+ All Categories
Home > Documents > arXiv:1612.00518v1 [cond-mat.stat-mech] 1 Dec 2016 · E ciency at the maximum power output for...

arXiv:1612.00518v1 [cond-mat.stat-mech] 1 Dec 2016 · E ciency at the maximum power output for...

Date post: 06-Jun-2020
Category:
Upload: others
View: 1 times
Download: 0 times
Share this document with a friend
9
Eciency at the maximum power output for simple two-level heat engine Sang Hoon Lee, 1 Jaegon Um, 2, 3 and Hyunggyu Park 1, 2 1 School of Physics, Korea Institute for Advanced Study, Seoul 02455, Korea 2 Quantum Universe Center, Korea Institute for Advanced Study, Seoul 02455, Korea 3 CCSS, CTP and Department of Physics and Astronomy, Seoul National University, Seoul 08826, Korea We introduce a simple two-level heat engine to study the eciency in the condition of the maximum power output, depending on the energy levels from which the net work is extracted. In contrast to the quasi-statically operated Carnot engine whose eciency reaches the theoretical maximum, recent research on more realistic engines operated in finite time has revealed other classes of eciency such as the Curzon-Ahlborn eciency maximizing the power output. We investigate yet another side with our heat engine model, which consists of pure relaxation and net work extraction processes from the population dierence caused by dierent transition rates. Due to the nature of our model, the time-dependent part is completely decoupled from the other terms in the generated work. We derive analytically the optimal condition for transition rates maximizing the generated power output and discuss its implication on general premise of realistic heat engines. In particular, the optimal engine eciency of our model is dierent from the Curzon-Ahlborn eciency, although they share the universal linear and quadratic coecients at the near-equilibrium limit. We further confirm our results by taking an alternative approach in terms of the entropy production at hot and cold reservoirs. PACS numbers: 05.70.Ln, 05.40.a, 05.20.y, 89.70.a I. INTRODUCTION The eciency of heat engines is a celebrated topic of clas- sical thermodynamics [1]. In particular, an elegant formula expressed only by hot and cold reservoir temperatures for the ideal quasi-static and reversible engine coined by Sadi Carnot has been an everlasting textbook example [2]. That ideal engine, however, is not the most ecient engine any more when we consider its power (the extracted work per unit time), which has added dierent types of optimal engine eciency such as the Curzon-Ahlborn eciency for some cases [3–5]. Following such steps, researchers have taken simple systems to investigate various theoretical aspects of underlying princi- ples of macroscopic thermodynamic engine eciency in de- tails [6–12] and its microscopic fluctuation [13–18]. In this paper, we introduce a simple two-level heat engine model to explore the condition for the maximum power. In our model, the time-dependent part is completely decoupled from the rest of the formulation, which makes the analysis considerably simpler. We derive analytically a parameter re- lation between transition rates at the maximum power for a given temperature ratio. We compare the functional form of the optimal eciency at the maximum power to previously known forms for some other cases. Our result shows a dif- ference from the Curzon-Ahlborn eciency, but shares the same asymptotic behavior up to the second order in a small eciency limit. We also take an alternative approach consid- ering the entropy production at the reservoirs, and discuss its implication. Generalization to multi-level engines is consid- ered, but a decoupling of the operating time does not happen, which makes the analytic investigation quite complex. II. TWO-LEVEL HEAT ENGINE Figure 1 illustrates our model. The two-level system is characterized by two discrete energy states composed of the ground state (E = 0) and the excited state (E = E 1 or E = E 2 , depending on the reservoir of consideration). The transition rates from the ground state to the excited state are denoted by q and , respectively, and their reverse processes by ˜ q and ˜ . We assume E 1 > E 2 and T 1 > T 2 . The system is at- tached to two dierent reservoirs: R 1 with temperature T 1 dur- ing time τ 1 , and R 2 with temperature T 2 during time τ 2 , and the adiabatic work extraction occurs in between. Although the amount of energy unit involving the work exchange is the same (W = E 1 - E 2 = W 0 ) in Fig. 1, the net positive work is achievable due to the dierence in the population of the excited states at the end of contact with R 1 and R 2 , which is determined by model parameters as presented in Sec. III. III. ENGINE EFFICIENCY A. Eciency as a function of model parameters The transition rates from the ground state to the excited state at reservoirs R 1 and R 2 are given as the following Ar- rhenius form, q/ ˜ q = e -E 1 /T 1 , / ˜ = e -E 2 /T 2 , (1) respectively (we let the Boltzmann constant k B = 1 for nota- tional convenience), thus the inequality 0 << q < 1/2 holds (< q is essential to get the positive amount of net work). The average amount of work extracted from the system at R 1 R 2 and that given to the system at R 2 R 1 considering the pop- ulation dierence are given by hWi = (E 1 - E 2 )P 1e , hW 0 i = (E 1 - E 2 )P 2e , (2) respectively, and |P 1 i = P 1e , P 1g T and |P 2 i = P 2e , P 2g T are the column vectors whose components represent the popula- arXiv:1612.00518v1 [cond-mat.stat-mech] 1 Dec 2016
Transcript
Page 1: arXiv:1612.00518v1 [cond-mat.stat-mech] 1 Dec 2016 · E ciency at the maximum power output for simple two-level heat engine Sang Hoon Lee,1 Jaegon Um,2,3 and Hyunggyu Park1,2 1School

Efficiency at the maximum power output for simple two-level heat engine

Sang Hoon Lee,1 Jaegon Um,2, 3 and Hyunggyu Park1, 2

1School of Physics, Korea Institute for Advanced Study, Seoul 02455, Korea2Quantum Universe Center, Korea Institute for Advanced Study, Seoul 02455, Korea

3CCSS, CTP and Department of Physics and Astronomy, Seoul National University, Seoul 08826, Korea

We introduce a simple two-level heat engine to study the efficiency in the condition of the maximum poweroutput, depending on the energy levels from which the net work is extracted. In contrast to the quasi-staticallyoperated Carnot engine whose efficiency reaches the theoretical maximum, recent research on more realisticengines operated in finite time has revealed other classes of efficiency such as the Curzon-Ahlborn efficiencymaximizing the power output. We investigate yet another side with our heat engine model, which consists ofpure relaxation and net work extraction processes from the population difference caused by different transitionrates. Due to the nature of our model, the time-dependent part is completely decoupled from the other terms inthe generated work. We derive analytically the optimal condition for transition rates maximizing the generatedpower output and discuss its implication on general premise of realistic heat engines. In particular, the optimalengine efficiency of our model is different from the Curzon-Ahlborn efficiency, although they share the universallinear and quadratic coefficients at the near-equilibrium limit. We further confirm our results by taking analternative approach in terms of the entropy production at hot and cold reservoirs.

