Vous êtes sur la page 1sur 18

PHYSICAL REVIEW E 91, 042140 (2015)

Stochastic approach to equilibrium and nonequilibrium thermodynamics

Tânia Tomé and Mário J. de Oliveira


Instituto de Fı́sica, Universidade de São Paulo, Caixa Postal 66318 05314-970 São Paulo, São Paulo, Brazil
(Received 5 January 2015; published 29 April 2015)
We develop the stochastic approach to thermodynamics based on stochastic dynamics, which can be discrete
(master equation) and continuous (Fokker-Planck equation), and on two assumptions concerning entropy. The rst
is the denition of entropy itself and the second the denition of entropy production rate, which is non-negative
and vanishes in thermodynamic equilibrium. Based on these assumptions, we study interacting systems with
many degrees of freedom in equilibrium or out of thermodynamic equilibrium and how the macroscopic laws are
derived from the stochastic dynamics. These studies include the quasiequilibrium processes; the convexity of the
equilibrium surface; the monotonic time behavior of thermodynamic potentials, including entropy; the bilinear
form of the entropy production rate; the Onsager coefcients and reciprocal relations; and the nonequilibrium
steady states of chemical reactions.

DOI: 10.1103/PhysRevE.91.042140 PACS number(s): 05.70.Ln, 05.10.Gg

I. INTRODUCTION of the equipartition of energy or may provide the macroscopic


properties directly [11].
The kinetic theory, introduced and developed in the second
The reasoning and examples given above lead us to presume
half of the 19th century by Clausius [1], Maxwell [2], and
that the macroscopic properties are obtained from microscopic
Boltzmann [3,4], aimed to derive the macroscopic properties
mechanics in two major steps: (1) from the underlying mechan-
of matter, which include the laws of thermodynamics, from the
ics to a probabilistic or stochastic approach and (2) from this
underlying microscopic movement, governed by the laws of
mechanics. In principle, this task can be achieved if we assume approach to the macroscopic properties. This is particularly
that the macroscopic laws are connected to the microscopic clear in the case of equilibrium thermodynamic properties,
laws. We cannot know a priori whether the connection exists which are derived from the Gibbs probability distribution,
or not. But, considering that it is an experimental fact that which in turn comes from the underlying mechanics, a step
the laws of mechanics, classical or quantum, are obeyed at not yet fully demonstrated and known as ergodic hypothesis.
the microscopic level by the same particles that constitute The rst step will not concern us here. The second step,
a macroscopic body, which obeys macroscopic laws, it is which is the purpose of the present paper, aims to derive
reasonable to assume that the connection exists. Once we the macroscopic properties, which include equilibrium and
assume this connection, the next task is to perform the actual nonequilibrium thermodynamic properties, from a stochastic
derivation of macroscopic laws from the microscopic laws of approach.
mechanics. This task was in fact undertaken by the founders The stochastic approach to equilibrium and nonequilibrium
of the kinetic theory and many macroscopic laws were in fact thermodynamics or, in short, stochastic thermodynamics,
derived. This includes the theorem of equipartition of energy, which is the second step of our scheme and the subject of
the Maxwell distribution of velocities [2], the Boltzmann the present paper, has been adopted by several authors and
H-theorem [3], and the Gibbs probability distribution [5]. become a consistent theory of nonequilibrium thermodynam-
However, many results cannot be said to have been derived ics [12–49]. An important step in this direction occurred
from pure mechanics alone [6,7]. A new ingredient was when Schnakenberg [17] introduced the stochastic denition
introduced in the course of derivation, namely the stochastic of entropy production rate which has a fundamental role in
behavior, in most cases in an implicit form. our approach in addition to the probabilistic denition of
The derivation from pure mechanics of the results just entropy itself, introduced by Boltzmann [4] and generalized
mentioned would be accomplished if we could show that the by Gibbs [5].
new ingredient, the stochastic behavior, is a consequence of Our approach here is based on the adoption of a Markovian
the microscopic mechanical motion, which is deterministic. stochastic evolution on a discrete or on a continuous space
At rst sight the random behavior seems to be in contradiction and on two assumptions concerning entropy. The rst is the
with a deterministic motion. However, the results coming denition of entropy itself and the second the denition of
from the theory of deterministic chaos [8] has proven that a entropy production rate. Based on these assumptions we will
deterministic motion can behave stochastically. In fact, the consider systems in equilibrium and out of thermodynamic
possibility of mapping chaotic dynamics into a stochastic equilibrium and how the macroscopic nonequilibrium laws
process has already been addressed [9]. The Gibbs probability can be derived from the stochastic dynamics, which is the
distribution, for instance, is believed to come from the second step mentioned above. We will treat some fundamental
underlying mechanics through a stochastic behavior, although issues that have barely been considered or that have not
there is no known general derivation from pure mechanics. been addressed in the context of stochastic thermodynamics.
In some cases, the derivation is known [10]. In other cases, This includes several thermodynamic results of systems in
such as a system of hard spheres, numerical simulations of equilibrium [50,51] and out of equilibrium [52–58] such as the
the equations of motion may, for instance, show the validity quasiequilibrium processes; the convexity of the equilibrium

1539-3755/2015/91(4)/042140(18) 042140-1 ©2015 American Physical Society


TÂNIA TOMÉ AND MÁRIO J. DE OLIVEIRA PHYSICAL REVIEW E 91, 042140 (2015)

surface in thermodynamic space; the monotonic time behavior following expression:


