Download as pdf or txt
Download as pdf or txt
You are on page 1of 7

Carnot efficiency is attainable in an irreversible process

Jae Sung Lee and Hyunggyu Park


School of Physics and Quantum Universe Center,
Korea Institute for Advanced Study, Seoul 02455, Korea
(Dated: November 24, 2016)
In thermodynamics, there exists a conventional belief that the Carnot efficiency is reachable
arXiv:1611.07665v1 [cond-mat.stat-mech] 23 Nov 2016

only when a process is reversible. However, there is no theorem proving that the Carnot efficiency
is impossible in an irreversible process. Here, we show that the Carnot efficiency is attainable in
an irreversible process through investigation of the Feynman-Smoluchowski ratchet (FSR). Thus,
this finding gives us a new possibility to develop a novel design of thermodynamic engines with
high efficiency regardless of the reversibility. Our result also answers the long-standing question of
whether the Carnot efficiency is possible in the FSR.

PACS numbers: 05.70.-a, 05.40.-a, 05.70.Ln

Thermodynamics is a field of science dealing with the T1 T2


relationship between energy, work, and heat [1]. It was pawl
practically initiated to develop a heat engine with high x0
efficiency. Here, the heat engine is a device transform-
ing heat energy into useful mechanical work. Therefore, F
one of the most interesting subjects in thermodynamics h y
is the study of the maximum possible efficiency attain- forward
able by a heat engine. The maximum efficiency of a heat vanes
backward x
engine operating in two thermal baths of different tem-
peratures T1 and T2 (T1 > T2 ) is fairly well understood; FIG. 1. Schematic of the FSR model. One-dimensional vanes
the efficiency cannot be greater than the Carnot efficiency (the pawl) move only horizontally (vertically), and are in con-
C = 1 T2 /T1 [2]. tact with thermal baths T1 (T2 ), respectively. x is the position
The efficiency reaches C when the process of the heat of one vane, y is the height from the bottom of the vanes to
engine is perfectly reversible [2]. Formally, defining Q1 the tip of the pawl, F is a constant external force, x0 is the
and Q2 as the average heat transferred from thermal distance between neighboring vanes, h is height of a vane, and
is angle of the pawl. The pawl is pulled down by a spring.
baths at temperatures T1 and T2 during one engine cy-
cle over a time duration cyc , respectively, then, the ef-
ficiency and the entropy production per cycle S are
i.e., S & O(1). In this case, the efficiency will also
defined as
approach C , which means that the Carnot efficiency is
|Q2 | |Q1 | |Q2 | reachable in an irreversible process. However, no such
1 , S + . (1)
|Q1 | T1 T2 example has yet been discovered, so it has been com-
monly misunderstood that C is only attainable in the
The 2nd law of thermodynamics guarantees that S 0,
reversible limit.
with the equality satisfied only for a reversible process.
In this work, we study the efficiency of the Feynman-
It is easy to see that = C for a reversible process, and
Smoluchowski ratchet (FSR) [4, 5] and find that there
it has been widely accepted that the Carnot efficiency is
indeed exists a limiting process satisfying S/|Q1 | 0
impossible for an irreversible process, since C > when
with non-vanishing entropy production. There are vari-
S > 0.
ous configurations of the FSR setup and here we use the
However, such exact reversible dynamics do not exist in
FSR model introduced by Sekimoto [6], which is illus-
the real world. Therefore, the attainability of the Carnot
trated in Supplemental Material (SM) Fig. 1. This FSR
efficiency should be decided through a limiting process.
can be modeled as shown in Fig. 1. The one-dimensional
It is convenient to rewrite Eq. (1) as
vanes move only horizontally and the pawl moves only
T2 S vertically. They are in contact with thermal baths T1
C = . (2) and T2 , respectively, and the ratcheting interaction oc-
|Q1 |
curs outside of the baths. x is the position of one vane
Approaching a reversible process means the limit S and y is the height from the bottom of vanes to the tip
0, which can be realized in a quasi-static process [3]. In of the pawl. x and y are stochastic variables due to ther-
this limit, the Carnot efficiency is reachable with zero en- mal noise. Since the pawl cannot penetrate the bottom,
tropy production. We note that there is another limiting y 0. The vanes and the pawl are pulled by the con-
process S/|Q1 | 0 with positive entropy production, stant external force F and the harmonic force ky, re-
2

