independent results can
be derived within a unifying framework.
Only very recently, Anza and Crutchfield [20–22] recognized that for such thermo-
dynamic considerations, so-called geometric quantum mechanics [50–52] are a uniquely
suited paradigm. In standard quantum theory, a quantum state is described by a density op-
erator ρ, which can be expanded in many different decompositions of pure states. However,
an often overlooked consequence is that, thus, the probabilistic interpretation of quantum
states is not unique. To remedy this issue, geometric quantum states [50–52] were introduced,
which are probability distributions on the manifold spanned by the quantum states. In
this sense, classical and quantum mechanics only differ in the geometric properties of the
underlying manifold.
We proceed by briefly outlining the main notions of geometric quantum mechanics,
which is well developed (cf. refs. [20–22,50–52] for a more complete exposition). In the
geometric approach, a pure quantum state |ψi is described as a point in a complex projective
space Vd ≡ C Pd−1 [51], where d is the dimension of the Hilbert space. Note that d can also
be infinite [50]. Here, z is the set of complex homogeneous coordinates in Vd , and z∗ is the
complex conjugate.
Hence, any pure state |ψi can be written as

d −1
|ψ(z)i = ∑ zα | eα i , (23)
α =0

where {eα }dα− 1


=0 is an arbitrary basis. The geometry of the manifold is determined by the
Fubini–Study metric [51]

1
ds2 = 2 ∑ gαγ∗ dzα dz∗γ ≡ ∑ ∂zα ∂z∗γ ln (z · z∗ )dzα dz∗γ , (24)
α,γ 2 α,γ

where we define gαγ∗ ≡ 41 ∂zα ∂zγ ln(z · z∗ ) and which allows to define a unique, unitarily
invariant volume element, dV ≡ det( g) dzdz∗ .
p
Entropy 2021, 23, 1107 6 of 13

It is easy to recognize that pure states are represented as generalized delta functions
on the projective space. In particular, for |ψ0 i ≡ |ψ(z0 )i, the corresponding geometric
quantum state becomes
δ ( z − z0 )
P (z) = δe(z − z0 ) ≡ p , (25)
det( g)
where we introduce the coordinate-covariant Dirac delta. Any (mixed) quantum state can
then be written as Z
ρ= dV P (z) |ψ(z)ihψ(z)| , (26)
Vd

where the geometric quantum states are given by

d  
∑ p j δe
p
P (z) = z − zj , (27)
j =1

p
and p j is again the eigenvalues of ρ and z j ≡ z( p j ).
We are now equipped to return to the expressions for the quantum and classical
ergotropies, Equations (7) and (20), respectively. We immediately recognize that to proceed,
we have to consider a generalization of the relative entropy to geometric quantum states.
In complete analogy to the classical case, we need to guarantee that the geometric quantum
states have the same support [53]. Hence, we introduce a geometric quantum generalization
of the conditional distribution to include a generalized transition probability distribution.
To this end, consider
d  
Pe (z) ≡ ∑ p j δe z − zsj , (28)
j =1

where now zsj ≡ z( s j ), and s j is an eigenstate of a density operator σ. The density


operator, ρe, corresponding to P
e (z) reads

ρe = U e† =
e ρU ∑ pj sj sj , (29)
j

where U
e is the “optimal” unitary maps.
The geometric relative entropy is then defined as
!
Pe (z)
  Z
D Pe ||S ≡ dV P
e (z) ln , (30)
Vd S(z)

where S is the geometric quantum


 state corresponding to σ (same as before). Moreover, we
have by construction D P ||S = S(ρe||σ ) = D (ρ||σ ), and we conclude that the geometric
e
relative entropy is identical in value to the classical relative entropy (5). Therefore, we can
write the quantum ergotropy (7) as
 
β E (ρ) = S(ρ||ρeq ) − D P e ||P eq , (31)

where P eq is the geometric quantum state corresponding to ρeq . In other words, the
quantum ergotropy is the difference of the relative entropies of the density operator and
the geometric quantum state with respect to ρeq .
Remarkably, also the classical case can be fully treated within the geometric approach.
To this end, note that for any classical distribution, we can construct the corresponding
geometric quantum state. Therefore, it now becomes a fair comparison to consider the
difference of quantum and classical ergotropy, ∆E ≡ E (ρ) − Eclass (ρ). It is not far-fetched
to realize that ∆E is the genuinely quantum contribution to the extractable work. A more
careful analysis of this contribution may be related to quantum correlations (see also
Entropy 2021, 23, 1107 7 of 13

ref. [17]), yet a thorough analysis is beyond the scope of the present discussion. Rather, the
remainder of this analysis is dedicated to an application of the gained insight to quantum
work relations.