of thermodynamic potentials, including entropy; the bilinear 
form of the entropy production rate; the Onsager coefcients S(t) = −kB Pi (t) ln Pi (t), (2)
and reciprocal relations; and the nonequilibrium steady states i
of chemical reactions. which is the extension of the equilibrium Boltzmann-Gibbs en-
The stochastic approach in continuous state space was used tropy to nonequilibrium situations, where kB is the Boltzmann
by Einstein [59], Smoluchowski [60], and Langevin [61] to constant.
explain Brownian motion. It was generalized to the case of The second assumption concerns the denition of the pro-
Brownian particles subject to an external force by Fokker [62], duction of entropy. This quantity should meet two fundamental
Smoluchowski [63], Planck [64], and Ornstein [65], and properties. It must be non-negative and vanish identically in
the equation governing the time evolution of the probability thermodynamic equilibrium. Following Schnakenberg [17],
distribution became known as the Fokker-Planck equation. we assume the following expression for the entropy production
Kramers [66] extended the Fokker-Planck equation to the case rate:
of a massive particle and studied the escape of a Brownian
particle over a potential barrier arriving at the Arrhenius kB  Wij Pj (t)
(t) = {Wij Pj (t) − Wj i Pi (t)} ln , (3)
factor. 2 ij Wj i Pi (t)
Markovian stochastic dynamics [67–70] has been used in
various problems in physics, chemistry, and biology, either which is clearly non-negative because each term is of the
in continuous or discrete state space. In the former case, form (x − y) ln(x/y). This form of entropy production rate has
the evolution of the probability distribution is governed by a been used by several authors [18,19,26–30,32,37,38,40,42,70]
Fokker-Planck equation and in the later by a master equation. within stochastic dynamics and applications.
We mention the study of chemical reactions [16,19,20,30,46],
population dynamics and epidemiology [71–73], and bio- B. Entropy ux
logical systems in general [15,28,43,45,74–78]. We wish to
mention particularly the stochastic models with many degrees Let us consider the time variation of the average of a state
of freedom such as the so-called stochastic lattice models function, such as energy, given by
usually used to describe phase transitions and criticality in 
physics, chemistry, and biology [26,79–86]. U (t) = Ei Pi (t). (4)
i

Using the master equation (1) it follows that


II. MASTER EQUATION
dU
A. Entropy and entropy production = u , (5)
dt
We assume that the system follows a microscopic stochastic
dynamics. More precisely, we assume that the system is where
described by a continuous-time Markovian stochastic process. 
u (t) = (Ei − Ej )Wij Pj (t) (6)
Considering a discrete space of states, this assumption posits
ij
that the time evolution equation is set up once the transition
rates are given. The transition rates play thus a fundamental is the total ux of energy from outside to the system.
role in the present approach and we may say that a system Equation (5) represents the conservation of energy.
is considered to be theoretically dened when this quantity Equation of the type (5) is valid for any conserved
is given a priori. Given the transition rates, the probability quantity and this is not the case of entropy. For instance,
Pi (t) of state i at time t is obtained by solving the evolution in a nonequilibrium stationary state the total ux of energy
equation, in this case a master equation, vanishes but not the total ux of entropy, which is nonzero
because entropy is continuously being produced. The equation
d 
Pi (t) = {Wij Pj (t) − Wj i Pi (t)}, (1) for the time variation of entropy S should be written as [55]
dt j dS
=  − , (7)
dt
where Wij denotes the transition rate from state j to state i.
In this section and the next we will consider transitions with where  is the ux of entropy from the system to the outside
the following property: If the rate Wij of the transition j → i and  is the entropy production per unit time, given by Eq. (3).
is nonzero, then the rate Wj i of the reverse transition i → j is It is common to write di S/dt and de S/dt for the entropy
also nonzero. Later, in the study of the Fokker-Planck equation, production rate and entropy ux, respectively, but we avoid
we will have the opportunity to treat the case in which the this terminology because these quantities are not in fact time
reverse transition rate may vanish. derivatives of any quantity.
As mentioned above, the derivation of the macroscopic Taking the time derivative of Eq. (2) and using the master
properties, including the laws of thermodynamics, is carried equation (1), we may write the time derivative of entropy as
out by the introduction of two assumptions concerning entropy. 
dS
The rst is the denition of entropy itself. The entropy S of a = kB {Wij Pj (t) − Wj i Pi (t)} ln Pi (t), (8)
system in equilibrium or out of equilibrium is taken to be the dt ij

042140-2
STOCHASTIC APPROACH TO EQUILIBRIUM AND . . . PHYSICAL REVIEW E 91, 042140 (2015)

or, in an equivalent form, which is the detailed balance condition that characterizes
dS  Pi (t) the thermodynamic equilibrium [14] and is equivalent to
= kB Wij Pj (t) ln . (9) microscopic reversibility.
dt ij
Pj (t) In the stationary state, that is, when the probability Pi is
independent of time, the right-hand side of (1) vanishes, that
Comparing with (7) we see that the right-hand side of this
is,
equation should equal  − . Using the denition of , given 
by (3), which we write in the form {Wij Pj − Wj i Pi } = 0, (16)
 Wij Pj (t) j
(t) = kB Wij Pj (t) ln , (10)
Wj i Pi (t) which we may call the global balance. The reversibility
ij
condition (15) thus can be understood as detailed balance
and, comparing with Eq. (9), we get the ux of entropy from condition because each term of the global balance equation
the system to outside, vanishes. Although the global balance is a necessary condition
 Wij for reversibility, it is not a sufcient condition.
(t) = kB Wij Pj (t) ln , (11) Considering that the equilibrium distribution Pie is known,
Wj i
ij the solution of (15) for the transition rate is
which is equivalent to  1/2
Pie
kB  Wij Wij = Kij , (17)
(t) = {Wij Pj (t) − Wj i Pi (t)} ln . (12) Pje
2 ij Wj i
where Kij is symmetric, that is, Kij = Kj i . The transition
The integration of (7) in a time interval will lead us to the rates for the various situation in which the system is found in
Clausius inequality. Indeed, from Eq. (7) we may write equilibrium in the stationary state can now be constructed. For
  an isolated system (microcanonical ensemble) the equilibrium
S = dt − dt. (13) probability distribution Pi is a constant whenever the energy
function Ei equals a given energy, say, U , and vanishes
If we identify the entropy ux  as the ratio between the heat otherwise. Therefore, in this case Wij = Kij when Ei = Ej
∫ dQ/dt ∫and the temperature T of the environment, then
ux and vanishes otherwise. In short, Wij = Wj i .
dt = − (dQ/T ). But the rst integral is non-negative For a system in contact with a heat reservoir (canonical
because  > 0 so ensemble) at temperature T , the equilibrium probability
 distribution is given by
dQ
S > , (14) 1 −βEi
T Pie =
e , (18)
Z
∫which is the Clausius inequality [87]. In equilibrium, S =
dQ/T , equality that was used by Clausius to∫ dene entropy. where β = 1/kB T , so in this case the transition rate fullls
The difference between S and the integral dQ/T , which the relation
is the production of entropy, represents, according to Clausius, Wij
the “uncompensated transformation” [87]. = e−β(Ei −Ej ) (19)
Wj i
In the recent literature it is common to use another
nomenclature for the entropy production , the entropy ux and is given by
, and the time derivative of entropy dS/dt. The quantities Wij = Kij e−β(Ei −Ej )/2 . (20)
that correspond to the time integral of these three quantities are
called, respectively, the total entropy change, the environment If, in addition, to be in contact with a heat reservoir, the
entropy change, and internal entropy change [44,48]. system is in contact with a reservoir of particles, then
1 −βEi +βμni
Pie = e , (21)
C. Thermodynamic equilibrium 
The microscopic denition of thermodynamic equilibrium, where μ is the chemical potential and ni is the number of
from the static point of view, is usually characterized in terms particles. In this case the transition rate fullls the relation
of the Gibbs probability distribution. From the dynamic point Wij
of view, the description of equilibrium by the Gibbs distribu- = e−β(Ei −Ej )+βμ(ni −nj ) (22)
Wj i
tion is necessary but not sufcient. There are examples [88–90]
of spin models that are described by the Gibbs distribution but and is given by
are not in thermodynamic equilibrium in the sense that entropy
Wij = Kij e−β(Ei −Ej )/2+βμ(ni −nj )/2 , (23)
is continuously being generated. From a dynamic point of
view, the thermodynamic equilibrium is characterized by the where, again, Kij = Kj i .
vanishing of the entropy production rate and, of course, by a
time-independent probability distribution. The vanishing of (3) D. The approach to equilibrium
gives
Let us consider the transient regime of a system that
Wij Pj = Wj i Pi , (15) approaches equilibrium. The time-dependent probability

