Coherent fluctuation relations: from the abstract to the concrete
Zoë Holmes
1
, Sebastian Weidt
2
, David Jennings
1,3,4
, Janet Anders
5
, and Florian Mintert.
1
1
Controlled Quantum Dynamics Theory Group, Imperial College London, London, SW7 2BW, United Kingdom.
2
Department of Physics and Astronomy, University of Sussex, Brighton BN1 9QH, United Kingdom.
3
Department of Physics, University of Oxford, Oxford, OX1 3PU, United Kingdom.
4
School of Physics and Astronomy, University of Leeds, Leeds, LS2 9JT, United Kingdom.
5
CEMPS, Physics and Astronomy, University of Exeter, Exeter, EX4 4QL, United Kingdom.
February 20, 2019
Recent studies using the quantum informa-
tion theoretic approach to thermodynamics
show that the presence of coherence in quan-
tum systems generates corrections to classi-
cal fluctuation theorems. To explicate the
physical origins and implications of such cor-
rections, we here convert an abstract frame-
work of an autonomous quantum Crooks rela-
tion into quantum Crooks equalities for well-
known coherent, squeezed and cat states. We
further provide a proposal for a concrete ex-
perimental scenario to test these equalities.
Our scheme consists of the autonomous evo-
lution of a trapped ion and uses a position
dependent AC Stark shift.
1 Introduction
The emergent field of quantum thermodynamics
seeks to extend the laws of thermodynamics and non-
equilibrium statistical mechanics to quantum sys-
tems. A central question is whether quantum me-
chanical phenomena, such as coherence and entan-
glement, generate corrections to classical thermal
physics.
A quantum information theoretic approach has
proven a fruitful means of incorporating genuinely
quantum mechanical effects into thermal physics [1].
For example, the consequences of quantum entangle-
ment for Landauer erasure [2], the thermodynamic
arrow of time [3] and thermalisation [4] have been
investigated. In addition, for incoherent quantum
systems, general criteria for state conversion [5], gen-
eralisations of the second laws [6], and limits to work
extraction protocols [7] have been established. More
recently, these results have been extended to coherent
quantum states [811]. Much of this research [511]
utilised the resource theory framework and in par-
ticular the concept of thermal operations [12, 13].
However, despite significant theoretical progress, the
results have been seen as rather abstract and una-
mendable to experimental implementation [14].
A separate line of enquiry has explored fluctuation
theorems, which can be seen as generalisations of the
second law of thermodynamics to non-equilibrium
processes [15]. They consider systems that are driven
out of equilibrium and establish exact relations be-
tween the resultant thermal fluctuations [16]. Ex-
perimental tests of fluctuation theorems, in partic-
ular the Jarzynski [17] and Crooks [18] equalities,
have been conducted in classical systems, including
stretched RNA molecules [19, 20], over-damped col-
loidal particles in harmonic potentials [21] and clas-
sical two-state systems [22], and quantum systems
such as trapped ions [23] and NMR systems [24]. An
experiment to test a quantum Jarzynski equality in
a weakly measured system using circuit QED has re-
cently been performed [25].
However, the use of two point energy measure-
ments or continual weak measurements to obtain a
work probability distribution reduces a system’s co-
herence with respect to the energy eigenbasis [2631].
This limits the extent to which these previous exper-
iments probe quantum mechanical phenomena that
arise from coherences.
A recent theoretical proposal of a new quantum
fluctuation relation [32] that connects naturally with
quantum information theory, models the work sys-
tem explicitly as a quantum battery. Rather than
implicitly appealing to an additional classical system
to drive the system out of equilibrium, the quan-
tum system and battery here evolve autonomously
under a time independent Hamiltonian [3336]. This
proposal does not require projective measurements
onto energy eigenstates and so does not destroy
coherences. As a result coherence actively con-
tributes to the Autonomous Quantum Crooks equal-
ity (AQC) [32], which will be defined in Eq. (8).
In this paper, we exploit the combined strengths
of the quantum information and fluctuation theorem
approaches to quantum thermodynamics [37, 38]. We
develop quantum Crooks equalities for a quantum
Accepted in Quantum 2018-02-12, click title to verify 1
arXiv:1806.11256v4 [quant-ph] 19 Feb 2019
harmonic oscillator battery that is prepared in op-
tical coherent states, squeezed states or cat states
and explore the physical content of these equalities.
In particular, we present the coherent state Crooks
equality, see Eq. (15), in which quantum corrections
to the classical Crooks equality can be attributed to
the presence of quantum vacuum fluctuations.
We propose an experiment to test the AQC
through which the role of coherence in thermal
physics can be probed. Using a trapped ion, one
can realise a two level system with a pair of the ion’s
internal energy levels and an oscillator battery with
the ion’s axial phonon mode. The motion of the ion
through an off-resonance laser beam induces a po-
sition dependent AC Stark shift on the internal en-
ergy levels. In this way an effectively time-dependent
Hamiltonian for the two level system can be realised.
2 The Classical and Quantum Crooks
Equalities
2.1 The classical Crooks equality
An initially thermal system, at temperature T , can
be driven from equilibrium with a time-dependent
Hamiltonian, changing from H
i
S
to H
f
S
. The classical
Crooks equality [18],
P
+
(W )
P
(W )
= exp
F
k
B
T
exp
W
k
B
T
, (1)
quantifies the ratio of the probability P
+
(W ) of the
work W done on a system in such a non-equilibrium
process, to the probability P
(W ) to extract the
work W in the time reversed process (where the
Hamiltonian is changed back from H
f
S
to H
i
S
). This
ratio is a function of the change in the system’s free
energy,
F
:
= F (H
f
S
) F (H
i
S
) with
F (H
S
)
:
= k
B
T ln
Tr
S
exp
H
S
k
B
T

,
(2)
and hence measurement of P (W ) and P (W ) for a
non-equilibrium process provides a means of inferring
free energy changes, an equilibrium property of a sys-
tem. More generally, Eq. (1) asserts that processes
that produce work are exponentially less likely than
processes that require work, irrespective of how the
change from H
i
S
to H
f
S
is realised.
2.2 The autonomous quantum Crooks equality
The realisation of a time-dependent Hamiltonian in
the classical Crooks equality, Eq. (1), necessarily im-
plies an interaction with a classical agent. The au-
tonomous framework [32] proposed by Johan Åberg
makes this control explicit by introducing a battery.
Specifically, the system evolves together with the bat-
tery according to the time-independent Hamiltonian,
H
SB
= H
S
B
+
S
H
B
+ V
SB
, (3)
comprised of the Hamiltonians H
S
and H
B
for sys-
tem and battery, and their interaction V
SB
.
For the initial and final Hamiltonians of the system
to be well defined, we need the system and battery
not to interact at these times. To ensure this, we
consider an interaction of the form
V
SB
= H
i
S
Π
i
B
+ H
f
S
Π
f
B
+ V
SB
, (4)
where Π
i
B
and Π
f
B
are projectors onto two orthog-
onal subspaces, R
i
and R
f
, of the battery’s Hilbert
space, and V
SB
has support only outside those two
subspaces, i.e. (X
S
Π
i
B
)V
SB
= (X
S
Π
f
B
)V
SB
= 0
for any system operator X
S
. The system Hamilto-
nian H
S
can always be absorbed into V
SB
, which we
will do in the following. Assuming the battery is ini-
tialised in a state in subspace R
i
only and evolves to a
final state in subspace R
f
only, the system Hamilto-
nian evolves from H
i
S
to H
f
S
, i.e. the system Hamilto-
nian is effectively time-dependent. Since the system
and battery are non-interacting in the regions R
i
and
R
f
, and since energy is globally conserved, the energy
required or produced by the system as its Hamilto-
nian evolves from H
i
S
to H
f
S
is necessarily provided
or absorbed by the battery.
The initially thermal system in the Crooks equal-
ity is incoherent in energy and so, to extend the rela-
tion to the quantum regime in this autonomous set-
ting, coherence
1
must begin in an additional non-
equilibrium quantum system. Here the coherence is
provided by the battery which we assume can be pre-
pared in superpositions of energy states. The evolu-
tion under H
SB
generates correlations between the
system and battery reducing the coherence that can
be attributed to the battery alone. Since the dynam-
ics conserves coherence globally, the coherence of the
final joint state of the system and battery is equal
to that initially provided by the battery. Thus this
framework allows us to describe the evolution of co-
herence under interactions with a thermal system.
1
We use the term ‘coherence’ in the sense of a ‘superposi-
tion of states belonging to different energy eigenspaces’. Such
energetic coherences are formally quantified in the thermal op-
erations framework [12, 13] and by measures such as the l
1
norm of coherence [39], C
l
1
(ρ)
:
=
P
j6=k
| hE
j
| ρ |E
k
i |, the sum
of the absolute value of the off diagonal elements of a state in
the energy eigenbasis.
Accepted in Quantum 2018-02-12, click title to verify 2
The difficulties surrounding how to define work in
the quantum regime [4045] are avoided by formu-
lating the AQC in terms of transition probabilities
between battery states. Considering a system ini-
tialised in the thermal state γ
i
and the battery ini-
tialised in the pure state |φ
i
i, the probability to ob-
serve the battery after a time-interval t in the state
|φ
f
i reads
P(φ
f
|φ
i
, γ
i
) = hφ
f
|Tr
S
[U(γ
i
|φ
i
ihφ
i
|)U
] |φ
f
i ,
(5)
with U = exp(iH
SB
t/~) the propagator of the sys-
tem and battery and the trace over the system de-
noted by Tr
S
. If we consider the transition probabil-
ity between energy eigenstates, the scheme consid-
ered here reduces to the usual two-point projective
energy measurement scheme [2628] and as such we
can recover the classical Crooks relation, Eq. (1). For
transition probabilities between more general quan-
tum states, such as between coherent states, squeezed
states and cat states [46] as we will consider, the post
measurement battery state has coherences in the en-
ergy eigenbasis and the measurement can induce a
non-classical updating of the state of the system.
The AQC relates the probabilities P(φ
f
|φ
i
, γ
i
) of a
forward process and P(ψ
i
|ψ
f
, γ
f
) of a reverse process
for two thermal system states
γ
x
exp
H
x
S
k
B
T
, with x = i, f (6)
and two pairs of pure initial and final battery states
satisfying
|ψ
f
i T exp
H
B
2k
B
T
|φ
f
i ,
|φ
i
i T exp
H
B
2k
B
T
|ψ
i
i .
(7)
The states |ψ
i
i and |ψ
f
i correspond to a time-
reversed process and are defined in terms of the time-
reversal operator
2
T [47]. This relationship between
the initial and final states of the forwards and re-
versed processes, Eq. (7), is forced by the deriva-
tion of the AQC, i.e. the structure arises natu-
rally when one reverses a general quantum process
in an autonomous setting. If the battery states
|φ
i
i and |φ
f
i are eigenstates of H
B
, the states |ψ
f
i
and |ψ
i
i of the reversed process are regular time-
reversed states. However, for any coherent super-
position of energy eigenstates the additional term
2
The time-reversal [47] operation T on a battery state |ψi
is defined as complex conjugation in the energy eigenbasis of
the battery, T |ψi = |ψ
i. We introduce T in more detail in
Appendix A.2.
exp(H
B
/2k
B
T ), which is linked with the Petz re-
covery map [48, 49], is essential to capture the influ-
ence of quantum coherence.
The AQC [32] reads
P(φ
f
|φ
i
, γ
i
)
P(ψ
i
|ψ
f
, γ
f
)
= exp
F
k
B
T
exp
˜
E
k
B
T
!
, (8)
where F is the change in the system’s equilibrium
free energy, Eq. (2), and thus the AQC, similarly
to the classical Crooks equality, provides a means of
inferring free energy changes. The term
˜
E(ψ
i
, φ
f
)
:
=
˜
E(ψ
i
)
˜
E(φ
f
) (9)
is a quantum mechanical generalisation of the energy
flow from the battery to the system. The function
˜
E(ψ)
:
= k
B
T ln
hψ|exp
H
B
k
B
T
|ψi
(10)
is an effective potential for the battery state |ψi that
specifies the relevant energy value within the fluctu-
ation theorem context. We provide a discussion of
its properties in Appendix A.7.
For the AQC to hold exactly we require |ψ
i
i and
|φ
f
i to have support only within subregions Π
i
B
and Π
f
B
respectively and for the battery Hamilto-
nian not to induce evolution between subregions, i.e.
(
B
Π
i
B
)H
B
Π
i
B
= 0 and (
B
Π
f
B
)H
B
Π
f
B
= 0. In
general, the second of these conditions does not hold
because a battery Hamiltonian that evolves states be-
tween subregions is required. However, despite this,
numerical simulations (discussed in Section 3.2) in-
dicate that any error can be made negligible as long
as |ψ
i
i and |φ
f
i are well localised in subregions Π
i
B
and Π
f
B
respectively. In Section 2.5 we discuss how
the accuracy of the equality can be quantified; how-
ever, the material is relatively technical and can be
skipped.
To make the predictions of the AQC, Eq. (8),
more concrete we now choose the battery to be
a quantum harmonic oscillator with Hamiltonian
H
B
=
P
n
~ω
n +
1
2
|nihn|. In order to derive
explicit expressions for the ratio of transition prob-
abilities, Eq. (8), we consider common states of
the harmonic oscillator, specifically coherent states,
squeezed states and cat states [46]. For each of these
three states we derive a specific AQC which is based
on the generalised energy flow, Eq. (9), and the re-
lation between the initial and final states of the for-
ward and reversed process, Eq. (7). While the bat-
tery Hamiltonian and states are fixed in the following
sections, the system Hamiltonians, H
i
S
and H
f
S
, and
the system and battery interaction, V
SB
, can still be
chosen at will.
Accepted in Quantum 2018-02-12, click title to verify 3
2.3 Coherent state Crooks equality
The physical content of the AQC is neatly illustrated
by considering transition probabilities between co-
herent states of the battery, |αi = exp(αa
α
a) |0i
with a
and a the creation and annihilation opera-
tors of the oscillator respectively, and α a complex
number. In the forwards process, we assume the
battery is initialised in |α
i
exp(χ)i with χ =
~ω
2k
B
T
and we are interested in the transition probability
P(α
f
|α
i
exp(χ), γ
i
) to the coherent state |α
f
i. Fol-
lowing Eq. (7), the initial and final states of the
reversed process are given by the coherent states
|α
f
exp(χ)i and |α
i
i. The symbol
denotes com-
plex conjugation in the Fock basis and arises from the
time reversal operation on a coherent state, T |αi =
|α
i.
The dimensionless parameter χ is the ratio of the
magnitude of quantum fluctuations,
~ω
2
, to the mag-
nitude of thermal fluctuations, k
B
T , and quantifies
the degree to which any regime is quantum mechan-
ical. In the limit in which thermal fluctuations dom-
inate, χ tends to 0, and the prefactor exp(χ) tends
to 1. Consequently, in this limit the reverse process
is the exact time-reversed process; however, in gen-
eral, the pairs of states considered in the forwards
and reverse process differ by an amount determined
by exp(χ).
As derived in more detail in B.1, the energy flow,
Eq. (9), takes the explicit form
˜
E(α
f
, α
i
) = k
B
T
|α
i
|
2
|α
f
|
2
)(1 exp(2χ)
.
(11)
In order to highlight similarities and differences to
the classical situation, it is instructive to explicitly
introduce the difference between the average energy
cost E
+
and gain E
of the forward and reverse
processes. We define the prefactor
q
:
=
˜
E
W
q
(12)
as the ratio between
˜
E and W
q
= (∆E
+
E
)/2.