(a) pawl-open (b) pawl-closed In the pawl-closed state (y h), the pawl completely
pawl forbids a backward hop of the vanes at x = nx0 . This
blockage is felt by the vanes as an infinite potential bar-
pawl
rier located at x = nx0 , as illustrated in Fig. 2(d). Note
that no energy is transferred to the pawl by this block-
ing collision because a horizontal force does not induce
h y h y any (vertical) y displacement. For nx0 < x < nx0 + a
(a = x0 h), there is no collision between the vanes and
vanes vanes the pawl, i.e., gv (x, y) = gp (x, y) = 0. Therefore, the
vanes feels only a linear potential of slope F in this re-
(c) (d) no-collision collision gion. For nx0 + a x < (n + 1)x0 , a collision can occur,
forward forward so an extra interaction potential will be felt by the vanes.
potential

potential
U0 In this collision region, some energy is transferred from
backward backward
the vanes to the pawl by the collision, which is eventually
Fx0 Fx0 dissipated as heat Qcol into the heat bath 2. For vanes
going over (n+1)x0 , the pawl will be lifted up to y = h by
nx0 (n+1)x0 nx0 nx0+a (n+1)x0 a vane. Then, the vanes will overcome an energy barrier
x x
of height U0 +F x0 with U0 = kh2 /2. In this one-step for-
FIG. 2. Schematics of the pawl-open and pawl-closed states. ward hopping process, the energy delivered to the pawl
(a) Pawl-open state (y > h). (b) Pawl-closed state (y h). from the vanes is U0 , which is dissipated as heat Qhop in
(c) Potential in the pawl-open state. Only a linear potential the heat bath 2, i.e., Qhop = U0 [7]. For later discussion
with slope F is felt by the vanes. (d) Potential in the pawl- we define the average time for one forward hop as hop .
closed state when nx0 x < (n + 1)x0 (n is an integer). We now consider the following limit:
An infinite potential wall at nx0 prevents a backward hop.
There are two regions: the no-collision (nx0 < x < nx0 + a) T2 < T1 U0 , F x0 . (5)
and collision regions (nx0 + a x < (n + 1)x0 ), depending
on whether a collision between the vanes and the pawl takes In the large U0 limit (large k or h), the FSR will almost
place. U0 is the potential energy of the pawl at y = h.
always be in the pawl-closed state due to huge energy bar-
riers. Along with very large F x0 , the vanes will rarely
spectively. Here, the direction against F is defined as reach the collision region against a very steep energy hill.
forward. x0 is the distance between neighboring vanes, Therefore, in the above limit, the vanes will spend most of
h is the height of a vane, and is the angle of the pawl. their time in the no-collision region (nx0 < x < nx0 + a).
Then, the corresponding Langevin equation can be writ- Then, Eq. (3), the dynamics of vanes, can be approxi-
ten as mately written as

vane:v = x,
mv = F gv (x, y) 1 v + 1 , (3) vane: v = x,
mv = F 1 v + 1 (x nx0 ), (6)
pawl:u = y,
mp u = gp (x, y) ky 2 u + 2 (y 0),(4)
with an infinite energy barrier at x = nx0 . Similarly,
where m and mp are the masses of the vanes and the pawl Eq. (4) can be practically written as
respectively, i is the damping coefficient of heat bath i,
and i (t) is the Gaussian noise of heat bath i at time t pawl: u = y,
mp u = ky 2 u + 2 (y 0), (7)
satisfying hi (t)j (t )i = 2i Ti ij (t t ) (the Boltzmann with an infinite energy barrier at y = 0. This implies that
constant is set to kB = 1). gv (x, y) and gp (x, y) denote the steady-state probability distributions of the vanes
the forces applied to the vanes and the pawl, respectively, and the pawl are almost the same as the equilibrium
which are generated when a vane and the pawl collide or distributions of the Langevin equations 6 and 7, respec-
interact. The detailed forms of the forces are given in tively, in the limit of Eq. (5). Hence, the probabilities for
SM Fig. 2. the pawl-open and pawl-closed states, po and pc , become
We define two states in this model: the pawl-open and
Z r r
pawl-closed states as shown in Fig. 2(a) and (b), respec- 2k ky 2
T2 UT0
po dy e 2T2
e 2 ,
tively. In the pawl-open state (y > h), both forward and h T 2 U 0
backward hopping movements of the vanes are possible. pc = 1 po 1, (8)
Here, one hop denotes movement of x from nx0 < x <
(n + 1)x0 (n is an integer) to n x0 < x < (n + 1)x0 respectively. This will be demonstrated numerically later
(n = n 1). Since gv (x, y) = 0 in this state, only a lin- for the case of small p0 . Note that all higher-order cor-
ear potential with slope F is felt by the vanes, as shown rections are exponentially small in U0 .
in Fig. 2(c). Furthermore, no energy is transferred from In this limit, we can estimate the power hW is and heat
the vanes to the pawl in this state.
dissipation rate into the heat bath 2 hQ2 is , where h is
3