6. Ergotropy from Conditional Thermal States


To this end, imagine a closed system that is driven by the variation of some external
control parameter. We denote the initial Hamiltonian by H A and the final Hamiltonian by
HB , and the average work is simply given by hW i = h HB i − h H A i. The maximum work the-
orem predicts that hW i is always larger than the work performed for quasistastic driving [7].
If the system was initially prepared in a thermal state, the quasistatic work is nothing but
the difference in Helmholtz free energy ∆F [54]. The difference of total work and free energy
difference is called irreversible work, and we have hWirr i = hW i − ∆F ≥ 0 [54]. Only rather
recently, it was recognized that a sharper inequality can be derived, for both quantum [23]
as well as classical [26] systems if the quantum work statistics are conditioned on the initial
state. Note that this corresponds to the one-time measurement approach, where only one
projective measurement is taken at the beginning of the process.
In particular, the following was shown [23,26]
 
eq
βhWirr i ≥ S $ B ||ρ B , (32)

eq
where ρ B = exp (− βHB )/ZB , and $ B is called the conditional thermal state [26]. It
reads [23]
exp (− β h B ( j A ))
$B ≡ ∑ Uτ | j A ih j A |Uτ† , (33)
j
Z ( B | A )

where | j A i is an eigenstate of the initial Hamiltonian H A . Further, Uτ is the unitary


evolution operator corresponding to driving the system from H A to HB , and

h B ( j A ) ≡ h j A |Uτ† HB Uτ | j A i . (34)

Finally, Z ( B| A) is the conditional partition function of $ B . Since the discovery of


Equation (32), the significance of the conditional thermal state has been somewhat obscure.
In ref. [23,25], the lower bound in Equation (32) was understood as some contribution to the
usable work that would have been destroyed by a second projective measurement. Yet, a
transparent interpretation is lacking.
Remarkably, it is not hard to see that $ B is a representation of the geometric canonical
ensemble as proposed by Anza and Crutchfield [20,22]. The geometric canonical ensemble
is defined as the geometric state that maximizes the corresponding Shannon entropy under
the usual boundary conditions [55]. Specifically, we have [20,22]

exp (− β h(z))
P( z ) ≡ , (35)
Z
where h(z) ≡ hψ(z)| H |ψ(z)i and the geometric partition function is
Z
Z≡ dV exp (− β h(z)) . (36)
Vd

Thus, to maintain the consistency of the presentation, we continue to employ the


geometric formulation of quantum states. Now, consider the geometric representation
of $ B Z
$B = dVPB (z) |ψ(z)ihψ(z)| (37)
Vd

and we have
exp (− β h B (z)) e


PB (z) = δ z − zj , (38)
j
Z ( B| A)
Entropy 2021, 23, 1107 8 of 13

where, as before, h B (z) ≡ hψ(z)| HB |ψ(z)i and the covariant Dirac delta is evaluated at
ψ(z j ) ≡ Uτ | j A i. Comparing Equations (35) and (38), we immediately recognize that the
PB (z) is nothing but the geometric canonical state evaluated on the quantum manifold.
The natural question arises as to whether any work can be extracted from the geometric
ensemble. To this end, consider the corresponding ergotropy (31)
   
eq e B ||Peq ,
β E ($ B ) = S $ B ||ρ B − D P B (39)

where, in complete analogy to the above, P


e B is given by

exp (− βh B ( j A )) e 

e B (z) ≡ eq
P δ z − zj , (40)
j
Z ( B| A)