As shown in B.1, we find that for coherent states the
prefactor takes the form
q (χ) =
1
χ
tanh(χ) (13)
and can be related to the average frequency, ω
T
, of
the oscillator battery in a thermal state, γ(H
B
)
exp
H
B
k
B
T
, at temperature T ,
~ω
T
:
= hH
B
i
γ(H
B
)
=
k
B
T
q(χ)
. (14)
0.5 1.0 1.5 2.0
χ
0.5
1.0
1.5
2.0
H
B
γ
B
ℏω
T
1
2
ℏω
h
2
mλ
th
2
k
B
T
Figure 1: The solid, dark blue line shows the average energy,
~ω
T
, of the oscillator in a thermal state at temperature T
as a function of χ. The red lines indicate the contribution
of thermal (dotted) and quantum (dashed) contributions to
~ω
T
as determined by Eq. (16). The grey line is the classical
limit in which the energy of the harmonic oscillator equals
k
B
T . Energies are given in units of k
B
T .
The coherent state AQC can thus be written as
P
α
f
α
i
exp(χ), γ
i
P
α
i
α
f
exp(χ), γ
f
= exp
F
k
B
T
exp
W
q
~ω
T
.
(15)
In this form, it is analogous to the classical Crooks
equality, Eq. (1), with the difference encoded in ω
T
.
The thermal frequency ω
T
can be understood in
terms of the thermal wavelength λ
th
=
h
p
[50] given
in terms of the momentum p of a particle. This is
a spatial scale that quantifies the regime in which
a thermal system should be treated as classical or
quantum. Broadly speaking, when λ
th
is small the
system is well localised and can be considered clas-
sical; however, quantum effects dominate when it is
larger.
The average energy of the oscillator is composed
of kinetic, potential, and vacuum fluctuation terms.
For an oscillator in a thermal state the kinetic and
potential energies are equal, and with the vacuum
energy E
vac
=
1
2
~ω, it follows that
~ω
T
=
h
2
2
th
+
1
2
~ω . (16)
This equality describes the splitting of ~ω
T
into its
thermal,
h
2
2
th
, and quantum,
1
2
~ω, contributions, as
is shown in Fig. 1.
In the classical limit the thermal wavelength λ
th
is
small and the dominant contribution to ~ω
T
is ther-
mal. In this limit ~ω
T
tends to k
B
T as predicted by
the equipartition theorem and in agreement with the
classical Crooks equality.
In the quantum limit of large λ
th
the thermal con-
tribution tends to zero and vacuum fluctuations dom-
inate. These vacuum fluctuation induce quantum
Accepted in Quantum 2018-02-12, click title to verify 4
corrections to ~ω
T
which increase monotonically with
χ. Consequently, the exponential dependence of the
ratio of transition probabilities on the difference in
energy, W
q
, is suppressed and the probability for the
reverse process is larger than expected classically. We
conclude that the presence of quantum coherence in
a sense makes irreversibility milder. As the dynam-
ics quantified by the AQC are unitary and so fully
reversible, the irreversibility quantified here stems
from the choice in the prepared and measured bat-
tery states [51].
2.4 Squeezed state and cat state Crooks equalities
Analogously to Eq. (15), we now discuss quantum
Crooks equalities that quantify the ratio of transi-
tion probabilities between pairs of squeezed states
of the battery and pairs of cat states of the bat-
tery. Squeezed displaced states, |r, αi
:
= exp(αa
α
a) exp(
r
2
(a
2
a
2
)) |0i, are the unbalanced mini-
mum uncertainty states where for a given squeezing
parameter r, the variance of the oscillator’s position
and momentum are rescaled as x = exp(r)
q
~
2
and p = exp(+r)
q
~
2
. Cat states |Cati |αi +
|βi are superpositions of coherent states |αi and |βi.
The coherent, squeezed, and cat state equalities
differ in the choice of preparation and measurement
states of the battery. The following tables list the
form of these battery states respectively for the for-
wards (|φ
i/f
i) and reverse (|ψ
i/f
i) processes.
Forwards Preparation Measurement
Coherent |α
i
e
χ
i |α
f
i
Squeezed |s
i
, µ
i
i |r
f
, α
f
i
Cat η
|α
i
|
2
χ
|α
i
e
χ
i + η
|β
i
|
2
χ
|β
i
e
χ
i |α
f
i + |β
f
i
Reverse Preparation Measurement
Coherent |α
f
e
χ
i |α
i
i
Squeezed |s
f
, µ
f
i |r
i
, α
i
i
Cat η
|α
f
|
2
χ
|α
f
e
χ
i + η
|β
f
|
2
χ
|β
f
e
χ
i |α
i
i + |β
i
i
In the above table the cat states are unnor-
malized for brevity and we have defined η
χ
:
=
exp
1
2
(1 exp(2χ))
, s
:
= tanh
1
(exp(2χ) tanh(r))
and µ
:
=
exp(χ)(1+tanh(r))
1+exp(2χ) tanh(r)
. We do not show the ex-
plicit expressions of
˜
E for the cat and squeezed
state equalities here as they are reasonably long and
uninstructive. However, Appendix B contains full
derivations and statements of the two equalities.
We compare the additional effect of squeezing to
the coherent state cases in Fig. 2 by plotting q(χ),
1 2 3 4 5
χ
0.2
0.4
0.6
0.8
1.0
q
(a)
1 2 3 4 5
χ
0.2
0.4
0.6
0.8
1.0
q
(b)
Figure 2: We plot q
:
=
˜
E
W
q
as a function of χ for squeezed
displaced states. As indicated by the grey Wigner plot in-
sets, in a) the prepared and measured squeezed displaced
battery state are displaced with respect to position only,
i.e. α
i,f
= <(α
i,f
) and µ
i,f
= <(µ
i,f
); whereas in b)
the state is displaced with respect to momentum only, i.e.
α
i,f
= =(α
i,f
) and µ
i,f
= =(µ
i,f
). In a) the red, blue and
turquoise lines plot r = 1, r = 0 and r = 1 respectively. In
b) the red, blue and turquoise lines plot r = 1, r = 0 and
r = 1 respectively. In both figures the grey line indicates the
classical limit in which q = 1.
the ratio of
˜
E and W
q
as defined in Eq. (12). As
is shown in dark blue, for coherent states q reduces
to 1 in the classical limit where χ tends to 0. How-
ever, q decreases monotonically for increasing χ, and
vanishes as χ tends to infinity.
When the battery is prepared and measured in
states that are displaced with respect to position and
squeezed with respect to momentum, we see in Fig.
2a that q behaves in a similar way to the coherent
state case but with a quicker initial rate of decrease.
When the battery is displaced with respect to posi-
tion and squeezed with respect to position, the be-
haviour of q for large χ again replicates the coherent
state case; however, there is now a regime of low χ
in which q is greater than the classical value of 1.
In Fig. 2b the prepared and measured states are dis-
placed with respect to momentum and we see identi-
cal behaviour to Fig. 2a except the roles of squeezing
with respect to position and momentum are reversed.
Thus we see that for a squeezed battery the prob-
ability for the reverse process is again larger than
expected classically in the limit of large χ. How-
ever, there is now an intermediary regime where the
probability for the reverse process is even more sup-
pressed than in the classical limit. In the presence of
squeezing, irreversibility can thus be stronger than
expected classically.
Cat states are inherently quantum mechanical and
do not reduce to a semi-classical state in the limit of
low χ. Consequently, the analogous plot to Fig. 2 for
cat states is not a straightforward correction to the
Accepted in Quantum 2018-02-12, click title to verify 5
classical limit. The plot is thus uninstructive and so
we do not include it here.
2.5 Quantifying accuracy of the AQC
The approximate nature of the AQC can be quanti-
fied by error bounds. Here we establish three such
error measures, D, and (1 R).
It is instructive to first note that Eq. (8) can be
rewritten as
Z
i
hψ
i
|exp
H
B
k
B
T
|ψ
i
iP (φ
f
|φ
i
, γ
i
)
Z
f
hφ
f
|exp
H
B
k
B
T
|φ
f
iP (ψ
i
|ψ
f
, γ
f
) 0
(17)
using the definition of the generalised energy flow,
Eq. (9), and the definition of the system’s free energy
which can be rephrased as exp(F/k
B
T ) = Z
f
/Z
i
in terms of the initial and final partition functions Z
i
and Z
f
. The degree of approximation is then made
precise by switching from a statement that the mag-
nitude of the left hand side of Eq (17) is approxi-
mately zero, to a statement that it is less than or
equal to some, hopefully small, error bound,
D
:
= |Z
i
hψ
i
|exp
H
B
k
B
T
|ψ
i
iP (φ
f
|φ
i
, γ
i
)
Z
f
hφ
f
|exp
H
B
k
B
T
|φ
f
iP (ψ
f
|ψ
f
, γ
f
) | .
(18)
It is possible to determine analytically the terms
hψ
i
|exp
H
B
k
B
T
|ψ
i
i and hψ
f
|exp
H
B
k
B
T
|ψ
f
i in D
for the coherent, squeezed and cat state equalities by
arranging their approximate forms.
The total error, , is the sum of the initial and final
factorisation errors,
= ||
i
SB
||
1
+ ||
f
SB
||
1
(19)
where ||||
1
= Tr
h
i
denotes the trace norm of
the operator . These factorisation errors
i,f
cap-
ture the extent to which the exponential of the total
Hamiltonian, H
SB
, factorises into the exponential of
effective battery and system Hamiltonians, H
B
and
H
i,f
S
, when acting on the battery states that are mea-
sured in subregions Π
i,f
B
(i.e. |ψ
f
i and |φ
f
i respec-
tively). More accurately, and as derived in Appendix
A, having defined the map
J
K
(σ) = e
K
2k
B
T
σe
K
2k
B
T
, (20)
the factorisation errors are defined as
i
SB
= J
H
SB
(
S
|ψ
i
ihψ
i
|) J
K
i
SB
(
S
|ψ
i
ihψ
i
|)
f
SB
= J
H
SB
(
S
|φ
f
ihφ
f
|) J
K
f
SB
(
S
|φ
f
ihφ
f
|)
(21)
where K
i
SB
= H
i
S
B
+
S
H
B
and K
f
SB
=
H
f
S
B
+
S
H
B
. As calculating involves ex-
ponentiating H
SB
, this generally needs to be done
numerically.
While Eq. (18) enables the AQC to be tested, D
and are not always reliable indicators of the ac-
curacy of the AQC. For example, D decreases as
the transition probabilities P decrease but this need
not correspond to the AQC holding more accurately.
Therefore, we also introduce 1 R as an alternative
means of quantifying its accuracy, with R the ratio
of the two sides of the AQC,
R
:
=
P(φ
f
|φ
i
, γ
i
)
P(ψ
i
|ψ
f
, γ
f
)
exp
˜
E F
k
B
T
!!
1
. (22)
The quantity 1 R equals 0 when the AQC holds
exactly.
3 Proposed Physical Implementation
To make concrete the quantum and thermal physics
governed by the AQC we specify a particular interac-
tion, Eq. (4), between system and battery for which
we expect the equality to hold and the procedure for
testing it. We start by considering an ‘ideal’ inter-
action that replicates the classical Crooks equality
setup most closely. We then take a broader perspec-
tive and suggest alternate potentials that could be
used to verify the physics quantified by the AQC.
3.1 Position dependent level shift
Following [32], we consider a two level system inter-
acting with a harmonic oscillator battery via
V
SB
= σ
z
S
E(x
B
) , (23)
where σ
z
S
= |eihe| |gihg| with |ei and |gi the ex-
cited and ground states of the two level system re-
spectively. E(x
B
) is an energetic level-shift that de-
pends on the position operator, x
B
, of the oscillator.
By choosing E(x) to be constant for x x
i
and for
x x
f
, two distinct effective Hamiltonians, H
i
S
and
H
f
S
in Eq. (4), can be realised for the two level sys-
tem. The independence of the AQC on the form of
V
SB
in Eq. (4) means that the choice of the form of
E(x) in the region x
i
< x < x
f
is arbitrary and so
for simplicity we assume a linear increase,
E(x) =
E
i
x x
i
E
f
E
i
x
f
x
i
(x x
i
) + E
i
x
i
< x < x
f
E
f
x x
f
.
(24)
Accepted in Quantum 2018-02-12, click title to verify 6
α
i
e
-Χ
α
f
H
S
i
=
E
i
σ
S
z
H
S
f
=
E
f
σ
S
z
x
i
x
f
(a) P (α
f
|α
i
exp(χ), γ
i
)
α
i
*
α
f
*
e
-Χ
H
S
i
=
E
i
σ
S
z
H
S
f
=
E
f
σ
S
z
x
i
x
f
(b) P(α
i
|α
f
exp(χ), γ
f
)
Figure 3: Diagrammatic representation of the protocol to
test the coherent state Crooks equality. The blue lines rep-
resent the ground and excited state of the system as a func-
tion of x. The solid Gaussians represent the coherent states
that are prepared at the start of the protocols (red) and the
measurements (grey) that are performed at the end of the
protocols. The dashed lines represent the evolved states.
The harmonic trap that drives the evolution is centered at
the midpoint of x
i
and x
f
.
This energy level splitting profile is shown in Fig.
3. Without loss of generality, we chose to center the
interaction region around x = 0 such that |x
i
| = |x
f
|.
In order to measure the transition probability,
P(φ
f
|φ
i
, γ
i
), for the forwards process one has to be
able to implement the following three steps:
1. The battery is prepared in a state |φ
i
i that is
localised such that hφ
i
|x
B
|φ
i
i < x
i
and the sys-
tem is prepared in a thermal state with respect
to its effective Hamiltonian H
i
S
= E
i
σ
z
S
.
2. The system and battery are allowed to evolve for
some time τ. The time τ is chosen such that the
final battery wavepacket is localised in the region
beyond x
f
in order to ensure that the effective
system Hamiltonian has changed to H
f
S
= E
f
σ
z
S
.
3. The battery is measured to determine whether
its evolved state is |φ
f
i.
The forwards transition probability, P(φ
f
|φ
i
, γ
i
), is
the relative frequency with which the battery is found
in the state |φ
f
i when the steps 1-3 are repeated a
large number of times. We discuss how the prepa-
ration and measurement steps might be practically
implemented in a trapped ion system in Section IV
B.
The procedure to obtain the reverse process transi-
tion probability, P(ψ
i
|ψ
f
, γ
f
), is entirely analogous.
The battery is prepared in the state |ψ
f
i in the region
beyond x
f
and the system is prepared in a thermal
state with respect to H
f
S
= E
f
σ
z
S
. A final measure-
ment, after evolving for time τ, determines whether
the battery is in state |ψ
i
i.
A sketch of the test for the coherent state Crooks
equality is shown in Fig. 3. The procedure for the co-
herent, squeezed, and cat state equalities differ only
in the choice of preparation and measurement states,
as specified by the table in 2.4.
3.2 Numerical calculations of the error bounds
Here we numerically determine the error in the ap-
proximate equalities. While coherent, squeezed and
cat states have support outside of the regions x < x
i
and x > x
f
, and so the AQC is not obeyed perfectly,
we find that there are regimes where the errors fall
below the expected experimental error margins. In
such regimes, which we characterise below, the AQC
is exact for all practical purposes.
We simulated the full quantum state evolution of
the system and harmonic oscillator battery under the
total Hamiltonian with the position dependent level
splitting, E(x), and used this to find the error mea-
sures D and R, as defined in Eq. (18) and Eq. (22).
We further numerically calculated the error bound ,
defined in Eq. (19). As expected we find that D for
the coherent, squeezed and cat states is less than ,
for a wide range of simulated parameters (k
B
T , ω,
α
i
, α
f
, r
i
, r
f
, β
i
, β
f
, E
i
, E
f
, x
i
, x
f
). Fig. 4 presents
some of these results.