042140-3
TÂNIA TOMÉ AND MÁRIO J. DE OLIVEIRA PHYSICAL REVIEW E 91, 042140 (2015)

distribution is the solution of the master equation (1) with tran- The grand-canonical distribution describes the contact
sition rates that satisfy the detailed balance and is appropriate of the system with a particle reservoir and with a heat
for each type of contact of the system with the environment. reservoir. The transition rate for this case is given by (23),
We treat rst the case of microcanonical distribution, which which, replaced in the expression (11) and using the master
describes an isolated system. In this case, as we have seen, equation (1), allows us to reach the following expression for
Wij = Wj i so the entropy ux (11) vanishes identically,  = the entropy ux:
0. Therefore,
1 dU μ dN
dS =− + , (32)
= , (24) T dt T dt
dt where U is the average energy, given by (4), and N is the
so average number of particles,
dS 
> 0. (25) N(t) = ni Pi (t). (33)
dt
i
That is, the entropy of an isolated system is a monotonically
increasing function of time. Taking into account that dS/dt =  − , we get
Next we consider the canonical distribution which describes dU dS dN
the contact of a system with a heat reservoir. The transition −T −μ = −T , (34)
dt dt dt
rate is given by (20), which, replaced in the entropy ux (11),
gives which can be written as
 dφ
 = −kB β {Wij Pj (t) − Wj i Pi (t)}Ei . (26) = −T , (35)
dt
ij
where φ = U − T S − μN is the grand thermodynamic poten-
Using the master equation (1), the ux of entropy can be tial and we have taken into account that T and μ are constant.
written in the form Since  > 0 it follows that dφ/dt 6 0.
1 dU Let us integrate equation (34) from an initial time t = t0 to
=− , (27) innity,
T dt
 ∞
where U is the average of energy, given by (4). Equation (27)
shows that the quantity  is proportional to dU/dt. Notice (U − U0 ) − T (S − S0 ) − μ(N − N0 ) = −T dt,
t0
that (27) implies that  vanishes in the equilibrium regime (36)
(t → ∞) as it should. from which follows the inequality
Equation (7) gives
(U − U0 ) − T (S − S0 ) − μ(N − N0 ) 6 0, (37)
dU dS
−T = −T . (28) because  > 0. Taking into account that, for large-enough
dt dt
times, the system reaches equilibrium at a temperature T
If we dene the free energy by F = U − T S and take into
and imposing that at t = t0 the system was in equilibrium,
account that T is constant, that is, it does not depend on time,
at a different temperature, say, T0 , we may conclude from the
we get
inequality (37) that U , S, and N make up a convex surface, in
dF accordance with equilibrium thermodynamics.
= −T , (29)
dt
so E. Quasiequilibrium
dF It is common to state the laws of equilibrium thermody-
6 0. (30)
dt namics in terms of thermodynamic processes. This seems at
That is, the free energy of a system in contact with a heat rst sight contradictory because a process implies a change
reservoir is a monotonically decreasing function of time. In in the thermodynamic state and thus a displacement from
other terms, the free energy decreases monotonically to its equilibrium. To overcome this problem, one introduces the
equilibrium value. quasistatic process, a process which is so slow that the system
Equation (30) is also the expression of the Boltzmann H- may be considered to be in equilibrium. We will show below
theorem [3]. Indeed, the Boltzmann H function is dened by that the production of entropy in this process is negligible so
in fact the system may be considered to be in equilibrium. In
 which sense the production is negligible will be shown below.
Pi (t) Let us consider a system in contact with a heat bath
H (t) = Pi (t) ln , (31)
i
Pi e and a particle reservoir whose temperature and chemical
potential, understood as control parameters, are slowly varying
where Pi e is the equilibrium canonical distribution given by in time. To describe this situation we assume a time-dependent
Eq. (18). It is straightforward to show that F = F0 + H /β, transition rate Wij (t) of the form (23), where Kij (t) may
where F0 does not depend on time. Therefore, the inequal- depend on time, that is,
ity (30) is equivalent to dH /dt 6 0, which is the Boltzmann
H-theorem. Wij (t) = Kij (t)e−β(Ei −Ej )/2+βμ(ni −nj )/2 , (38)