PACS numbers: 05.70.Ln, 05.40.a, 05.20.y, 89.70.a

I. INTRODUCTION

The efficiency of heat engines is a celebrated topic of clas-sical thermodynamics [1]. In particular, an elegant formulaexpressed only by hot and cold reservoir temperatures for theideal quasi-static and reversible engine coined by Sadi Carnothas been an everlasting textbook example [2]. That idealengine, however, is not the most efficient engine any morewhen we consider its power (the extracted work per unit time),which has added different types of optimal engine efficiencysuch as the Curzon-Ahlborn efficiency for some cases [3–5].Following such steps, researchers have taken simple systemsto investigate various theoretical aspects of underlying princi-ples of macroscopic thermodynamic engine efficiency in de-tails [6–12] and its microscopic fluctuation [13–18].

In this paper, we introduce a simple two-level heat enginemodel to explore the condition for the maximum power. Inour model, the time-dependent part is completely decoupledfrom the rest of the formulation, which makes the analysisconsiderably simpler. We derive analytically a parameter re-lation between transition rates at the maximum power for agiven temperature ratio. We compare the functional form ofthe optimal efficiency at the maximum power to previouslyknown forms for some other cases. Our result shows a dif-ference from the Curzon-Ahlborn efficiency, but shares thesame asymptotic behavior up to the second order in a smallefficiency limit. We also take an alternative approach consid-ering the entropy production at the reservoirs, and discuss itsimplication. Generalization to multi-level engines is consid-ered, but a decoupling of the operating time does not happen,which makes the analytic investigation quite complex.

II. TWO-LEVEL HEAT ENGINE

Figure 1 illustrates our model. The two-level system ischaracterized by two discrete energy states composed of the

ground state (E = 0) and the excited state (E = E1 or E = E2,depending on the reservoir of consideration). The transitionrates from the ground state to the excited state are denotedby q and ε, respectively, and their reverse processes by q andε. We assume E1 > E2 and T1 > T2. The system is at-tached to two different reservoirs: R1 with temperature T1 dur-ing time τ1, and R2 with temperature T2 during time τ2, andthe adiabatic work extraction occurs in between. Althoughthe amount of energy unit involving the work exchange is thesame (W = E1 − E2 = W ′) in Fig. 1, the net positive workis achievable due to the difference in the population of theexcited states at the end of contact with R1 and R2, which isdetermined by model parameters as presented in Sec. III.

III. ENGINE EFFICIENCY

A. Efficiency as a function of model parameters

The transition rates from the ground state to the excitedstate at reservoirs R1 and R2 are given as the following Ar-rhenius form,

q/q = e−E1/T1 ,

ε/ε = e−E2/T2 ,(1)

respectively (we let the Boltzmann constant kB = 1 for nota-tional convenience), thus the inequality 0 < ε < q < 1/2 holds(ε < q is essential to get the positive amount of net work). Theaverage amount of work extracted from the system at R1 → R2and that given to the system at R2 → R1 considering the pop-ulation difference are given by

〈W〉 = (E1 − E2)P1e ,

〈W ′〉 = (E1 − E2)P2e ,(2)

respectively, and |P1〉 =(P1e, P1g

)Tand |P2〉 =

(P2e, P2g

)Tare

the column vectors whose components represent the popula-

arX

iv:1

612.

0051

8v1

[co

nd-m

at.s

tat-

mec

h] 1

Dec

201

6

Page 2: arXiv:1612.00518v1 [cond-mat.stat-mech] 1 Dec 2016 · E ciency at the maximum power output for simple two-level heat engine Sang Hoon Lee,1 Jaegon Um,2,3 and Hyunggyu Park1,2 1School

2

R1

E1

T1

q

R2

E2

T2

✏0

q

0✏

during duringstochastic Markov processes

⌧1 ⌧2

|P1i(t1 = 0) |P1i(t1 = ⌧1) = |P2i(t2 = 0) |P2i(t2 = ⌧2)

|P2i(t2 = ⌧2) = |P1i(t1 = 0)

W = E1 � E2

W 0 = E1 � E2

Q1 Q2

hWnetiT1

sh =T2

T1sc +

hWnetiT1

sh =E1

T1

sc

G�1 (sc)

sc

shFIG. 1. Schematic illustration of our simple two-level heat engine, composed of two energy levels coupled with two heat reservoirs R1 and R2.

tions of excited and ground states (in that order) at the end ofthe contact with R1 and R2, respectively. We then take the nor-malization convention P1e+P1g = P2e+P2g = q+q = ε+ε = 1,expressing the conservation of total population. For notationalconvenience, we define the function X(T, r) of temperature Tand transition rate r as

X(T, r) ≡ T ln(r/r) . (3)

Then, the average amount of heat to the system from R1 andthat from the system to R2 are

〈Q1〉 = (P1e − P2e)X(T1, q)〈Q2〉 = (P1e − P2e)X(T2, ε) ,

(4)

respectively, based on the Schnakenberg entropy productionfor stochastic processes [19–21]. The average total entropyproduction during one cycle is given by the entropy change ofthe reservoir,

〈∆S 〉 = −〈Q1〉

T1+〈Q2〉

T2

= (P1e − P2e)[ln(ε/ε) − ln(q/q)

].