For the coherent states and squeezed states, with
real α
i
and real α
f
= α
i
, the inset of Fig. 4a shows
that 1 R tends to 0 as α
i
is increased. As is shown
in red and blue respectively, for a given α
i
squeezing
the position variance of the prepared and measured
states decreases 1 R; while, squeezing the momen-
tum variance increases 1 R. Coherent states with
also a complex displacement α corresponding to fi-
nite momentum result in similar behaviour, as we
have explicitly verified. For the cat state equality,
we firstly consider final states of the form |ψ
i
i
|−αi + |−(α + 1)i and |φ
f
i |αi + |α + 1i. On
doing so, as is shown by the red line in the inset
of Fig. 4b, we find that, as with the coherent state
case, 1 R tends to 0 as α is increased. However, for
the cat states |αi+|−αi, we find that 1R = 0.59
for all α, as shown by blue line. In contrast to the
Accepted in Quantum 2018-02-12, click title to verify 7
x
f
x
i
x
f
x
i
4 6 8
-α
i
= α
f
0.2
0.4
0.6
1 - R
2 3 4 5 6 7 8
-α
i
= α
f
10
-32
10
-22
10
-2
D , ϵ
(a) Coherent State and Squeezed State Equalities
4 6 8
α
0.2
0.4
0.6
1 - R
3 4 5 6 7 8
α
10
-28
10
-23
10
-18
10
-13
10
-8
0.001
D , ϵ
(b) Cat State Equality
Figure 4: The solid and dashed lines of the main plots
show the error measures D and respectively, and the in-
sets plot the error measure 1 R, for evolution under the
total Hamiltonian with the position dependent level split-
ting interaction and a harmonic oscillator battery. In (a)
these are plotted as a function of α
i
= α
f
with the coher-
ent state Crooks equality data shown in grey and squeezed
state Crooks equality data shown in blue (r = 1) and red
(r = 1) (as defined in the table in 2.4). In (b) the data for
cat states prepared and measured straddling the interaction,
|ψ
i
i = |φ
f
i |αi+|−αi, is shown in blue and for cat states
prepared and measured on a single side of the interaction re-
gion, |ψ
i
i |−αi+ |−(α + 1)i and |φ
f
i |αi+ |α + 1i, is
shown in red. In this simulation we have used the following
parameters: x
i
= 4, x
f
= 4, ~ω = 1, E
i
= 1, E
f
= 2 and
k
B
T = 1. The displacement parameters, α
i,f
, and the posi-
tions x
i
and x
f
, are given in units of dimensionless position
X =
p
2~
x.
other states we simulated that are initially displaced
from the origin, these cat states are symmetrically
delocalised such that the average battery position is
0. In particular, there may be a significant probabil-
ity to find the battery in both regions, x < x
i
and
x > x
f
.
We conclude that decreasing the overlap of the pre-
pared and measured battery states from the interac-
tion region increases the accuracy of the AQC. More-
over, the convergence to the exact limit is reached
quickly. For example, we see in Fig. 4a that for coher-
ent states with |α
i
| = |α
f
| 6 the error measures are
very small, i.e. D, , (1 R) < 10
6
. Such discrep-
ancies are lower than experimental error margins and
consequently the approximate nature of these equal-
ities can effectively be disregarded when the prepa-
ration and measurement states are chosen appropri-
ately.
The behaviour of (1 R) in Fig. 4 makes phys-
ical sense. Fundamentally, the approximate nature
of the AQC arises because, by considering the au-
tonomous evolution under a time-independent inter-
acting Hamiltonian, the system never strictly has a
well-defined local Hamiltonian. The crux of the issue
is the battery Hamiltonian induces evolution between
subregions with different effective Hamiltonians, H
i
S
and H
f
S
. As a result, the thermal state of the system
at the start of the forward and reverse protocols is not
well defined. However, when the battery is initially
prepared with support in a single region far from the
interaction region the system can nonetheless prop-
erly thermalise with respect to its initial Hamilto-
nian. In this limit the AQC is exact.
3.3 Alternative position dependent level splittings
Thus far we have specified one particular choice in
E(x), Eq. (24), to induce an effective change in sys-
tem Hamiltonian. Here we outline how alternative
choices in E(x) can also be used to probe the physics
quantified by the AQC.
In Eq. (24) we took E(x) to be constant in the
regions x x
i
and x x
f
to ensure that the initial
and final Hamiltonians are well-defined and so the
AQC holds to a high degree of approximation. How-
ever, there is more flexibility as to the choice of E(x)
if, rather than aiming to directly verify the AQC, we
instead focus simply on detecting the quantum de-
viation from the classical Crooks equality that the
AQC predicts.
The deviation of the AQC from the classical
Crooks equality is encapsulated by the prefactor
q
:
=
˜
E/W
q
, as per Eq. (12), that appears in the
Accepted in Quantum 2018-02-12, click title to verify 8
-5 5
x
-1
1
E(x)
0.5 1.0 1.5 2.0 2.5
χ
0.2
0.4
0.6
0.8
1.0
q
Figure 5: The predicted quantum deviation q deduced from
the transition probabilities between coherent states, using
Eq.(25), for different position dependent level splittings,
E(x). The turquoise circles show the inferred q for the usual
‘flat ends’ potential, the purple squares show q for a sin(x)
2
potential, and the red triangles show q for a purely linear
potential. The blue line is the predicted analytical form of
q =
tanh(χ)
χ
(also sketched in blue in Fig. 2). The grey line
indicates the classical limit in which q = 1. The inset plots
the level splittings simulated. In this simulation we have cho-
sen the following physically plausible parameters, x
i
= 2,
x
f
= 2, E
i
= 1MHz, E
f
= 2MHz, α
i
= 5 and α
f
= 4.
The displacement parameters α
i,f
and the positions, x
i
and
x
f
, are given in units of dimensionless position.
term exp
q
W
q
k
B
T
of the predicted ratio of transition
probabilities. The prefactor q can always be inferred
from the transition probabilities using
q =
k
B
T
W
q
ln
P(φ
f
|φ
i
, γ
i
)
P(ψ
i
|ψ
f
, γ
f
)
exp
F
k
B
T
!
; (25)
however, when the AQC does not hold exactly, it will
deviate from
˜
E/W
q
.
We numerically simulated the evolution of a co-
herent state for different choices of E(x) and used
the obtained transition probabilities to infer q, which
we plot in Fig. 5. The analytic form of q for coher-
ent states, Eq. (13), is indicated by the dark blue
line. As shown by the turquoise circles, we find for
E(x) with the ‘flat ends’, Eq. (24), that q closely
replicates the predicted analytic form. The change
in effective Hamiltonian from E
i
σ
z
S
to E
f
σ
z
S
can be
approximately replicated by using a sin(x)
2
function
and choosing α
i
and α
f
to sit in the troughs and
peaks of the potential where the effective Hamilto-
nian is approximately constant. As shown in purple,
for this choice in E(x) the behaviour of q reasonably
closely replicates the analytic expression. If we abol-
ish the flat ends entirely and instead enact a change
in effective Hamiltonian using a solely linear poten-
tial then, as shown by the red triangles, we find that
q deviates further from the predicted analytic form
but is still of a similar functional form.
As such, we conclude that it should be possible
to detect quantum deviations using sinusoidal and
linear potentials as well as plausibly other potentials
that have yet to be explored. This flexibility as to the
choice of E(x) opens up many possible realisations to
verify the physics quantified by the AQC.
4 Experimental Outline: Trapped Ion and
AC Stark Shift
To make the proposal more concrete we will now out-
line a potential experimental implementation utilis-
ing an ion confined to a linear Paul trap. A pair of
internal electronic levels and the elongated phonon
mode represent the system and battery respectively.
For an experimental verification of the AQC to be
convincing and practical we have the following six
conditions.
1. Practically, we require forwards and reverse
transition probabilities of a measurable order of
magnitude, i.e.
10
2
. P(φ
f
|φ
i
, γ
i
) . 1 10
2
and
10
2
. P(ψ
i
|ψ
f
, γ
f
) . 1 10
2
.
2. To practically simulate a non-trivial thermal sys-
tem state we require non-negligible populations
of both the ground and excited states. This re-
quires the initial and final level splittings, 2E
i
and 2E
f
respectively, to be a similar order of
magnitude to the temperature, i.e.
2E
i
k
B
T 2E
f
.
3. To non-trivially test the AQC we require a mod-
erately strong system-battery interaction. This
is ensured as long as the change in the splitting
of the system energy levels is of a similar order
of magnitude as the trap frequency, i.e.
|2E
f
2E
i
| ~ω .
4. To cleanly test the AQC, the decoherence and
heating rates of the battery need to be consider-
ably less than the trap frequency. If long-lived
internal levels are used to represent the two-level
system then the primary constraint is the heat-
ing rate of the phonon mode battery, ν
therm
, and
thus we need,
ω ν
therm
.
Accepted in Quantum 2018-02-12, click title to verify 9
Figure 6: Sketch of trapped ion proposed implementation.
The ion (red wavepacket) is at the start of the evolution
stage of the forwards protocol. It is displaced in an elongated
trap (grey line) that drives its evolution. An off-resonance
laser beam propagates perpendicularly across one side of the
trap (blue oval). The laser has an intensity profile that is
sloped on the edges and flat through the center (blue line
in oval). The trapped ion experiences a position dependent
AC Stark shift (pair of blue lines) as it travels autonomously
through the laser beam. The red dashed line indicates the
spread of the ion and the black dashed line the length of the
trap. The preparation and measurement stages are sketched
in Fig. 7.
5. The interaction needs to be designed in such
a way that the ion can be prepared in a state
comfortably in the regions x < x
i
and x > x
f
with minimal overlap with the interaction re-
gion, x
i
< x < x
f
.
The simplest means of inducing our proposed
Hamiltonian with the level splitting specified by
Eq. (24) would be to use a pair of Zeeman split inter-
nal electronic levels and a position dependent mag-
netic field, B(x) E(x). However, it is challenging
to obtain the required magnetic field environment for
currently obtainable ion heights and corresponding
motional heating rates.
We instead advocate using an AC Stark shift,
generated by an off resonant laser field, to induce
a position dependent level splitting. While the
magnetic field realisation is perfectly autonomous,
strictly speaking this implementation allows the ex-
change of photons (and thus energy) between the ion
and the light field, and therefore is not perfectly au-
tonomous. However, the energy exchange quickly de-
cays for increasing detuning and for experimentally
achievable Rabi-frequencies and detunings the sys-
tem can nonetheless be considered autonomous for
all practical purposes.
To replicate Eq. (24) as closely as possible we re-
quire the laser beam to have an intensity profile with
a flat region corresponding to x > x
f
. In the region
x < x
i
there is no laser beam and the Hamiltonian is
H
i
S
, while in the region x > x
f
a flat laser beam pro-
file induces a constant AC Stark shift realising H
f
S
.
In the region x
i
< x < x
f
the tail of the laser beam
changes the Hamiltonian from H
i
S
to H
f
S
. In Fig. 6 we
sketch this for a smoothed top hat potential. How-
ever, in general the tails of the intensity profile need
not be smoothed as there are no constraints on the
interaction in the region x
i
and x
f
.
The AC Stark shift typically changes the average
energy of the system as well as the splitting between
the energy levels. This change in average energy and
the smoothed change in gradient at x
i
and x
f
(as
shown in Fig. 6) requires a minor change to V
SB
as
specified in Eq. (23) and (24) but makes no pertinent
difference to the underlying thermal physics being
investigated.
While we focus on the tapered beam implementa-
tion to maintain a closer resemblance to the classi-
cal Crooks equality setup, it is also possible, as dis-
cussed in Section 3.3, to detect the quantum devi-
ation predicted by the AQC using a level splitting
that varies sinusoidally or linearly with position. An
AC Stark shift that varies sinusoidally with position
could be realised using a standing wave generated
by two counter propagating lasers [52, 53]. A mag-
netic field that increases strength linearly with po-
sition B(x) x could be used to realise a Zeeman
shift that increases linearly with position [5456].
4.1 Experimental Parameters
To support the plausibility of our proposal we present
a set of potential parameters that satisfy the five re-
quirements listed above and could as such enable the
AQC to be verified.
We propose using a single
171
Yb
+
ion trapped in
a linear Paul trap
3
with an axial secular frequency
of 0.3 MHz [58]. Such secular frequencies typically
experience heating rates of the order of 40 phonons
per second [59], thereby satisfying requirement 4. Re-
quirements 2 and 3 limit us to a pair of internal levels
with a separation of the order of MHz and thus we
are constrained to using hyperfine levels. We pro-
pose using the F = 1, M
F
= 1 and M
F
= 1 levels
of the
2
S
1/2
ground state. A magnetic field can be
applied to lift the degeneracy of the F = 1 mani-
fold [58] and set E
i
with typical splittings ranging
between approximately 0.5 MHz and 10 MHz.
3
The radial trapping strength must be large enough to
freeze out the radial force on the ion that is induced by the
magnetic component of the laser [57].
Accepted in Quantum 2018-02-12, click title to verify 10
We suggest using a laser field with an intensity of
4.5 Wmm
2
that is red detuned from the 369nm
2
S
1/2
-
2
P
1/2
transition by 0.1×10
14
Hz. Using the methods
and data from [60, 61], this is expected to induce an
AC Stark shift on the M
F
= 1 and M
F
= 1 states
of -0.2 MHz and -1.2 MHz respectively.
The position variance of
171
Yb
+
for an axial secu-
lar frequency of 0.3 MHz is 0.01 µm [46]. Therefore,
to satisfy requirement 5, a top hat potential with the
size of the ‘flat’ region being greater than 0.05 µm
is required. There is some flexibility in how accu-
rately the beam profile needs to be shaped as our
simulations indicate that 1 R scales linearly
4
with
a gradient across the ideally ‘flat’ region of the beam.
Requirement 1 gives rise to a trade off between the
spatial scale on which the beam can be engineered
and the temperature regime that can be probed. As
the distance x
f
x
i
is the minimum the ion must
travel to realise the change in Hamiltonian H
i
S
to
H
f
S
, this sets the minimum possible displacement of
the initial and final states. When the ion oscillates
over large distances, its position variance is relatively
small. In this limit, it is more challenging to choose
the initial and final ion states such that there is a
significant overlap between the measured states and
the evolved states in both the forwards and reverse
processes. It is easier to ensure there is a signifi-
cant overlap in the high temperature regime because
the pair of states prepared in the reverse process are
approximately the time reverse of the pair of states
in the forwards process. Our simulations indicate
that the k
B
T 25 ~ω regime can be probed with
x
f
x
i
= 1µm. The experiment would be feasible
with a larger x
f
x
i
; however, the shorter the spatial
scale on which the ion oscillates, the lower the tem-
perature regime it would be possible to probe and
the greater the χ induced quantum deviations can
be detected
5
.
4.2 Experimental Techniques
The procedure outlined in 3.1 is realisable for trapped
ions using the following currently available experi-
4
The exact constant of proportionality depends on the tem-
perature, the magnitude of the Stark shift and the battery
wavepacket parameters; however, it is generally of the order of
1 R
δI
I
where δI is the change in laser intensity over a dis-
tance of 0.1 µm and I is the maximum laser intensity in that
region. The sensitivity to an intensity gradient is lower for
higher temperatures, smaller Stark shifts and battery states
that are prepared comfortably outside the interaction region.
5
It is possible to realise standing waves and linear mag-
netic field gradients on the scale of 10
2
µm [52, 56]. Thus
the small spatial scales of these alternative approaches could
enable larger values of χ to be more easily probed.
mental techniques.
Preparation of two-level system. A thermal
state of the internal electronic energy levels can be
modeled using ‘pre’-processing [62]. Specifically, a
thermal ‘ensemble’ of N = N
g
+N
e
ions can be mod-
eled by running the protocol on a single ion N
g
and
N
e
times (where N
e
/N
g
= exp(2E
i
/k
B
T )) with the
ion initialised in the ground state and excited states
respectively. The ion can be prepared in the ground
state or excited state by using an appropriate se-
quence of laser and microwave pulses. Simulating the
thermal state in this way makes it straightforward to
ensure that we are in the low temperature limit with
respect to the trapping frequency.
Preparation of oscillator states. The first step
to generate coherent, squeezed and cat states is to
prepare the ion in the motional ground state using
sideband cooling. A coherent state can then be gen-
erated by shifting the trap center non-adiabatically
resulting in the effective displacement of the ion with
respect to the new trapping potential [46] (this is
sketched in Fig. 7). A squeezed displaced state
can similarly be generated from the motional ground
state by first generating a squeezed vacuum state us-
ing a non-adiabatic drop in the trap frequency and
then displacing it using the non-adiabatic shift of the
trap center [46].
