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

Similar documents
Efficiency at Maximum Power in Weak Dissipation Regimes

Thermodynamics for small devices: From fluctuation relations to stochastic efficiencies. Massimiliano Esposito

Introduction to Stochastic Thermodynamics: Application to Thermo- and Photo-electricity in small devices

Nonequilibrium Thermodynamics of Small Systems: Classical and Quantum Aspects. Massimiliano Esposito

J. Stat. Mech. (2011) P07008

Unified Approach to Thermodynamic Optimization of Generic Objective Functions in the Linear Response Regime

arxiv: v2 [cond-mat.stat-mech] 26 Jan 2012

Efficiency of Inefficient Endoreversible Thermal Machines

Optimal quantum driving of a thermal machine

arxiv: v1 [cond-mat.stat-mech] 28 Sep 2017

arxiv: v1 [cond-mat.stat-mech] 6 Nov 2015

Introduction to a few basic concepts in thermoelectricity

Order and Disorder in Open Systems

Entropy and the second law of thermodynamics

Carnot Theory: Derivation and Extension*

Notes on Thermodynamics of Thermoelectricity - Collège de France - Spring 2013

arxiv: v1 [cond-mat.stat-mech] 10 Feb 2010

Efficiency at maximum power of thermally coupled heat engines

Thermo-economics of an irreversible solar driven heat engine

Bounds on thermal efficiency from inference

SUSTAINABLE EFFICIENCY OF FAR-FROM-EQUILIBRIUM SYSTEMS

PHYSICS 715 COURSE NOTES WEEK 1

PERFORMANCE ANALYSIS OF SOLAR DRIVEN HEAT ENGINE WITH INTERNAL IRREVERSIBILITIES UNDER MAXIMUM POWER AND POWER DENSITY CONDITION

General formula for the efficiency of quantum-mechanical analog of the Carnot engine

8.044 Lecture Notes Chapter 5: Thermodynamcs, Part 2

Quantum heat engine using energy quantization in potential barrier

Performance of Discrete Heat Engines and Heat Pumps in Finite. Time. Abstract

Second law, entropy production, and reversibility in thermodynamics of information

Engine efficiency at maximum power, entropy production. and equilibrium thermodynamics

Quantifying Dissipation

Reversibility. Processes in nature are always irreversible: far from equilibrium

Entropy production and time asymmetry in nonequilibrium fluctuations

Irreversibility and the arrow of time in a quenched quantum system. Eric Lutz Department of Physics University of Erlangen-Nuremberg

Quantum-Mechanical Carnot Engine

CENTER FOR NONLINEAR AND COMPLEX SYSTEMS. Como - Italy

General Formula for the Efficiency of Quantum-Mechanical Analog of the Carnot Engine

arxiv: v1 [cond-mat.stat-mech] 6 Mar 2008

Carnot efficiency at divergent power output

Enhancing Otto-mobile Efficiency via Addition of a Quantum Carnot Cycle

Adiabats and entropy (Hiroshi Matsuoka) In this section, we will define the absolute temperature scale and entropy.

In this report, I discuss two different topics of thermodynamics and explain how they have

Nonequilibrium thermodynamics at the microscale

1. Second Law of Thermodynamics

Irreversible Processes

arxiv: v2 [cond-mat.stat-mech] 9 Jul 2012

First Law showed the equivalence of work and heat. Suggests engine can run in a cycle and convert heat into useful work.

Stability of Tsallis entropy and instabilities of Rényi and normalized Tsallis entropies: A basis for q-exponential distributions

12 The Laws of Thermodynamics

Title of communication, titles not fitting in one line will break automatically

arxiv: v1 [quant-ph] 14 Apr 2015

CHAPTER V. Brownian motion. V.1 Langevin dynamics

The Second Law of Thermodynamics

Reversible Processes. Furthermore, there must be no friction (i.e. mechanical energy loss) or turbulence i.e. it must be infinitely slow.

1. Second Law of Thermodynamics

UNDERSTANDING BOLTZMANN S ANALYSIS VIA. Contents SOLVABLE MODELS

Chapter 3. The Second Law Fall Semester Physical Chemistry 1 (CHM2201)

Thermodynamics of nuclei in thermal contact

arxiv: v3 [quant-ph] 24 Nov 2014

Entropy and the Second and Third Laws of Thermodynamics

PHYSICAL REVIEW E 88, (2013)

Imperial College London BSc/MSci EXAMINATION May 2008 THERMODYNAMICS & STATISTICAL PHYSICS

UNIVERSITY OF SOUTHAMPTON

Physics 53. Thermal Physics 1. Statistics are like a bikini. What they reveal is suggestive; what they conceal is vital.

A Simple Model for Carnot Heat Engines

October 18, 2011 Carnot cycle - 1

Photon steam engines. Work can be extracted from a single heat bath at the boundary between classical and quantum thermodynamics

18.13 Review & Summary

IV. Classical Statistical Mechanics

(# = %(& )(* +,(- Closed system, well-defined energy (or e.g. E± E/2): Microcanonical ensemble

Chapter 16 Thermodynamics

Lecture 2 Entropy and Second Law

Resource Theory of Work and Heat

Summarizing, Key Point: An irreversible process is either spontaneous (ΔS universe > 0) or does not occur (ΔS universe < 0)

arxiv:physics/ v4 [physics.class-ph] 27 Mar 2006

Introduction. Statistical physics: microscopic foundation of thermodynamics degrees of freedom 2 3 state variables!