(5)

Eqs. (2) and (4) ensure the energy conservation or the first lawof thermodynamics 〈W〉 − 〈W ′〉 = 〈Q1〉 − 〈Q2〉, consideringEq. (1). The average net work extracted from the system is

〈Wnet〉 = 〈W〉−〈W ′〉 = (P1e−P2e)[X(T1, q) − X(T2, ε)

], (6)

and the efficiency is given by the ratio

η =〈Wnet〉

〈Q1〉= 1 −

X(T2, ε)X(T1, q)

, (7)

independent of τ1 and τ2, and η approaches ηC = 1 − T2/T1(the Carnot efficiency [1, 2]) when ε ' q, and meaningful onlyfor q > ε, or 〈Wnet〉 > 0.

Now let us consider the explicit form of populations at theexcited states at end of each reservoir contact process, whose

time evolution is given by the following linear differentialequation system for given q and ε values,

d|P1〉

dt1=

[−q qq −q

]|P1〉 ,

d|P2〉

dt2=

[−ε εε −ε

]|P2〉 ,

(8)

where 0 ≤ t1 ≤ τ1 and 0 ≤ t2 ≤ τ2 are the intermediatetime spent in contact with R1 and R2, respectively. As thepopulations do not change during the adiabatic work extrac-tion (supply) processes, we get the circular boundary con-dition as P1e(t1 = 0, t2 = τ2) = P2e(t1 = τ1, t2 = τ2) andP2e(t1 = τ1, t2 = 0) = P1e(t1 = τ1, t2 = τ2). Thus, the solutionat t1 = τ1 and t2 = τ2 is given by

P1e =q(1 − e−τ1 ) + ε(1 − e−τ2 )e−τ1

1 − e−(τ1+τ2) ,

P2e =ε(1 − e−τ2 ) + q(1 − e−τ1 )e−τ2

1 − e−(τ1+τ2) ,

(9)

and limτ1,τ2→∞ P1e = q and limτ1,τ2→∞ P2e = ε as expected.With τ1 = τ2 = τ/2, we obtain the average net work as

〈Wnet〉 =(q − ε)(1 − e−τ/2)2

1 − e−τ[X(T1, q) − X(T2, ε)

], (10)

so the monotonically increasing factor (1−e−τ/2)2/(1−e−τ) forthe time scale τ is decoupled from the rest of the formula andonly plays the role of an overall factor. It is important to notethat the decoupling holds regardless of the τ1 = τ2 condition;the overall factor becomes (1 − e−τ1 )(1 − e−τ2 )/(1 − e−τ1+τ2 ).The average power output 〈P〉 is given by

〈P〉 =(q − ε)(1 − e−τ/2)2

τ(1 − e−τ)[X(T1, q) − X(T2, ε)

], (11)

which decreases monotonically with τ. Therefore, from nowon, we discard the time dependence altogether and focus onother parameters, i.e., denoting

〈Wnet〉 ≡ 〈P〉 ≡ (q − ε)[X(T1, q) − X(T2, ε)

], (12)

Page 3: arXiv:1612.00518v1 [cond-mat.stat-mech] 1 Dec 2016 · E ciency at the maximum power output for simple two-level heat engine Sang Hoon Lee,1 Jaegon Um,2,3 and Hyunggyu Park1,2 1School

3

(a) (b)<Wnet>(τ → ∞), T1 = 1, T2 = 1/2

0.1 0.2 0.3 0.4 0.5q

0.1

0.2

0.3

0.4

0.5

ε

0 0.005 0.01 0.015 0.02 0.025 0.03 0.035 0.04

η, T1 = 1, T2 = 1/2

0.1 0.2 0.3 0.4 0.5q

0.1

0.2

0.3

0.4

0.5

ε

0 0.05 0.1 0.15 0.2 0.25 0.3 0.35 0.4 0.45 0.5

(c) (d)<Wnet>(τ → ∞), T1 = 1, T2 = 9/10

0.1 0.2 0.3 0.4 0.5q

0.1

0.2

0.3

0.4

0.5

ε

0

0.0002

0.0004

0.0006

0.0008

0.001

0.0012

η, T1 = 1, T2 = 9/10

0.1 0.2 0.3 0.4 0.5q

0.1

0.2

0.3

0.4

0.5

ε 0 0.01 0.02 0.03 0.04 0.05 0.06 0.07 0.08 0.09 0.1

FIG. 2. (a) The average net work limτ→∞〈Wnet〉 and (b) efficiency η for T1 = 1 and T2 = 1/2, and (c) 〈Wnet〉(τ → ∞) and (d) η for T1 = 1 andT2 = 9/10. For better visibility focused on the 〈Wnet〉 ≥ 0 regime, we set all of the negative values as 0.

without considering the overall factor involving τ for nota-tional convenience. Numerically, we obtain the net workand efficiency for (q, ε) combination, as shown in Fig. 2. InSec. III B, we derive the condition for the efficiency at themaximum power output.

B. Efficiency at the maximum power output

1. The condition for the maximum power output

For a given T2/T1 value, the maximum power output con-dition for the two-variable function is

∂〈P〉∂q

∣∣∣∣∣q=q∗,ε=ε∗

=∂〈P〉∂ε

∣∣∣∣∣q=q∗,ε=ε∗

= 0 , (13)

which leads to

1 −X(T2, ε

∗)X(T1, q∗)

=q∗ − ε∗

q∗(1 − q∗) ln[(1 − q∗)/q∗], (14a)

and

1 −X(T2, ε

∗)X(T1, q∗)

=(T2/T1)(q∗ − ε∗)

ε∗(1 − ε∗) ln[(1 − q∗)/q∗], (14b)

from Eq. (12). By eliminating the left-hand side of Eqs. (14a)and (14b), we obtain the following simple relation

T2q∗(1 − q∗)T1ε∗(1 − ε∗)

= 1 , (15a)

or

ε∗ =12

[1 − U(ηC , q∗)

], (15b)

with

U(ηC , q∗) ≡√

4ηCq∗(1 − q∗) + (1 − 2q∗)2 . (16)

By substituting ε∗ as a function of q∗ in Eq. (15b) to Eq. (14a)or Eq. (14b), we get the optimum condition

ln(

1 − q∗

q∗

)−

T2

T1ln

[1 + U(ηC , q∗)1 − U(ηC , q∗)

]

q∗ −12

+12

U(ηC , q∗)

q∗(1 − q∗)= 0 .

