You are on page 1of 11

Efficiency statistics of a quantum Otto cycle

Zhaoyu Fei,1 Jin-Fu Chen,1, 2, 3 and Yu-Han Ma1, ∗


1
Graduate School of China Academy of Engineering Physics,
No. 10 Xibeiwang East Road, Haidian District, Beijing, 100193, China
2
Beijing Computational Science Research Center, Beijing 100193, China
3
School of Physics, Peking University, Beijing, 100871, China
The stochastic efficiency [G. Verley et al., Nat. Commun. 5, 4721 (2014)] was introduced to
evaluate the performance of energy-conversion machines in micro-scale. However, such an efficiency
generally diverges when no heat is absorbed while work is produced in a thermodynamic cycle. As
arXiv:2109.12816v1 [cond-mat.stat-mech] 27 Sep 2021

a result, any statistical moments of the efficiency do not exist. In this study, we come up with a
different version of the definition for the stochastic efficiency which is always finite. Its mean value is
equal to the conventional efficiency, and higher moments characterize the fluctuations of the cycle.
In addition, the fluctuation theorems are re-expressed via the efficiency. For working substance
satisfying the equipartition theorem, we clarify that the thermodynamic uncertainty relation for
efficiency is valid in an Otto engine. To demonstrate our general discussions, the efficiency statistics
of a quantum harmonic-oscillator Otto engine is systematically investigated. The probability that
the stochastic efficiency surpasses the Carnot efficiency is explicitly obtained. This work may shed
new insight for optimizing micro-machines with fluctuations.

PACS numbers:

I. INTRODUCTION ill-defined [9]. Thus, one fails to evaluate the performance


of the heat engine by the moments of this version of
stochastic efficiency.
For a heat engine operating between a hot and cold
reservoir, the conventional efficiency is defined by the
ratio of the output work and the heat absorbed from
the hot reservoir, which characterizes the performance
of the engine. As the size of the engine decreases, the To avoid such weirdness, meanwhile to evaluate the
thermal fluctuations [1, 2] and quantum fluctuations [3, fluctuations in a practical heat engine, we come up with
4] become more significant. From the point view of a different version of the stochastic efficiency. Then,
stochastic thermodynamics, the work, heat, and entropy the fluctuation theorems [1–4, 12, 13] is re-expressed via
of microscopic systems are all stochastic quantities. the stochastic efficiency. Moreover, the thermodynamic
Hence, at a microscopic level, it is natural to expect uncertainty relation (TUR) [14, 15] for efficiency is
that the efficiency, introduced to evaluate the ability of investigated. For working substance satisfying the
energy conversion for various thermal machines, is also a equipartition theorem, we obtain the TUR for a quantum
stochastic quantity. Otto cycle in the quasistatic limit (a general proof) and
in a finite-time Otto cycle (numerical simulations). As a
Recently, the stochastic efficiency, defined as the ratio
specific example, we apply our version of the stochastic
of the stochastic output work and the stochastic heat
efficiency to study a quantum harmonic-oscillator Otto
absorbed from the hot reservoir in a cycle, has been
cycle. We find that both the probability that the
widely studied for classical heat engines [5–9]. This
stochastic efficiency surpasses the Carnot efficiency and
stochastic efficiency is also applied to a quantum Otto
the probability that the stochastic efficiency is negative
cycle in Refs. [10, 11]. However, such a definition of
increase as the temperatures of the reservoirs decrease.
the stochastic efficiency seems weird for the following
three reasons: (1) The mean value of the stochastic
efficiency is not equal to the conventional efficiency in
general. In contrary, the mean values of stochastic
work, heat, and entropy are equal to their counterparts This paper is arranged as follows. In Sec. II,
in the conventional thermodynamics; (2) The efficiency we introduce the quantum Otto cycle and the joint
approaches infinity with a non-zero probability. Such a distribution of input work and absorbed heat from the
result is due to the possibility that no heat is absorbed hot reservoir. In Sec. III, a different version of stochastic
from the hot reservoir while work is produced in one efficiency is given. The fluctuation theorems and the
realization of the cycle [11]; (3) Due to the divergent TUR are also re-expressed via the stochastic efficiency.
efficiency distribution, any moments of the efficiency are In Sec. IV, we demonstrate our general discussions in
a quantum Otto cycle with the harmonic oscillator
being the working substance. And we systematically
investigate the statistics of the stochastic efficiency.
∗ Electronic address: yhma@gscaep.ac.cn Section V is the summary and discussion.
2