eq
and now z j ≡ z(| jB i), where | jB i is the eigenstate of the final Hamiltonian HB . Thus,
exploiting Equation (12), we can write the sharpened maximum work theorem (32) as
 
eq
βhWirr i ≥ βEi ($ B ) + C($ B ) + S L(τB )||ρ B , (41)

where τB is the coherence-invariant state of $ B . In conclusion, realizing that the conditional


thermal state (33) is nothing but a representation of the geometric canonical ensemble, the
physical interpretation of the sharpened maximum work theorem (32) becomes apparent.
The lower bound on the irreversible work has three contributions, namely the incoherent
ergotropy and the quantum coherences stored in the conditional thermal state, and the
eq
population mismatch between $ B and ρ B . Therefore, we conclude that the conditional
thermal state provides an informational contribution from its coherence to the second law.
From the fact that the classical and quantum ergotropy share the same geometric relative
entropy, we emphasize that thermodynamics based on geometric quantum mechanics is a
unified approach to the quantum-to-classical limit.

7. Conclusions
In conclusion, motivated by the desire to express the maximally extractable work
in a form independent of the optimization over unitary operations, we have obtained
several results. Given that the quantum ergotropy can be expressed as the difference of the
quantum and classical relative entropies, we identified three distinct contributions to the
coherent ergotropy, of which the relative entropy of coherence and the population mismatch
between thermal state and fully decohered state are the most important. This insight was
extended to classical systems, in which inhomogeneities in the energy distribution play
the role of quantum coherences. To quantify how much work can be extracted from
classical states, we introduced the classical ergotropy, and we postulated that the genuine
quantum contribution to the ergotropy is given by the difference of the quantum and
classical expressions. Our analysis provides a consistent approach to maximum work
extraction in both quantum and classical systems. In particular, we have not only shown
that classical inhomogeneities play the role of “classical coherence", but also that work can
be extracted that is quantified by the classical ergotropy. This was solidified by exploiting
the geometric approach to quantum mechanics, in which quantum and classical states
can be treated in a unified framework. As an application, we demonstrated that the
recently introduced notion of “conditional thermal state” actually belongs to the family of
geometric canonical ensembles and that, hence, the corresponding sharpened maximum
work theorem becomes easy to interpret. This demonstrates that understanding quantum as
well as classical ergotropies is an essential pillar of modern thermodynamics with a myriad
of potential applications. Finally, these results demonstrate that the geometric approach
can be regarded as a methodology of unifying the quantum and classical approaches to the
second law of thermodynamics.
Entropy 2021, 23, 1107 9 of 13

Author Contributions: A.S. and S.D. contributed to all aspects of this work. Conceptualization, A.S.
and S.D.; Formal analysis, A.S. and S.D.; Investigation, A.S. and S.D.; Supervision, S.D.; Validation,
A.S. and S.D.; Writing—original draft, A.S.; Writing—review & editing, A.S. and S.D. All authors
have read and agreed to the published version of the manuscript.
Funding: A.S. was supported by the U.S. Department of Energy, the Laboratory Directed Research
and Development (LDRD) program and the Center for Nonlinear Studies at LANL. He is now
supported by the internal R&D from Aliro Technologies, Inc. S.D. acknowledges support from the
U.S. National Science Foundation under Grant No. DMR-2010127.
Institutional Review Board Statement: Not applicable.
Informed Consent Statement: Not applicable.
Data Availability Statement: Not applicable.
Acknowledgments: We would like to thank Fabio Anza, Christopher Jarzynski, and Kanupriya
Sinha for the insightful discussions.
Conflicts of Interest: The authors declare no conflict of interest.