(17)

Furthermore, the condition in Eq. (17) leads to the followingform of ηop from Eq. (7),

ηop =

q∗ −12

+12

U(ηC , q∗)

q∗(1 − q∗) ln[(1 − q∗)/q∗]. (18)

It is also straightforward to show that this point (q∗, ε∗) is in-deed a maximum point by investigating the second derivativesof the power.

In order to calculate the efficiency for given T2/T1 at themaximum power, first find the q∗ value satisfying Eq. (17) and

Page 4: arXiv:1612.00518v1 [cond-mat.stat-mech] 1 Dec 2016 · E ciency at the maximum power output for simple two-level heat engine Sang Hoon Lee,1 Jaegon Um,2,3 and Hyunggyu Park1,2 1School

4

0

0.05

0.1

0.15

0.2

0 0.2 0.4 0.6 0.8 1

q*(ηc→0) = ε*(ηc→0)

q*(ηc=1)op

timal

tran

sitio

n ra

tes

ηc

q*ε*

ηc→0 and 1 asymptotes

FIG. 3. Numerically found q∗ and ε∗ values satisfying Eq. (17), as afunction of ηC = 1−T2/T1, along with the q∗(ηC → 0) = ε∗(ηC → 0)and q∗(ηC = 1) values presented in Sec. III B 2. ε∗(ηC = 1) = 0 (thehorizontal axis). The ηC → 0 asymptote indicates Eq. (19) up tothe η2

C term, and the ηC → 1 asymptote indicates Eq. (25) up to the(1 − ηC) term with the coefficients given by Eqs. (27) and (28).

q

✏ = q

net power < 0

⌘C from 0 to 1

q⇤(⌘C ! 0) = ✏⇤(⌘C ! 0) ' 0.083 222

q⇤(⌘C = 1) ' 0.217 812✏⇤(⌘C = 1) = 0

FIG. 4. Illustration of the optimal transition rates (q∗, ε∗) for the max-imum power output as the T2/T1 value varies.

substitute the q∗ value to Eq. (18). As Eq. (17) is a transcen-dental equation, the closed-form solution for ηop is unattain-able.

2. Asymptotic behaviors obtained from series expansion

The upper bound for q∗ is given by the condition ηC = 1,satisfying ln[(1 − q∗)/q∗] = 1/(1 − q∗) and q∗(ηC = 1) '0.217 812 found numerically and ε∗(ηC = 1) = 0 exactly fromEq. (15b). ηC = 0 always satisfies Eq. (17) regardless of q∗

values, so finding the optimal q∗ is meaningless (in fact, whenηC = 0, the operating regime for the engine is shrunk to theline q = ε and there cannot be any positive work). Therefore,

0

0.2

0.4

0.6

0.8

1

0 0.2 0.4 0.6 0.8 1

ηop

ηc

at (q*, ε*)ηCA = 1−√1−ηc

ηc/(2−ηc)ηc/2

ηc→1 asymptote

0.88

0.92

0.96

1

0.97 0.98 0.99 1

ηop

ηc

FIG. 5. The efficiency at the maximum power ηop as the function ofthe Carnot efficiency ηC in Eq. (18) using numerically found optimalq∗ values, along with various asymptotic cases: the Curzon-Ahlbornefficiency ηCA in Eq. (23), the upper bound ηC/(2−ηC) and the lowerbound ηC/2 in Ref. [22], and the ηC → 1 asymptote for ηC ≥ 0.8.The inset shows the region 0.97 < ηC < 1.

let us examine the case ηC ' 0 using the series expansion ofq∗ with respect to ηC , as

q∗ = a0 + a1ηC + a2η2C + a3η

3C + O

(η4

C

). (19)

Substituting Eq. (19) into Eq. (17) and expanding the left-handside with respect to ηC again, we obtain

c1ηC + c2η2C + c3η

3C + O

(η4

C

)= 0 , (20)

where cn describes the relation among a0, · · · , an−1, each ofwhich should be identically zero to satisfy Eq. (20). Lettingthe linear coefficient c1 be zero yields

21 − 2a0

= ln(

1 − a0

a0

), (21)

from which the lower bound for q∗(ηC → 0) = a0 = ε∗(ηC →

0) ' 0.083 222 found numerically [limηC→0 U(ηC , q∗) = 1 −2q∗, thus ε∗(ηC → 0) = q∗(ηC → 0) by Eq. (15b)]. Figure 3shows the numerical solution (q, ε) = (q∗, ε∗) as a function ofηC , where the asymptotic behaviors derived above hold whenηC ' 0 and ηC ' 1. It seems that q∗ is monotonically increasedand ε∗ is monotonically decreased, as ηC is increased, i.e.,q∗min = q∗(ηC → 0), q∗max = q∗(ηC = 1), ε∗min = 0, and ε∗max =

ε∗(ηC → 0). Figure 4 illustrates the situation on the (q, ε)plane. The linear coefficient a1 in Eq. (19) can be written interms of a0 when we let c2 = 0 in Eq. (20), and the quadraticcoefficient a2 in Eq. (19) can also be written in terms of a0alone, by letting c3 = 0 in Eq. (20) and using the relationsin Eqs. (21) and a1 expressed by a0 terms, which are wellconsistent with the numerical solution as shown in Fig. 3.