II. THE JOINT DISTRIBUTION OF WORK


AND HEAT IN A QUANTUM OTTO CYCLE Cold Heat Bath

As illustrated in Fig. 1, we consider a quantum Otto


cycle which involves four strokes: two adiabatic processes Cold Isochoric
and two isochoric processes [16, 17]. In the adiabatic
Adiabatic Adiabatic
compression (expansion) process, the Hamiltonian of the
working substance is changed from H(λ0 ) to H(λ1 ) (from Compression Expansion
H(λ1 ) to H(λ0 )) during time τc (τh ) through a time- Hot Isochoric
dependent parameter λ. In the two isochoric processes,
during time th (tc ), the working substance contacts a
hot (cold) reservoir at the inverse temperature βh (βc ) Hot Heat Bath
with fixed λ. For simplicity, we assume that a complete
thermalization is achieved in the two isochoric processes.
Namely, the working substance is thermal equilibrium
with the corresponding reservoir at the end of each FIG. 1: The schematic of a finite-time quantum Otto cycle.
isochoric process.
The stochastic work and heat in a quantum Otto
cycle are defined under the two-point measurement III. STOCHASTIC EFFICIENCY
scheme [10]. At time t = 0, τc , τc + th , τc + th + τh
(t = 0 is the initial time of the adiabatic compression A. Definition
process), we apply the projective measurements of
energy on the working substance according to the We define the stochastic efficiency of a (classical)
corresponding instantaneous Hamiltonian. Then, the quantum heat engine as
stochastic work wc (we ) in the adiabatic compression
(expansion) process, and the stochastic absorbed heat w
η=− , (3)
q in the hot isochoric process are defined as hqi
where h·i denotes the mean value over numerous
1
wc = Em − En0 measurements, i.e.,
q = Ek1 − Em
1 Z
(1)
hqi = dwdqP (w, q)q. (4)
we = El0 − Ek1 ,
For heat engines, the denominator is always non-zero
(hqi > 0), so the stochastic efficiency η in Eq. (3) is finite.
where 1
En0 , Em , Ek1 , El0are the measured energy of the four
Moreover, it follows from Eq. (4) that hηi = − hwi / hqi,
projective measurements corresponding to the times t =
which is just the conventional efficiency.
0, τc , τc + th , τc + th + τh respectively (n, m, k, l denote
From the joint distribution P (w, q), the distribution of
the corresponding quantum numbers). Thus, the joint
the stochastic efficiency P (η) is obtained by
probability distribution P (w, q) of the total stochastic
input work w = wc + we , and the stochastic absorbed
Z
heat from the hot reservoir q is given by P (η) = dwdqP (w, q)δ(η + w/ hqi). (5)

X The fluctuation of the stochastic efficiency is determined


1
P (w, q) = δ(w − Em + En0 − El0 + Ek1 )δ(q − Ek1 + Em 1 by the output work, which characterizes the reliability of
)
n,m,k,l the heat engine.
0
−βc En −βh Ek1 For practical calculation of the joint distribution of
2 2 e work and heat, we show the characteristic function of
× |1 hm |Uc | ni0 | |0 hl |Uh | ki1 | 0 ,
Z (βc )Z 1 (βh ) P (w, q) in the following
(2)
χ(u, v) ≡ eiuw+ivq = χc (u, v)χh (u, v),


(6)
where Uc , Uh are the unitary evolution operators where
corresponding to the compression and expansion
Tr[Uc† e(iu−iv)H(λ1 ) Uc e(−iu−βc )H(λ0 ) ]
processes, |ji0(1) is the eigenstate of the Hamiltonian χc (u, v) = , (7)
Z 0 (βc )
H(λ0 ) (H(λ1 )) and Z 0 (βc ) = Tr[e−βc H(λ0 ) ], Z 1 (βh ) =
Tr[e−βh H(λ1 ) ] are the partition functions corresponding
to the equilibrium states at t = 0 and t = τc + th Tr[Ue† eiuH(λ0 ) Ue e(−iu+iv−βh )H(λ1 ) ]
χh (u, v) = . (8)
respectively. Z 1 (βh )
3

The cumulant moments of work and heat are obtained between the relative fluctuation of the efficiency and the
from χ(u, v), such as the average input work dissipation quantified through the entropy production
in a cycle. When hδsi → 0, f (hδsi) ≈ 2/ hδsi, which
∂ ln χ(u, v) reproduces the TUR [14, 15] for the efficiency.
hwi = −i , (9)
∂u
u=v=0 For the spectra of the working substance with scale
property, i.e., En1 = En0 / [10] ( is n-independent), the
the average heat absorbed form the hot reservoir
general expression of the joint characteristic function
(Eq. (6)) is obtained in the quasistatic limit (see

∂ ln χ(u, v)
hqi = −i , (10) Appendix B). From Eqs. (9),(11), we obtain
∂v
u=v=0

and the variance of input work


 
σc σh
hwi = ηO Tc − ,
∂ 2 ln χ(u, v)
1 − ηO 1 − ηC
∆w2 = − , (11)   (15)
∂u2 2 2 2 Cc Ch
∆w = ηO kB Tc + ,