Cat states can be generated using laser pulses that
entangle the internal electronic states and motional
states of the ion [46]. Recently a cat state separated
by 259nm has been achieved [63] and with this ap-
paratus it would be possible to generate states sepa-
rated by µm as required here.
Measurement of oscillator states. A measure-
ment to determine the overlap of the final state of the
phonon modes, |ψ
final
i, with some test state, |ψ
test
i,
can be achieved by performing the inverse of the rel-
evant preparation process, U
test
B
, and then measur-
ing to determine whether the battery is in the mo-
tional ground state |0i, i.e. the overlap hψ
test
|ψ
final
i =
h0|U
test
B
|ψ
final
i. For example, to determine the over-
lap of some final state with the coherent state |α
test
i
we would have U
test
B
= D(α
test
)
= D(α
test
) (this is
sketched in Fig. 7).
6
6
There is some flexibility on how fast the measurement
needs to be performed. The inverse process U
test
needs to be
performed at time τ or when the ion returns to the same place
after another complete rotation, i.e. (2n + 1)τ n Z
+
. The
measurement to determine whether the ion is in the motional
ground state need only be performed before there is a substan-
Accepted in Quantum 2018-02-12, click title to verify 11
(a) Preparation
(b) Measurement
Figure 7: Sketch of preparation and measurement stages of
the forwards protocol of the coherent state trapped ion im-
plementation. (a) To prepare a coherent state the ion is first
prepared in the motional ground state of a harmonic trapping
potential, shown here in blue. The potential center is then
shifted non-adiabatically resulting in the effective displace-
ment of the ion with respect to the new trapping potential,
shown here in dark grey. (b) To find the overlap between
the evolved state of the ion (the dashed red wavepacket)
and some test coherent state (the blue Gaussian), the trap
potential is first shifted such that were the evolved state pre-
cisely in the test state, then the evolved state would be in
the ground state with respect to the new trapping potential.
This shifted trapping potential is shown in blue. A projective
measurement is then performed to find the overlap between
the evolved state and the ground state of the new trapping
potential. In both figures the blue oval in the background is
the laser beam that induces the AC Stark shift.
A filtering scheme can be used to perform a pro-
jective measurement on a single ion that answers
the binary question ‘is the phonon in the state n =
m
test
?’ [30, 64]. Choosing m
test
= 0, this method
can be used to determine whether the ion is in the
motional ground state. The filtering scheme uses the
dependence of the Rabi frequency of the red and blue
sidebands on phonon number.
tial probability for the state to be disturbed by decoherence
or heating.
5 Conclusions and Outlook
The AQC is derived from a simple set of physical
principles and in virtue of this the coherent, squeezed
and cat state Crooks equalities are both natural and
general. Specifically, the derivation of the AQC only
assumes that the system and battery evolve under a
microscopically energy conserving and time reversal
invariant unitary and that the system and battery
obey a ‘factorisability’ condition. (This condition,
defined in Appendix A, characterises the extent to
which the system and battery are independent sub-
systems at the start and end of the protocol.) Given
that these are a natural set of assumptions, shared
with other derivations of fluctuation theorems, it is
perhaps intriguing that for coherent states, which are
often viewed as the most classical of the motional
quantum states, we find the AQC can be written in
the satisfyingly compact form of Eq. (15). The de-
pendence of the coherent state Crooks equality on
the thermal de Broglie wavelength is a non-classical
feature, arising from coherent superpositions between
energy eigenstates.
The coherent, squeezed and cat state Crooks equal-
ities can be viewed as setting out quantum correc-
tions to the classical Crooks equality. In particu-
lar, the coherent state Crooks equality is its low-
est order extension to quantum states. The coher-
ent state equality effectively reduces to the classi-
cal Crooks equality in the classical limit but by in-
creasing the ratio of quantum fluctuations to ther-
mal fluctuations it is possible to smoothly interpo-
late from the classical to the quantum regime. In
the general quantum case the probability for the re-
verse process is greater than expected classically and
correspondingly irreversibility is apparently softened.
The squeezed state and cat state Crooks equalities
are higher order extensions to the classical Crooks
equality. The equalities are correspondingly more
complex and there are regimes in which irreversibil-
ity is apparently strengthened. It would be interest-
ing to investigate whether the squeezed and cat state
equalities can be written in a more compact form by
relating them to appropriate thermal wavelengths.
Our proposal can be seen to provide a means to ex-
perimentally test resource theories for quantum ther-
modynamics. As in the protocol the system starts in
a thermal state and energy is globally conserved, the
operation on the battery is a thermal operation for
a system with a changing Hamiltonian [38]. Conse-
quently, the proposed experiment allows us to probe
coherent features of the resource framework, which is
noteworthy as these frameworks [8, 12, 13] have, with
the exception of a few recent developments [65, 66],
Accepted in Quantum 2018-02-12, click title to verify 12
remained abstract.
More broadly, we have taken a highly mathemati-
cal result from within the quantum information theo-
retic approach to thermodynamics and both explored
its physical content and an experimental implemen-
tation. We hope our research encourages more such
attempts to physically ground recent quantum ther-
modynamics theory results.
Acknowledgements
The authors thank Johan Åberg for comments on a
draft, Tom Hebdige, Erick Hinds Mingo and Chris
Ho for numerous helpful discussions and Stephen
Barnett for highlighting the radial force induced on
the ion by an off resonant laser. This research was
supported in part by the COST network MP1209
“Thermodynamics in the quantum regime.” ZH is
supported by Engineering and Physical Sciences Re-
search Council Centre for Doctoral Training in Con-
trolled Quantum Dynamics. SW is supported by the
U.K. Quantum Technology hub for Networked Quan-
tum Information Technologies, EP/M013243/1. DJ
is supported by the Royal Society. JA acknowledges
support from Engineering and Physical Sciences Re-
search Council, Grant EP/M009165/1, and the Royal
Society. FM acknowledges support from Engineer-
ing and Physical Sciences Research Council, Grant
EP/P024890/1.
References
[1] Sai Vinjanampathy and Janet Anders.
Quantum thermodynamics.
Contemporary
Physics, 57(4):545–579, 2016. DOI:
10.1080/00107514.2016.1201896.
[2] Lídia del Rio, Johan Åberg, Renato Renner, Os-
car Dahlsten, and Vlatko Vedral. The thermo-
dynamic meaning of negative entropy. Nature,
474:61, Jun 2011. DOI: 10.1038/nature10123.
[3] David Jennings and Terry Rudolph. Entangle-
ment and the thermodynamic arrow of time.
Phys. Rev. E, 81:061130, Jun 2010. DOI:
10.1103/PhysRevE.81.061130.
[4] Sandu Popescu, Anthony J. Short, and Andreas
Winter. Entanglement and the foundations of
statistical mechanics. Nature Physics, 2:754,
Oct 2006. DOI: 10.1038/nphys444.
[5] Michal Horodecki and Jonathan Oppen-
heim. Fundamental limitations for quantum
and nanoscale thermodynamics. Nature
Communications, 4:2059, Jun 2013. DOI:
10.1038/ncomms3059.
[6] Fernando Brandão, Michał Horodecki, Nelly Ng,
Jonathan Oppenheim, and Stephanie Wehner.
The second laws of quantum thermodynam-
ics. Proceedings of the National Academy of
Sciences, 112(11):3275–3279, 2015. ISSN 0027-
8424. DOI: 10.1073/pnas.1411728112.
[7] Johan Åberg. Truly work-like work ex-
traction via a single-shot analysis. Nature
Communications, 4:1925, Jun 2013. DOI:
10.1038/ncomms2712.
[8] Gilad Gour, David Jennings, Francesco
Buscemi, Runyao Duan, and Iman Marvian.
Quantum majorization and a complete set of en-
tropic conditions for quantum thermodynamics.
Nature Communications, 9(1):5352, 2018. ISSN
2041-1723. DOI: 10.1038/s41467-018-06261-7.
[9] Matteo Lostaglio, David Jennings, and Terry
Rudolph. Description of quantum coherence
in thermodynamic processes requires constraints
beyond free energy. Nature Communications, 6:
6383, Mar 2015. DOI: 10.1038/ncomms7383.
[10] Johan Åberg. Catalytic coherence. Phys.
Rev. Lett., 113:150402, Oct 2014. DOI:
10.1103/PhysRevLett.113.150402.
[11] Kamil Korzekwa, Matteo Lostaglio, Jonathan
Oppenheim, and David Jennings. The extrac-
tion of work from quantum coherence. New
Journal of Physics, 18(2):023045, 2016. DOI:
10.1088/1367-2630/18/2/023045.
[12] D. Janzing, P. Wocjan, R. Zeier, R. Geiss, and
Th. Beth. Thermodynamic cost of reliability
and low temperatures: Tightening landauer’s
principle and the second law. International
Journal of Theoretical Physics, 39(12):2717–
2753, Dec 2000. ISSN 1572-9575. DOI:
10.1023/A:1026422630734.
[13] Fernando G. S. L. Brandão, Michał Horodecki,
Jonathan Oppenheim, Joseph M. Renes, and
Robert W. Spekkens. Resource theory of
quantum states out of thermal equilibrium.
Phys. Rev. Lett., 111:250404, Dec 2013. DOI:
10.1103/PhysRevLett.111.250404.
[14] Nicole Yunger Halpern. Information and
Interaction: Eddington, Wheeler, and the
Limits of Knowledge, chapter Toward Physical
Realizations of Thermodynamic Resource Theo-
ries, pages 135–166. Springer International Pub-
lishing, Cham, 2017. ISBN 978-3-319-43760-6.
DOI: 10.1007/978-3-319-43760-6_8.
[15] Michele Campisi, Peter Hänggi, and Peter
Talkner. Colloquium: Quantum fluctuation re-
Accepted in Quantum 2018-02-12, click title to verify 13
lations: Foundations and applications. Rev.
Mod. Phys., 83:771–791, Jul 2011. DOI:
10.1103/RevModPhys.83.771.
[16] Peter Hänggi and Peter Talkner. The other
qft. Nature Physics, 11:108, Feb 2015. DOI:
10.1038/nphys3167.
[17] C. Jarzynski. How does a system re-
spond when driven away from thermal equilib-
rium? Proceedings of the National Academy
of Sciences, 98(7):3636–3638, 2001. ISSN 0027-
8424. DOI: 10.1073/pnas.081074598.
[18] Gavin E. Crooks. Entropy production fluctua-
tion theorem and the nonequilibrium work rela-
tion for free energy differences. Phys. Rev. E,
60:2721–2726, Sep 1999. DOI: 10.1103/Phys-
RevE.60.2721.
[19] Jan Liphardt, Sophie Dumont, Steven B. Smith,
Ignacio Tinoco, and Carlos Bustamante. Equi-
librium information from nonequilibrium mea-
surements in an experimental test of jarzynski’s
equality. Science, 296(5574):1832–1835, 2002.
ISSN 0036-8075. DOI: 10.1126/science.1071152.
[20] D. Collin, F. Ritort, C. Jarzynski, S. B. Smith,
I. Tinoco Jr, and C. Bustamante. Verification of
the crooks fluctuation theorem and recovery of
rna folding free energies. Nature, 437:231, Sep
2005. DOI: 10.1038/nature04061.
[21] V. Blickle, T. Speck, L. Helden, U. Seifert, and
C. Bechinger. Thermodynamics of a colloidal
particle in a time-dependent nonharmonic po-
tential. Phys. Rev. Lett., 96:070603, Feb 2006.
DOI: 10.1103/PhysRevLett.96.070603.
[22] O.-P. Saira, Y. Yoon, T. Tanttu, M. Möttö-
nen, D. V. Averin, and J. P. Pekola. Test
of the jarzynski and crooks fluctuation rela-
tions in an electronic system. Phys. Rev. Lett.,
109:180601, Oct 2012. DOI: 10.1103/Phys-
RevLett.109.180601.
[23] Shuoming An, Jing-Ning Zhang, Mark Um,
Dingshun Lv, Yao Lu, Junhua Zhang, Zhang-Qi
Yin, H. T. Quan, and Kihwan Kim. Experimen-
tal test of the quantum jarzynski equality with
a trapped-ion system. Nature Physics, 11:193,
Dec 2014. DOI: 10.1038/nphys3197.
[24] Tiago B. Batalhão, Alexandre M. Souza,
Laura Mazzola, Ruben Auccaise, Roberto S.
Sarthour, Ivan S. Oliveira, John Goold, Gabriele
De Chiara, Mauro Paternostro, and Roberto M.
Serra. Experimental reconstruction of work dis-
tribution and study of fluctuation relations in
a closed quantum system. Phys. Rev. Lett.,
113:140601, Oct 2014. DOI: 10.1103/Phys-
RevLett.113.140601.
[25] M. Naghiloo, J. J. Alonso, A. Romito, E. Lutz,
and K. W. Murch. Information gain and loss
for a quantum maxwell’s demon. Phys. Rev.
Lett., 121:030604, Jul 2018. DOI: 10.1103/Phys-
RevLett.121.030604.
[26] H. Tasaki. Jarzynski Relations for Quan-
tum Systems and Some Applications. eprint
arXiv:cond-mat/0009244, September 2000.
URL https://arxiv.org/abs/cond-mat/
0009244.
[27] J. Kurchan. A Quantum Fluctuation Theorem.
eprint arXiv:cond-mat/0007360, July 2000.
URL https://arxiv.org/abs/cond-mat/
0007360.
[28] Peter Talkner and Peter Hänggi. The
tasaki–crooks quantum fluctuation theo-
rem. Journal of Physics A: Mathematical
and Theoretical, 40(26):F569, 2007. DOI:
10.1088/1751-8113/40/26/F08.
[29] Tameem Albash, Daniel A. Lidar, Milad Mar-
vian, and Paolo Zanardi. Fluctuation theorems
for quantum processes. Phys. Rev. E, 88:032146,
Sep 2013. DOI: 10.1103/PhysRevE.88.032146.
[30] Gerhard Huber, Ferdinand Schmidt-Kaler, Se-
bastian Deffner, and Eric Lutz. Employing
trapped cold ions to verify the quantum jarzyn-
ski equality. Phys. Rev. Lett., 101:070403, Aug
2008. DOI: 10.1103/PhysRevLett.101.070403.
[31] P. G. Di Stefano, J. J. Alonso, E. Lutz, G. Falci,
and M. Paternostro. Nonequilibrium thermody-
namics of continuously measured quantum sys-
tems: A circuit qed implementation. Phys. Rev.
B, 98:144514, Oct 2018. DOI: 10.1103/Phys-
RevB.98.144514.
[32] Johan Åberg. Fully quantum fluctuation theo-
rems. Phys. Rev. X, 8:011019, Feb 2018. DOI:
10.1103/PhysRevX.8.011019.
[33] C. Jarzynski and O. Mazonka. Feynman’s
ratchet and pawl: An exactly solvable model.
Phys. Rev. E, 59:6448–6459, Jun 1999. DOI:
10.1103/PhysRevE.59.6448.
[34] Mischa P. Woods, Ralph Silva, and Jonathan
Oppenheim. Autonomous quantum machines
and finite-sized clocks. Annales Henri Poincaré,
20(1):125–218, Jan 2019. ISSN 1424-0661. DOI:
10.1007/s00023-018-0736-9.
[35] Sebastian Deffner and Christopher Jarzynski.
Information processing and the second law of
thermodynamics: An inclusive, hamiltonian ap-
proach. Phys. Rev. X, 3:041003, Oct 2013. DOI:
10.1103/PhysRevX.3.041003.