With the relations of coefficients in hand, we find theasymptotic behavior of ηop in Eq. (18) by expanding it withrespect to ηC after substituting q∗ as the series expansion of

Page 5: arXiv:1612.00518v1 [cond-mat.stat-mech] 1 Dec 2016 · E ciency at the maximum power output for simple two-level heat engine Sang Hoon Lee,1 Jaegon Um,2,3 and Hyunggyu Park1,2 1School

5

ηC in Eq. (19). Then,

ηop =12ηC +

18η2

C +7 − 24a0 + 24a2

0

96(1 − 2a0)2 η3C + O

(η4

C

). (22)

With this method, we are able to find the coefficients in termsof a0 up to an arbitrary order in principle. We would liketo emphasize that the expansion form of ηop in Eq. (22) hasexactly the same coefficients up to the quadratic term to thoseof the Curzon-Ahlborn efficiency [3–5] defined as

ηCA = 1 −√

T2/T1 = 1 −√

1 − ηC , (23)

with the expansion form

ηCA =12ηC +

18η2

C +1

16η3

C +5

128η4

C + O(η5C) , (24)

when ηC ' 0. As a result, numerically found ηop by solvingEq. (17) and substituting the q∗ value to Eq. (18), and ηCAshare a very similar functional form for ηC . 1/2, as shown inFig. 5. In fact, the linear term ηC/2 and quadratic term η2

C/8are naturally from the strong coupling between the thermo-dynamic fluxes and the symmetry between the reservoirs (aswe will check in Sec. III C, the reservoir symmetry is relatedto the symmetry in the entropy production at the hot and coldreservoir and holds only approximately in our model) [23, 24].The third order coefficient (' 0.077 492) in Eq. (22), however,is clearly different from 1/16 for the ηCA. In other words, thedeviation from ηCA for ηop enters from the third order that hasnot been theoretically investigated yet. Indeed, ηop deviatesfrom ηCA for ηC & 1/2, until they coincide at ηC = 1. There-fore, the efficiency ηop of our model at maximum power outputis different from ηCA.

For ηC ' 1, we need to consider the logarithmic correc-tion due to the functional form, based on the numerical evi-dence also shown in Fig. 5. In contrast to the linear heat con-duction for the Curzon-Ahlborn endoreversible engine [3–5],our model has an exponential or Boltzmann type of relaxationprocess. We believe that this different functional form of heatconduction process results in the different types of singular-ity at ηC ' 1: the algebraic singularity of ηCA in Eq. (23) atηC = 1 with the infinite slope, and the logarithmic singular-ity in our case. We take the singular series expansion of thefunctional form in Eq. (17) near ηC = 1 as

q∗ =q∗max + bln(1 − ηC) ln(1 − ηC)

+ b1(1 − ηC) + O[(1 − ηC)2

].

(25)

It is possible to consider other types of terms such as (1 −ηC) ln2(1 − ηC), but we will check that it is enough to pre-dict the functional form of ηop, consistent with an alternativeapproach from entropy-production-based analysis provided inSec. III C. If we take only the zeroth order term, we obtain theidentity

11 − q∗max

= ln(

1 − q∗max

q∗max

), (26)

R1

E1

T1

q

R2

E2

T2

✏0

q

0✏

during duringstochastic Markov processes

⌧1 ⌧2

|P1i(t1 = 0) |P1i(t1 = ⌧1) = |P2i(t2 = 0) |P2i(t2 = ⌧2)

|P2i(t2 = ⌧2) = |P1i(t1 = 0)

W = E1 � E2

W 0 = E1 � E2

Q1 Q2

hWnetiT1

sh =T2

T1sc +

hWnetiT1

sh =E1

T1

sc

G�1 (sc)

sc

sh

FIG. 6. The entropy relation between sh and sc, given by Eq. (31) andthe linear relation in Eq. (33) representing the first law of thermody-namics. The maximum value of 〈Wnet〉 (the intercept of the linearrelation on the vertical axis times T1) is achieved when the line be-comes the tangential one of the curve, as illustrated here.

exactly at ηC = 1 that is already mentioned in the first part ofthis subsection. Similar to the ηC ' 0 case, by letting each co-efficient be zero, we find the relations among the coefficientsas

bln = q∗max(1 − q∗max)2 , (27)

and

b1 = q∗max(1 − q∗max)2 {1 + ln[q∗max(1 − q∗max)]

}, (28)

which are well consistent with the numerical solution asshown in Fig. 3.

Again, the asymptotic behavior of ηop in Eq. (18) for ηC ' 1can be deduced from the series expansion in terms of (1 −ηC) > 0, using Eqs. (25), (27), and (28), which is

ηop =1 + (1 − q∗max)(1 − ηC) ln(1 − ηC)+ (1 − q∗max) ln[q∗max(1 − q∗max)](1 − ηC)

+ O[(1 − ηC)2

],

(29)

based on the relations in Eqs. (27) and (28) [the same proce-dure as the one leading to (22)]. As shown in Fig. 5, however,the asymptotic form only holds in a rather limited range of ηCvery close to unity, indicating the necessity to taking higherorder terms into account for more accurate asymptotic behav-ior.

C. The entropy production relation

In Ref. [10], it is argued that the necessary and sufficientcondition for the Curzon-Ahlborn efficiency at the maximumpower is that the entropy productions at the hot and cold reser-voirs (denoted by sh and sc, respectively, ) should be relatedby a specific functional form, namely, sh = F (sc) whereF (x) = x/(1 + ζx) with the system-specific constant ζ.

Page 6: arXiv:1612.00518v1 [cond-mat.stat-mech] 1 Dec 2016 · E ciency at the maximum power output for simple two-level heat engine Sang Hoon Lee,1 Jaegon Um,2,3 and Hyunggyu Park1,2 1School

6

The entropy production in our model is given by

sh =〈Q1〉

T1≡ (q − ε)

E1

T1,

sc =〈Q2〉

T2≡ (q − ε)

E2

T2,

(30)

where we again discard the common explicit time dependentterm (1−e−τ/2)2/(1−e−τ), which does not affect the followingdiscussion for notational convenience. Given T1 and T2 andputting E1 as a constant, we obtain the entropy relation givenby

sh =E1

T1

sc

G−1 (sc), (31)

where G−1 is the inverse function of G defined from the rela-tion in Eq. (30), sc = G(E2/T2). Note that sh is an increasingfunction with respect to sc while dsh/dsc is a decreasing one,so sh(sc) is an increasing and concave function of sc as il-lustrated in Fig. 6. Therefore, the unique [guaranteed by theconcavity of sh(sc)] optimal entropy production denoted bys∗c, which makes the power reach its maximum value, is deter-mined by

dsh

dsc

∣∣∣∣∣sc=s∗c

=E1

T1

ε∗ε∗

q − ε + G−1 (s∗c

)ε∗ε∗

=T2

T1, (32)

where T2/T1 comes from the thermodynamic first law,

sh =T2

T1sc +

〈Wnet〉

T1. (33)

The parameter ε∗ obtained from Eq. (32) is still a function ofE1 or q. By optimizing the entropy production with respect toq, we find the same optimal q∗ and ε∗ as those in the previoussection.