u=v=0
q (1 − ηO )2 (1 − ηC )2
2
where ∆(·) = h·2 i − h·i denotes the standard
deviation. where, Tc (Th ) is the temperature of the cold (hot)
reservoir, kB is the Boltzmann constant, ηO = 1−, σc ≡
Ec /Tc (σh ≡ Eh /Th ), Ec (Eh ) is the internal energy of
B. Fluctuation theorems the working substance corresponding to the equilibrium
state at t = 0 (t = τc + th ), and Cc ≡ ∂Ec /∂Tc
(Ch ≡ ∂Eh /∂Th ) is the heat capacity at constant volume.
Fluctuation theorems indicate the equality relation
In addition, the the average heat absorbed from the hot
in a general nonequilibrium process. According to
reservoir and the average entropy production of the cycle
Ref. [18], the fluctuation theorems are reexpressed via the
follow as hqi = − hwi /ηO and hδsi = βc hqi (ηC − ηO ),
efficiency for the Hamiltonian of the working substance
respectively. If the working substance statisfies the
involving time-reversal symmetry:
equipartition theorem Ec ∝ kB Tc and Eh ∝ kB Th in
D E the high-temperature limit, one has
e−δs(η,q) = 1, (12)
∆η 2 ∆w2
 
1 1 − ηO 1 − ηC 2
2 = 2 = + ≥ (16)
hηi hwi hδsi 1 − ηC 1 − ηO hδsi
P (−η hqi , q) = PR (η hqi , −q)eδs(η,q) , (13)

where δs(η, q) = βc (ηC q − η hqi) is the total stochastic with the equal sign saturated at ηO = ηC . The
entropy production expressed in terms of η and q, and inequality (16) is consistent with the TUR in steady
ηC = 1 − βh /βc is the Carnot efficiency. The subscript states [14, 15] or in a specific Otto cycle [22].
R denotes the reverse process of the cycle (the clockwise Moreover, due to the third law of thermodynamics,
direction in Fig. 1). Then, using the Jensen’s inequality hδsi → 0 in the low-temperature limit. Using the
ehxi ≤ hex i, we have hηi ≤ ηC for heat engines (hqi > 0), property of the function f (x) in Eq. (14), the TUR for
which is the second law of thermodynamics. It is worth the efficiency is also reproduced in the low-temperature
mentioning that this inequality is not sharp. In fact, limit. Consequently, we expect that the TUR for
for quantum systems without energy-level crossing when efficiency is valid for an arbitrary temperature under
changing the parameter λ, we obtain a sharper inequality these conditions. For a finite-time cycle, we numerically
hηi ≤ ηO as a result of the minimum work principle [19], study the TUR in a specific model below.
where ηO is the Otto efficiency, i.e., the efficiency of an
Otto cycle in the quasistatic limit (see Appendix A).

IV. QUANTUM HARMONIC-OSCILLATOR


C. Thermodynamic uncertainty relation HEAT ENGINE

Since the fluctuation theorems always imply the In this section, we illustrate our general discussions
genralized TUR [20], it follows from Eq. (13) that above with a specific example: a quantum harmonic-
oscillator being the working substance of the Otto cycle.
∆η 2 The frequency is changed from ω0 to ω1 (ω1 > ω0 )
2 ≥ f (hδsi), (14)
hηi in the adiabatic compression process. Then, according
to Refs. [21, 23], χc (u, v) and χh (u, v) of the joint
where f (x) = csch2 [g(x/2)], and g(x) is the inverse characteristic function in Eq. (6) are explicitly obtained
function of xtanhx. Equation (14) expresses a trade-off as (see Appendix C for detailed derivation)
4

 
βc ω 0 − 12
χc (u, v) = 2 sinh {2 cos [(u − v) ω1 ] cos [(u − iβc ) ω0 ] + 2Qc sin [(u − v) ω1 ] sin [(u − iβc ) ω0 ] − 2} , (17)
2

and

 
βh ω1 − 21
χh (u, v) = 2 sinh {2 cos (uω0 ) cos [(u − v − iβh ) ω1 ] + 2Qh sin (uω0 ) sin [(u − v − iβh ) ω1 ] − 2} , (18)
2

where Qc(h) ≥ 1 is the corresponding non-adiabatic Substituting Eqs. (19) and (20) into Eq. (3), the average
factor [24, 25]. The equal sign is hold when the quantum efficiency is obtained as
adiabatic condition is satisfied. In the following, we
study the efficiency statistics of the Otto cycle in different
circumstances.
(ω1 − ω0 Qh ) ϑh − (ω1 Qc − ω0 ) ϑc
hηi = (22)
ω1 (ϑh − Qc ϑc )
A. Average efficiency and efficiency distribution of
the heat engine
In the quasistatic limit, the non-adiabatic factors Qc,h =
1, and the average efficiency is the Otto efficiency ηO =
The average output work and average absorbed heat
1 −  ( = ω0 /ω1 ). Then, Eq. (22) is further expressed
per cycle can be obtained using Eqs. (9),(10),(17),(18) as
with ηO as

1
− hwi = − [(ω1 Qc − ω0 ) ϑc − (ω1 − ω0 Qh ) ϑh ] , (19)
2 P
 α=h,c ϑα (Qα − 1)
and hηi = ηO − . (23)
ϑh − Qc ϑc

ω1
hqi = (ϑh − Qc ϑc ) , (20) This result means that the non-adiabatic effect
2 decreases the average efficiency of the engine, which is
where demonstrated in Fig. 2.
On the other hand, the variance of the efficiency is
2
βh ω1 βc ω0 ∆η 2 = ∆w2 / hqi , where the variance of work is obtained
ϑh ≡ coth , ϑc ≡ coth . (21) from Eq. (11) as
2 2