denotes the steady-state average. These can be written Using Eqs. (9), (10), and (12), we calculate the effi-
as ciency and entropy production in the limit of Eq. (5)
and mp /m 1. First, the efficiency is given by
hW is = (rf rb )F x0 ,
hW is (1 e )F x0
hQ 2 is = hQ col is + hQ hop is , (9) (, ) = ,(13)
is + hQ 2 is
hW (1 e )F x0 + U0
where rf and rb are the rates of forward and backward
hopping, respectively. The power hW is is the work rate where rb /rf e with
of lifting the load hanging from the axle of the vanes. The r
U0 U0 + F x0 No T2
heat dissipation can be separated into two terms based = and = . (14)
on collisions and hopping, as discussed before. T2 T1 Nc U0
First, consider the rates in the pawl-closed state. Since To be a useful heat engine (positive work extraction
the dynamics are almost identical to the equilibrium against the load), 1 e should be larger than zero.
state, the rate of forward hopping that overcomes an en- Moreover, since F x0 > 0, we have the condition for as
ergy barrier of height U0 + F x0 can be estimated from
the Arrhenius rate equation as rf,c Nc pc e(F x0 +U0 )/T1 ln < < m , (15)
where Nc is the hopping-attempt frequency [8, 9]. The
backward hopping rate rb,c is simply zero in this case. where
In the pawl-open state, both forward and backward hops  
1 1
are possible, but the forward hopping rate rf,o is exponen- m = U0 . (16)
T2 T1
tially smaller ( eF x0 /kB T1 ) than the backward hopping
rate rb,o for large F . So, the backward hopping rate will For fixed U0 , T1 , and T2 , we find the maximum effi-
be almost identical to the hopping-attempt frequency, ciency by varying F x0 (or ) in the range of Eq. (15):
i.e., rb,o No po with exponentially small corrections. d()/d|= = 0. The result is
Note that No and Nc can differ, but this difference will    
not be very large. The vanes spend most of time in the 1
1 ln m

ln m + ln 1 ln m + O ,
pawl-closed state and become fully relaxed. As the pawl 2 m 2 m
opens for a very short period ( hop po 1/U0 ), we (17)
expect that the vane statistics does not deviate signifi- which is well inside of the range of Eq. (15). Plugging
cantly from the full relaxed one. Therefore, No /Nc can Eq. (17) into Eq. (13), it is easy to see
be reasonably assumed to be a constant of O(1). Then,
we have T22 ln U0
(
, ) C , (18)
U0 +F x0
2T1 U0

rf = rf,o + rf,c rf,c Nc e T1
,
r where we used rb /rf |= = e 1/U0 . This clearly
T2 UT0 shows that the Carnot efficiency C can be reached in the
rb = rb,o + rb,c = rb,o No e 2. (10)
U0 limit of Eq. (5) with F x0 U0 (T1 /T2 1) (T1 /2) ln U0
U0 1

Now consider hQ 2 is . Since Qhop = U0 , hQ hop is = 21 ln m ) and rf Nc e T2 + 2 ln U0 .