[36] Juliette Monsel, Cyril Elouard, and Alexia Auf-
fèves. An autonomous quantum machine to mea-
Accepted in Quantum 2018-02-12, click title to verify 14
sure the thermodynamic arrow of time. npj
Quantum Information, (1):59, 2018. ISSN 2056-
6387. DOI: 10.1038/s41534-018-0109-8.
[37] Nicole Yunger Halpern, Andrew J P Garner,
Oscar C O Dahlsten, and Vlatko Vedral. In-
troducing one-shot work into fluctuation rela-
tions. New Journal of Physics, 17(9):095003,
2015. DOI: 10.1088/1367-2630/17/9/095003.
[38] Álvaro M. Alhambra, Lluis Masanes, Jonathan
Oppenheim, and Christopher Perry. Fluctu-
ating work: From quantum thermodynamical
identities to a second law equality. Phys. Rev.
X, 6:041017, Oct 2016. DOI: 10.1103/Phys-
RevX.6.041017.
[39] T. Baumgratz, M. Cramer, and M. B. Ple-
nio. Quantifying coherence. Phys. Rev. Lett.,
113:140401, Sep 2014. DOI: 10.1103/Phys-
RevLett.113.140401.
[40] A. E. Allahverdyan and Th. M. Nieuwen-
huizen. Fluctuations of work from quantum
subensembles: The case against quantum work-
fluctuation theorems. Phys. Rev. E, 71:066102,
Jun 2005. DOI: 10.1103/PhysRevE.71.066102.
[41] Peter Talkner, Eric Lutz, and Peter Hänggi.
Fluctuation theorems: Work is not an observ-
able. Phys. Rev. E, 75:050102, May 2007. DOI:
10.1103/PhysRevE.75.050102.
[42] Peter Talkner and Peter Hänggi. Aspects of
quantum work. Phys. Rev. E, 93:022131, Feb
2016. DOI: 10.1103/PhysRevE.93.022131.
[43] P. Kammerlander and J. Anders. Coherence
and measurement in quantum thermodynam-
ics. Scientific Reports, 6:22174, Feb 2016. DOI:
10.1038/srep22174.
[44] Martí Perarnau-Llobet, Elisa Bäumer, Karen V.
Hovhannisyan, Marcus Huber, and Antonio
Acin. No-go theorem for the characterization of
work fluctuations in coherent quantum systems.
Phys. Rev. Lett., 118:070601, Feb 2017. DOI:
10.1103/PhysRevLett.118.070601.
[45] Matteo Lostaglio. Quantum fluctuation theo-
rems, contextuality, and work quasiprobabili-
ties. Phys. Rev. Lett., 120:040602, Jan 2018.
DOI: 10.1103/PhysRevLett.120.040602.
[46] D. Leibfried, R. Blatt, C. Monroe, and
D. Wineland. Quantum dynamics of single
trapped ions. Rev. Mod. Phys., 75:281–324, Mar
2003. DOI: 10.1103/RevModPhys.75.281.
[47] Leslie E. Ballentine. Quantum mechanics: A
modern development. World scientific, 1998.
[48] Dénes Petz. Sufficient subalgebras and the rela-
tive entropy of states of a von neumann algebra.
Communications in Mathematical Physics, 105
(1):123–131, Mar 1986. ISSN 1432-0916. DOI:
10.1007/BF01212345.
[49] Hyukjoon Kwon and M. S. Kim. Fluctuation
Theorem for an Arbitrary Quantum Process.
ArXiv e-prints, art. arXiv:1810.03150, Octo-
ber 2018. URL https://arxiv.org/abs/1810.
03150.
[50] Charles Kittel and Herbert Kroemer. Thermal
Physics. W. H. Freeman, 1980.
[51] C. L. Clarke and I. J. Ford. Dissipation pro-
duction in a closed two-level quantum system
as a test of the obversibility of the dynam-
ics. ArXiv e-prints, June 2018. URL https:
//arxiv.org/abs/1806.11055.
[52] C. T. Schmiegelow, H. Kaufmann, T. Ruster,
J. Schulz, V. Kaushal, M. Hettrich, F. Schmidt-
Kaler, and U. G. Poschinger. Phase-stable
free-space optical lattices for trapped ions.
Phys. Rev. Lett., 116:033002, Jan 2016. DOI:
10.1103/PhysRevLett.116.033002.
[53] D. von Lindenfels, O. Gräb, C. T. Schmiegelow,
V. Kaushal, J. Schulz, F. Schmidt-Kaler, and
U. G. Poschinger. A spin heat engine coupled to
a harmonic-oscillator flywheel. ArXiv e-prints,
August 2018. URL https://arxiv.org/abs/
1808.02390.
[54] Florian Mintert and Christof Wunderlich. Ion-
trap quantum logic using long-wavelength radi-
ation. Phys. Rev. Lett., 87:257904, Nov 2001.
DOI: 10.1103/PhysRevLett.87.257904.
[55] M. Johanning, A. Braun, N. Timoney, V. El-
man, W. Neuhauser, and Chr. Wunderlich. Indi-
vidual addressing of trapped ions and coupling
of motional and spin states using rf radiation.
Phys. Rev. Lett., 102:073004, Feb 2009. DOI:
10.1103/PhysRevLett.102.073004.
[56] K. Lake, S. Weidt, J. Randall, E. D. Standing,
S. C. Webster, and W. K. Hensinger. Genera-
tion of spin-motion entanglement in a trapped
ion using long-wavelength radiation. Phys. Rev.
A, 91:012319, Jan 2015. DOI: 10.1103/Phys-
RevA.91.012319.
[57] E. A. Hinds and Stephen M. Barnett. Momen-
tum exchange between light and a single atom:
Abraham or minkowski? Phys. Rev. Lett.,
102:050403, Feb 2009. DOI: 10.1103/Phys-
RevLett.102.050403.
[58] S. Weidt, J. Randall, S. C. Webster, K. Lake,
A. E. Webb, I. Cohen, T. Navickas, B. Lekitsch,
A. Retzker, and W. K. Hensinger. Trapped-
ion quantum logic with global radiation fields.
Phys. Rev. Lett., 117:220501, Nov 2016. DOI:
10.1103/PhysRevLett.117.220501.
Accepted in Quantum 2018-02-12, click title to verify 15
[59] S. Weidt, J. Randall, S. C. Webster, E. D.
Standing, A. Rodriguez, A. E. Webb, B. Lek-
itsch, and W. K. Hensinger. Ground-state cool-
ing of a trapped ion using long-wavelength radi-
ation. Phys. Rev. Lett., 115:013002, Jun 2015.
DOI: 10.1103/PhysRevLett.115.013002.
[60] A Roy, S De, Bindiya Arora, and B K Sa-
hoo. Accurate determination of black-body ra-
diation shift, magic and tune-out wavelengths
for the 6s 1/2 to 5d 3/2 clock transition in yb
+. Journal of Physics B: Atomic, Molecular
and Optical Physics, 50(20):205201, 2017. DOI:
10.1088/1361-6455/aa8bae.
[61] Kevin J. Weatherill. A C02 laser lattice
experiment for cold atoms. PhD thesis, Durham
University, 2007.
[62] Nathanaël Cottet, Sébastien Jezouin, Landry
Bretheau, Philippe Campagne-Ibarcq, Quentin
Ficheux, Janet Anders, Alexia Auffèves, Rémi
Azouit, Pierre Rouchon, and Benjamin Huard.
Observing a quantum maxwell demon at
work. Proceedings of the National Academy of
Sciences, 114(29):7561–7564, 2017. ISSN 0027-
8424. DOI: 10.1073/pnas.1704827114.
[63] K. G. Johnson, J. D. Wong-Campos, B. Neyen-
huis, J. Mizrahi, and C. Monroe. Ultrafast cre-
ation of large schrödinger cat states of an atom.
Nature Communications, 8(1):697, 2017. ISSN
2041-1723. DOI: 10.1038/s41467-017-00682-6.
[64] Gerhard Thomas Huber. Quantum
thermodynamics with trapped ions. PhD
thesis, Universität Ulm, 2011.
[65] Niels Lörch, Christoph Bruder, Nicolas Brun-
ner, and Patrick P Hofer. Optimal work ex-
traction from quantum states by photo-assisted
cooper pair tunneling. Quantum Science
and Technology, 3(3):035014, 2018. DOI:
10.1088/2058-9565/aacbf3.
[66] Á. M. Alhambra, M. Lostaglio, and C. Perry.
Heat-Bath Algorithmic Cooling with Thermal
Operations. ArXiv e-prints, July 2018. URL
https://arxiv.org/abs/1807.07974.
[67] Bryan W. Roberts. When we do (and do not)
have a classical arrow of time. Philosophy
of Science, 80(5):1112–1124, 2013. DOI:
10.1086/674001.
[68] B. W. Roberts. Three myths about time reversal
in quantum theory. ArXiv e-prints, July 2016.
URL https://arxiv.org/abs/1607.07388.
[69] Hector Manuel Moya-Cessa and Francisco Soto-
Eguibar. Introduction to Quantum Optics. Rin-
ton Press, 2011.
[70] Christopher Gerry and Peter Knight.
Introductory Quantum Optics. Cambridge
UP, 2005.
A Derivation of the AQC
The AQC is derived in [32] by combining two properties that we will call ‘global invariance’ and ‘factorisability’.
Global invariance is a property of any unitary evolution that is energy conserving and time reversal invariant.
Factorisability is a property that quantifies when two parts of a global system can be considered well defined
subsystems. The battery states considered in the AQC are parameterized by a map known as the Gibbs map.
To derive the AQC we will first introduce these four concepts: the Gibbs map, the quantum time reversal
operation, global invariance and factorisability.
A.1 Gibbs Map
The Gibbs map is a mathematical generalisation of the thermal state that turns out to be a convenient means
of parameterising the battery states.
Definition 1. Gibbs map.
Given a system with Hamiltonian H at temperature T , the action of the Gibbs map, G
H
, on a state ρ is,
G
H
(ρ)
:
=
exp
H
2k
B
T
ρ exp
H
2k
B
T
˜
Z
H
(ρ)
, with
˜
Z
H
(ρ)
:
= Tr
exp
H
k
B
T
ρ
.
(26)
This map arises naturally as a quantum-mechanical version of the Crooks reversal of a Markov process,
and is intimately linked with the Petz recovery map for general quantum states [48, 49]. It also has a range
Accepted in Quantum 2018-02-12, click title to verify 16
of physically natural properties. When the energy of the input state is completely certain, as in an energy
eigenstate |E
k
i, the Gibbs map has no effect and
G
H
(|E
k
ihE
k
|) = |E
k
ihE
k
|.
(27)
However, when the energy of the input state is completely uncertain, as in a maximally mixed state or in an
equal superposition, the state output by the Gibbs map has thermally distributed populations. Additionally,
the map is non-dephasing and affects the energy populations in the same way irrespective of the coherent
properties the state. In this way, the Gibbs map takes a maximally mixed state, ρ
:
=
1
N
, to the thermal
state,
G
H
(ρ ) = γ(H)
:
=
1
Z
H
exp
H
k
B
T
, with
Z
H
:
= Tr
exp
H
k
B
T

,
(28)
and an equal superposition of energy states, |ψ
+
i
:
=
1
N
P
N
k=1
|E
k
i, to the coherent thermal state (a pure
state with the same energy populations as the thermal state),
G
H
(|ψ
+
ihψ
+
|) = |ψ
γ
ihψ
γ
|, with
|ψ
γ
i =
1
Z
H
N
X
k
exp
E
k
2k
B
T
|E
k
i .
(29)
As such we see that the Gibbs maps makes a state crudely ‘as thermal as possible’ subject to the constraints
imposed by the input state. However, this loose claim is not intended to be taken literally but rather as a
signpost towards the map’s deeper physical significance.
It is worth noting that while the Gibbs map is invertible for bounded Hamiltonians; for unbounded Hamilto-
nians, the normalisation constant
˜
Z
H
(ρ) is not guaranteed to be finite and so the map is not strictly invertible.
However, this does not cause any difficulties in our work because for coherent, squeezed and cat states
˜
Z
H
(ρ)
is finite. A full analysis of the invertibility of the Gibbs map is outside the scope of this paper.
A.2 Time Reversal
In broad terms the time reversal operation enacts a motion reversal. This can be visualised as taking a video
of the dynamics of a process and running it in reverse. While there are numerous subtleties involved with
making this intuitive picture precise [67, 68]; in general terms the classical time reversal operation amounts
to the replacement of t by t in all pertinent physical variables describing the dynamics of a system.
As time is a parameter rather than observable in quantum mechanics, the quantum time reversal operation
acts indirectly on t via its action on the position and momentum operators. To maintain correspondence
with classical time reversals, the position operator x is required to remain invariant under a reversal while the
momentum operator p changes sign, i.e. T(x) = x and T(p) = p where T(X) to denotes the time-reversal
of any quantum operator X. In virtue of the time reversal being a symmetry operation, T is additionally
required to be length preserving in the sense that Tr[|ψihψ|] = Tr[T(|ψihψ|)] for any state |ψi. These two
requirements entail that T is an anti-unitary operator [47].
The simplest choice in anti-unitary operation to represent T is complex conjugation. This is the ‘text-
book’ [47] quantum time reversal operation; however, the anti-linearity of conjugation can make it arduous to
work with. On Hermitian operators, and therefore on any quantum state or observable, the transpose oper-
ation is equivalent to complex conjugation. As the AQC quantifies transition probabilities between quantum
states, we are free to use this equivalence and represent T via the transpose.
Definition 2. Quantum time reversal.
The quantum time reversal operation T on any quantum operator X is defined as
T(X)
:
= X
T
,
(30)
where
T
is the transpose operation. It follows that the time reversal operation T on any pure state |ψi is
given by
T |ψi = |ψ
i , (31)
where
denotes complex conjugation.
Accepted in Quantum 2018-02-12, click title to verify 17
As the transpose operation is basis dependent, so are T and T . The physical situation dictates the appro-
priate choice in basis. Specifically, the basis is chosen to ensure that the time reversed battery states drive
the change in Hamiltonian from H
f
S
back to H
i
S
and to ensure that the system and battery Hamiltonian, and
therefore the system and battery evolution operator, are time reversal invariant. For the osccilator battery
and two level system implementation discussed in the main text we take the time reversal operation of the
battery with respect to the Fock basis, {|nihn|}, and the time reversal of the system with respect to the Pauli
z basis, {|eihe|, |gihg|}. This flips the sign of the momentum of any battery oscillator state while leaving its
position unchanged and ensures that the thermal states of the system are time reversal invariant.
The derivation of the AQC makes use of two properties of T. The first of these is the fact that as probabilities
are scalars they are invariant under T. The second is that T reverses the order of products of operators,
T(AB) = T(B)T(A).
A.3 Global Invariance
Global invariance is a property of transition probabilities for a single system with a time reversal invariant
Hamiltonian and an energy conserving and time reversal invariant unitary evolution. As such, it applies to
a large class of systems because most pertinent physical situations are time reversal invariant and all are
ultimately unitary and energy conserving when the total setup considered is made sufficiently large. Global
invariance is the starting point to derive a large family of quantum fluctuation theorems and plays an analogous
role to detailed balance for classical fluctuation theorems.
An energy conserving evolution is here defined to be an evolution that commutes with the Hamiltonian, i.e.
[U, H] = 0. This is stronger than requiring the average energy of the setup to remain constant as a result of
the evolution. In particular, it rules out the creation of non-degenerate superpositions of energy states from
energy eigenstates [11].
For the autonomous evolution of a system under a time independent Hamiltonian H, the strict energy
conservation condition is trivially satisfied as [exp(iHt/~), H] = 0 and the time reversal invariance of the
evolution operator follows from the time reversal invariance of the Hamiltonian, i.e. T(H) = H entails
T (exp(iHt/~)) = exp(iHt/~).
Theorem 1. Global Invariance.