042140-4
STOCHASTIC APPROACH TO EQUILIBRIUM AND . . . PHYSICAL REVIEW E 91, 042140 (2015)

where Kij (t) = Kj i (t), so


T
Wij (t) e−β(Ei −μni )
= −β(Ej −μnj ) , (39)
Wj i (t) e
S
where β = 1/kB T and T (t) depends on time and the chemical µ
U
potential μ(t) also depends on time. We assume, moreover,
N
that dβ/dt = α and dμ/dt = γ are small and are both of the
same order of magnitude. FIG. 1. A path in the T ,μ space and the corresponding trajectory
Replacing Eq. (39) in expression (11) for the entropy ux in the thermodynamic space S,U,N. If the variations in T and μ are
and after some straightforward algebraic manipulation we very slow, then the trajectory in the thermodynamic space approaches
reach again the result and remains on a certain surface which has the property of convexity
and is identied as the thermodynamic equilibrium surface. The
1 dU μ dN
=− + . (40) portions of the trajectory outside and on the surface are represented
T dt T dt by dashed and solid lines, respectively.
Now dS/dt =  −  so
dU dS dN Let us take a look at the thermodynamic space spanned by
=T +μ − T . (41)
dt dt dt the variables S, U , and N. From the solution of the master
equation, we may determine these quantities as a function of
Let us now nd the solution of the master equation. To this
time by using the denitions (2), (4), and (33). The evolution
end, we write the solution as
of the system may be represented by a trajectory of a point in
Pi (t) = Pi∗ (t) + Ai (t), (42) this space, as shown in Fig. 1. The representative point will
describe a generic trajectory in this space. But if T and μ start
where to vary very slowly, the trajectory, according to the result (47),
1 will approach and remain on a certain surface of this space, as
Pi∗ (t) = exp{−β(Ei − μni )}, (43) seen in Fig. 1. According to (47), the surface is represented by
(β,μ)
the equation
and (β,μ) is a time-dependent quantity such that Pi∗ (t) is
normalized at any time and Ai is small when compared to Pi∗ . dU = T dS + μdN, (48)
It is important to bear in mind that although Pi∗ (t) obeys the
relation so the temperature T of the thermal reservoir becomes
identied as the tangent to the surface U (S,N) along the S
Wij (t)Pj∗ (t) = Wj i (t)Pi∗ (t), (44) direction, T = ∂U/∂S, and thus can be interpreted as the
temperature of the system. Similarly, the chemical potential
and can be interpreted as a probability distribution, it is not the
of the particle reservoir becomes identied as the tangent to
solution of the master equation, given by (1). The substitution
the surface U (S,N) along the N direction, μ = ∂U/∂N, and
of Pi∗ on the master equation (1) makes the right-hand side
thus can be interpreted as the chemical potential of the system.
equal to zero but not the left-hand side. Replacing Eq. (42) into
Notice that, according to the inequality (37), this surface has
the master equation, we get, up to rst order in the perturbation
the property of convexity.
Ai ,
We should remark that, far from equilibrium, the tempera-
d ∗  ture of the system cannot be dened because S, U , and N are
Pi (t) = {Wij (t)Aj (t) − Wj i (t)Ai (t)}. (45)
dt not connected by relation (48). The same can be said about the
j
free energy of systems far from equilibrium. Notice, however,
Now that the quantity F = U − T S, dened previously and called
dPi∗ ∂Pi∗ ∂Pi∗ free energy, is not properly a property of the system because
= α+ γ, (46) T is the temperature of the reservoir and not the temperature
dt ∂β ∂μ of the system, since it cannot be dened. In equilibrium or
which, in view of Eq. (45), implies that the perturbation Aj (t) quasiequilibrium, however, it becomes a well-dened quantity
is of the order of α and γ . From the expression (40) for the as much as the temperature. It is worth mentioning in addition
entropy ux , it follows that  is also of the order α and that according to the approach just presented, the control
γ . On the other hand, if we consider the expression (3) for parameters should be the thermodynamic variables known as
the entropy production , it follows that  is of of second thermodynamic eld variables [51].
order in α and γ . Therefore, in the quasiequilibrium regime,
in which we consider only terms up to rst order in α and γ , F. Fluxes and forces
the relation dS/dt =  −  becomes dS/dt = −, that is,
the production of entropy vanishes when compared with the We consider here the contact of a system with two distinct
ux of entropy. Using this result it follows from (41) that the reservoirs. To treat this situation properly, we assume that each
following thermodynamic relation holds: pair of states (i,j ) is either associated to the rst reservoir or to
the second reservoir or to neither of them. In other words, the
dU dS dN set of pairs (i,j ) is partitioned into three subsets, associated to
=T +μ . (47)
dt dt dt the rst reservoir, to the second reservoir and neither of them,

042140-5
TÂNIA TOMÉ AND MÁRIO J. DE OLIVEIRA PHYSICAL REVIEW E 91, 042140 (2015)

which we denote by A, B, and C, respectively. The transition of the uxes:


rates associated to the reservoirs 1 and 2 are denoted by Wij1
Ju = Luu Xu + Lun Xn , (57)
and Wij2 , respectively, and are assumed to be of the same form
of (23), that is,
Jn = Lnu Xu + Lnn Xn . (58)
Wijr = Kijr e−βr (Ei −Ej )/2+βr μr (ni −nj )/2 , (49)
The coefcients Luu , Lun , Lnu , and Lnn are the Onsager
for r = 1,2, where Kijr is symmetric as before. In addition, Kij1 coefcients. According to Onsager, the cross coefcients are
depends on T1 and μ1 and is nonzero only if (i,j ) ∈ A, and equal, Lun = Lnu , which is the Onsager reciprocal relation. In
Kij2 depends on T2 and μ2 and is nonzero only if (i,j ) ∈ B. We the following we will derive expressions for these coefcients
are denoting by T1 and μ1 and T2 and μ2 the temperatures and and prove the reciprocal relation.
chemical potentials of the two reservoirs and βr = 1/kB Tr . We will suppose that T1 and μ1 are xed and let T2 →
The full transition rate is given by T1 and μ2 → μ1 . Let Pie be the probability distribution
corresponding to the equilibrium case, given by
Wij = Wij0 + Wij1 + Wij2 , (50) 1 −β1 (Ei −μ1 ni )
Pie = e . (59)
where Wij0 may be nonzero only if (i,j ) ∈ C. In this case it is 
nonzero if Ei = Ej and Ni = Nj , in which case Wij0 = Wj0i . The transition rate Wije obeys the detailed balance
In the following, we consider the stationary regime for Wije Pje = Wjei Pie (60)
which the stationary probability distribution Pi fullls the
global balance (16) but not the detailed balance. In the present and is given by
case  =  and using the expression (11) we may write
Wije = Kije e−β1 (Ei −Ej )/2+β1 μ1 (ni −nj )/2 , (61)
 Wijr
 = kB Wijr Pj ln , (51) where Kije = Kij1 + Kij∗ + Kij0 and Kij∗ equals Kij2 when T2 →
r=1,2 ij
Wjri T1 and μ2 → μ1 and Kij0 = Wij0 .
The stationary solution Pi of the master equation (16) is
where the rst summation runs only over r = 1,2. The terms written as
corresponding to r = 0 vanish because Wij0 = 0 or because
Wij0 = Wj0i . Replacing expression (49) in this equation, the Pi = Pie (1 + ai Xu + bi Xn ), (62)
entropy production can be written as up to linear term in Xu and Xn . Replacing into the expres-
 sions (53) and (54) we get the Onsager coefcients in the
 = kB Wijr Pj βr [(Ej − Ei ) − μr (nj − ni )]. (52) form
r=1,2 ij
1 1 e
Luu = Wij Pj (aj − ai )(Ei − Ej ), (63)
Now the ux of energy Ju and the ux of particles Jn from 2 ij
reservoir 1 into the system are given by
 1 1 e
Ju = Wij1 Pj (Ei − Ej ), (53) Lun = Wij Pj (bj − bi )(Ei − Ej ), (64)
2 ij
ij
 1 1 e
Jn = Wij1 Pj (ni − nj ). (54) Lnu = Wij Pj (aj − ai )(ni − nj ), (65)
ij
2 ij