( =
rf,c U0 . Qint originates from the energy transfer due to At = ,
the entropy production during one hopping
the collision between a vane and the pawl in the collision period S becomes
region of the pawl-closed state. In our elastic collision !
model (SM Fig. 2), one can easily show that the trans- 1 hQ 2 is hQ 1 is 1
S = hop hSis ln U0 , (19)
ferred energy per collision is proportional to mp /m for rf T2 T1 2
mp /m 1. The collision frequency cannot be calcu-
lated simply in our setup, nevertheless we may expect where hop ( 1/rf ) is the average time for one forward
s is the steady state entropy production rate, and
hop, hSi
hQ col is = F (mp /m), (11) hQ1 is = hQ 2 is + hW
is is the average heat transfer rate
from heat bath 1. Since Eq. (19) can be very large, the
where F (z) is a vanishing function as z goes to zero. FSR operates definitely in an irreversible process, while
Equation (11) will be checked numerically later. Then, retaining the Carnot efficiency. In terms of Eq. (2), both
we can make hQ col is arbitrarily small compared to S and |Q1 | during one hopping period diverge, but in
hQ hop is , i.e. hQ col is hQ hop is , by taking an appro- a different manner to S ln U0 and |Q1 | U0 , thus
priately small value of mp /m. Therefore, in this limit, its ratio approaches zero in the U0 limit (see also
we get Eq. (18)), which is in contrast to the conventional re-
versible Carnot engine. We note that hW is rf U0 and
hQ 2 is hQ hop is rf U0 , (12)
hSis rf ln U0 are vanishingly small due to exponentially
4

101 h = 10. Since U0 = kh2 /2 = 100 is much larger than


100
the thermal energy, no forward or backward hops can
occur within our simulation time ( = 7.5 107 ), so x
P(x)

10-1
10-2 remains between 0 and x0 when n = 0. Figures 3(a)
10-3
(a) F=2 (b) F=20 and (c) show the probability distributions of x and y for
0 1 2 3 0 0.2 0.4 0.6 0.8 1 F = 2, respectively. They show clear deviations from
x x the equilibrium distributions (solid lines), as expected
100 for small F . However, for F = 20, we can see perfect
agreement in Fig. 3(b) and (d). We also checked the va-
10-1
lidity of Eq. (11). To facilitate more collisions, a lighter
P(y)

10-2 load (small F ) was used, by setting F = 1, with x0 = 2,


10-3 (c) F=2 (d) F=20 k = 100 and h = 1. Since U0 = 50 is still too large,
0 1 2 3 0 1 2 3
no hopping occurs ( = 2 109 ) and the heat dissipa-
y y tion into heat reservoir 2 is solely from energy transfer
10-1 through collisions: hQ 2 is = hQ col is . Figure 3(e) shows
(e) the log-log plot for hQ col is versus mp /m, which shows
a power-law scaling of hQ col is with an exponent 0.27,
<Qcol>s

which confirms Eq. (11). It is infeasible to measure the


efficiency numerically for large U0 because hop grows ex-
.