Consider a Hamiltonian H and evolution U such that: T(H) = H, T(U) = U and [U, H] = 0. Global
invariance relates a forwards and reverse transition probability for such a system. In the forwards process,
the system is prepared in a state G
H
(ρ
i
), evolves under U and a binary POVM measurement is performed
with POVM elements {ρ
f
, ρ
f
}. The evolved state, UG
H
(ρ
i
)U
, collapses onto ρ
f
with the probability
P (ρ
f
|G
H
(ρ
i
))
:
= Tr[ρ
f
UG
H
(ρ
i
)U
] . (32)
In the reverse process the state G
H
(T(ρ
f
)) evolves under U and on measurement collapses onto T(ρ
i
) with
the probability
P (T(ρ
i
)|G
H
(T(ρ
f
)))
:
= Tr[T(ρ
i
)UG
H
(T(ρ
f
))U
] . (33)
The ratio of these transition probabilities is,
P (ρ
f
|G
H
(ρ
i
))
P (T(ρ
i
)|G
H
(T(ρ
f
)))
=
˜
Z
H
(T(ρ
f
))
˜
Z
H
(ρ
i
)
. (34)
The proof of global invariance starts with the definition of the forwards transition probability, and makes
use of the definition of T in combination with the fact that T(H) = H and T(U) = U ,
P (ρ
f
|G
H
(ρ
i
))
:
= Tr[ρ
f
UG
H
(ρ
i
)U
] = Tr[T
ρ
f
UG
H
(ρ
i
)U
] = Tr[U
G
H
(T(ρ
i
))UT(ρ
f
)] . (35)
If we now substitute in the definition of the Gibbs map, Eq. (26), we have
P (ρ
f
|G
H
(ρ
i
)) =
1
˜
Z
H
(ρ
i
)
Tr
U
exp
H
2k
B
T
T(ρ
i
) exp
H
2k
B
T
UT(ρ
f
)
,
(36)
Accepted in Quantum 2018-02-12, click title to verify 18
which can be reordered using the cyclic nature of the trace operation and the fact that [U, H] = 0,
P (ρ
f
|G
H
(ρ
i
)) =
1
˜
Z
H
(ρ
i
)
Tr
T(ρ
i
) exp
H
2k
B
T
UT(ρ
f
)U
exp
H
2k
B
T

=
1
˜
Z
H
(ρ
i
)
Tr
T(ρ
i
)U exp
H
2k
B
T
T(ρ
f
) exp
H
2k
B
T
U
.
(37)
The proof is completed by reusing the definition of the Gibbs map and the definition of the reverse transition
probability,
P (ρ
f
|G
H
(ρ
i
)) =
˜
Z
H
(T(ρ
f
))
˜
Z
H
(ρ
i
)
Tr
h
T(ρ
i
)UG
H
(T(ρ
f
))U
i
=
˜
Z
H
(T(ρ
f
))
˜
Z
H
(ρ
i
)
P (T(ρ
i
)|G
H
(T(ρ
f
))) .
(38)
The AQC is derived by applying global invariance to a bipartite setup. By evaluating the forwards and
reverse transition probabilities, P (ρ
f
|G
H
(ρ
i
)) and P (T(ρ
i
)|G
H
(T(ρ
f
))), with respect to a carefully chosen
Hamiltonian H and states ρ
i
and ρ
f
, they can be use to quantify a Crooks-like scenario where an initially
thermal system is driven from equilibrium with a time dependent Hamiltonian with the energy supplied by
a quantum battery. However, for the initial thermal states of the system to be well defined, the system and
battery need to be effectively non-interacting at the start and end of the process. This requirement is captured
by an additional assumption that we call the ‘factorisability’ of the system and battery.
A.4 Factorisability
Factorisability characterises the extent to which multiple, potentially interacting, systems can be considered
independent subsystems. It is defined in terms of the map
J
H
(σ)
:
= exp
H
2k
B
T
σ exp
H
2k
B
T
(39)
for any operator σ. This map is, on any quantum state ρ, the unnormalised version of the Gibbs map, i.e.
J
H
(ρ) =
˜
Z
H
(ρ) G
H
(ρ).
Definition 3. Factorisability.
A composite system with a Hamiltonian H
AB
in the state ρ
A
ρ
B
is factorisable if
J
H
AB
(ρ
A
ρ
B
) = J
H
A
(ρ
A
) J
H
B
(ρ
B
)
(40)
for a pair of Hamiltonians H
A
and H
B
of subsystems A and B respectively.
If a composite system is factorisable then the Gibbs map and the Gibbs map normalisation term, Eq. (26),
decompose into terms acting on the separate subsystems, i.e. if a composite system with a Hamiltonian H
AB
in the state ρ
A
ρ
B
is factorisable then
G
H
AB
(ρ
A
ρ
B
) = G
H
A
(ρ
A
) G
H
B
(ρ
B
) and
˜
Z
H
AB
(ρ
A
ρ
B
) =
˜
Z
H
A
(ρ
A
)
˜
Z
H
B
(ρ
B
) .
(41)
To ensure that system and battery are effectively non-interacting at the start and end of non-equilibrium
process, we require that given an appropriate choice in battery states their Hamiltonian, H
SB
, factorises into
an initial or final system Hamiltonian, H
i
S
or H
f
S
, and battery Hamiltonian, H
B
. Specifically we assume that
the following factorisability condition holds for the system and battery.
Assumption 1. Factorisability of the system and battery.
There exist a pair of pure battery states, |Ψ
i
i and |Ψ
f
i, such that the system and battery Hamiltonian H
SB
factorises as
J
H
SB
(ρ |Ψ
i
ihΨ
i
|) = J
H
i
S
(ρ ) J
H
B
(|Ψ
i
ihΨ
i
|) and
J
H
SB
(ρ |Ψ
f
ihΨ
f
|) = J
H
f
S
(ρ ) J
H
B
(|Ψ
f
ihΨ
f
|)
(42)
where H
i
S
and H
f
S
are Hamiltonians of the system and H
B
is a Hamiltonian of the battery.
Accepted in Quantum 2018-02-12, click title to verify 19
This condition essentially encodes that the system is a genuine system with concrete initial and final
Hamiltonians. The AQC follows directly from global invariance, Eq. (34), when the factorisability condition,
Eq. (42), is obeyed.
A.5 Main Derivation
To see how the AQC follows from the factorisability of the system and battery, we set |Ψ
i
ihΨ
i
| = T(|ψ
i
ihψ
i
|)
and |Ψ
f
ihΨ
f
| = T(|φ
f
ihφ
f
|) and for convenience define
σ
i
= ρ T (|ψ
i
ihψ
i
|) and
T(σ
f
) = ρ T(|φ
f
ihφ
f
|) .
(43)
The factorisability condition, Eq. (42), entails that the Gibbs map acting on σ
i
and T(σ
f
), Eq. (43),
decomposes as
G
H
SB
(σ
i
) = γ
i
|φ
i
ihφ
i
| and
G
H
SB
(T(σ
f
)) = γ
f
|ψ
f
ihψ
f
|
(44)
where we have used the shorthand γ
k
γ(H
k
S
) for k = i, f and with
|φ
i
i T exp
H
B
2k
B
T
|ψ
i
i
|ψ
f
i T exp
H
B
2k
B
T
|φ
f
i
(45)
as in the main text. In this way the factorised Gibbs map generates both the thermal states of the system,
Eq. (6), and the temperature dependent operation that parameterises the battery states, Eq. (7), that, as we
will see, appear in the AQC.
We can use the decomposition of the Gibbs map, Eq (44), to convert a transition probability of the system
and battery to a transition probability of the battery alone. The forwards transition probability P (σ
f
|G
H
(σ
i
)),
Eq (32), of the system and battery reduces to P(φ
f
|φ
i
, γ
i
), the transition probability of the battery from |φ
i
i
to |φ
f
i for a system initialised in the thermal state γ
i
,
P (σ
f
|G
H
SB
(σ
i
)) = Tr
h
(ρ |φ
f
ihφ
f
|) U G
H
SB
(ρ T(|ψ
i
ihψ
i
|)) U
i
=
1
N
hφ
f
|Tr
S
h
U(γ
i
|φ
i
ihφ
i
|)U
i
|φ
f
i
:
=
1
N
P(φ
f
|φ
i
, γ
i
) .
(46)
(The factor
1
N
is a normalisation term from the maximally mixed system state ρ and cancels out in the final
fluctuation theorem.) Similarly, the reverse global transition probability P (T(σ
i
) |G
H
SB
(T(σ
f
)) reduces to
the transition probability of the battery for the reverse process P(ψ
i
|ψ
f
, γ
f
), in which the system is initialised
in the thermal state γ
f
,
P (T(σ
i
) |G
H
SB
(T(σ
f
)) = Tr
h
(ρ |ψ
i
ihψ
i
|)U G
H
SB
(ρ T(|φ
f
ihφ
f
|)) U
i
=
1
N
hψ
i
|Tr
S
h
U(γ
f
|ψ
f
ihψ
f
|)U
i
|ψ
i
i
:
=
1
N
P(ψ
i
|ψ
f
, γ
f
) .
(47)
It also follows from the factorisation of the system and battery that the Gibbs map normalisation terms
decompose as
˜
Z
H
SB
(σ
f
) =
1
N
Z
f
×
˜
Z
φ
f
and
˜
Z
H
SB
(σ
i
) =
1
N
Z
i
×
˜
Z
ψ
i
,
(48)
Accepted in Quantum 2018-02-12, click title to verify 20
where we have used the shorthand Z
k
Z
H
k
S
for k = i, f for the partition functions of the system and
˜
Z
φ
f
˜
Z
H
B
(|φ
f
ihφ
f
|) and
˜
Z
ψ
i
˜
Z
H
B
(T(|ψ
i
ihψ
i
|)) =
˜
Z
H
B
(|ψ
i
ihψ
i
|) as shorthand for the Gibbs map normalisation
terms of the battery.
The AQC now follows directly from substituting the factorised global transition probabilities, Eq. (46) and
Eq. (47), and the factorised Gibbs map normalisation terms, Eq. (48), into global invariance, Eq. (34),
P(φ
f
|φ
i
, γ
i
)
P(ψ
i
|ψ
f
, γ
f
)
=
Z
f
Z
i
˜
Z
φ
f
˜
Z
ψ
i
.
(49)
The derivation is completed by rewriting the ratio of the partition functions in terms of the change in free
energy of the system,
F
:
= F (H
f
S
) F (H
i
S
) with
F (H)
:
= k
B
T ln (Z
H
) ,
(50)
and by analogously defining and the generalised energy flow in terms of the Gibbs map normalisation terms,
˜
E
:
=
˜
E(ψ
i
)
˜
E(φ
f
) with
˜
E(ψ)
:
= k
B
T ln
˜
Z
ψ
= k
B
T ln
hψ|exp
H
B
k
B
T
|ψi
.
(51)
The AQC can thus be stated as
P(φ
f
|φ
i
, γ
i
)
P(ψ
i
|ψ
f
, γ
f
)
= exp
F
k
B
T
exp
˜
E
k
B
T
!
. (52)
Relationship to Ref. [32] The above equality, Eq. (52), is Eq.(38) of Ref. [32] but written in the limit
where factorisability holds exactly and for an initially thermal system and a battery that is prepared in a pure
state. Our battery plays the role of both the energy reservoir and control system in Ref. [32].
A.6 Autonomy and approximate factorisability
To realise an autonomous change in system Hamiltonian we consider a system and a battery with a Hamiltonian
of the form
H
SB
= H
i
S
Π
i
B
+ H
f
B
Π
f
B
+ V
SB
+
S
H
B
, (53)
where Π
i
B
and Π
f
B
are projectors onto two orthogonal subspaces, R
i
and R
f
, of the battery’s Hilbert space,
and V
SB
has support only outside those two subspaces, i.e. (X
S
Π
i
B
)V
SB
= (X
S
Π
f
B
)V
SB
= 0 for any
system operator X
S
. When the battery is initialised in a state in subspace R
i
only and evolves to a final
state in subspace R
f
only, the system Hamiltonian evolves from H
i
S
to H
f
S
. The reverse change in system
Hamiltonian H
f
S
to H
i
S
is similarly realised by a battery initialised in R
i
that evolves to subspace R
f
.
This Hamiltonian, Eq. (53), is exactly factorisable on the states σ
i
and σ
f
, Eq. (43), when: (i) the battery
state T |ψ
i
i has support solely in R
i
and the battery state T |φ
f
i has support solely in R
f
, and (ii) the battery
Hamiltonian does not induce evolution between subregions, i.e. (
B
Π
i
B
)H
B
Π
i
B
= 0 and (
B
Π
f
B
)H
B
Π
f
B
=
0. Under these two restrictions it is straightforward to verify that
J
H
SB
(σ
i
) = exp
H
i
S
Π
i
B
+
S
H
B
2k
B
T
!
(ρ T(|φ
f
ihφ
f
|)) exp
H
i
S
Π
i
B
+
S
H
B
2k
B
T
!
= J
H
i
S
(ρ ) J
H
B
(T(|φ
f
ihφ
f
|))
(54)
and similarly for J
H
SB
(T(σ
f
)). Thus in this limit the AQC holds exactly.
However, to realise an autonomous change in system Hamiltonian, the evolution of the battery between
subspaces R
i
and R
f
must be generated by the unitary evolution of the system and battery under H
SB
, i.e.
by the propagator U = exp(iH
SB
t/~). Consequently, in practise H
B
must induce evolution between the
Accepted in Quantum 2018-02-12, click title to verify 21
subregions R
i
and R
f
and so (
B
Π
i
B
)H
B
Π
i
B
6= 0 and (
B
Π
f
B
)H
B
Π
f
B
6= 0. As a result, the system and
battery factorisability condition, Eq. (42), does not hold exactly for autonomous dynamics and consequently
nor does the AQC.
The AQC is thus strictly only ever approximate. Its approximate nature stems from the fact that when con-
sidering the autonomous dynamics of a system and battery under a time-independent interacting Hamiltonian,
the system never strictly has a well-defined local Hamiltonian. This tension is implicit to all fluctuation the-
orems and arises from the impossibility of perfectly delineating the individual energies of interacting systems.
Nonetheless, the tension is rarely a practical problem because it is generally possible to identify approximately
well defined subsystems with their own concrete Hamiltonians.
The AQC holds to a high degree of approximation whenever it is possible to identify the system and battery
as approximately well defined subsystems with approximately well defined Hamiltonians at the start and end
of the forwards and reverse protocols. This idea is captured by the claim that with an appropriate choice in the
battery states |ψ
i
i and |φ
f
i it is possible to ensure that factorisability holds to a high degree of approximation,
i.e.
J
H
SB
(σ
i
) ' J
H
i
S
(ρ ) J
H
B
(T(|ψ
i
ihψ
i
|)) and
J
H
SB
(T(σ
f
)) ' J
H
f
S
(ρ ) J
H
B
(T(|φ
f
ihφ
f
|)) .
(55)
From which it follows that with an appropriate choice in the battery states |ψ
i
i and |φ
f
i the AQC also holds
to a high degree of approximation
P(φ
f
|φ
i
, γ
i
)
P(ψ
i
|ψ
f
, γ
f
)
' exp
F
k
B
T
exp
˜
E
k
B
T
!
. (56)
The error-bounded AQC, Eq. (18), is an inequality that is obeyed when even when the factorisability
condition does not hold exactly. The extent to which factorisability holds can be quantified by the difference
between left and right hand side of the system and battery factorisability condition, Eq. (55). We thus define
the initial and final factorisation errors,
i
SB
and
f
SB
, as
i
SB
:
= J
H
SB
(
S
|ψ
i
ihψ
i
|) J
H
i
S
(
S
) J
H
B
(|ψ
i
ihψ
i
|) and
f
SB
:
= J
H
SB
(
S
|φ
f
ihφ
f
|) J
H
f
S
(
S
) J
H
B
(|φ
f
ihφ
f
|).
(57)
The proof of the error-bounded AQC proceeds in the same manner as the proof of the AQC except that we
keep track of the error generated from factorisability holding imperfectly. We start with the global invariance
condition for the states σ
i
and σ
f
, Eq. (43),
˜
Z
H
SB
(ρ |ψ
i
ihψ
i
|) P (ρ |φ
f
ihφ
f
||G
H
SB
(ρ T(|ψ
i
ihψ
i
|)))
=
˜
Z
H
SB
(ρ |φ
f
ihφ
f
|) P (ρ |ψ
i
ihψ
i
||G
H
SB
(ρ T(|φ
f
ihφ
f
|))) .