CHAPTER - 12 THERMODYNAMICS

Elements of Statistical Mechanics

Entropy and irreversibility in gas dynamics. Joint work with T. Bodineau, I. Gallagher and S. Simonella

MASSACHUSETTS INSTITUTE OF TECHNOLOGY Physics Department Statistical Physics I Spring Term 2013 Notes on the Microcanonical Ensemble

Chapter 12. The Laws of Thermodynamics. First Law of Thermodynamics

Reversibility, Irreversibility and Carnot cycle. Irreversible Processes. Reversible Processes. Carnot Cycle

MACROSCOPIC VARIABLES, THERMAL EQUILIBRIUM. Contents AND BOLTZMANN ENTROPY. 1 Macroscopic Variables 3. 2 Local quantities and Hydrodynamics fields 4

Microscopic Deterministic Dynamics and Persistence Exponent arxiv:cond-mat/ v1 [cond-mat.stat-mech] 22 Sep 1999

Chapter 16 The Second Law of Thermodynamics

Gibbs Paradox Solution

Chapter 20 The Second Law of Thermodynamics

Comment on: Sadi Carnot on Carnot s theorem.

The Story of Spontaneity and Energy Dispersal. You never get what you want: 100% return on investment

Effective Temperatures in Driven Systems near Jamming

International Physics Course Entrance Examination Questions

arxiv: v3 [cond-mat.stat-mech] 4 May 2016

OPTIMIZATION OF AN IRREVERSIBLE OTTO AND DIESEL CYCLES BASED ON ECOLOGICAL FUNCTION. Paraná Federal Institute, Jacarezinho, Paraná, Brasil.

Maxwell's Demon in Biochemical Signal Transduction

arxiv:physics/ v2 [physics.class-ph] 18 Dec 2006

Fluctuation Theorem for a Small Engine and Magnetization Switching by Spin Torque

Thermodynamics. 1.1 Introduction. Thermodynamics is a phenomenological description of properties of macroscopic systems in thermal equilibrium.

Removing the mystery of entropy and thermodynamics. Part 3

arxiv:cond-mat/ v1 4 Aug 2003

Entropy production fluctuation theorem and the nonequilibrium work relation for free energy differences

arxiv: v3 [cond-mat.stat-mech] 19 Mar 2015

Transcription:

Efficiency 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 2455 Korea 2 Quantum Universe Center Korea Institute for Advanced Study Seoul 2455 Korea 3 CCSS CTP and Department of Physics and Astronomy Seoul National University Seoul 8826 Korea respectively and P 1 = ( P 1e P 1g ) T and P2 = ( P 2e P 2g ) T are the column vectors whose components represent the populaarxiv:1612.518v1 [cond-mat.stat-mech] 1 Dec 216 We introduce a simple two-level heat engine to study the efficiency in the condition of the maximum power output depending on the energy levels from which the net work is extracted. In contrast to the uasi-statically operated Carnot engine whose efficiency reaches the theoretical maximum recent research on more realistic engines operated in finite time has revealed other classes of efficiency such as the Curzon-Ahlborn efficiency 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 difference caused by different 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 efficiency of our model is different from the Curzon-Ahlborn efficiency although they share the universal linear and uadratic coefficients at the near-euilibrium limit. We further confirm our results by taking an alternative approach in terms of the entropy production at hot and cold reservoirs. PACS numbers: 5.7.Ln 5.4.a 5.2.y 89.7.a I. INTRODUCTION The efficiency of heat engines is a celebrated topic of classical thermodynamics [1]. In particular an elegant formula expressed only by hot and cold reservoir temperatures for the ideal uasi-static and reversible engine coined by Sadi Carnot has been an everlasting textbook example [2]. That ideal engine however is not the most efficient engine any more when we consider its power (the extracted work per unit time) which has added different types of optimal engine efficiency such as the Curzon-Ahlborn efficiency for some cases [3 5]. Following such steps researchers have taken simple systems to investigate various theoretical aspects of underlying principles of macroscopic thermodynamic engine efficiency in details [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 relation between transition rates at the maximum power for a given temperature ratio. We compare the functional form of the optimal efficiency at the maximum power to previously known forms for some other cases. Our result shows a difference from the Curzon-Ahlborn efficiency but shares the same asymptotic behavior up to the second order in a small efficiency limit. We also take an alternative approach considering the entropy production at the reservoirs and discuss its implication. Generalization to multi-level engines is considered but a decoupling of the operating time does not happen which makes the analytic investigation uite complex. II. TWO-LEVEL HEAT ENGINE ground state (E = ) 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 and ɛ respectively and their reverse processes by and ɛ. We assume E 1 > E 2 and > T 2. The system is attached to two different reservoirs: R 1 with temperature during 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 ) in Fig. 1 the net positive work is achievable due to the difference 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. Efficiency 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 Arrhenius form / = e E 1/ ɛ/ ɛ = e E 2/T 2 respectively (we let the Boltzmann constant k B = 1 for notational convenience) thus the ineuality < ɛ < < 1/2 holds (ɛ < 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 population difference are given by W = (E 1 E 2 )P 1e W = (E 1 E 2 )P 2e (1) (2) Figure 1 illustrates our model. The two-level system is characterized by two discrete energy states composed of the