As G−1(sc) is not a linear function of sc, we do not have therelation sh = F (sc) mentioned before, so the fact that ηop ,ηCA is consistent with Ref. [10]. However, we reveal that itis possible to find the regime where the entropy production ofour model approximately follows the functional form F (x),which indeed corresponds to the ηop ' ηCA regime, as weshow in the following section.

1. The linear regime

First, we take the regime where ∆T = T1 − T2 � T1. Thenfrom Eq. (32), one can see the solution ∆E∗ = E1 − E∗2 � E1

or s∗c � 1, which allows the small sc expansion of G−1 up tothe linear order as

G−1(sc) ' G−1(0) +dG−1

dsc

∣∣∣∣∣∣sc=0

sc =E1

T1+ ζ(E1/T1)sc , (34)

where the constant ζ(E1/T1) is given by

ζ = 2(

T1

E1

)2 [1 + cosh

(E1

T1

)]. (35)

Inserting Eq. (34) to the entropy relation in Eq. (31), we findthat the entropy production for hot and cold reservoirs actuallyfollows the functional form F (x) = x/(1 + ζx) for ηC → 0,which explains the Curzon-Ahlborn-like behavior for ηC → 0.We have already checked that ηop ' ηCA as ηC → 0 dueto the same linear and quadratic coefficients from the seriesexpansion in Sec. III B 2, but the entropy production relationsuggests that there could be a deeper relation between ourmodel and engines with the optimal efficiency of ηCA thanthe reasons for the linear and quadratic coefficients, namely,the strong coupling between the thermodynamics fluxes andthe reservoir symmetry, respectively. In fact, the implicationof the reservoir symmetry in the expression F −1(x) = F (−x)holds only for ηC → 0.

We emphasize that the behavior of efficiency ηop ' ηCA inηC → 0 is independent of the value q. However we can opti-mize 〈Wnet〉with respect to q as following. The optimal condi-tion to maximize 〈Wnet〉 for a given T2/T1 value is equivalentto minimize ζ, because

〈Wnet〉 =T1

ζ

1 − √T2

T1

2

∝ 1/ζ , (36)

from the condition of tangent T2/T1 = 1/(1+ζs∗c)2. The condi-tion for the minimum value of ζ for given E1/T1 is, by takingthe derivative of the functional form in Eq. (35) with respectto E1/T1 and using the relation q/(1 − q) = e−E1/T1 ,

21 − 2q

= ln(

1 − qq

), (37)

which is equivalent to the condition for the zeroth order termof q∗(ηC → 0) given by Eq. (21). In other words, the opti-mal condition, at least for the lowest order of ηC at ηC ' 0,can also be derived from the functional form of entropy pro-duction given by Eqs. (31) and (35), further supporting theconsistency of our result.

2. The logarithmic regime

The other extreme regime is T2 � T1, or the ηC → 1 limit,where the solution satisfying Eq. (32) exists in the region ofE∗2/T2 � 1 or sc � 1. In this limit we rewrite the relationbetween sc and E2/T2 in Eq. (30) as

E2

T2'

sc

q

(1 +

1q

e−E2/T2

). (38)

Using the above, we obtain G−1(sc) up to the order of(sc/q2)e−sc/q

G−1(sc) 'sc

q+

sc

q2 e−sc/q . (39)

Inserting Eq. (39) to Eq. (31), the entropy production relationreads

sh =E1

T1

q2

q + e−sc/q, (40)

Page 7: arXiv:1612.00518v1 [cond-mat.stat-mech] 1 Dec 2016 · E ciency at the maximum power output for simple two-level heat engine Sang Hoon Lee,1 Jaegon Um,2,3 and Hyunggyu Park1,2 1School

7

0

E2

✏q

E1

Em

q1m

qm ✏m

✏2m

R1 R2

FIG. 7. The energy levels for each reservoir in the three-level heatengine, along with the transition rates, are illustrated. The work isextracted by adjusting the most upper level (the second excited state)between the reservoir contacts.

and using the condition

dsh

dsc

∣∣∣∣∣sc=s∗c

=E1

T1

qe−sc/q

(q + e−sc/q)2 'E1e−sc/q

qT1=

T2

T1, (41)

we derive the efficiency at the maximum power,

ηop ' 1−T1

E1(1 − ηC)

[1 +

T1

E1(1 − ηC)

]× ln

[E1

qT1 (1 − ηC)

].

(42)

Optimizing the power with respect to q or E1 is equivalentto the maximizing the saturation value of Eq. (40), which islimsc→∞ sh = qE1/T1. This condition yields the functionalform q∗max should satisfy, which is equivalent to Eq. (26). Us-ing the result E1/T1 = (1 − q∗max)−1, we obtain ηop in the (q, ε)space given by

ηop ' 1+(1 − q∗max

)(1 − ηC)

[1 +

(1 − q∗max

)(1 − ηC)

]× ln

[q∗max

(1 − q∗max

)(1 − ηC)

],

(43)

which is consistent with the result in the previous section; infact, Eq. (43) includes higher order terms, namely, (1 − ηC)2

and (1−ηC)2 ln(1−ηC), than Eq. (29). We also emphasize thatthis consistency justifies the series expansion form in Eq. (25).

Based on the analysis above for the regime where T2 � T1,we reach the conclusion that there cannot be a q value makingthe higher order terms than the linear term in the expansionof G−1 for any given T1 and T2 values. In other words, oneagain can see the systematic difference between our modeland the models belonging to the Curzon-Ahlborn efficiency atthe maximum power.

IV. EXTENSION TO MULTI-LEVEL HEAT ENGINE

Finally, we would like to remark on the possible extensionto multi-level systems, i.e., systems with more than two levels,which are more general cases. In that framework, our two-level heat engine can be taken as a simplified one consideringonly the ground and first excited states. For simplicity, again

we assume two heat reservoirs R1 and R2, which are charac-terized by the temperatures T1 and T2, and contacted with thesystem during the time τ1 and τ2, respectively.