 
2 2
ω 2 ω X
∆w2 = 1 (1 − ) ϑ2h + ϑ2c − 2 + 1 ϑ2h Q2h − 1 2 + ϑ2c Q2c − 1 −  (Qα − 1) ϑ2α − 1  ,
   
(24)
4 2
α=h,c

In the quasistatic limit, the variance of efficiency It is shown in Fig. 3 (a) that, in the quasistatic limit
accordingly becomes (Qc,h = 1), the efficiency fluctuation decreases as the
temperature increases, which is consistent with the TUR
(1 − )
2
ϑ2h + ϑ2c − 2
 for efficiency (Eq. (16)) since hδsi increases with the
2
∆ηadi = 2 . (25) temperature increases. In the non-adiabatic case, the
(ϑh − ϑc ) efficiency fluctuation as the function of Qc and Qh is
It is worth mentioning that the efficiency fluctuation illustrated in Fig. 3 (b), which reflects the enhancement
does not vanish for a quantum harmonic-oscillator Otto of the fluctuation due to the non-adiabatic driving.
cycle in the quasistatic limit, while the fluctuation of the The efficiency distribution is obtained with χ (u, v) by
previous version of the stochastic efficiency vanishes in the discrete Fourier transform. With different chosen
this case [10]. parameters, we plot the efficiency distribution in Fig. 4
5

(a)

0.5 2
10
0.4
101
0.3
100
0.2
1
1 0.7
1.1 0.5
1.1 10-1
1.2 1.2 0.3
1 100
0.1 10
102

FIG. 2: Average efficiency as the function of Qc and Qh . In (b)


this figure, we use ηO = 0.5, Th = 1 ηC = 0.8, and ω1 = 1.

(adiabatic case) and Fig. 5 (non-adiabatic case) with the 1.4


black dots. As comparisons, the Otto efficiency and
Carnot efficiency are respectively represented with the 1.2
blue dash-dotted line and the red dotted line. And we
show the probability that the efficiency of a stochastic 1
Otto cycle surpasses the Carnot efficiency in the figure.
In addition, one can infer that lower temperature leads 0.8
1.2
to greater probability of the heat engine surpassing the 1.2
Carnot efficiency. Meanwhile, the lower temperature 1.1 1.1
increases the probability of the engine to be useless, 1 1
namely, the engine outputs negative work.
FIG. 3: Efficiency fluctuation. (a) Efficiency fluctuation
as the function of temperature Th with different ηO in the
B. Finite-time performance of the heat engine quasistatic limit. (b) Efficiency fluctuation as the function of
Qc and Qh , where ηO = 0.5 and Th = 1 are chosen. In this
To further explore the finite-time performance of the figure, ηC = 0.8 and ω1 = 1 are fixed.
cycle, we first analyze the explicit time dependence of
the non-adiabatic factors for a specific protocol. For
an adiabatic process with frequency changed from ωi to In the following, we adopt the protocol of Eq. (26)
ωf during time t ∈ [0, τ ], the time dependence of the for the finite-time adiabatic processes in the Otto cycle,
frequency of the harmonic oscillator is [22, 26, 27] then we use of the explicit form of Q(τ ) given in Eq. (27)
to study the power at maximum efficiency (PME) and
ωi efficiency at maximum power (EMP) of the cycle. In this
ω(t) = . (26) sense, the non-adiabatic factors become Qc(h) = Q(τc(h) ),
(ωi /ωf − 1)t/τ + 1
and then the average power hP (τc , τh )i ≡ − hwi /(τh +
Then, the non-adiabatic factor Q(τ ) is obtained as (See τc ) and the efficiency hη(τc , τh )i of the Otto engine are
Appendix D for detailed derivation) respectively
√ 
1 − cos a2 τ 2 − 1 ln(ωf /ωi )
Q(τ ) = 1 + , (27) ω1 {[Q(τc ) − ] ϑc − [1 − Q(τh )] ϑh }
a2 τ 2 − 1 hP (τc , τh )i = ,
2 τh + τc
where (29)
2ωf ωi and
a≡ . (28)
ωf − ωi
P
As shown in Fig. 6, the non-adiabatic factor Q(τ ) (blue  α=h,c ϑα [Q(τα ) − 1]
hη(τc , τh )i = ηO − , (30)
solid line) oscillates with the driving time τ , reflecting the ϑh − Q(τc )ϑc
quantum coherence effect in the non-adiabatic transition.
The orange dashed line represents Q(τ ) = 1, which is where the total duration of the two isochoric processes,
achieved for the quantum adiabatic driving or with some i.e., tc + th , is assumed to be much smaller than τc + τh ,
special values of τ [26, 27]. and is thus ignored.
6

(a) (a)

(b) (b)