2 W = E 1 E 2 R 1 P 2 i(t 2 = 2 )= P 1 i(t 1 = ) R 2 E 1 P 1 i(t 1 = ) P 1 i(t 1 = 1 )= P 2 i(t 2 = ) P 2 i(t 2 = 2 ) during 1 during 2 stochastic Markov processes T 2 Q 1 W = E 1 E 2 Q 2 FIG. 1. Schematic illustration of our simple two-level heat engine composed s h s h = T 2 s c + hw of two energy levels coupled with two heat reservoirs R 1 and R 2. neti tions of excited and ground states (in that order) at the end of the contact with R 1 and R 2 respectively. We then take the normalization convention P 1e +P 1g = P 2e +P 2g = + = ɛ+ ɛ = 1 expressing the conservation of total population. For notational convenience we define the function X(T r) of temperature T and transition rate r as X(T r) ThW ln( r/r) net i. (3) Then the average amount of heat to the system from R 1 and that from the system to R 2 are Q 1 = (P 1e P 2e )X( ) Q 2 = (P 1e P 2e )X(T 2 ɛ) respectively based on the Schnakenberg entropy production for stochastic processes [19 21]. The average total entropy production during one cycle is given by the entropy change of the reservoir S = Q 1 + Q 2 T 2 = (P 1e P 2e ) [ ln( ɛ/ɛ) ln( /) ]. Es. (2) and (4) ensure the energy conservation or the first law of thermodynamics W W = Q 1 Q 2 considering E. (1). The average net work extracted from the system is W net = W W = (P 1e P 2e ) [ X( ) X(T 2 ɛ) ] (6) and the efficiency is given by the ratio (4) (5) η = W net Q 1 = 1 X(T 2 ɛ) X( ) (7) independent of τ 1 and τ 2 and η approaches η C = 1 T 2 / (the Carnot efficiency [1 2]) when ɛ and meaningful only for > ɛ or W net >. Now let us consider the explicit form of populations at the excited states at end of each reservoir contact process whose E 2 time evolution is given by the following linear differential euation system for given and ɛ values [ ] s h = E 1 s d P c 1 = P G 1 (s dt c ) 1 1 [ ] (8) d P 2 ɛ ɛ = P dt 2 ɛ ɛ 2 where t 1 τ 1 and t 2 τ 2 are the intermediate time spent in contact with R 1 and R 2 respectively. As the populations do not change during the adiabatic work extraction (supply) processes we get the circular boundary condition as P 1e (t 1 = s c t 2 = τ 2 ) = P 2e (t 1 = τ 1 t 2 = τ 2 ) and P 2e (t 1 = τ 1 t 2 = ) = P 1e (t 1 = τ 1 t 2 = τ 2 ). Thus the solution at t 1 = τ 1 and t 2 = τ 2 is given by P 1e = (1 e τ 1 ) + ɛ(1 e τ 2 )e τ 1 1 e (τ 1+τ 2 ) P 2e = ɛ(1 e τ 2 ) + (1 e τ 1 )e τ 2 1 e (τ 1+τ 2 ) and lim τ1 τ 2 P 1e = and lim τ1 τ 2 P 2e = ɛ as expected. With τ 1 = τ 2 = τ/2 we obtain the average net work as W net = ( ɛ)(1 e τ/2 ) 2 1 e τ [ X(T1 ) X(T 2 ɛ) ] (1) so the monotonically increasing factor (1 e τ/2 ) 2 /(1 e τ ) for the time scale τ is decoupled from the rest of the formula and only plays the role of an overall factor. It is important to note that the decoupling holds regardless of the τ 1 = τ 2 condition; the overall factor becomes (1 e τ 1 )(1 e τ 2 )/(1 e τ 1+τ 2 ). The average power output P is given by P = ( ɛ)(1 e τ/2 ) 2 τ(1 e τ ) (9) [ X(T1 ) X(T 2 ɛ) ] (11) which decreases monotonically with τ. Therefore from now on we discard the time dependence altogether and focus on other parameters i.e. denoting W net P ( ɛ) [ X( ) X(T 2 ɛ) ] (12)