First, let us take the three-level system, where we considerthe ground, first excited, and second excited states for eachreservoir. We further simplify the situation by differentiatingonly the second excited states of the reservoirs, namely E1for R1 and E2 for R2, and the common value Em for the firstexcited state, as depicted in Fig. 7. The transition rates aredenoted by q (the ground state to E1 in R1), qm (the groundstate to Em in R1), q1m (Em to E1 in R1), ε (the ground state toE2 in R2), εm (the ground state to Em in R2), and ε2m (Em to E2in R2); their reverse transition rates are q, qm, q1m, ε, εm, andε2m, respectively. Applying the Arrhenius form, we obtain therelations

q/q = e−E1/T1 ,

qm/qm = e−Em/T1 ,

q1m/q1m = e−(E1−Em)/T1 ,

ε/ε = e−E2/T2 ,

εm/εm = e−Em/T2 ,

ε2m/ε2m = e−(E2−Em)/T2 .

(44)

As in the two-level case, the net amount of work from thepopulation difference in different energy levels (only for thesecond excited states in this case) is 〈Wnet〉 = (P1e −P2e)(E1 −

E2), where P1e and P2e refer to the population of E1 in R1and E2 in R2, respectively. The heat exchange, on the otherhand, should take the Em level into account. As a result, theefficiency for the three-level system is given by

η =(P1e − P2e)(E1 − E2)

(P1e − P2e)E1 + (P1m − P2m)Em, (45)

where the term involving Em, unless it vanishes, represents theextra heat exchange that cannot be used in the work extraction.

In contrast to the two-level case, the temporal part (involv-ing τ1 and τ2) is not factorized in the functional form of〈Wnet〉/τ = (P1e − P2e)(E1 − E2)/τ for this three-level case,so we cannot focus solely on the thermodynamics parameters.As shown in Fig. 8, the overall functional shape of power out-put 〈Wnet〉/τ varies over τ and the maximum value of poweroutput occurs at different values of (q∗, ε∗) depending on τ.Therefore, we conclude that the two-level system of our maininterest is a special case that we can analyze deeply to obtainthe insight presented so far. Moreover, for the three-level sys-tem, even at the limit q ' ε (which corresponds to the equilib-rium or reversible limit for the two-level system, representedby the equilibrium distribution of population), there cannot bethe equilibrium condition given by

e−Em/T1/Z = e−Em/T2/Z ,

e−E1/T1/Z = e−E2/T2/Z ,(46)

with the partition function Z = 1 + e−Em/T1 + e−E1/T1 =

1 + e−Em/T2 + e−E2/T2 , unless Em = 0. Hence, the conditionof strong coupling between thermodynamic fluxes is also vio-lated, which results in the linear coefficient of the efficiency

Page 8: arXiv:1612.00518v1 [cond-mat.stat-mech] 1 Dec 2016 · E ciency at the maximum power output for simple two-level heat engine Sang Hoon Lee,1 Jaegon Um,2,3 and Hyunggyu Park1,2 1School

8

(a) (b)T1 = 1.0, T2 = 0.1, Em = 0.5, τ = 1.0

0.1 0.2 0.3 0.4

q

0.1

0.2

0.3

0.4

ε

0

0.005

0.01

0.015

0.02

0.025

0.03T1 = 1.0, T2 = 0.1, Em = 0.5, τ = 10.0

0.1 0.2 0.3 0.4

q

0.1

0.2

0.3

0.4

ε

0

0.002

0.004

0.006

0.008

0.01

0.012

FIG. 8. The average power output 〈Wnet〉/τ for the three-level system with the parameters T1 = 1, T2 = 0.1, Em = 0.5, and (a) τ = 1 [themaximum value of 〈Wnet〉/τ occurs at q∗ = 0.215(5) and ε∗ = 0.005(5)], and (b) τ = 10.0 [q∗ = 0.185(5) and ε∗ = 0.010(5)]. For bettervisibility focused on the 〈Wnet〉/τ ≥ 0 regime, we set all of the negative values as 0. We take the normalization convention q + q = qm + qm =

q1m + q1m = ε + ε = εm + ε = ε2m + ε2m = 1.

ηop at the maximum power in terms of ηC different from1/2 [23] when Em > 0, and we have numerically verified thefact as well, where we have obtained ηop for the parameters qand ε, for given Em and τ values. In contrast, recall that for thetwo-level system, the strong coupling between thermodynam-ics always holds, and the reservoir symmetry approximatelyholds for ηC → 0 (corresponding to q∗ ' ε∗ for the maximumpower output). The necessity for the extra heat in Eq. (45)also prevents the ηop from reaching unity at the other limitingcase ηC → 1, which we have also verified numerically.

V. CONCLUSIONS AND DISCUSSION

We have demonstrated that our simple two-level heat en-gine model has a nontrivial parameter relation for the effi-ciency at the maximum power output. Thanks to the simplic-ity of our model composed of the two-level system, the time-dependent term only plays the role of an overall factor, so wehave focused on the relative transition rates for a given tem-perature ratio of reservoirs. Based on numerical solutions andanalytically driven asymptotic behaviors, we have shown thatthe optimal efficiency ηop for maximum power output in our

model is clearly different from the Curzon-Ahlborn efficiencyηCA [3–5], although they share the same asymptotic behaviorup to the quadratic term when ηC ' 0 [23, 24]. We have dis-cussed its implication, in conjunction with the relation of theentropy production at the hot and cold reservoirs.

We have focused on the average thermodynamic quanti-ties to yield the macroscopic efficiency in this paper, but itwould be possible to consider the stochastic efficiency [13–18] by adopting more specific protocols involved in the heatand work transfer, which can be a future work, along with thequantum effects [25–33]. For comprehensive understanding,we would also need the full consideration of multi-level sys-tems sketched in Sec. IV here, which we leave as future work.

Note added.—Just before the submission of this paper, wehave learned that a recent contribution by Toral et al. [34] in-dependently reports the same form of ηop in Ising spin systemsor exclusion processes.

ACKNOWLEDGMENTS

We thank Hyun-Myung Chun, Jae Dong Noh, Hee JoonJeon, and Sang Wook Kim for fruitful discussions and com-ments.