Appendix A. Proof of Equation (5)


Proof of Equation (5). In this section, we provide an alternative proof of Equation (5)
different from the method in ref. [47]. Actually, the procedure of the proof is similar to that
of quantum ergotropy, which was introduced in ref. [2].
Let us consider the quantum relative entropy
n o
S(UρU † ||σ) ≡ tr{ρ ln (ρ)} − tr UρU † ln (σ ) . (A1)

In order to identify the specific U such that S(UρU † ||σ ) is minimized, we now
parametrize a variation as δU = (δX )U, where δX is an arbitrary infinitesimal, anti-
Hermitian operator, i.e., (δX )† = −δX. Hence, we may write
n h io
δS(UρU † ||σ) = −tr δX UρU † , ln (σ) = 0 (A2)

Then, a solution to Equation (A2) is given by


Z
e≡
U ∑|si ih pi | , (A3)
i

for which we immediately obtain


Z
S(Uρ
e Ue † ||σ ) = ∑ pi ln ( pi /si ) ≡ D(ρ||σ) . (A4)
i

Therefore, we conclude that the minimum of the quantum relative entropy under all
unitary transformations of ρ is nothing but the classical relative entropy of its distribution
of eigenvalues.

Appendix B. Proof of Equation (7)


Proof of Equation (7). In this section, we prove Equation (7). Let E (ρ) be the quantum
ergotropy of a quantum state
Z
ρ= ∑ pi | pi ih pi | with p i ≥ p i +1 , (A5)
i
Entropy 2021, 23, 1107 10 of 13

and ρeq be the Gibbs state

e− βH e− βEj
Z
ρ eq
=
Z
= ∑ Z
Ej Ej with E j ≤ E j +1 (A6)
j

with an arbitrary inverse temperature β. Then, quantum ergotropy is given by


Z Z
E (ρ) = ∑ pi Ej | Ej | pi |2 − ∑ pi Ei . (A7)
i,j i

The quantum relative entropy S(ρ||ρeq ) is explicitly written as

S(ρ||ρeq ) = tr
Z{
ρ ln (ρ)} + βtrZ{ρH } + ln ( Z )
= ∑ pi ln ( pi ) + β ∑ pi Ej | Ej | pi |2 + ln ( Z ) , (A8)
i i,j

and the classical relative entropy with respect to the eigenvalue distributions is
Z Z
D (ρ||ρeq ) = ∑ pi ln ( pi ) + β ∑ pi Ei + ln (Z) . (A9)
i i

Therefore, we can obtain Equation (7)

βE (ρ) = S(ρ||ρeq ) − D (ρ||ρeq ) . (A10)

Note that β−1 (S(ρ||ρeq ) − D (ρ||ρeq )) makes E (ρ) independent of β.

Appendix C. Non-Negativity of Divergence-like Quantity in Thermodynamic Scenario


In this section, let us explain why the divergence-like quantity introduced in Equation (16)
takes only non-negative values in the thermodynamic scenario, while in general, i could take
negative values.
Consider the relative entropy of the probability distribution

p B| A (Γ0 , Γ ) = p (Γ0 |Γ ) p A (Γ ) (A11)

with respect to a certain probability distribution r B (Γ0 ). Let us write


Z Z
p B (Γ0 ) ≡ dΓp B| A (Γ0 , Γ) = dΓp(Γ0 |Γ) p A (Γ) . (A12)

Then, we have

p B| A (Γ0 , Γ )
Z
!
 
D p B| A ||r B = dΓ0 dΓp B| A (Γ0 , Γ) ln
r B (Γ0 )
Z Z (A13)
 
0 0 0 0 0 0
dΓ dΓp B| A (Γ , Γ) ln p B| A (Γ , Γ) − dΓ p B (Γ ) ln r B (Γ ) .

=

Because of the vanishing conditional entropy due to the Liouville’s equation,


Z
dΓ0 dΓp B| A (Γ0 , Γ) ln p(Γ0 |Γ) = 0 ,

H( B| A) ≡ − (A14)

and the normalization of the conditional probability distribution in Equation (14), we have
Z   Z
dΓ0 dΓp B| A (Γ0 , Γ) ln p B| A (Γ0 , Γ) = dΓp A (Γ) ln ( p A (Γ)) . (A15)
Entropy 2021, 23, 1107 11 of 13

Therefore, we have
  Z Z
D p B| A ||r B = dΓp A (Γ) ln ( p A (Γ)) − dΓ0 p B (Γ0 ) ln r B (Γ0 ) .