3 <Wnet>(τ ) T1 = 1 T2 = 1/2.5.4.35.3.25.2.15.1.5.4.3.1 η T1 = 1 T2 = 1/2 (b).5.4.3.1.1.3.4.5 η T1 = 1 T2 = 9/1.1.3.4.5 <Wnet>(τ ) T1 = 1 T2 = 9/1 (d).5.5.12.4.1.4.8.3.3.6.4.1.1.9.8.7.6.5.4.3.2.1 (c).5.45.4.35.3 5.15.1.5 (a).1.2.1.3.4.5.1.3.4.5 FIG. 2. (a) The average net work limτ hwnet i and (b) efficiency η for = 1 and T 2 = 1/2 and (c) hwnet i(τ ) and (d) η for = 1 and T 2 = 9/1. For better visibility focused on the hwnet i regime we set all of the negative values as. without considering the overall factor involving τ for notational convenience. Numerically we obtain the net work and efficiency for ( ) combination as shown in Fig. 2. In Sec. III B we derive the condition for the efficiency at the maximum power output. or = 1. Efficiency at the maximum power output The condition for the maximum power output For a given T 2 / value the maximum power output condition for the two-variable function is hpi hpi = = (13) = = = = which leads to X(T 2 ) 1 = X( ) (1 ) ln[(1 )/ ] (14a) and 1 X(T 2 ) (T 2 / )( ) = X( ) (1 ) ln[(1 )/ ] (14b) from E. (12). By eliminating the left-hand side of Es. (14a) and (14b) we obtain the following simple relation T 2 (1 ) = 1 (1 ) (15a) (15b) with U(ηC ) B. 1 1 U(ηC ) 2 4ηC (1 ) + (1 2 )2. (16) By substituting as a function of in E. (15b) to E. (14a) or E. (14b) we get the optimum condition! " # T2 1 + U(ηC ) 1 ln ln T1 1 U(ηC ) (17) 1 1 + U(ηC ) 2 2 =. (1 ) Furthermore the condition in E. (17) leads to the following form of ηop from E. (7) ηop 1 1 + U(ηC ) 2 2 =. (1 ) ln[(1 )/ ] (18) It is also straightforward to show that this point ( ) is indeed a maximum point by investigating the second derivatives of the power. In order to calculate the efficiency for given T 2 / at the maximum power first find the value satisfying E. (17) and

4 optimal transition rates.15.1.5 * * η c and 1 asymptotes *(η c ) = *(η c ) *(η c =1).4.6.8 1 η c η op 1.8.6.4 η op 1.96.92.88.97.98.99 1 η c at (* *) η CA = 1 1 η c η c /(2 η c ) η c /2 η c 1 asymptote.4.6.8 1 η c FIG. 3. Numerically found and ɛ values satisfying E. (17) as a function of η C = 1 T 2 / along with the (η C ) = ɛ (η C ) and (η C = 1) values presented in Sec. III B 2. ɛ (η C = 1) = (the horizontal axis). The η C asymptote indicates E. (19) up to the ηc 2 term and the η C 1 asymptote indicates E. (25) up to the (1 η C ) term with the coefficients given by Es. (27) and (28). net power < C from to 1 = ( C! ) = ( C! ) '.83 222 ( C = 1) ' 17 812 ( C = 1) = FIG. 4. Illustration of the optimal transition rates ( ɛ ) for the maximum power output as the T 2 / value varies. substitute the value to E. (18). As E. (17) is a transcendental euation the closed-form solution for η op is unattainable. 2. Asymptotic behaviors obtained from series expansion The upper bound for is given by the condition η C = 1 satisfying ln[(1 )/ ] = 1/(1 ) and (η C = 1) 17 812 found numerically and ɛ (η C = 1) = exactly from E. (15b). η C = always satisfies E. (17) regardless of values so finding the optimal is meaningless (in fact when η C = the operating regime for the engine is shrunk to the line = ɛ and there cannot be any positive work). Therefore FIG. 5. The efficiency at the maximum power η op as the function of the Carnot efficiency η C in E. (18) using numerically found optimal values along with various asymptotic cases: the Curzon-Ahlborn efficiency η CA in E. (23) the upper bound η C /(2 η C ) and the lower bound η C /2 in Ref. [22] and the η C 1 asymptote for η C.8. The inset shows the region.97 < η C < 1. let us examine the case η C using the series expansion of with respect to η C as = a + a 1 η C + a 2 η 2 C + a 3η 3 C + O ( η 4 C). (19) Substituting E. (19) into E. (17) and expanding the left-hand side with respect to η C again we obtain c 1 η C + c 2 η 2 C + c 3η 3 C + O ( η 4 C) = (2) where c n describes the relation among a a n 1 each of which should be identically zero to satisfy E. (2). Letting the linear coefficient c 1 be zero yields ( ) 2 1 a = ln (21) 1 2a from which the lower bound for (η C ) = a = ɛ (η C ).83 222 found numerically [lim ηc U(η C ) = 1 2 thus ɛ (η C ) = (η C ) by E. (15b)]. Figure 3 shows the numerical solution ( ɛ) = ( ɛ ) as a function of η C where the asymptotic behaviors derived above hold when η C and η C 1. It seems that is monotonically increased and ɛ is monotonically decreased as η C is increased i.e. min = (η C ) max = (η C = 1) ɛ min = and ɛ max = ɛ (η C ). Figure 4 illustrates the situation on the ( ɛ) plane. The linear coefficient a 1 in E. (19) can be written in terms of a when we let c 2 = in E. (2) and the uadratic coefficient a 2 in E. (19) can also be written in terms of a alone by letting c 3 = in E. (2) and using the relations in Es. (21) and a 1 expressed by a terms which are well consistent with the numerical solution as shown in Fig. 3. With the relations of coefficients in hand we find the asymptotic behavior of η op in E. (18) by expanding it with respect to η C after substituting as the series expansion of a