The substitution of (53) and (54) into (52) and the use of the 1 1 e
Lnn = Wij Pj (bj − bi )(ni − nj ), (66)
global balance condition (16) allow us to write the entropy 2 ij
production rate in the bilinear form [52,53,55]
where we have used the detailed balance condition (60). In
 = Xu Ju + Xn Jn , (55) the form given by Eqs. (64) and (65) we cannot tell whether
the coefcients Lnu and Lun are equal. Next we perform a
where Xu and Xn are the thermodynamic forces transformation to nd expressions that will show that these
coefcients are indeed equal to each other.
1 1 μ1 μ2
Xu = − , Xn = − , (56) Replacing (62) into (16), and expanding the result up to
T2 T1 T1 T2 linear terms in Xu and Xn , we end up with the following
conjugated to the ux of energy and particles, respectively. equations for ai and bi :
 1  ∗ e
Wije Pje (aj − ai ) + Wij Pj (Ej − Ei ) = 0, (67)
G. Onsager coefcients j
kB j
When T2 = T1 and μ2 = μ1 , that is, when Xu = 0 and  1  ∗ e
Xn = 0, the uxes Ju and Jn vanish. Therefore, up to linear Wije Pje (bj − bi ) + Wij Pj (nj − ni ) = 0, (68)
kB j
terms in Xu and Xn we expect the following linear behavior j

042140-6
STOCHASTIC APPROACH TO EQUILIBRIUM AND . . . PHYSICAL REVIEW E 91, 042140 (2015)

where Wij∗ equals Wij2 when T2 → T1 and μ2 → μ1 and is which describes the contact with a heat reservoir at temperature
given by T , where Kij0 is symmetric.
At the stationary state, the entropy production rate equals
Wij∗ = Kij∗ e−β1 (Ei −Ej )/2+β1 μ1 (ni −nj )/2 . (69) the ux of entropy and is given by
Multiplying (67) by Ei and by ai and summing in i we are   Wijrk
led to two equations from which we may obtain the following  = kB Wijrk Pj ln , (75)
expression for Luu : r=0,1,2 k ij
Wjrki

1  ∗ e kB  e e which follows from the general expression (11). The substitu-


Luu = Wij Pj (Ej − Ei )2 − W P (aj − ai )2 . tion of (73) and (74) into this expression gives
2kB ij 2 ij ij j
1    rk  
(70) = Wij Pj μrk nki − nkj , (76)
T r=1,2 k ij
Multiplying (68) by ni and by bi and summing in i we are lead
to two equations from which we may obtain the following where the terms involving the energy vanish. Taking into
expression for Lnn : account that the ux Jk of particles of type k, from the reservoir
1 to the system, is given by
1  ∗ e kB  e e   
Lnn = Wij Pj (nj − ni )2 − W P (bj − bi )2 . Wij1k Pj nki − nkj , (77)
2kB ij 2 ij ij j Jk =
ij
(71)
and using the total balance equation (16), we may write again
Multiplying (67) by ni and (68) by ai and summing in i, we the entropy production rate in the bilinear form
get two equations from which we reach an expression for Lnu . 
Similarly, multiplying (67) by bi and (68) by Ei and summing = Xk Jk , (78)
k
in i, we get an expression for Lun which is equal to Lnu ,
proving the reciprocal relation. The expression for these two where
quantities is given by 1 1 
Xk = μ − μ2k (79)
1  ∗ e T k
Lun = Lnu = Wij Pj (Ej − Ei )(nj − ni )
2kB ij is the thermodynamic force associated to species k.
For the case of two types of particles, it follows from
kB  e e expression (78) that X1 J1 + X2 J2 > 0 because  > 0. If
− Wij Pj (bj − bi )(aj − ai ). (72)
2 ij X1 < 0, it is possible to have J1 > 0, as long as X2 J2 >
|X1 |J1 , so the ux of particles of type 1 will occur against the
It is worth mentioning that in the course of derivation of chemical potential gradient. This is, for instance, a mechanism
these expressions, we have made use of the detailed balance for the active transport across a cell membrane. A simple
condition, which is thus a necessary condition to prove the model [74] of this type of transport is examined next.
reciprocal relation. However, the expressions for the Onsager A cell membrane is assumed to have a certain number of
coefcients do not depend on the equilibrium distribution alone channels through which two types of molecules may cross the
but depend also on the deviations ai and bi . membrane from the exterior to the interior of the cell. The
channels function independent of each other so it sufces
H. Several species of particles to consider just one of them. A channel may be open to
the exterior, understood as reservoir 1, or to the interior,
We will now treat the case of a system composed by understood as reservoir 2, and can be either empty or hold
several types of particles in contact with two particle reservoirs, a molecule A or two molecules, one A and another B. The
denoted by 1 and 2. In the steady state, uxes of particles of the possible states and transitions are shown in Fig. 2.
various types will be established between the two reservoirs.
Each reservoir is in fact a set of reservoirs, one for each type
of particles. As before, denoting by Ei the energy of state i E EA EAB
and by nki the number of particles of species k in state i, the 1 3 5
rate of the transition j → i associated to the reservoir r and
species k is given by
r k k
Wijrk = Kijrk e−β[(Ei −Ej )−μk (ni −nj )]/2 , (73) 2 4 6
I IA IAB
where μrk is the chemical potential of species k associated
to reservoir r and Kijrk is symmetric. The reservoirs are also FIG. 2. Transition diagram for a model for active transport across
thermal reservoirs with a common temperature T and β = a cell membrane. The circles represent the possible states of a channel
1/kB T . A transition rate that is not associated to any reservoir and the bonds represent the possible transitions. The possible states of
is denoted by Wij0 and is assumed to be of the form the channel are (a) open to exterior and empty (E), holding a molecule
(EA), or holding two molecules (EAB); (b) open to interior and empty
Wij0 = Kij0 e−β(Ei −Ej )/2 , (74) (I), holding a molecule (IA), or holding two molecules (IAB).