(58)
and substitute in the definition of the factorisability errors
i
SB
and
f
SB
, Eq. 57,
Z
i
˜
Z
ψ
i
P(φ
f
|φ
i
, γ
i
) + Tr[(
S
|φ
f
ihφ
f
|)UT(
i
SB
)U
]
= Z
f
˜
Z
φ
f
P(ψ
i
|ψ
f
, γ
f
) + Tr[(
S
|ψ
i
ihψ
i
|)UT(
f
SB
)U
] ,
(59)
which on rearranging and using the triangle inequality gives
|Z
i
˜
Z
ψ
i
P(φ
f
|φ
i
, γ
i
) Z
f
˜
Z
φ
f
P(ψ
i
|ψ
f
, γ
f
)|
|Tr[(
S
|φ
f
ihφ
f
|)UT(
i
SB
)U
]| + |Tr[(
S
|ψ
i
ihψ
i
|)UT(
f
SB
)U
]| .
(60)
It follows from the Cauchy Schwartz identity that
Tr[(U
(
S
|φ
f
ihφ
f
|)U)
i
SB
] ||T(
i
SB
)||
1
= ||
i
SB
||
1
and
Tr[(U
(
S
|ψ
i
ihψ
i
|)U)
f
SB
] ||T(
f
SB
)||
1
= ||
f
SB
||
1
.
(61)
Thus we are left with the error-bounded AQC,
|Z
i
˜
Z
ψ
i
P(φ
f
|φ
i
, γ
i
) Z
f
˜
Z
φ
f
P(ψ
i
|ψ
f
, γ
f
)| ||
f
SB
||
1
+ ||
i
SB
||
1
:
= .
(62)
Accepted in Quantum 2018-02-12, click title to verify 22
Non-autonomous variant. It is possible to obtain an equality of the same general form as the AQC,
Eq. (52), for a non-autonomous setup where the evolution is induced by an applied unitary U rather than
by evolution under H
SB
, i.e. U 6= exp(iH
SB
t/~). This is in fact the central result of Ref. [32]. In the
non-autonomous variant an additional control system C is introduced and the total Hamiltonian (up to
technicalities concerning time reversals) is of the form,
H
CSB
= |C
i
ihC
i
| H
i
S
B
+ |C
f
ihC
f
| H
f
S
B
+ V
CSB
+
C
S
H
B
(63)
where these control states are orthogonal, hC
f
|C
i
i = 0, and V
CSB
has orthogonal support to both |C
i
i and
|C
f
i, i.e. V
CSB
(|C
i
ihC
i
|X
S
X
B
) = V
CSB
(|C
f
ihC
f
|X
S
X
B
) = 0 for any system and battery operators
X
S
and X
B
. A change in system Hamiltonian H
i
S
to H
f
S
is induced by applying a unitary that evolves the
control from the state |C
i
i to |C
f
i. By requiring the applied unitary to be energy conserving and time reversal
invariant, global invariance holds as before. Furthermore, as the above Hamiltonian, Eq. (63), is effectively
non-interacting as long as the control is prepared in the state |C
i
i, or the state |C
f
i, the factorisability
condition holds exactly. Given that these two conditions are satisfied the rest of the derivation runs identically
to that of the AQC. The resulting quantum Crooks equality holds exactly and differs from the AQC only by
the particular definition of the transition probabilities. It follows that the coherent, squeezed and cat state
equalities can also be used for this non-autonomous setup.
The role of the thermal bath. In this paper we have treated the thermal bath as implicit: a bath is
a means of preparing the system in the thermal state but we do not model it. Effectively this amounts to
assuming that the system and bath are sufficiently weakly interacting at the start of the protocols that they
can be considered well defined independent systems at these times.
More rigorously, we can think of the system discussed in this paper as an enlarged system consisting of
some small system of interest (s) that is driven by a change in Hamiltonian and a thermal bath (b) with a
constant Hamiltonian H
b
, i.e. H
i
S
= H
i
s
H
b
and H
f
S
= H
f
s
H
b
. The AQC can be re-derived explicitly
including the thermal bath if an additional factorisability condition is assumed to hold between the system
and bath. The factorisability condition will hold if their shared Hamiltonian is non-interacting in the regions
R
i
and R
f
, i.e.
H
sbB
= H
i
s
H
b
Π
i
B
+ H
f
s
H
b
Π
f
B
+ V
sbB
+
sb
H
B
. (64)
Under these circumstances, the influence of the bath ‘factorises out’. The resulting equality explicitly quantifies
the transition probabilities of the battery, when a system is driven with a change in Hamiltonian, in the
presence of a thermal bath. This quantum Crooks equality takes exactly the same form as the usual AQC,
Eq. (8), but with the relevant transition probabilities replaced by
P(φ
f
|φ
i
, γ
i
)
:
= hφ
f
|Tr
sb
h
U
sbB
(γ(H
f
s
) γ(H
b
) |φ
i
ihφ
i
|)U
sbB
i
|φ
f
i and
P(ψ
i
|ψ
f
, γ
f
)
:
= hψ
i
|Tr
sb
h
U
sbB
(γ(H
f
s
) γ(H
b
) |ψ
f
ihψ
f
|)U
sbB
i
|ψ
i
i .
(65)
A worthwhile extension to our research would be to explicitly simulate the bath in the experimental proposal
and numerical analysis. One means of doing so would be to investigate the master equation fluctuation theorem
that is proposed in the final version of [32]. Alternatively one could try and incorporate the bath using the
quantum jump approach explored in [36].
A.7 Properties of generalised energy flow
Fundamentally, the term
˜
E appears in the AQC because the energy of a general quantum state is not well
defined and instead some statistical estimate of the states’ energy is required. The precise form of
˜
E is forced
by the fluctuation theorem approach of comparing a forwards and reverse process, i.e. by the structure of the
derivation. In a contrast to Eq. (10) in the main text, we here treat the Hamiltonian H and temperature T
as variables,
˜
E(ψ, H, T )
:
= k
B
T ln
hψ|exp
H
k
B
T
|ψi
, (66)
Accepted in Quantum 2018-02-12, click title to verify 23
to enable us to discuss the properties of the function
˜
E more generally. An analysis of
˜
E provides some
physical support for our interpretation of
˜
E as a quantum generalisation of the energy flow between the
system and the battery. In particular, the following properties hold:
1.
˜
E(ψ, H + δ, T ) =
˜
E(ψ, H, T ) + δ.
2.
˜
E(ψ, λH, T ) = λ
˜
E(ψ, H, λT ).
3. For an energy eigenstate, |E
k
i,
˜
E is simply the associated eigenstate energy,
˜
E(E
k
, H, T ) = E
k
.
4. In the high temperature limit
˜
E tends to the average energy of the state, lim
T →∞
˜
E(ψ, H, T ) = hψ|H |ψi.
5. In the general case
˜
E is less than the average energy:
˜
E(ψ, H, T ) hψ|H |ψi.
6.
˜
E is independent of the the phase of the state
˜
E(ψ, H, T ) =
˜
E
ψe
, H, T
φ <.
Properties 1 and 2 verify that
˜
E scales as one would expect an energy measure to scale. Property 3
is intuitive because when the energy of a state is well-defined there is no need to statistically estimate it.
Moreover, property 3 is required to regain the classical limit. Property 6 ensures that in the absence of
interactions,
˜
E is constant in time which is a desirable condition for a statistical estimate of the energy of a
quantum state.
B Derivation of the coherent, squeezed and cat state AQCs
B.1 Coherent state AQC
The coherent state AQC is derived by considering a harmonic oscillator battery, i.e. H
B
= ~ω
a
a +
1
2
=
P
n
~ω
n +
1
2
|nihn|, and transition probabilities between coherent states. The measured states |φ
f
i and
|ψ
i
i are set to the coherent states |α
f
i and |α
i
i respectively,
|φ
f
i = |α
f
i and
|ψ
i
i = |α
i
i ,
(67)
where |αi
:
= exp( |α|
2
/2)
P
n=0
α
n
n!
|ni. We calculate |φ
i
i and |ψ
f
i using the relation between states |φ
i,f
i
and |ψ
i,f
i as specified in Eq. (7). This requires calculating the effect of applying the temperature dependent
operation exp
H
B
2k
B
T
= exp
χ
a
a +
1
2

where χ =
~ω
2k
B
T
, to a coherent state. Using operator algebra
we find that
|φ
i
i =
1
q
˜
Z
α
i
exp(−|α
i
|
2
/2)
X
n
exp
χ
a
a +
1
2

α
n
i
n!
|ni = |α
i
exp (χ)i and
|ψ
f
i = T
1
q
˜
Z
α
f
exp(−|α
f
|
2
/2)
X
n
exp
χ
a
a +
1
2

α
n
f
n!
|ni
= |α
f
exp (χ)i .
(68)
Additionally, given the definition of
˜
E in Eq. (9), it can be shown using operator algebra that
exp
˜
E
k
B
T
!
=
˜
Z
α
f
˜
Z
α
i
=
exp
(
−|α
f
|
2
/2
)
exp(−|α
f
exp(χ)|
2
/2)
2
exp(−|α
i
|
2
/2)
exp(−|α
i
exp(χ)|
2
/2)
2
= exp
(|α
f
|
2
|α
i
|
2
)(exp(2χ) 1)
.
(69)
The error-bounded coherent state AQC follows from Eq. (67), Eq. (68) and Eq. (69) and the error bounded
AQC, Eq. (18).
Accepted in Quantum 2018-02-12, click title to verify 24
To compare the coherent state Crooks equality to the classical Crooks equality, Eq. 1, we rewrite
˜
E in
terms of the difference between the average energy change of the battery in the forwards and reverse processes,
W
q
:
= (∆E
+
E
)/2, with
E
+
:
= hφ
i
|H
B
|φ
i
i hφ
f
|H
B
|φ
f
i and
E
:
= hψ
f
|H
B
|ψ
f
i hψ
i
|H
B
|ψ
i
i .
(70)
Since energy is globally conserved, W
q
is also the difference between the average energy changes of the system
in the forwards and reverse processes. When the battery is prepared and measured in energy eigenstates, W
q
reduces from an average to a well defined energy change. Specifically, if |ψ
i,f
i = |E
i,f
i then it follows from
Eq. (45) that |φ
i,f
i = |E
i,f
i and therefore
W
q
=
1
2
((E
i
E
f
) (E
i
E
f
))
1
2
(W (W )) = W (71)
where (E
i
E
f
) W is the energy gained (lost) by the system (battery). Beyond this classical limit, W
q
is
a generalisation of W that accounts for coherence by considering the battery’s average energy and allows for
the fact that the AQC quantifies different average energy gains/losses in the forwards and reverse processes
(i.e. that in general E
+
6= E
).
By calculating E
+
and E
for coherent states,
E
+
:
= (exp(2χ)|α
i
|
2
|α
f
|
2
)~ω (72)
E
:
= (exp(2χ)|α
f
|
2
|α
i
|
2
)~ω , (73)
we can rewrite
˜
E for coherent states, Eq. 69, in terms of W
q
,
˜
E = k
B
T (|α
i
|
2
|α
f
|
2
)(1 exp(2χ)) =
1
χ
1 exp(2χ)
1 + exp(2χ)
1
2
(∆E
+
E
) =
1
χ
1 exp(2χ)
1 + exp(2χ)
W
q
.
(74)
Thus the coherent state AQC can be written as
P (α
f
|α
i
exp(χ), γ
i
)
P
α
i
|α
f
exp(χ), γ
f
= exp
F
k
B
T
exp
q (χ)
W
q
k
B
T
,
(75)
where we have defined the quantum prefactor,
q (χ)
:
=
1
χ
1 exp(2χ)
1 + exp(2χ)
=
1
χ
tanh(χ) . (76)
The quantum prefactor, Eq. (76), can be related to the average energy of a thermal battery hH
B
i
γ(H
B
)
:
=
Tr[H
B
γ(H
B
)]. We calculate the explicit algebraic form of hH
B
i
γ(H
B
)
and find that
hH
B
i
γ(H
B
)
=
X
n
~ω(n + 1/2)
exp
~ω(n+
1
2
)
k
B
T
Z
H
B
=
1
2
~ω
1 + exp(2χ)
1 exp(2χ)
=
k
B
T
q(χ)
.
(77)
Rearranging we are left with
q (χ) =
k
B
T
hHi
γ(H
B
)
:
=
k
B
T
~ω
T
. (78)
where we have implicitly defined the ‘thermal frequency’ ω
T
as the average frequency of a harmonic oscillator
in a thermal state. The rewritten coherent state Crooks equality in the main text, Eq. (15), follows from
Eq. (75) and Eq. (78).
Accepted in Quantum 2018-02-12, click title to verify 25
Comment on classical limit. For the coherent state Crooks equality, the classical limit corresponds to
χ =
~ω
2k
B
T
0. In this limit, the core physics of the coherent state Crooks equality is equivalent to that of the
classical Crooks equality; however, there are technical distinctions resulting from their different formulations.
In particular, the coherent state Crooks equality quantifies state transitions of the battery rather than energy
changes of the system and these are related many-to-one. However, in the classical limit the coherent state
equality quantifies transitions between battery states with sharp energies and as energy is globally conserved
this corresponds to the sharp energy changes of the system quantified by the classical Crooks equality. As
such, battery states can be viewed as pointer states to determine the energy flow in or out of the system
during the protocols.
B.2 Cat state Crooks equality
We take a cat state to be an equal superposition of two arbitrary coherent states, |Cati (|αi+ |βi). This is
a more general definition than that sometimes seen in the literature where a cat state is taken to be only a
superposition of a pair of coherent states |αi and |−αi. To derive a cat state AQC the measured states |φ
f
i
and |ψ
i
i are set to
|φ
f
i = |Cat
f
i
:
=
1
q
2 + 2<(exp(
1
2
(|α
f
|
2
+ |β
f
|
2
2β
f
α
f
)))
(|α
f
i + |β
f
i) and
|ψ
i
i = |Cat
i
i
:
=
1
q
2 + 2<(exp(
1
2
(|α
i
|
2
+ |β
i
|
2
2β
i
α
i
)))
(|α
i
i + |β
i
i) ,
(79)
where the normalisation term is calculated using the fact that the overlap between two coherent states |αi
and |βi is given by hβ|αi = exp(
1
2
(|α|
2
+ |β|
2
2β
α)). The prepared states |φ
i
i and |ψ
f
i are calculated
using the relation between states |φ
i,f
i and |ψ
i,f
i as specified in Eq. (7). Using operator algebra we find that
|φ
i
i
:
= exp
χ
a
a +
1
2

|Cat
i
i
exp
χa
a
|α
i
i + exp
χa
a
|β
i
i
=
1
q
˜
Z
Cat
i
η
|α
i
|
2
χ
|exp(χ)α
i
i + η
|β
i
|
2
χ
|exp(χ)β
i
i
and
|ψ
f
i =
1
q
˜
Z
Cat
f
η
|α
f
|
2
χ
|exp(χ)α
f
i + η
|β
f
|
2
χ
|exp(χ)β
f
i
(80)
where we have defined
η
χ
:
= exp
1
2
(1 exp(2χ))
(81)
and the normalisation term
˜
Z
Cat
i
(and equivalently for
˜
Z
Cat
f
) takes the form
˜
Z
Cat
i
=
η
2|α
i
|
2
χ
+ η
2|β
i
|
2
χ
+ η
|α
i
|
2
+|β
i
|
2
χ
2<
exp
1
2
exp(2χ)(|β
i
|
2
+ |α
i
|
2
2β
i
α
i
)

2 + 2<(exp(
1
2
(|α
i
|
2
+ |β
i
|
2
2β
i
α
i
)))
.
(82)
The normalisation terms
˜
Z
Cat
i
and
˜
Z
Cat
f
also provide the exp
˜
E
k
B
T
term. It follows from the definition of
˜
E in Eq. (9) that
exp
˜
E
k
B
T
!