during 1 during 2 stochastic Markov processes 5 η C in E. (19). Then η op = 1 2 η C + 1 8 η2 C + 7 24a + 24a 2 η 3 96(1 2a ) 2 C + O ( ) ηc 4. (22) With this method we are able to find the coefficients in terms of a up to an arbitrary order in principle. We would like to emphasize that the expansion form of η op in E. (22) has exactly the same coefficients up to the uadratic term to those of the Curzon-Ahlborn efficiency [3 5] defined as with the expansion form η CA = 1 T 2 / = 1 1 η C (23) hw net i s h s h = T 2 s c + hw neti s h = E 1 s c G 1 (s c ) s c η CA = 1 2 η C + 1 8 η2 C + 1 16 η3 C + 5 128 η4 C + O(η5 C ) (24) when η C. As a result numerically found η op by solving E. (17) and substituting the value to E. (18) and η CA share a very similar functional form for η C 1/2 as shown in Fig. 5. In fact the linear term η C /2 and uadratic term ηc 2 /8 are naturally from the strong coupling between the thermodynamic fluxes and the symmetry between the reservoirs (as we will check in Sec. III C the reservoir symmetry is related to the symmetry in the entropy production at the hot and cold reservoir and holds only approximately in our model) [23 24]. The third order coefficient (.77 492) in E. (22) however is clearly different from 1/16 for the η CA. In other words the deviation from η CA for η op enters from the third order that has not been theoretically investigated yet. Indeed η op deviates from η CA for η C 1/2 until they coincide at η C = 1. Therefore the efficiency η op of our model at maximum power output is different from η CA. For η C 1 we need to consider the logarithmic correction due to the functional form based on the numerical evidence also shown in Fig. 5. In contrast to the linear heat conduction for the Curzon-Ahlborn endoreversible engine [3 5] our model has an exponential or Boltzmann type of relaxation process. We believe that this different functional form of heat conduction process results in the different types of singularity at η C 1: the algebraic singularity of η CA in E. (23) at η C = 1 with the infinite slope and the logarithmic singularity in our case. We take the singular series expansion of the functional form in E. (17) near η C = 1 as = max + b ln (1 η C ) ln(1 η C ) + b 1 (1 η C ) + O [ (1 η C ) 2]. (25) It is possible to consider other types of terms such as (1 η C ) ln 2 (1 η C ) but we will check that it is enough to predict the functional form of η op consistent with an alternative approach from entropy-production-based analysis provided in Sec. III C. If we take only the zeroth order term we obtain the identity ( 1 1 ) 1 = ln max max (26) max FIG. 6. The entropy relation between s h and s c given by E. (31) and the linear relation in E. (33) representing the first law of thermodynamics. The maximum value of W net (the intercept of the linear relation on the vertical axis times ) is achieved when the line becomes the tangential one of the curve as illustrated here. exactly at η C = 1 that is already mentioned in the first part of this subsection. Similar to the η C case by letting each coefficient be zero we find the relations among the coefficients as and b ln = max(1 max) 2 (27) b 1 = max(1 max) 2 { 1 + ln[ max(1 max)] } (28) which are well consistent with the numerical solution as shown in Fig. 3. Again the asymptotic behavior of η op in E. (18) for η C 1 can be deduced from the series expansion in terms of (1 η C ) > using Es. (25) (27) and (28) which is η op =1 + (1 max)(1 η C ) ln(1 η C ) + (1 max) ln[ max(1 max)](1 η C ) + O [ (1 η C ) 2] (29) based on the relations in Es. (27) and (28) [the same procedure as the one leading to (22)]. As shown in Fig. 5 however the asymptotic form only holds in a rather limited range of η C very close to unity indicating the necessity to taking higher order terms into account for more accurate asymptotic behavior. C. The entropy production relation In Ref. [1] it is argued that the necessary and sufficient condition for the Curzon-Ahlborn efficiency at the maximum power is that the entropy productions at the hot and cold reservoirs (denoted by s h and s c respectively ) should be related by a specific functional form namely s h = F (s c ) where F (x) = x/(1 + ζ x) with the system-specific constant ζ.