FIG. 4: Efficiency distribution in adiabatic case with Qh = FIG. 5: Efficiency distribution in non-adiabatic case with
Qc = 1. The probability distribution of efficiency is plotted Qh = Qc = 1.2. The probability distribution of efficiency is
with the black dots. The Otto efficiency and Carnot efficiency plotted with the black dots. The Otto efficiency and Carnot
are respectively represented with the blue dash-dotted line efficiency are respectively represented with the blue dash-
and the red dotted line. The (gray) area in the left side dotted line and the red dotted line. The (gray) area in the
denotes the negative work output regime; the (light red) area left side denotes the negative work output regime; the (light
in the right side of the red dotted line represents the regime of red) area in the right side of the red dotted line represents the
η > ηC . The parameters are chosen as: (a) Th = 10, Tc = 2; regime of η > ηC . The parameters are chosen as: (a)Th = 10,
(b) Th = 1, Tc = 0.2. In this figure, we choose ω0 = 0.5, Tc = 2; (b) Th = 1, Tc = 0.2. In this figure, we choose
ω1 = 1, and Carnot efficiency is fixed at 0.8. ω0 = 0.5, ω1 = 1, and Carnot efficiency is fixed at 0.8.

Since Qc(h) = 1 can be achieved within finite time, 1.25


the average efficiency of some cycles approach the Otto
efficiency ηO with non-vanishing power. These cycles 1.2
happen to have the special operation time sets (τc∗ , τh∗ )
1.15
corresponding to Qh (τh∗ ) = Qc (τc∗ ) = 1. With the help
of Eq. (27), one finds the special operation time follows 1.1
as
1.05
p
a2 τ 2 − 1 ln  = 2nπ, n = 1, 2, 3..., (31) 1
0 10 20 30 40
namely,

r FIG. 6: Time dependence of the non-adiabatic factor. In


ηO 1  nh,c π 2 this figure, the blue solid curve represents Q(τ ) in Eq. (D7),

τh,c = + . (32) the orange dashed line is Q(τ ) = 1. The initial and final
ω0 4 ln 
frequencies of the harmonic oscillator in the adiabatic process
are chosen as ω0 = 0.5 and ω1 = 1.
Therefore, the PME is
7

(a)
0.6 12
0.5 10
0.4
8
0.3
6
0.2
4
0.1
2
0
0 0.1 0.2 0.3 0.4 0.5 0.6 0
1 1.05 1.1 1.15 1.2

FIG. 7: Efficiency at maximum power of the Otto engine (b)


as the function of ηO . In this figure, ηO /ηC = 0.8 is fixed.
The blue solid curve represents the EMP of the Otto engine. 6
The black dash-dotted curve is the upper bound for EMP,
5
η+ = 2ηO /(3 − ηO ), of the Otto cycle obtained in Ref. [31]
without considering the oscillation of work. The red dotted 4
line denotes the Otto efficiency.
3

ω0 ω1 (ϑh − ϑc ) 1
P (τc∗ , τh∗ ) = q . (33)
P
2 α=h,c 1/4 + (nα π/ ln )
2 0
1 1.05 1.1 1.15 1.2

It should be noted that in the usual finite-time


thermodynamic cycles, the PME generally approaches FIG. 8: Thermodynamic uncertainty relation for the
zero [28–31]. Here, thanks to the special protocol we efficiency with different non-adiabatic factors. (a) Th =
have chosen to realize the quantum adiabatic process in 1, 2, 10, ηO = 0.5; (b) Th = 10, ηO = 0.5, 0.6, 0.7. In this
finite time, the current quantum Otto cycle outputs non- figure, Qc = 1, ηC = 0.8 and ω1 = 1 are fixed parameters.
zero or even relatively large power (comparable to the
maximum power) when the Otto efficiency is reached.
Obviously, the maximum P (τc∗ , τh∗ ) is indicates that the oscillation of the output work (due
to quantum coherence) are conducive to improving the
ω0 ω1 (ϑh − ϑc ) EMP.
Pmax (τc∗ , τh∗ ) = q , (34)
2
2 1 + (2π/ ln )
C. Thermodynamic uncertainty relation (TUR) for
which is achieved at nc = nh = 1. Besides, the efficiency
second largest and third largest power are reached at
(nc = 1, nh = 2) and (nc = 2, nh = 2), respectively.
Because the spectra of a quantum harmonic oscillator
For (nα π/ ln )2  1/4, Eq. (33) can be approximated
have scale property and the system follows the
as
equipartition theorem in the high-temperature limit, we
conclude that in the quasistatic limit, the TUR (Eq. (16))
ω0 ω1 ln  (ϑh − ϑc ) is valid according to the discussions in Sec. III C.
P (τc∗ , τh∗ ) ≈ , (35) For the non-adiabatic driving cycle, the results are
2π (nc + nh )
shown in Fig.  8. The TUR (Eq. (16)) is still valid
which shows that P (τc∗ , τh∗ ) is a monotonically decreasing since ∆η
2
/ 2
increases monotonically with Qh and
hηi2 hδSi
quasi-continuous function of nc and nh .
Qc . Here, without loss of generality, we take Qh
In addition, the EMP of this Otto engine as the
as the independent variable in the figure. On the
function of ηO is illustrated in Fig. 7. As shown in
contrary, the TUR may be violated due to the incomplete
this figure, the EMP of our cycle (blue solid curve) is
thermalization in the isochoric processes [22]. One can
found to surpass the upper bound, η+ = 2ηO /(3 − ηO )
conclude
 from  Fig. 8(a) that higher temperature makes
(black dashed curve), of the Otto cycle’s EMP without ∆η 2 2
considering the oscillation of the output work [31]. This hηi2
/ hδSi lower. Moreover, as shown in Fig. 8(b),
8