042140-7
TÂNIA TOMÉ AND MÁRIO J. DE OLIVEIRA PHYSICAL REVIEW E 91, 042140 (2015)

Denoting by wij the rate of the transition j → i, then w31 , state j to state i, then
w13 , w53 , and w35 are associated to the reservoir 1, whereas  
nki = nkj and nki = nkj , k  = k, (85)
w42 , w24 , w64 , and w46 , associated to reservoir 2. In this simple
model, the energies of the state are assumed to be the same so, because we are assuming that the k-th reservoir causes a change
according to (73), they hold the following relations: in the number of particles of type k but causes no changes in
w31 1 w53 1 the number of particles of the other types.
= eβμA , = eβμB , (80) When the system is in thermodynamic equilibrium with the
w13 w35
reservoirs, the probability distribution describing the system
w42 2 w64 2
is the Gibbs distribution
= eβμA , = eβμB , (81)
w24 w46 1 ∑ k
Pie = e−βEi +β k μk ni , (86)
where μkA and μkB are the chemical potentials of molecules A 
and B associated to reservoir k. The other rates are not related where β = 1/kB T .
to the reservoirs and are symmetric, w21 = w12 , w43 = w34 , ̂ijk describing the contact of
To set up the transition rate W
and w65 = w56 . Assuming that chemical potentials are given, the system with the k reservoir we use the detailed balance
the model has seven independent transition rates. condition with respect to the probability distribution (86),
At the stationary state, the probability distribution Pi obeys ̂ijk
W Pie
the global balance equation = , (87)
̂jki
W Pje

(wij Pj − wj i Pi ) = 0, (82) where i and j are states such that condition (85) is fullled, so
j
̂ijk
W k k

but do not obey the detailed balance condition, which means = e−β(Ei −Ej )+βμk (ni −nj ) , (88)
̂jki
W
that wij Pj − wj i Pi = 0 in general. The uxes JA and JB
of molecules A and B, respectively, from the exterior to the which leads us to the following form:
interior, are given by ̂ijk e−β(Ei −Ej )/2+βμk (nki −nkj )/2 ,
̂ijk = K
W (89)
JA = w13 P3 − w31 P1 , (83) ̂ijk is symmetric, that is, K
̂ijk = K
̂jki , and is positive or
where K
vanishes according to whether the condition (85) is fullled.
JB = w35 P5 − w53 P3 , (84) The total transition rate W ̂ij , due to the contact with all
reservoirs, is written as the sum
and are nonzero because detailed balance does not hold. The q

entropy production rate is  = XA JA + XB JB , where XA = ̂ij =
W ̂ijk .
W (90)
(μ1A − μ2A )/T and XB = (μ1B − μ2B )/T . By an appropriate k=1
choice of the transition rates, it is thus possible to have a ux
of particles B against its chemical potential gradient [74], that Notice that at most one of the q terms on the right-hand side
is, it is possible to have JB > 0 and XB < 0, as long as the can be nonzero.
ux of particles A agrees with the gradient of its chemical Let us consider now the occurrence of chemical reactions.
potential, that is, XA JA > 0. The number of particles of each species will vary not only
because of the contact with the reservoirs but also because
of the reactions. We consider the occurrence of r reactions
III. CHEMICAL REACTIONS described by the chemical equations
q
A. Equilibrium 
We will be concerned in this section with a system com- νk Bk = 0,  = 1,2, . . . ,r, (91)
k=1
posed by q species of particles that react among themselves
according to r reactions. The system is in contact with a heat where Bk denotes the chemical formula of species k and νk
reservoir and may be closed to particles or may be open and are the stoichiometric coefcients, which are negative for the
exchange particles with the environment. This last situation is reactants and positive for the products of the reaction. If the
carried out by placing the system with particle reservoirs. We -th reaction causes a change from state j to state i, then
will treat in the following the more general open case. The nki − nkj = νk or nki − nkj = −νk . (92)
results for the closed case will readily be obtained from the
results of the open case by formally imposing the vanishing of To set up the transition rate W̃ij describing the change
the particle ux. caused by the -th reaction we assumed that it obeys the
The system is placed in contact with q particle reservoirs, Arrhenius equation [91,92]
one for each type of particle. Each particle reservoir is also
a thermal reservoir. The k-th reservoir exchanges heat, at W̃ij
= e−β(Ei −Ej ) . (93)
temperature T , and only particles of type k at a chemical W̃ji
potential μk . Notice that all reservoirs are at the same
The most general form of the transition rate is
temperature T . The number of particle of species k in state
i is denoted by nki . If the k-th reservoir causes a change from W̃ij = K̃ij e−β(Ei −Ej )/2 , (94)