6 The entropy production in our model is given by s h = Q 1 ( ɛ) E 1 s c = Q 2 T 2 ( ɛ) E 2 T 2 (3) where we again discard the common explicit time dependent term (1 e τ/2 ) 2 /(1 e τ ) which does not affect the following discussion for notational convenience. Given and T 2 and putting E 1 as a constant we obtain the entropy relation given by s h = E 1 s c G 1 (s c ) (31) where G 1 is the inverse function of G defined from the relation in E. (3) s c = G(E 2 /T 2 ). Note that s h is an increasing function with respect to s c while ds h /ds c is a decreasing one so s h (s c ) is an increasing and concave function of s c as illustrated in Fig. 6. Therefore the uniue [guaranteed by the concavity of s h (s c )] optimal entropy production denoted by s c which makes the power reach its maximum value is determined by ds h ds c sc =s c = E 1 ɛ ɛ ɛ + G 1 ( s c ) ɛ ɛ = T 2 (32) where T 2 / comes from the thermodynamic first law s h = T 2 s c + W net. (33) The parameter ɛ obtained from E. (32) is still a function of E 1 or. By optimizing the entropy production with respect to we find the same optimal and ɛ as those in the previous section. As G 1 (s c ) is not a linear function of s c we do not have the relation s h = F (s c ) mentioned before so the fact that η op η CA is consistent with Ref. [1]. However we reveal that it is possible to find the regime where the entropy production of our model approximately follows the functional form F (x) which indeed corresponds to the η op η CA regime as we show in the following section. 1. The linear regime First we take the regime where T = T 2. Then from E. (32) one can see the solution E = E 1 E2 E 1 or s c 1 which allows the small s c expansion of G 1 up to the linear order as G 1 (s c ) G 1 () + dg 1 ds c sc = where the constant ζ(e 1 / ) is given by ζ = 2 ( T1 E 1 ) 2 [ 1 + cosh s c = E 1 + ζ(e 1 / )s c (34) ( E1 )]. (35) Inserting E. (34) to the entropy relation in E. (31) we find that the entropy production for hot and cold reservoirs actually follows the functional form F (x) = x/(1 + ζx) for η C which explains the Curzon-Ahlborn-like behavior for η C. We have already checked that η op η CA as η C due to the same linear and uadratic coefficients from the series expansion in Sec. III B 2 but the entropy production relation suggests that there could be a deeper relation between our model and engines with the optimal efficiency of η CA than the reasons for the linear and uadratic coefficients namely the strong coupling between the thermodynamics fluxes and the reservoir symmetry respectively. In fact the implication of the reservoir symmetry in the expression F 1 (x) = F ( x) holds only for η C. We emphasize that the behavior of efficiency η op η CA in η C is independent of the value. However we can optimize W net with respect to as following. The optimal condition to maximize W net for a given T 2 / value is euivalent to minimize ζ because W net = ζ 1 T2 2 1/ζ (36) from the condition of tangent T 2 / = 1/(1+ζs c) 2. The condition for the minimum value of ζ for given E 1 / is by taking the derivative of the functional form in E. (35) with respect to E 1 / and using the relation /(1 ) = e E 1/ ( ) 2 1 1 2 = ln (37) which is euivalent to the condition for the zeroth order term of (η C ) given by E. (21). In other words the optimal condition at least for the lowest order of η C at η C can also be derived from the functional form of entropy production given by Es. (31) and (35) further supporting the consistency of our result. 2. The logarithmic regime The other extreme regime is T 2 or the η C 1 limit where the solution satisfying E. (32) exists in the region of E2 /T 2 1 or s c 1. In this limit we rewrite the relation between s c and E 2 /T 2 in E. (3) as E 2 s c T 2 (1 + 1 ) e E 2/T 2. (38) Using the above we obtain G 1 (s c ) up to the order of (s c / 2 )e s c/ G 1 (s c ) s c + s c 2 e s c/. (39) Inserting E. (39) to E. (31) the entropy production relation reads s h = E 1 2 + e s c/ (4)