(A16)

This could be negative. However, in the thermodynamic scenario, from the second
law of thermodynamics, the final entropy H( B) has to be greater than or equal to the initial
entropy H( A)
Z Z
dΓ p A (Γ) ln ( p A (Γ)) ≥ dΓ0 p B (Γ0 ) ln p B (Γ0 ) .

H( A) ≤ H( B) ⇐⇒ (A17)

Hence, we have

p B (Γ0 )
  Z  
D p B| A ||r B ≥ dΓ0 p B (Γ0 ) ln = D ( p B ||r B ) ≥ 0 . (A18)
r B (Γ0 )

Therefore, the divergence-like quantity in Equation (A13) becomes non-negative


 
D p B| A ||r B ≥ 0 . (A19)

Appendix D. Proof of Equation (19)


Proof of Equation (19). In this section, we provide a proof for Equation (19). A variation
in ξ can be written as
δξ ≡ δΓ0 · ∇Γ0 ξ + δΓ · ∇Γ ξ , (A20)
eq
where we replace Γ00 with Γ0 without loss of generality. From the expression of p B and the
vanishing conditional entropy due to the Liouvillian evolution, we obtain
  Z Z
eq
δD QB| A || p B = β dΓ dΓ0 p A (Γ) EB (Γ0 ) δξ , (A21)

eq
where we use the explicit  for p B . We can find that the variation of the relative
 expression
eq
entropy vanishes, δD QB| A || p B = 0, for ξ (Γ0 |Γ) = δ(Γ0 − Γ); therefore, we obtain
h  i  
eq eq
min D QB| A || p B = D p A || p B . (A22)
ξ

References
1. Gellhorn, E. The emotions and the ergotropic and trophotropic systems. Psychol. Forsch. 1970, 34, 67–94. [CrossRef]
2. Allahverdyan, A.E.; Balian, R.; Nieuwenhuizen, T.M. Maximal work extraction from finite quantum systems. EPL (Europhys.
Lett.) 2004, 67, 565. [CrossRef]
3. Pusz, W.; Woronowicz, S.L. Passive states and KMS states for general quantum systems. Commun. Math. Phys. 1978, 58, 273–290.
[CrossRef]
4. Koukoulekidis, N.; Alexander, R.; Hebdige, T.; Jennings, D. The geometry of passivity for quantum systems and a novel
elementary derivation of the Gibbs state. Quantum 2021, 5, 411. [CrossRef]
5. Górecki, J.; Pusz, W. Passive states for finite classical systems. Lett. Math. Phys. 1980, 4, 433. [CrossRef]
6. Daniëls, H.A.M. Passivity and equilibrium for classical Hamiltonian systems. J. Math. Phys. 1981, 22, 843. [CrossRef]
7. Deffner, S.; Campbell, S. Quantum Thermodynamics; Morgan and Claypool Publishers: San Rafael, CA, USA, 2019.
8. Goold, J.; Huber, M.; Riera, A.; del Rio, L.; Skrzypczyk, P. The role of quantum information in thermodynamics—A topical review.
J. Phys. A Math. Theor. 2016, 49, 143001. [CrossRef]
9. Perarnau-Llobet, M.; Bäumer, E.; Hovhannisyan, K.V.; Huber, M.; Acin, A. No-Go Theorem for the Characterization of Work
Fluctuations in Coherent Quantum Systems. Phys. Rev. Lett. 2017, 118, 070601. [CrossRef]
10. Levy, A.; Lostaglio, M. A quasiprobability distribution for heat fluctuations in the quantum regime. Phys. Rev. X Quantum 2020,
1, 010309. [CrossRef]
Entropy 2021, 23, 1107 12 of 13