2
 


when Qh → 1, ∆η hηi2
2
/ hδSi is closer to 1 in the case Namely, δwc(e) ≡ wc(e) − wc(e) adi ≥ 0. Thus, the
with ηO = 0.7. This is consistent with the average efficiency of a quantum Otto cycle with the
 discussions
 complete thermalization satisfies
∆η 2 2
in Sec. III C that the condition for hηi2 / hδSi → 1 is
ηO → ηC . hwi hwc iadi + hwe iadi + δwc + δwe
hηi = − =−
hqi hqiadi − δwc
(A1)
V. SUMMARY AND DISCUSSION hwc iadi + hwe iadi
≤− = ηO ,
hqiadi
In this paper, we come up with a new definition of the
stochastic efficiency for heat engine in micro scale. The for hqiadi > − hwc iadi − hwe iadi > 0, and δwc ≥ 0, δwe ≥
moments of the efficiency always exist, and its mean value 0, where hqiadi denotes the heat absorbed from the hot
is equal to the conventional efficiency. Moreover, the reservoir in the quasistatic limit.
fluctuation theorems are resexpressed via the efficiency.
For spectra of the working substance with scale property,
the statistics of the efficiency is fully determined by Appendix B: The joint characteristic function for a
the partition functions of the working substance in the quantum Otto cycle in the quasistatic limit
quasistatic limit. Importantly, we reveal the connection
between the TUR and the equipartition theorem.
According to Eq. (6), the expression of the
For a quantum Otto cycle with a harmonic oscillator
joint characteristic function χ(u, v) is obtained by a
being the working substance, we obtain the exact
transformation of the characteristic function of work
expression of the joint characteristic function of work and
χc (u) ≡ χc (u, 0) and χh (u) ≡ χh (u, 0), i.e.,
heat. We find that the Otto efficiency can be reached
with a finite output power (the power at maximum
efficiency) with some special duration and the EMP χc (u, v) = χc (u)|u→u−v,βc →βc +iv
(B1)
surpasses the upper bound obtained in Ref. [31]. χh (u, v) = χh (u)|βh →βh −iv .
The theoretical predictions of current study can be
tested on some state-of-art experiments, such as the with scale property (En1 = En0 /), in the quasistatic limit,
Brownian particle system [8] and trapped ion system [33]. the expressions of the characteristic function χc (u) and
As a direct extention, similarly to the stochastic χh (u) read
efficiency defined here, the coefficient of performance
of a refrigerator can be defined as the ratio of the X e−βc En0 1 0
stochastic released heat to the average input work. Then, χc (u) = eiu(En −En )
Z 0 (βc )
the statistics of a stochastic refrigerator can be further n
discussed. Besides, it is expected that the many-body X e−[βc −iu(−1 −1)]En0
effect of the working substance [24, 26, 34–37] and the = (B2)
n
Z 0 (βc )
influences of the control protocols for the cycle [38–40]
on the efficiency statistics and TUR will be taken into Z [βc − iu(−1 − 1)]
0
= ,
consideration in future investigations. Z 0 (βc )

and
Acknowledgments
X e−βh En1 0 1
χh (u) = eiu(En −En )
This work is supported by the National Natural
n
Z 1 (βh )
Science Foundation of China (NSFC) (Grants No.
X e−[βc −iu(−1)]En1
11534002, No. 11875049, No. U1730449, No. U1530401, = (B3)
and No. U1930403), and the National Basic Research n
Z 1 (βh )
Program of China (Grants No. 2016YFA0301201). Y. 1
Z [βh − iu( − 1)]
H. Ma is supported by the China Postdoctoral Science = .
Foundation (Grant No. BX2021030). Z 1 (βh )

Then, it follows from Eq. (B1) that the joint


Appendix A: Proof of hηi ≤ ηO characteristic function χ(u, v) reads

χ(u, v) =
As a result of the minimum work principle [19], when
the energy levels do not cross during the driving, the Z 0 [βc + iv − i(u − v)(−1 − 1)]Z 1 [βh − iv − iu( − 1)]
.
average work under finite-time driving hwc (e)i is not Z 0 (βc )Z 1 (βh )
less than it under quantum adiabatic driving hwc (e)iadi . (B4)
9

Appendix C: The joint characteristic function for a Then, the characteristic functions of work χc (u), χh (u)
quantum harmonic oscillator heat engine (see Appendix B) reads [21, 23]