10-2 ponentially with U0 . Instead, we obtained efficiency data


at U0 = 5, which are presented in SM Fig. 3. This result
also supports our theory.
10-3 10-2 10-1 100 In conclusion, we have described a heat engine that
mp/m can operate with the Carnot efficiency in an irreversible
process. The key observation is that the entropy pro-
FIG. 3. Numerical results. (A) and (B ) show the probability duction divergence is weaker than the divergence of the
distributions of x for F = 2 and F = 20, respectively. (C ) heat currents to attain the Carnot efficiency. Therefore,
and (D) show the probability distributions of y for F = 2 this finding gives us a new possibility to develop a novel
and F = 20, respectively. Solid curves denote equilibrium
distributions of Eqs. 6 and 7, respectively. (E ) Log-log plot
design of thermodynamic engines, especially for micro-
of hQ col is versus mp /m. The dashed line is a guide line with scopic ones actively studied recently [1416], with high
slope 0.27. efficiency regardless of the reversibility. Note that our
conclusion is different from the result of Ref. [17] which
shows that the efficiency can be close to Carnot efficiency
at finite entropy production rate for long but finite time.
diverging hopping period hop , which is similar to an or-
However, our results are consistent with the recent rig-
dinary Carnot engine operating in a quasi-static way.
orous bound claiming that power should go to zero when
We performed a numerical simulation to check the the efficiency approaches the Carnot efficiency [18].
validity of our theory. In this simulation, we numeri- Our result also answers the long-standing debate on
cally integrated the Langevin Eqs. (3) and (4), using a the attainability of the Carnot efficiency in the FSR.
second-order integrator [10]. To implement the interac- Feynman first claimed that the efficiency of the FSR
tion forces gv (x, y) and gp (x, y), we assumed that colli- can reach C in a reversible process in the 1960s [5].
sions between a vane and the pawl were elastic and in- However, in 1996, Parrodo and Espa nol [19] criticized
stantaneous (see SM Fig. 2). For convenience, we used Feynmans argument, stating that the reversible pro-
dimensionless variables byprescaling time, length, and en- cess was in fact not reversible, and that irreversible heat
ergy in units of 0 /k0 , T2 /k0 , and T2 , respectively. flows must exist in the process [20]. They also claimed
Here 0 and k0 are constants with dimensions of the that, when a system is connected to two different ther-
damping coefficient and the spring R constant, respectively. mal baths simultaneously, the dynamics should be al-
Heat can be calculated as Q1 = 0 dt v (1 v + 1 ) and ways irreversible, thus the Carnot efficiency is unattain-
R
Q2 = 0 dt u(2u+2 ) during , where denotes the able. Ever since, the FSR efficiency has been discussed
Stratonovich integral [6, 1113]. Then, hQ i is Qi / in in a great number of studies [2123], and the claim made
a steady state. For convenience, we take T1 = 2, T2 = 1, by Parrodo and Espa nol is generally accepted [2426].
m = 10, 1 = 2 = 1, and = 45 . However, our conclusion clearly disproves this claim and
We first check whether the vanes and the pawl are al- shows that the FSR can attain the Carnot efficiency even
most always in equilibrium without the collision forces though the dynamics are irreversible.
described by Eqs. (6) and (7), respectively, in the limit This research was supported by the NRF grant No.
of Eq. (5). We set mp = 0.1, x0 = 11, k = 2, and 2011-35B-C00014 (JSL).
5

Rold
[14] I. A. Martnez, E. an, L. Dinis, D. Petrov, J. M. R.
Parrondo, and R.A. Rica, Nature Phys. 12, 67 (2016).
[15] V. Blickle and C. Bechinger, Nature. Phys. 8, 143 (2012).
[1] E. Fermi, Thermodynamics, Dover edition (Courier Cor- [16] J. V. Koski, V. F. Maisi, J. P. Pekola, and D. V. Averin,
poration, 1959). Proc. Natl. Acad. Sci. 111, 13786 (2014).
[2] C. Kittel and H. Kroemer, Thermal physics, 2nd Ed. (W. [17] M. Polettini, G. Verley, and M. Esposito, Phys. Rev.
H. Freeman and Company, 1980), Ch. 8. Lett. 114, 050601 (2015).
[3] F. L. Curzon and B. Ahlborn, Am. J. Phys. 43, 22(1975). [18] N. Shiraishi, K. Saito, and H. Tasaki, preprint
[4] M. von Smoluchowski, Phys. Zeitschr. 13, 1069 (1912) . arXiv:1605.00356v1 (2016).
[5] R. P. Feynman, The Feynman Lectures on Physics, Vol. [19] J. M. R. Parrondo and P. Espa nol, Am. J. Phys. 64, 1125
1. (Massachusetts, USA: Addison-Wesley) (1963), Ch. 46. (1996) .
[6] K. Sekimoto, Prog. Theor. Phys. 130, 17 (1998) . [20] This was also argued independently and demonstrated
[7] In principle, Qhop can be larger than U0 . However, in numerically in K. Sekimoto, Phys. Soc. Jpn. 66, 1234
the U0 limit, the probability for Qhop > U0 is (1997).
exponentially small. Thus, it is sufficient to set Qhop = U0 [21] M. O. Magnasco and G. Stolovitzky, J. Stat. Phys. 93,
in the zeroth-order calculation. 615 (1998).
[8] P. Hanggi, P. Talkner, and M. Borkovec, Rev. Mod. Phys. [22] P. Reimann, Phys. Rep. 361, 57 (2002) and references
62, 251 (1990). therein.
[9] S. A. Arrhenius, Z. Phys. Chem. 4,226 (1889). [23] P. Hanggi P and F. Marchesoni, Rev. Mod. Phys. 81, 387
[10] E. Vanden-Eijnden and G. Ciccotti, Chem. Phys. Lett. (2009) and references therein.
429, 310 (2006). [24] C. Van den Broeck, Carnot efficiency revisited. Adv.
[11] T. Hatano and S. -I. Sasa, Phys. Rev. Lett. 86, 3463 Chem. Phys. 135, 189 (2007).
(2001). [25] I. Derenyi and R. D. Astumian, Phys. Rev. E 59, R6219
[12] J. D. Noh and J. -M. Park, Phys. Rev. Lett. 108, 240603 (1999).
(2012). [26] T. Hondou, K. Sekimoto, Phys. Rev. E 62, 6021 (2000).
[13] J. S. Lee, C. Kwon, and H. Park, Phys. Rev. E 87,
020104(R) (2013).
6