11. Santos, J.P.; Céleri, L.C.; Landi, G.T.; Paternostro, M. The role of quantum coherence in non-equilibrium entropy production. NPJ
Quantum Inf. 2019, 5, 23. [CrossRef]
12. Niedenzu, W.; Mukherjee, V.; Ghosh, A.; Kofman, A.G.; Kurizki, G. Quantum engine efficiency bound beyond the second law of
thermodynamics. Nat. Commun. 2018, 9, 165. [CrossRef] [PubMed]
13. Cherubim, C.; Brito, F.; Deffner, S. Non-Thermal Quantum Engine in Transmon Qubits. Entropy 2019, 21, 545. [CrossRef]
[PubMed]
14. Francica, G.; Binder, F.; Guarnieri, G.; Mitchison, M.; Goold, J.; Plastina, F. Quantum Coherence and Ergotropy. Phys. Rev. Lett.
2020, 125, 180603. [CrossRef]
15. Çakmak, B. Ergotropy from coherences in an open quantum system. Phys. Rev. E 2020, 102, 042111. [CrossRef]
16. Francica, G.; Goold, J.; Plastina, F.; Paternostro, M. Daemonic ergotropy: Enhanced work extraction from quantum correlations.
NPJ Quantum Inf. 2017, 3, 12. [CrossRef]
17. Touil, A.; Çakmak, B.; Deffner, S. Second law of thermodynamics for quantum correlations. arXiv 2021, arXiv:2102.13606.
18. Smith, A.M. Studies in Nonequilibrium Quantum Thermodynamics. Ph.D. Thesis, University of Maryland, College Park, MD,
USA, 2019.
19. Smith, A.; Sinha, K.; Jarzynski, C. (to be published).
20. Anza, F.; Crutchfield, J.P. Geometric Quantum State Estimation. arXiv 2020, arXiv:2008.08679.
21. Anza, F.; Crutchfield, J.P. Beyond Density Matrices: Geometric Quantum States. arXiv 2020, arXiv:2008.08682.
22. Anza, F.; Crutchfield, J.P. Geometric Quantum Thermodynamics. arXiv 2020, arXiv:2008.08683.
23. Deffner, S.; Paz, J.P.; Zurek, W.H. Quantum work and the thermodynamic cost of quantum measurements. Phys. Rev. E 2016,
94, 010103(R). [CrossRef]
24. Beyer, K.; Luoma, K.; Strunz, W.T. Work as an external quantum observable and an operational quantum work fluctuation
theorem. Phys. Rev. Res. 2020, 2, 033508. [CrossRef]
25. Sone, A.; Liu, Y.X.; Cappellaro, P. Quantum Jarzynski Equality in Open Quantum Systems from the One-Time Measurement
Scheme. Phys. Rev. Lett. 2020, 125, 060602. [CrossRef] [PubMed]
26. Sone, A.; Deffner, S. Jarzynski equality for stochastic conditional work. J. Stat. Phys 2021, 183, 11. [CrossRef]
27. Allahverdyan, A.E.; Nieuwenhuizen, T.M. Fluctuations of work from quantum subensembles: The case against quantum
work-fluctuation theorems. Phys. Rev. E 2005, 71, 066102. [CrossRef] [PubMed]
28. Kurchan, J. A Quantum Fluctuation Theorem. arXiv 2001, arXiv:cond-mat/0007360.
29. Tasaki, H. Jarzynski Relations for Quantum Systems and Some Applications. arXiv 2000, arXiv:cond-mat/0009244.
30. Talkner, P.; Lutz, E.; Hänggi, P. Fluctuation theorems: Work is not an observable. Phys. Rev. E. 2007, 75, 050102(R). [CrossRef]
31. Huber, G.; Schmidt-Kaler, F.; Deffner, S.; Lutz, E. Employing Trapped Cold Ions to Verify the Quantum Jarzynski Equality. Phys.
Rev. Lett. 2008, 101, 070403. [CrossRef]
32. Campisi, M.; Hänggi, P.; Talkner, P. Colloquium: Quantum fluctuation relations: Foundations and applications. Rev. Mod. Phys.
2011, 83, 771. [CrossRef]
33. Deffner, S.; Lutz, E. Nonequilibrium Entropy Production for Open Quantum Systems. Phys. Rev. Lett. 2011, 107, 140404.
[CrossRef]
34. Kafri, D.; Deffner, S. Holevo’s bound from a general quantum fluctuation theorem. Phys. Rev. A 2012, 86, 044302. [CrossRef]
35. Mazzola, L.; De Chiara, G.; Paternostro, M. Measuring the Characteristic Function of the Work Distribution. Phys. Rev. Lett. 2013,
110, 230602. [CrossRef] [PubMed]
36. Dorner, R.; Clark, S.R.; Heaney, L.; Fazio, R.; Goold, J.; Vedral, V. Extracting Quantum Work Statistics and Fluctuation Theorems
by Single-Qubit Interferometry. Phys. Rev. Lett. 2013, 110, 230601. [CrossRef]
37. Roncaglia, A.J.; Cerisola, F.; Paz, J.P. Work Measurement as a Generalized Quantum Measurement. Phys. Rev. Lett. 2014,
113, 250601. [CrossRef]
38. Batalhão, T.B.; Souza, A.M.; Mazzola, L.; Auccaise, R.; Sarthour, R.S.; Oliveira, I.S.; Goold, J.; De Chiara, G.; Paternostro, M.; Serra,
R.M. Experimental Reconstruction of Work Distribution and Study of Fluctuation Relations in a Closed Quantum System. Phys.
Rev. Lett. 2014, 113, 140601. [CrossRef]
39. An, S.; Zhang, J.N.; Um, M.; Lv, D.; Lu, Y.; Zhang, J.; Yin, Z.Q.; Quan, H.T.; Kim, K. Experimental test of the quantum Jarzynski
equality with a trapped-ion system. Nat. Phys. 2015, 11, 193–199. [CrossRef]
40. Deffner, S.; Saxena, A. Jarzynski Equality in P T -Symmetric Quantum Mechanics. Phys. Rev. Lett. 2015, 114, 150601. [CrossRef]
[PubMed]
41. Deffner, S.; Saxena, A. Quantum work statistics of charged Dirac particles in time-dependent fields. Phys. Rev. E 2015, 92, 032137.
[CrossRef] [PubMed]
42. Talkner, P.; Hänggi, P. Aspects of quantum work. Phys. Rev. E 2016, 93, 022131. [CrossRef]
43. Gardas, B.; Deffner, S.; Saxena, A. Non-hermitian quantum thermodynamics. Sci. Rep. 2016, 6, 23408. [CrossRef] [PubMed]
44. Bartolotta, A.; Deffner, S. Jarzynski Equality for Driven Quantum Field Theories. Phys. Rev. X 2018, 8, 011033. [CrossRef]
45. Gardas, B.; Deffner, S. Quantum fluctuation theorem for error diagnostics in quantum annealers. Sci. Rep. 2018, 8, 17191.
[CrossRef] [PubMed]
46. Touil, A.; Deffner, S. Information Scrambling versus Decoherence—Two Competing Sinks for Entropy. PRX Quantum 2021,
2, 010306. [CrossRef]
Entropy 2021, 23, 1107 13 of 13