042140-8
STOCHASTIC APPROACH TO EQUILIBRIUM AND . . . PHYSICAL REVIEW E 91, 042140 (2015)

where the prefactor is symmetric, that is, K̃ij = K̃ji , and is Using the master equation (1), we see that the average
positive if condition (92) is fullled and vanishes otherwise. number of particles Nk of type k,
The transition rate W̃ij due to all reactions is written as the 
Nk (t) = nki Pi (t), (99)
sum
i
r
 evolves as
W̃ij = W̃ij . (95) 
dNk  
=1 = Wij Pj nki − nkj . (100)
dt ij
Notice that at most one of the r terms on the right-hand side
can be nonzero. According to (96), the transition rate Wij has two parts, one
The full transition rate Wij describing the r reactions as related to the reservoirs, which is Ŵij , and the other related to
well as the contact with the q reservoirs is given by the chemical reactions, which is W̃ij , so (100) can be written
in the form
̂ij .
Wij = W̃ij + W (96) 
dNk  
= W̃ij Pj nki − nkj + k , (101)
Again, just one of the two terms on the right-hand side can be dt ij
nonzero.
In equilibrium, detailed balance should be obeyed for each where k is given by
  
one of the transition rates on the right-hand side of (96). We k = ̂ij Pj nki − nkj
W (102)
have seen that this is the case of the transition rates W ̂ijk ,
ij
related to the contact with each reservoir, when the probability
distribution is that given by (86). It sufces, therefore, to and describes the ux of particle from the k-th reservoir to
impose detailed balance to the transition rate associated to the system. The contact of the system with the k-th reservoir,

each chemical reaction. To this end we compare the ratio (93) described by the transformation (85), causes no changes in nki ,

with the ratio k = k. As a consequence,
̂ijk (nki  − nkj  ) = 0, k = k  .
W (103)
Pie ∑
−β(Ei −Ej )+β k νk μk
e =e , (97)
Pj Using (90) and the result (103), the ux of particle (102) is
written as
obtained from (86) and valid when the rst of the two   
k = ̂ijk Pj nki − nkj .
W (104)
conditions in (92) is fullled. The condition of detailed balance
ij
is obeyed when the two ratios are equal to each other, that is,
when The summation in the right-hand side of (101) describes the
 change in the number of particle due to the chemical reactions.
νk μk = 0, (98) To describe properly this part, which corresponds to the
k creation and annihilation of particles caused by the reactions,
it is convenient to use a new set of variables in the place of
for each reaction . The same conclusion is obtained if we
the set of variables nki , k = 1,2, . . . ,q. The new variables are
use the second of the two conditions in (92). Equation (98) is
denoted by σi ,  = 1,2, . . . ,r and xi ,  = r + 1,r + 2, . . . ,q
the well-known equilibrium condition that should be fullled
and dened by the linear transformation
when chemical reactions take place in a system [50,51].
r

In the presence of chemical reactions and in equilibrium,
the chemical potentials of the chemical species cannot be nki = xik + νk σi , k = 1,2, . . . ,q, (105)
independent but are related by (98). In other words, the =1

equilibrium occurs only when the chemical potentials μk of where the quantities xik , k = 1,2 . . . ,r are not variables but
the particle reservoirs are tuned so (98) is fullled. Otherwise, arbitrary constants chosen to be the same for all i. If a
the system will be out of equilibrium, as we shall see next. transformation nki → nkj is performed according to the -th
chemical reaction (91), described by the transformation (92),
B. Nonequilibrium regime
the variables xik remains unchanged, that is, xjk = xik . As a
consequence of this invariance,
Let us now suppose that the condition (98) is not obeyed.  
In this case the detailed balance condition is not fullled W̃ij xik − xjk = 0. (106)
and the system cannot be in equilibrium. Each reaction is In addition, according to the transformation (92), the vari-
shifted either to the products or to the reactants. That is, for ables σim , m = , associated to the other reactions remain
a given reaction, either the products are being created and the unchanged, σjm = σim , and, as a consequence,
reactants being annihilated (forward reaction) or the reactants  
are being created and the products being annihilated (backward W̃ijm σi − σj = 0, m = . (107)
reaction). In this nonequilibrium regime the time variation in Using (105) and the results (106) and (107), we obtain
the number of particles has two parts: one due to the ux of    
particles from the reservoirs and the other due the creation and W̃ij Pj nki − nkj = νk χ , (108)
annihilation caused by the reactions. ij 

042140-9
TÂNIA TOMÉ AND MÁRIO J. DE OLIVEIRA PHYSICAL REVIEW E 91, 042140 (2015)

where the quantity χ is Using (110), we reach the result


   dS 1 dU  A 1  dNk
χ = W̃ij Pj σi − σj . (109) =+ − χ − μk , (119)
ij dt T dt 
T T k dt

The variation in the number of particles then can be written as where A is the De Donder afnity [52],

dNk  A = − νk μk , (120)
= νk χ + k . (110)
dt 
k

associated to the -th chemical reaction.


When χ > 0, the -th chemical reaction is shifted to the
In the stationary state dS/dt = 0, dU/dt = 0, and
right, in the direction of the products. When χ < 0, it is
dNk /dt = 0, and we reach the following expression for the
shifted to the left, in the direction of the reactants. The extent
production of entropy in the stationary state [52,55,57,58]:
of reaction ξ is dened as the average of σi ,
 A
 = χ , (121)
ξ = σi Pi . (111) T

i
or, in the equivalent form,
From the master equation and using the properties (103)
and (107), we get  A dξ
= , (122)
T dt
dξ 
= χ , (112)
dt equation originally introduced by De Donder [52]. In equilib-
and we may conclude that the quantity χ is the rate of the rium there is no production of entropy and the afnities vanish,
extent of the -th reaction. A = 0, in accordance with (98), and the rate in which the -th
The time variation of the internal energy is written as reaction proceeds vanish as well, χ = dξ /dt = 0.
In a nonequilibrium stationary state, a ux of particles
dU  is continuously taking place, which sustains the chemical
= Wij Pj (Ei − Ej ). (113)
dt reactions. The quantities k and χ are nonzero, in general.
ij
At the same time there is a ux of heat toward the system,
Let us now consider the time variation of entropy characterizing an endothermic reaction, or from the system,
characterizing an exothermic reaction. To understand this
dS
=  − , (114) situation we write down the variation in the energy, given
dt by Eq. (113) in the form
where the ux  is given by (3), which, by the use of (96), (95), dU 
and (90), is given by = R + u , (123)
dt 
 ̂ijk
W
 = kB ̂ijk Pj ln
W where R is the energy delivered by the -th reaction per unit
̂jki
W time, given by
ij k