For a harmonic oscillator with time-dependent


frequency in an adiabatic process during time [0, τ ], the
Hamiltonian is
p2 1
H(t) = + mω(t)2 x2 . (C1)
2m 2

 
βω0 − 12
χc (u) = 2 sinh {2 cos(uω1 ) cos[(u − iβc )ω0 ] + 2Qc sin(uω1 ) sin[(u − iβc )ω0 ] − 2} , (C2)
2

 
βω1 − 21
χh (u) = 2 sinh {2 cos(uω0 ) cos[(u − iβh )ω1 ] + 2Qh sin(uω0 ) sin[(u − iβh )ω1 ] − 2} , (C3)
2

where with the initial value {y1 (0), y2 (0), ẏ1 (0), ẏ2 (0)} =
{1, 0, 0, ω0 }. Similarly, the expression of Qh is given
ẏ1 (τc )2 + ẏ2 (τc )2
 
ω1 2 2 by the replacement: ω0 ↔ ω1 , t ∈ [0, τc ] → t ∈
Qc = y1 (τc ) + y2 (τc ) + ,
2ω0 ω12 [τc + th , τc + th + τh ].
(C4)
the overhead dot denotes the time derivative, y1 and y2
are the two general solutions of the classical harmonic
oscillator, i.e.,
Then, the expression of χ(u, v) is obtained by χ(u, v) =
ÿ(t) + ω(t)2 y(t) = 0, (C5) χc (u, v)χh (u, v), where (Eq. (B1))

χc (u, v) = χc (u)|u→u−v,βc →βc +iv


 
βω0 −1
= 2 sinh {2 cos[(u − v)ω1 ] cos[(u − iβc )ω0 ] + 2Qc sin[(u − v)ω1 ) sin[(u − iβc )ω0 ] − 2} 2 ,
2

χh (u, v) = χh (u)|βh →βh −iv


 
βω1 −1
= 2 sinh {2 cos(uω0 ) cos[(u − v − iβh )ω1 ] + 2Qh sin(uω0 ) sin[(u − v − iβh )ω1 ] − 2} 2 .
2

Appendix D: Explicit Time-dependence of the


non-adiabatic factor
d→ − →

φ (t) = M(t) φ (t). (D1)
dt
For the specific driving protocol in Eq. (26), the time- Here,
dependence of the non-adiabatic factor in Eq. (27) can
be directly calculated from its definition (Eq. (C4)) [26].

− T
Here, we present another approach with respect to the φ (t) ≡ hH(t)i hL(t)i hD(t)i , (D2)
internal energy of the working substance. It follows from
Ref. [22] that the evolution of the harmonic oscillator h·i denotes the ensemble average respect to the density
in an adiabatic process during time t ∈ [0, τ ] can be matrix of the oscillator, T denotes the matrix transpose.
described by a linear differential equation as H(t) = p2 /(2m) + mω 2 (t)x2 /2, L(t) = p2 /(2m) −
10

mω 2 (t)x2 /2, and D(t) = ω(t)(xp+px)/2 are respectively where a = 2ωi ωf /(ωf − ωi ). Thus, the internal energy of
the Hamiltonian, the Lagrangian, and the generator of the system at t = τ is
the scale transformation. The time-dependent matrix
reads
 
ω̇/ω 2 −ω̇/ω 2 0 √
M(t) =  −ω̇/ω 2 ω̇/ω 2 −2  . (D3) a2 τ 2 − cos( a2 τ 2 − 1 ln(ωf /ωi )) ωf
hH(τ )i = hH(0)i
0 2 ω̇/ω 2 a2 τ 2 − 1 ωi
(D6)
The general solution of Eq. (D1) follows as Consequently, the non-adiabatic factor is obtained by [24]

Z τ 

− →

φ (τ ) = T← exp M(t)dt φ (0), (D4) √ 
0 1 − cos( a2 τ 2 − 1 ln(ωf /ωi ))
hH(τ )i
Q(τ ) = =1+ ,
where T← denotes the time-ordered operation. For hHiadi a2 τ 2 − 1
the specific protocol in Eq. (26), the matrix M(t) is (D7)
independent of t. For the thermal equilibrium initial where hHiadi = hH(aτ → ∞)i = hH(0)i ωf /ωi is the

− T internal energy of the system at the end of the process
state, φ (0) = hH(0)i 0 0 , we find
under quantum adiabatic driving. In the short-time limit
aτ → 0, and long-time limit aτ → ∞, it is easy to check
 ω −√1−a2 τ 2  ω √1−a2 τ 2
" #
that
 ω

− ωf f
ωi + ωf −2a2 τ 2
i i
 
2(a" 2 τ 2 −1)
√  ω 2√1−a2 τ 2
 # 
1−a2 τ 2
 


 ω 1−
f
 1− ωf 
 hH(0)i ,
φ (τ ) =  ωi i
 √  2
2 1−a 2 2
" τ [ln(ωf /ωi )]
1−√1−a2 τ 2  ω √1−a2 τ 2 2
 # 



f
1− ωf
 lim Q(τ ) = 1 + , lim Q(τ ) = 1. (D8)
aτ →0 2 aτ →∞
 
ωi i

2(a2 τ 2 −1)
(D5)