47. Nielsen, M.A.; Chuang, I.L. Quantum Computation and Quantum Information: 10th Anniversary Edition, 10th ed.; Cambridge
University Press: New York, NY, USA, 2011.
48. Łobejko, M. The tight Second Law inequality for coherent quantum systems and finite-size heat baths. Nat. Commun. 2021,
12, 918. [CrossRef]
49. Baumgratz, T.; Cramer, M.; Plenio, M. Quantifying Coherence. Phys. Rev. Lett. 2014, 113, 140401. [CrossRef]
50. Ashtekar, A.; Schilling, T.A. Geometrical Formulation of Quantum Mechanics. In On Einstein’s Path; Springer: New York, NY,
USA, 1999; pp. 23–65.
51. Bengtsson, I.; Zyczkowski, K. Geometry of Quantum States; Cambridge University Press: New York, NY, USA, 2017.
52. Cariñena, J.F.; Clemente-Gallardo, J.; Marmo, G. Geometrization of quantum mechanics. Theor. Math. Phys. 2007, 152, 894.
[CrossRef]
53. Anza, F.; Crutchfield, J. in preparation.
54. Deffner, S.; Lutz, E. Generalized Clausius Inequality for Nonequilibrium Quantum Processes. Phys. Rev. Lett. 2010, 105, 170402.
[CrossRef]
55. Jaynes, E.T. Information Theory and Statistical Mechanics. Phys. Rev. 1957, 106, 620. [CrossRef]

You might also like