T1 vanes T2
y
x

pawl
F

-sU(y) y

x pawl

vanes
cross-sectional image

FIG. 4. (Supplemental Material Fig.1) Schematic of the FSR. The FSR consists of two components: vanes and a pawl.
Both are in contact with different thermal baths of temperatures T1 and T2 , and their ratcheting interaction occurs outside of
the baths as shown in the boxed area. Ratcheting is achieved by interaction between the symmetric vanes and an angled pawl
in our representation, while it takes place between a ratchet wheel with angled teeth and a simple pawl in the original FSR [5].
However, both provide essentially the same rectifying function. Since only rotational motion is allowed, the dynamics of the
vanes and the pawl can be described by their angles x and y, respectively, which are stochastic variables due to thermal noise.
A restoring force U pulls down the pawl and a constant load F hangs on the axle of the vanes.
7

(a)
gp G

-gv
gv

-G -gp

(b)

-G=-gv G=gv

FIG. 5. (Supplemental Material Fig.2) Schematic of the elastic collision model. (a) Collision occurred when vanes move
in the forward direction. When a vane and the pawl collide together, action and reaction forces denoted by G and G acting
on the pawl and a vane respectively, which are orthogonal to the inclined plane of the pawl, are produced. gv (= G cos ) and
gp (= G sin ) are its horizontal and vertical components, respectively. Since only vertical (horizontal) motion is allowed for
the pawl (vanes) by fixed boundary conditions, the horizontal (vertical) component of G should be canceled out by boundary
forces and their motions are affected only by gv (gp ) as described in Eq. (4) (Eq. (5)) in the main text. Collision is almost
instantaneous, so the velocity change during collision are governed by action-reaction forces only: mv gv and mp u gp .
With = 45 , we get mv + mp u 0, which looks like a momentum-conservation law in one dimensional motion of two particles
with mass m and mp . Assuming the elastic collision, we have the energy conservation equation as dt d
[mv 2 /2 + mp u2 /2] = 0.
Using these two equations, we can calculate velocities after collision without any detailed knowledge of action-reaction forces.
(b) Collision occurred when vanes move in the backward direction. In this case, only horizontal forces are produced, which
does not affect the motion of the pawl, i.e., gp = 0. To the vane, gv produces an infinite potential barrier.

0.3

0.2
0.01
0.1
(F)

0.1
mp/m=1
0

-0.1

0 0.2 0.4 0.6 0.8 1


-1 -1
Fx0U0 (T1/T2-1)

FIG. 6. (Supplemental Material Fig.3) Plot of the efficiency as a function of F . The horizontal axis is rescaled as
F x0 U01 (T1 /T2 1)1 = 1 /m . As U0 = 5 is rather small, many forward hops occur as well as a lot of collision events.
At finite mp /m, hQ col is is not negligible and affects the efficiency negatively. The solid line is drawn by Eq. (14) in the main
text at the value of U0 = 5, which should be correct only in the U0 and mp /m 0 limits. Thus, this line is not the
asymptotic line as mp /m approaches zero, but can be used as an upper-bound guide line. As expected, the efficiency increases
and seems to approach the vicinity of the guide line as mp /m decreases. This strongly supports that our theory works well.
One data point is obtained from averaging efficiencies in 10 steady states and we used = 2 1011 .

You might also like