=
hφ
f
|exp (2χ) |φ
f
i
hψ
i
|exp (2χ) |ψ
i
i
=
˜
Z
Cat
f
˜
Z
Cat
i
.
(83)
As such, the cat state AQC follows from Eq.(B12 -B16). (We do not state the explicit form of exp
˜
E
k
B
T
in
terms of α
i,f
, β
i,f
, χ and η
χ
as the expression does not simplify.) The error-bounded cat state Crooks equality
follows from the error bounded autonomous Crooks equality, Eq. (62), with
˜
Z
i,f
defined in Eq. (82).
Accepted in Quantum 2018-02-12, click title to verify 26
B.3 Squeezed state Crooks equality
The squeezed state AQC quantifies transition probabilities between battery oscillator states that are not only
displaced but also squeezed. As such, the measured states |φ
f
i and |ψ
i
i are set to the squeezed displaced
states |α
f
, r
f
i and |α
i
, r
i
i respectively, i.e.
|φ
f
i = |α
f
, r
f
i = D(α
f
)S(r
f
) |0i and
|ψ
i
i = |α
i
, r
i
i = D(α
i
)S(r
i
) |0i ,
(84)
where we have introduced the displacement operator D(α) = exp(αa
α
a), the squeeze operator S(r) =
exp(
r
2
(a
2
a
2
)), and for simplicity we assume from the outset that the squeezing parameters r
i
and r
f
are
real. We calculate |φ
i
i and |ψ
f
i using the relation between states |φ
i,f
i and |ψ
i,f
i as specified in Eq. (7). This
amounts to calculating the effect of applying the operator exp
χa
a
to a squeezed displaced state, i.e.
|φ
i
i =
1
q
˜
Z
α
i
,r
i
exp
χ
a
a +
1
2

D(α
i
)S(r
i
) |0i D(µ
i
)S(t
i
) |0i = |µ
i
, t
i
i and
|ψ
f
i =
1
q
˜
Z
α
f
,r
f
exp
χ
a
a +
1
2

D(α
f
)S(r
f
) |0i D(µ
f
)S(t
f
) |0i = |µ
f
, t
f
i .
(85)
We introduce µ
i,f
and t
i,f
to denote the displacement and squeezing parameters respectively of the prepared
states. (That the states |φ
i
i and |ψ
f
i are also squeezed displaced states is a non-trivial result of the calcula-
tion.) To derive the AQC we will also need to calculate the exp
˜
E
k
B
T
term. It follows from the definition of
˜
E in Eq. (9) that this can be written as
exp
˜
E
k
B
T
!
=
˜
Z
α
f
,r
f
˜
Z
α
i
,r
i
=
(exp(χa
a)D(α
f
)S(r
f
) |0i)
exp(χa
a)D(α
f
)S(r
f
) |0i
(exp(χa
a)D(α
i
)S(r
i
) |0i)
exp(χa
a)D(α
i
)S(r
i
) |0i
.
(86)
To derive the AQC we thus need to calculate µ, s and
˜
Z
α,r
.
B.3.1 Useful Identities
To calculate µ, s and N
α,r
we will make use of the following equalities:
1. exp(ma
a) exp(na
) exp(ma
a) = exp(n exp(m)a
)
Proof. This is derived using the Taylor expansion of the exponential followed by the Hadamard lemma [69],
exp(ma
a) exp(na
) exp(ma
a) =
X
k
(n exp(ma
a)a
exp(ma
a))
k
k!
=
X
k
(na
exp(m))
k
k!
= exp(n exp(m)a
) .
2. exp(ma) exp(na
) = exp(mn) exp(na
) exp(ma)
Proof. This is derived using the Baker-Campbell-Hausdorff formula [69] followed by the Zassenhaus formula [69],
exp(ma) exp(na
) = exp(ma + na
) exp
mn
2
= exp(na
) exp(ma) exp(mn) .
3. exp(ma
2
) exp(na
) = exp(mn
2
) exp(na
) exp(ma
2
) exp(2mna)
Proof. This is again derived using the Baker-Campbell-Hausdorff formula followed by the Zassenhaus formula,
exp(ma
2
) exp(na
) = exp
mn
2
6
exp(ma
2
+ mna + na
) = exp(mn
2
) exp(na
) exp(ma
2
) exp(2mna)
Accepted in Quantum 2018-02-12, click title to verify 27
4. exp
1
2
(ma
2
+ na
2
)
=
1
p
cos
(
mn
)
exp
1
2
p
n
m
tan(
mn)a
2
exp
ln (cos(
mn)) a
a
exp
1
2
p
m
n
tan(
mn)a
2
Proof. To derive this equality it is helpful to first introduce the operators K
+
:
=
1
2
a
2
, K
:
=
1
2
a
2
, K
z
:
=
1
4
(aa
+ a
a). These operators obey the commutation relations [K
+
, K
] = 2K
z
and [K
z
, K
±
] = ±K
±
. Next,
we introduce the ansatz
f
:
= exp((mK
+
+ nK
)t) = exp(pK
+
) exp(qK
z
) exp(sK
) .
(87)
To find the coefficients p, q and s we differentiate f with respect to t,
f
0
= (mK
+
+ nK
)f = (p
0
K
+
+ q
0
exp(pK
+
)K
z
exp(pK
+
) + s
0
exp(pK
+
) exp(qK
z
)K
exp(qK
z
) exp(pK
+
)) f.
(88)
To simplify this expression we use the Hadamard lemma to obtain the following relations,
exp(pK
+
)K
z
exp(pK
+
) = K
z
pK
+
and
exp(pK
+
) exp(qK
z
)K
exp(qK
z
) exp(pK
+
) = exp(pK
+
) exp(q)K
exp(pK
+
) = exp(q)(K
2pK
z
+ p
2
K
+
) .
(89)
Eq. (89) can be substituted into the right hand side of Eq. (88), from which it then follows that
mK
+
+ nK
= p
0
K
+
+ q
0
(K
z
+ pK
+
) + s
0
exp(q)(K
2pK
z
+ p
2
K
+
) .
(90)
Equating coefficients in Eq. (90) we obtain a set of differential equations,
p
0
q
0
p + s
0
exp(q)p
2
= m
q
0
2ps
0
exp(q) = 0
s
0
exp(q) = n ,
which can be solved, subject to the constraint that p(0) = q(0) = s(0) = 0, to find that
p =
m
n
tan
mnt
q = 2ln
cos
mnt

s =
n
m
tan
mnt
.
(91)
Identity 4 is obtained by substituting this solution, Eq. (91), back into the ansatz, Eq. (87).
5. exp
1
2
ma
2
+ na
2

=
p
cos(
mn) exp
1
2
p
m
n
tan(
mn)a
2
exp
ln (cos(
mn)) a
a
exp
1
2
p
n
m
tan(
mn)a
2
Proof. The proof here follows the same method as for identity 4. Using the ansatz f = exp((mK
+
+ nK
)t) =
exp(pK
) exp(qK
z
) exp(sK
+
) it is possible to derive the same differential equations as in Eq. (91) but with
q q.
6. exp
ma
2
exp
na
2
=
1
14mn
exp
n
14mn
a
2
exp
ln
1
14mn
a
a

exp
m
14mn
a
2
Proof. This is proven using identities 4 and 5. Let p =
1
2
p
m
n
tan(
mn), q =
1
2
p
n
m
tan(
mn) and exp(2M) =
cos(
mn)
2
=
1
14pq
, it follows that
M exp(pa
2
) exp(ln(M )a
a) exp(qa
2
) =
1
M
exp(qa
2
) exp(ln(M )a
a) exp(pa
2
) ,
which can be rearranged to give
exp
pa
2
exp
qa
2
= M exp(ln(Ma
a) exp(qa
2
) exp(pa
2
) exp(ln(M )a
a)
= M exp(q exp(2M)a
2
) exp(2ln(M a
a) exp(p exp(2M )a
2
) .
The substitution p p then gives identity 6.
Accepted in Quantum 2018-02-12, click title to verify 28
B.3.2 Main Derivation
We are now in a position to calculate the effect of the operator exp
χa
a
on a squeezed displaced state,
exp
χa
a
D(α)S(r) |0i = exp
χa
a
exp(αa
α
a) exp
r
2
a
2
a
2
|0i . (92)
We first use the standard factorised form of the displacement and squeeze operators [70],
D(α)S(r) |0i =
exp
1
2
|α|
2
p
cosh(r)
exp
αa
exp (α
a) exp
1
2
tanh(r)a
2
|0i (93)
to rewrite Eq. (92),
= exp
χa
a
exp
1
2
|α|
2
p
cosh(r)
exp
αa
exp (α
a) exp
1
2
tanh(r)a
2
|0i .
(94)
Eq. (94) is then rewritten using identity 3,
=
exp
1
2
|α|
2
p
cosh(r)
exp
χa
a
exp
αa
exp
1
2
tanh(r)α
2
exp
α
tanh(r)a
exp
1
2
tanh(r)a
2
|0i ,
(95)
followed by identity 1,
=
exp(
1
2
(|α|
2
α
2
tanh(r)))
p
cosh(r)
exp
exp (χ) (α + α
tanh(r))a
exp
1
2
exp (2χ) tanh(r)a
2
|0i .
(96)
By comparing Eq. (96) with itself in the limit that χ = 0, i.e. the limit in which exp
χa
a
is not applied
and Eq. (96) is simply a rewritten version of a squeezed displaced state, we identify the following modified
displacement and squeeze coefficients µ and s,
<(µ) = <(α)
exp (χ) (1 + tanh(r))
1 + exp (2χ) tanh(r)
,
=(µ) = =(α)
exp (χ) (1 tanh(r)
1 exp (2χ) tanh(r))
and
tanh(s) = exp (2χ) tanh(r) .
(97)
It remains to calculate the exp
˜
E
k
B
T
term. This amounts to calculating,
˜
Z
α,r
=
exp
χa
a
D(α)S(r) |0i
exp
χa
a
D(α)S(r) |0i
(98)
The above equation can be immediately simplified using Eq. (96)
=
exp
(|α|
2
<(α)
2
tanh(r))
cosh(r)
h0|exp
ma
2
exp (na) exp
n
a
exp
m
a
2
|0i
(99)
where
m =
1
2
exp (2χ) tanh(r) and
n = exp (χ) (α
+ α tanh(r)) .
(100)
Accepted in Quantum 2018-02-12, click title to verify 29
Using ‘Useful Identities’ 6, 3, 1, and 2 in succession this can be rewritten to give,
h0|exp
ma
2
exp (na) exp
m
a
exp
n
a
|0i
=
1
r
h0|exp (na) exp
m
exp(2r)
a
2
exp
ln (exp(2r)) a
a
exp
m
exp(2M)
a
2
exp
n
a
|0i
=
1
M
exp
2<(n
2
m)
exp(2M)
!
h0|exp (na) exp(ln(exp(2M))a
a) exp
n
a
|0i
=
1
M
exp
2<(n
2
m)
exp(2M)
!
h0|exp
n
M
a
exp
n
M
a
|0i
=
1
M
exp
2<(n
2
m)
exp(2M)
!
exp
|n|
2
exp(2M)
!
(101)
where M =
1
14|m|
2
. It follows that,
˜
Z
α,r
:
=
exp((|α|
2
<(α
2
) tanh(r)))
cosh(r)
p
1 tanh(r)
2
exp (4χ)
exp
exp (2χ) (|α|
2
(1 + tanh(r)
2
) + 2 tanh(r)<(α
2
)) exp (2χ) tanh(r)
<(α
2
) + <(α
2
) tanh(r)
2
+ 2|α|
2
tanh(r)
1 tanh(r)
2
exp (4χ)
!
,
(102)
which can be simplified if we assume that the displacement parameters are real, α
i,f
= <(α
i,f
),
˜
Z
α,r
:
=
exp
−|α|
2
(1 tanh(r)
cosh(r)
p
1 tanh(r)
2
exp (4χ)
exp
|α|
2
(1 + tanh(r))
2
(1 tanh(r) exp (2χ)) exp (2χ)
1 tanh(r)
2
exp (4χ)
!
.
(103)
This completes the derivation of the squeezed state AQC. The error-bounded squeezed state Crooks equality
follows from the error bounded AQC, Eq. (18).
C Dynamics of proposal
Analysis of the dynamics of the harmonic oscillator state under the proposed interaction, Eq. (23), verifies
that the oscillator behaves as a battery. The interaction is diagonal in the system energy eigenbasis and thus
the total Hamiltonian can be written in the form
H
SB
= |eihe| H
e
B
+ |gihg| H
g
B
, (104)
where H
e
B
= H
B
+ E(x
B
) and H
g
B
= H
B
E(x
B
). Consequently, H
SB
does not induce transitions between
the system energy levels. At the start of the forwards protocol the two-level system and oscillator are prepared
in the state
ρ
SB
(0) = (p
e
|eihe|+ p
g
|gihg|) |φ
i
ihφ
i
| , (105)
where
p
e
p
g
= exp
2E
i
k
B
T
. The system and battery then evolve under H
SB
, for some time t, to the state
ρ
SB
(t) = p
e
|eihe| exp(itH
e
B
/~) |φ
i
ihφ
i
|exp(itH
e
B
/~) + p
g
|gihg| exp(itH
g
B
/~) |φ
i
ihφ
i
|exp(itH
g
B
/~).
(106)
As such the oscillator state evolves into two non-equally weighted components, |φ
e
i
(t)i
:
= exp(itH
e
B
/~) |φ
i
i
and |φ
g
i
(t)i
:
= exp(itH
g
B
/~) |φ
i
i, correlated with the system being prepared in the excited and ground state
respectively, i.e.
ρ
SB
(t) = p
e
|eihe| |φ
e
B
(t)ihφ
e
B
(t)| + p
g
|gihg| |φ
g
B
(t)ihφ
g
B
(t)| .
(107)
The oscillator state in the forwards protocol is prepared such that its average position is initially in the
region of narrow splitting, i.e. hφ
i
|x
B
|φ
i
i < x
i
. In this case |φ
e
B
(t)ihφ
e
B
(t)| travels up a potential hill and by
energy conservation its average energy decreases by an amount E
f
E
i
as it travels from x
i
to x
f
. Conversely,
|φ
g
B
(t)ihφ
g
B
(t)| travels down a potential hill and its average energy increases by an amount E
f
E
i
.
Accepted in Quantum 2018-02-12, click title to verify 30
(a) Initial State (b) Evolved State
Figure 8: These are plots of the Wigner function of the a) initial oscillator state and b) final oscillator state after evolution
under our proposed Hamiltonian. In these plots the following parameters are chosen somewhat arbitrarily: m = 1, ~ω = 1,
E
i
= 1, E
f
= 21 and the interaction region extends from x
i
= 2 to x
f
= 2. The oscillator starts in a coherent state
centered around x = 9. These plots correspond to the case in which the system is prepared in the excited state and so
show the component of the wavepacket that travels up the potential hill: |φ
e
i
(t)i
:
= exp(iH
e
B
t) |exp(χ)αi. The oscillator
wavepacket is slowed as it evolves through the potential hill and the interaction squeezes the coherent state and leave small
residual ripples trailing behind the bulk of the Wigner function. The Wigner function is negative in the troughs of these
ripples.
Numerical simulations reveal that the oscillator state wavepacket is distorted as it evolves through the
interaction region. In Fig. 8 we see that a coherent oscillator states is squeezed and interference ‘ripples’ are
generated by the interaction. The Wigner function of these ‘ripples’ is in places negative. Negative values of
a quasi probability distribution are hard to explain classically so this is a signature of quantum mechanical
phenomena. Furthermore, numerical simulations indicate that when the final state is approximated by a
coherent state with the same average position and momentum as the complete final wavepacket the error-
bounded coherent state AQC does not always hold. As such, we see that even when the battery is prepared in
a coherent state, the most classical of the motional quantum states, it does not remain in this approximately
classical state. Moreover, the quantum distortion effects are essential to the coherent state AQC being obeyed.
Accepted in Quantum 2018-02-12, click title to verify 31