[1] K. Huang, Statistical Mechanics (John Wiley & Sons, NewYork, 1963).

[2] S. Carnot, Reflexions Sur La Puissance Motrice Du Feu Et SurLes Machines Propres A Developper Cette Puissance (Bache-lier Libraire, Paris, 1824).

[3] P. Chambadal, Les Centrales Nuclaires (Armand Colin, Paris,1957).

[4] I. I. Novikov, Efficiency of an atomic power generating installa-tion, At. Energy (N.Y.) 3, 1269 (1957); The efficiency of atomicpower stations, J. Nucl. Energy 7, 125 (1958).

[5] F. L. Curzon and B. Ahlborn, Efficiency of a Carnot engine atmaximum power output, Am. J. Phys. 43, 22 (1975).

[6] J. Hoppenau, M. Niemann, and A. Engel, Carnot process witha single particle, Phys. Rev. E 87, 062127 (2013).

[7] K. Proesmans, C. Driesen, B. Cleuren, and C. Van den Broeck,Efficiency of single-particle engines, Phys. Rev. E 92, 032105(2015).

[8] J. Um, H. Hinrichsen, C. Kwon, and H. Park, Total cot of oper-ating an information engine, New J. Phys. 17, 085001 (2015).

[9] V. Holubec and A. Ryabov, Efficiency at and near maximum

Page 9: arXiv:1612.00518v1 [cond-mat.stat-mech] 1 Dec 2016 · E ciency at the maximum power output for simple two-level heat engine Sang Hoon Lee,1 Jaegon Um,2,3 and Hyunggyu Park1,2 1School

9

power of low-dissipation heat engines, Phys. Rev. E 92, 052125(2015).

[10] J.-M. Park, H.-M. Chun, and J. D. Noh, Efficiency at maximumpower and efficiency fluctuations in a linear Brownian heat en-gine model, Phys. Rev. E 94, 012127 (2016).

[11] A. Ryabov and V. Holubec, Maximum efficiency of steady-stateheat engines at arbitrary power, Phys. Rev. E 93, 050101(R)(2016).

[12] N. Shiraishi, K. Saito, and H. Tasaki, Universal trade-off rela-tion between power and efficiency for heat engines, Phys. Rev.Lett. 117, 190601 (2016).

[13] G. Verley, M. Esposito, T. Willaert, and C. Van den Broeck, Theunlikely Carnot efficiency, Nat. Commun. 5, 4721 (2014).

[14] G. Verley, T. Willaert, C. Van den Broeck, and M. Esposito,Universal theory of efficiency fluctuations, Phys. Rev. E 90,052145 (2014).

[15] T. R. Gingrich, G. M. Rotskoff, S. Vaikuntanathan, and P. L.Geissler, Efficiency and large deviations in time-asymmetricstochastic heat engines, New J. Phys. 16, 102003 (2014).

[16] K. Proesmans, B. Cleuren, and C. Van den Broeck, Stochasticefficiency for effusion as a thermal engine, Europhys. Lett. 109,20004 (2015).

[17] K. Proesmans and C. Van den Broeck, Stochastic efficiency:five case studies, New J. Phys. 17, 065004 (2015).

[18] M. Polettini, G. Verley, and M. Esposito, Efficiency statisticsat all times: Carnot limit at finite power, Phys. Rev. Lett. 114,050601 (2015).

[19] J. Schnakenberg, Network theory of microscopic and macro-scopic behavior of master equation systems, Rev. Mod. Phys.48, 571 (1976).

[20] D. Andrieux and P. Gaspard, Fluctuation theorem and Onsagerreciprocity relations, J. Chem. Phys. 121, 6167 (2004).

[21] U. Seifert, Entropy production along a stochastic trajectory andan integral fluctuation theorem, Phys. Rev. Lett. 95, 040602

(2005).[22] M. Esposito, R. Kawai, K. Lindenberg, and C. Van den Broeck,

Efficiency at maximum power of low-dissipation Carnot en-gines, Phys. Rev. Lett. 105, 150603 (2010).

[23] C. Van den Broeck, Thermodynamic efficiency at maximumpower, Phys. Rev. Lett. 95, 190602 (2005).

[24] M. Esposito, K. Lindenberg, and C. Van den Broeck, Univer-sality of efficiency at maximum power, Phys. Rev. Lett. 102,130602 (2009).

[25] H. E. D. Scovil and E. O. Schulz-DuBois, Three-level masers asheat engines, Phys. Rev. Lett. 2, 262 (1959).

[26] J. E. Geusic, E. O. Schulz-DuBois, and H. E. D. Scovil, Quan-tum equivalent of the carnot cycle, Phys. Rev. 156, 343 (1967).

[27] C. M. Bender, D. C. Brody, and B. K. Meister, Quantum me-chanical Carnot engine, J. Phys. A 33, 4427 (2000).

[28] S. Abe and S. Okuyama, Similarity between quantum mechan-ics and thermodynamics: Entropy, temperature, and Carnot cy-cle, Phys. Rev. E 83, 021121 (2011).

[29] S. Abe, Maximum-power quantum-mechanical Carnot engine,Phys. Rev. E 83, 041117 (2011).

[30] J. Wang, J. He, and X. He, Performance analysis of a two-statequantum heat engine working with a single-mode radiation fieldin a cavity, Phys. Rev. E 84, 041127 (2011).

[31] R. Uzdin, Am. Levy, and R. Kosloff, Equivalence of quantumheat machines, and quantum-thermodynamic signatures, Phys.Rev. X 5, 031044 (2015).

[32] H. J. Jeon and S. W. Kim, Optimal work of the quantum Szilardengine under isothermal processes with inevitable irreversibil-ity, New J. Phys. 18, 043002 (2016).

[33] S. Liu and C. Ou, Maximum power output of quantum heatengine with energy bath, Entropy 18, 205 (2016).

[34] R. Toral, C. Van den Broeck, D. Escaff, and K. Lindenberg,Stochastic thermodynamics for Ising chain and symmetric ex-clusion process, e-print arXiv:1611.08188.


Recommended