[1] U. Seifert, Rep. Prog. Phys. 75, 126001 (2012). Rev. E 76, 031105 (2007).
[2] K. Sekimoto, Stochastic Energetics (Spinger, Berlin, [17] Y. Rezek and R. Kosloff, New J. Phys. 8, 83 (2006).
2010). [18] N. A. Sinitsyn, J. Phys. A: Math. Theor. 44, 405001
[3] M. Esposito, U. Harbola, and S. Mukamel, Rev. Mod. (2011).
Phys. 81, 1665 (2009). [19] A. E. Allahverdyan, and Th. M. Nieuwenhuizen, Phys.
[4] M. Campisi, P. Hänggi, and P. Talkner, Rev. Mod. Phys. Rev. E 71, 046107 (2005).
83, 771 (2011). [20] A. M. Timpanaro, G. Guarnieri, J. Goold, and G. T.
[5] G. Verley, M. Esposito, T. Willaert, and C. Van den Landi, Phys. Rev. Lett. 123, 090604 (2019).
Broeck, Nat. Commun. 5, 4721 (2014). [21] S. Deffner, and E. Lutz, Phys. Rev. E 77, 021128 (2008).
[6] G. Verley, T. Willaert, C. Van den Broeck, and M. [22] S. Lee, M. Ha, and H. Jeong, Phys. Rev. E 103, 022136
Esposito,Phys. Rev. E 90, 052145 (2014). (2021).
[7] S. K. Manikandan, L. Dabelow, R. Eichhorn, and S. [23] Z. Y. Fei, and H. T. Quan, Phys. Rev. Research 1, 033175
Krishnamurthy, Phys. Rev. Lett. 122, 140601 (2019). (2019).
[8] I. A. Martínez, É. Roldán, L. Dinis, D. Petrov, J. M. R. [24] J. Jaramillo, M. Beau, and A. del Campo, New J. Phys.
Parrondo, and R. A. Rica, Nature Phys 12, 67-70 (2016). 18 075019, (2019).
[9] M. Polettini, G. Verley, and M. Esposito, Phys. Rev. Lett [25] K. Husimi, Prog. Theor. Phys. 9(4) 381 (1953).
114, 050601 (2015). [26] M. Beau, J. Jaramillo, and A. del Campo, Entropy 2016,
[10] T. Denzler, and E. Lutz, Phys. Rev. Research 2, 18, 168.
032062(R) (2020). [27] J. F. Chen, C. P. Sun, and H. Dong, Phys. Rev. E 100,
[11] T. Denzler, and E. Lutz New J. Phys. 23 075003 (2021). 032144 (2019).
[12] C. Jarzynski, Annu. Rev. Condens. Matter Phys. 2, 329 [28] N. Shiraishi, K. Saito, and H. Tasaki, Phys. Rev. Lett.
(2011). 117, 190601 (2016).
[13] G. E. Crooks, Phys. Rev. E 60, 2721 (1999). [29] Y. H. Ma, D. Xu, H. Dong, and C. P. Sun, Phys. Rev. E
[14] A. C. Barato and U. Seifert, Phys. Rev. Lett. 114, 158101 98, 042112 (2018).
(2015). [30] H. Yuan, Y. H. Ma, and C. P. Sun, arXiv:2107.11342
[15] P. Pietzonka, A. C. Barato, and U. Seifert, Phys. Rev. E (2021).
93, 052145 (2016). [31] J. F. Chen, C. P. Sun, and H. Dong, Phy. Rev. E 100,
[16] H. T. Quan, Y. X. Liu, C. P. Sun, and F. Nori, Phys. 062140 (2019).
11

[32] G. Jiao, Y. Xiao, J. He, Y. Ma, and J. Wang, New J. Phys. Rev. Lett. 120, 100601 (2018).
Phys. 23 063075 (2021). [37] J. Chen, H. Dong, and C. P. Sun, Phys. Rev. E 98,
[33] O. Abah, J. RoBnagel, G. Jacob, S. Deffner, F. Schmidt- 062119 (2018).
Kaler,K. Singer, and E. Lutz, Phys. Rev. Lett. 109, [38] Y. H. Ma, D. Xu, H. Dong, and C. P. Sun, Phys. Rev. E
203006 (2012). 98, 022133 (2018).
[34] Y. H. Ma, S. H. Su, and C. P. Sun, Phys. Rev. E 96, [39] Y. H. Ma, R. X. Zhai, J. F. Chen, C. P. Sun, and H.
022143 (2018). Dong, Phys. Rev. Lett. 125, 210601 (2020).
[35] Z. Fei, N. Freitas, V. Cavina, H. T. Quan, and M. [40] K. Brandner and K. Saito. Phys. Rev. Lett. 124, 040602
Esposito, Phys. Rev. Lett. 124, 170603 (2020). (2020).
[36] J. Bengtsson, M. Nilsson Tengstrand, A. Wacker, P.
Samuelsson, M. Ueda, H. Linke, and S. M. Reimann,

You might also like