7 E 1 R 1 R 2 1m m 2m E m m FIG. 7. The energy levels for each reservoir in the three-level heat engine along with the transition rates are illustrated. The work is extracted by adjusting the most upper level (the second excited state) between the reservoir contacts. and using the condition ds h ds c sc =s c E 2 = E 1 e sc/ ( + e s c/ ) E 1e sc/ = T 2 (41) 2 we derive the efficiency at the maximum power η op 1 T [ 1 (1 η C ) 1 + T ] 1 (1 η C ) E 1 E 1 [ ] E 1 ln. (1 η C ) (42) Optimizing the power with respect to or E 1 is euivalent to the maximizing the saturation value of E. (4) which is lim sc s h = E 1 /. This condition yields the functional form max should satisfy which is euivalent to E. (26). Using the result E 1 / = (1 max) 1 we obtain η op in the ( ɛ) space given by η op 1+ ( 1 max) (1 ηc ) [ 1 + ( 1 ) max (1 ηc ) ] ln [ ( max 1 ) max (1 ηc ) ] (43) which is consistent with the result in the previous section; in fact E. (43) includes higher order terms namely (1 η C ) 2 and (1 η C ) 2 ln(1 η C ) than E. (29). We also emphasize that this consistency justifies the series expansion form in E. (25). Based on the analysis above for the regime where T 2 we reach the conclusion that there cannot be a value making the higher order terms than the linear term in the expansion of G 1 for any given and T 2 values. In other words one again can see the systematic difference between our model and the models belonging to the Curzon-Ahlborn efficiency at the maximum power. IV. EXTENSION TO MULTI-LEVEL HEAT ENGINE Finally we would like to remark on the possible extension to multi-level systems i.e. systems with more than two levels which are more general cases. In that framework our twolevel heat engine can be taken as a simplified one considering only the ground and first excited states. For simplicity again we assume two heat reservoirs R 1 and R 2 which are characterized by the temperatures and T 2 and contacted with the system during the time τ 1 and τ 2 respectively. First let us take the three-level system where we consider the ground first excited and second excited states for each reservoir. We further simplify the situation by differentiating only the second excited states of the reservoirs namely E 1 for R 1 and E 2 for R 2 and the common value E m for the first excited state as depicted in Fig. 7. The transition rates are denoted by (the ground state to E 1 in R 1 ) m (the ground state to E m in R 1 ) 1m (E m to E 1 in R 1 ) ɛ (the ground state to E 2 in R 2 ) ɛ m (the ground state to E m in R 2 ) and ɛ 2m (E m to E 2 in R 2 ); their reverse transition rates are m 1m ɛ ɛ m and ɛ 2m respectively. Applying the Arrhenius form we obtain the relations / = e E 1/ m / m = e E m/ 1m / 1m = e (E 1 E m )/ ɛ/ ɛ = e E 2/T 2 ɛ m / ɛ m = e E m/t 2 ɛ 2m / ɛ 2m = e (E 2 E m )/T 2. (44) As in the two-level case the net amount of work from the population difference in different energy levels (only for the second excited states in this case) is W net = (P 1e P 2e )(E 1 E 2 ) where P 1e and P 2e refer to the population of E 1 in R 1 and E 2 in R 2 respectively. The heat exchange on the other hand should take the E m level into account. As a result the efficiency for the three-level system is given by η = (P 1e P 2e )(E 1 E 2 ) (P 1e P 2e )E 1 + (P 1m P 2m )E m (45) where the term involving E m unless it vanishes represents the extra heat exchange that cannot be used in the work extraction. In contrast to the two-level case the temporal part (involving τ 1 and τ 2 ) is not factorized in the functional form of W net /τ = (P 1e P 2e )(E 1 E 2 )/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 output W net /τ varies over τ and the maximum value of power output occurs at different values of ( ɛ ) depending on τ. Therefore we conclude that the two-level system of our main interest is a special case that we can analyze deeply to obtain the insight presented so far. Moreover for the three-level system even at the limit ɛ (which corresponds to the euilibrium or reversible limit for the two-level system represented by the euilibrium distribution of population) there cannot be the euilibrium condition given by e E m/ /Z = e E m/t 2 /Z e E 1/ /Z = e E 2/T 2 /Z (46) with the partition function Z = 1 + e E m/ + e E 1/ = 1 + e E m/t 2 + e E 2/T 2 unless E m =. Hence the condition of strong coupling between thermodynamic fluxes is also violated which results in the linear coefficient of the efficiency

8 (a) = 1. T 2 =.1 E m =.5 τ = 1..3 (b) = 1. T 2 =.1 E m =.5 τ = 1..12.4.3.1.25.2.15.1.5.4.3.1.1.8.6.4.2.1.3.4.1.3.4 FIG. 8. The average power output W net /τ for the three-level system with the parameters = 1 T 2 =.1 E m =.5 and (a) τ = 1 [the maximum value of W net /τ occurs at = 15(5) and ɛ =.5(5)] and (b) τ = 1. [ =.185(5) and ɛ =.1(5)]. For better visibility focused on the W net /τ regime we set all of the negative values as. We take the normalization convention + = m + m = 1m + 1m = ɛ + ɛ = ɛ m + ɛ = ɛ 2m + ɛ 2m = 1. η op at the maximum power in terms of η C different from 1/2 [23] when E m > and we have numerically verified the fact as well where we have obtained η op for the parameters and ɛ for given E m and τ values. In contrast recall that for the two-level system the strong coupling between thermodynamics always holds and the reservoir symmetry approximately holds for η C (corresponding to ɛ for the maximum power output). The necessity for the extra heat in E. (45) also prevents the η op from reaching unity at the other limiting case η C 1 which we have also verified numerically. V. CONCLUSIONS AND DISCUSSION We have demonstrated that our simple two-level heat engine model has a nontrivial parameter relation for the efficiency at the maximum power output. Thanks to the simplicity of our model composed of the two-level system the timedependent term only plays the role of an overall factor so we have focused on the relative transition rates for a given temperature ratio of reservoirs. Based on numerical solutions and analytically driven asymptotic behaviors we have shown that the optimal efficiency η op for maximum power output in our model is clearly different from the Curzon-Ahlborn efficiency η CA [3 5] although they share the same asymptotic behavior up to the uadratic term when η C [23 24]. We have discussed its implication in conjunction with the relation of the entropy production at the hot and cold reservoirs. We have focused on the average thermodynamic uantities to yield the macroscopic efficiency in this paper but it would be possible to consider the stochastic efficiency [13 18] by adopting more specific protocols involved in the heat and work transfer which can be a future work along with the uantum effects [25 33]. For comprehensive understanding we would also need the full consideration of multi-level systems sketched in Sec. IV here which we leave as future work. Note added. Just before the submission of this paper we have learned that a recent contribution by Toral et al. [34] independently reports the same form of η op in Ising spin systems or exclusion processes. ACKNOWLEDGMENTS We thank Hyun-Myung Chun Jae Dong Noh Hee Joon Jeon and Sang Wook Kim for fruitful discussions and comments. [1] K. Huang Statistical Mechanics (John Wiley & Sons New York 1963). [2] S. Carnot Réflexions Sur La Puissance Motrice Du Feu Et Sur Les Machines Propres À Développer Cette Puissance (Bachelier Libraire Paris 1824). [3] P. Chambadal Les Centrales Nuclaires (Armand Colin Paris 1957). [4] I. I. Novikov Efficiency of an atomic power generating installation At. Energy (N.Y.) 3 1269 (1957); The efficiency of atomic power stations J. Nucl. Energy 7 125 (1958). [5] F. L. Curzon and B. Ahlborn Efficiency of a Carnot engine at maximum power output Am. J. Phys. 43 22 (1975). [6] J. Hoppenau M. Niemann and A. Engel Carnot process with a single particle Phys. Rev. E 87 62127 (213). [7] K. Proesmans C. Driesen B. Cleuren and C. Van den Broeck Efficiency of single-particle engines Phys. Rev. E 92 3215 (215). [8] J. Um H. Hinrichsen C. Kwon and H. Park Total cot of operating an information engine New J. Phys. 17 851 (215). [9] V. Holubec and A. Ryabov Efficiency at and near maximum

9 power of low-dissipation heat engines Phys. Rev. E 92 52125 (215). [1] J.-M. Park H.-M. Chun and J. D. Noh Efficiency at maximum power and efficiency fluctuations in a linear Brownian heat engine model Phys. Rev. E 94 12127 (216). [11] A. Ryabov and V. Holubec Maximum efficiency of steady-state heat engines at arbitrary power Phys. Rev. E 93 511(R) (216). [12] N. Shiraishi K. Saito and H. Tasaki Universal trade-off relation between power and efficiency for heat engines Phys. Rev. Lett. 117 1961 (216). [13] G. Verley M. Esposito T. Willaert and C. Van den Broeck The unlikely Carnot efficiency Nat. Commun. 5 4721 (214). [14] G. Verley T. Willaert C. Van den Broeck and M. Esposito Universal theory of efficiency fluctuations Phys. Rev. E 9 52145 (214). [15] T. R. Gingrich G. M. Rotskoff S. Vaikuntanathan and P. L. Geissler Efficiency and large deviations in time-asymmetric stochastic heat engines New J. Phys. 16 123 (214). [16] K. Proesmans B. Cleuren and C. Van den Broeck Stochastic efficiency for effusion as a thermal engine Europhys. Lett. 19 24 (215). [17] K. Proesmans and C. Van den Broeck Stochastic efficiency: five case studies New J. Phys. 17 654 (215). [18] M. Polettini G. Verley and M. Esposito Efficiency statistics at all times: Carnot limit at finite power Phys. Rev. Lett. 114 561 (215). [19] J. Schnakenberg Network theory of microscopic and macroscopic behavior of master euation systems Rev. Mod. Phys. 48 571 (1976). [2] D. Andrieux and P. Gaspard Fluctuation theorem and Onsager reciprocity relations J. Chem. Phys. 121 6167 (24). [21] U. Seifert Entropy production along a stochastic trajectory and an integral fluctuation theorem Phys. Rev. Lett. 95 462 (25). [22] M. Esposito R. Kawai K. Lindenberg and C. Van den Broeck Efficiency at maximum power of low-dissipation Carnot engines Phys. Rev. Lett. 15 1563 (21). [23] C. Van den Broeck Thermodynamic efficiency at maximum power Phys. Rev. Lett. 95 1962 (25). [24] M. Esposito K. Lindenberg and C. Van den Broeck Universality of efficiency at maximum power Phys. Rev. Lett. 12 1362 (29). [25] H. E. D. Scovil and E. O. Schulz-DuBois Three-level masers as heat engines Phys. Rev. Lett. 2 262 (1959). [26] J. E. Geusic E. O. Schulz-DuBois and H. E. D. Scovil Quantum euivalent of the carnot cycle Phys. Rev. 156 343 (1967). [27] C. M. Bender D. C. Brody and B. K. Meister Quantum mechanical Carnot engine J. Phys. A 33 4427 (2). [28] S. Abe and S. Okuyama Similarity between uantum mechanics and thermodynamics: Entropy temperature and Carnot cycle Phys. Rev. E 83 21121 (211). [29] S. Abe Maximum-power uantum-mechanical Carnot engine Phys. Rev. E 83 41117 (211). [3] J. Wang J. He and X. He Performance analysis of a two-state uantum heat engine working with a single-mode radiation field in a cavity Phys. Rev. E 84 41127 (211). [31] R. Uzdin Am. Levy and R. Kosloff Euivalence of uantum heat machines and uantum-thermodynamic signatures Phys. Rev. X 5 3144 (215). [32] H. J. Jeon and S. W. Kim Optimal work of the uantum Szilard engine under isothermal processes with inevitable irreversibility New J. Phys. 18 432 (216). [33] S. Liu and C. Ou Maximum power output of uantum heat engine with energy bath Entropy 18 25 (216). [34] R. Toral C. Van den Broeck D. Escaff and K. Lindenberg Stochastic thermodynamics for Ising chain and symmetric exclusion process e-print arxiv:1611.8188.