 W̃ij R = W̃ij Pj (Ei − Ej ), (124)
+ kB W̃ij Pj ln . (115)
ij 
W̃ji ij

and u is the heat ux to the system, given by


Substituting the rates (88) and (93) into this equation we 
get u = ̂ijk Pj (Ei − Ej ).
W (125)
 ij k
 = kB Wij Pj (−β)(Ei − Ej ) ∑
ij
In the stationary state dU/dt = 0 and u = −  R . If
   u < 0, the reactions are exothermic. If u > 0, they are
+ kB ̂ijk Pj βμk nki − nkj .
W (116) endothermic. Notice that there is no contribution to the entropy
ij k production rate coming from the ux of heat because the
temperatures of the reservoirs are the same.
Using Eqs. (113) and (104) we may write the ux of entropy Although we have considered a system in contact with one
as reservoir for each type of particle, the formulas can easily be
1 dU 1  adapted to the case in which the system is closed to some
=− + μk k . (117) types of particles. If the system is closed to particles of type k,
T dt T k
then it sufces to formally set μk = 0 and W ̂ijk = 0 so k = 0
Substituting into (114) and taking into account Eqs. (113), we for this species. It is worth mentioning that in the case of
get a closed system, when there is no ux of particles from the
environment, k = 0 for all species, and using equation (118),
dS 1 dU 1  we see that dF /dt = −T , where F = U − T S is the free
=+ − μk k . (118)
dt T dt T k energy, so dF /dt 6 0. Therefore, the chemical reactions occur

042140-10
STOCHASTIC APPROACH TO EQUILIBRIUM AND . . . PHYSICAL REVIEW E 91, 042140 (2015)

in a direction such that the variations in the number of particles We assume an expansion of the form
will decrease the free energy [93].  r q

As an example of the approach just developed we analyze a  
e
Pi = Pi 1 + Ri A + aik μk , (134)
system with four species of particles and two reactions, which
=1 k=r+1
are
where μk = μk − μ∗k .
B1 + B2 = B3 , B3 = B2 + B4 , (126) ̂ijk , given by (89), around its value at
We also expand W
and represent the Michaelis-Menten mechanism in which a equilibrium,
the substrate B1 is converted, in two steps, into the product B4
by the action of an enzyme. The substrate B1 reacts with the W ̂ijk e−β(Ei −Ej )/2+βμ∗k (nki −nkj )/2 ,
̂ij∗k = K (135)
enzyme B2 giving rise to a complex B3 which in turn breaks up
to get
into the product B4 and the enzyme B2 . It is assumed that both
{   }
reactions have reverses. The system is assumed to be closed ̂ijk = W
W ̂ij∗k 1 + βμk nki − nkj /2 . (136)
to the particles B2 and B3 and is in contact with reservoirs of
particles of type B1 and B4 . The transition rate W̃ij needs no expansion because this
Using formula (120), and bearing in mind that we should quantity is also its value at equilibrium since it does not depend
set μ2 = 0 and μ3 = 0 in this formula, the afnities A1 and on the chemical potentials.
A2 associated to the two reactions are given by Replacing the expansions (134) and (136) into the global
balance equation (16), we get
A1 = μ1 , A2 = −μ4 . (127)
 r

The variations in the number of particles of each species are Wij∗ Pje A (Rj  − Ri )
dN1 dN2 j =1
= −χ1 + 1 , = −χ1 + χ2 , (128)
dt dt  q

+ Wij∗ Pje (aj k − aik )μk
dN3 dN4
= χ1 − χ2 , = χ2 + 4 . (129) j k=r+1
dt dt   
+ ̂ij∗ Pje βμk nki − nkj = 0,
W (137)
In the stationary state, χ1 = χ2 = 1 = −4 , so, us-
j k
ing (121), the entropy production rate is found to be
χ1 where
= (μ1 − μ4 ). (130) 
T ̂ij∗ =
W ̂ij∗k ,
W ̂ij∗ + W̃ij .
and Wij∗ = W (138)
We may now draw the following conclusion for the case of k
a nonequilibrium steady-state situation, for which  > 0. If
μ1 > μ4 , then χ1 > 0 and χ4 > 0 so the two reaction equations Using (105) and taking into account relation (133), we see that
are shifted to the right, establishing a continuous annihilation  q

   
of particles of type B1 , which come from reservoir B1 because μk nki − nkj = μk xik − xjk
1 > 0, and production of particles of type B4 , which go to k k=r+1
reservoir B4 because 4 < 0. r
  
− A σi − σj , (139)
C. Onsager coefcients =1

In the nonequilibrium stationary state but close to equilib- which is replaced in (137) to get an expression linear in A and
rium we may expand the rates of the extents of reaction χ in μk . Since the coefcients of A and μk in this expression
terms of the afnities A to get should vanish, we obtain
    
χ = Lm Am , (131) Wij∗ Pje (aj k − aik ) + ̂ij∗ Pje xik − xjk = 0, (140)
W
m j j
where Lm are the Onsager coefcients. They obey the
valid for r + 1 6 k 6 q, and
reciprocal relations, which we demonstrate next.
   
We start by expanding the stationary probability distribution Wij∗ Pje (Rj  − Ri ) − β ̂ij∗ Pje σi − σj = 0,
W
Pi , that satises the global balance equation (16), around the j j
equilibrium distribution Pie given by
(141)
1 ∑ ∗ k
Pie
= e−βEi +β k μk ni , (132) valid for 1 6  6 r. These last two equations determine aik
Z
and Ri .
where the chemical potentials μ∗k obey the equilibrium Let us consider now the expansion of the rate of the extent
condition of reaction χ , given by (109). Replacing the expansion (134)

νk μ∗k = 0. (133) into (109), we get an expression linear in Am and μk . The
k coefcient of Am is the Onsager coefcient Lm which is

042140-11
TÂNIA TOMÉ AND MÁRIO J. DE OLIVEIRA PHYSICAL REVIEW E 91, 042140 (2015)

given by forces Fi are also conservative and the random forces are set
   up in such a way that in thermodynamic equilibrium they will
Lm = W̃ij Pje (Rj m − Rim ) σi − σj . (142) lead to the Gibbs canonical distribution.
ij Using the Itô interpretation, we can show that the Langevin
Next we use equation (141) to write the Onsager coefcient equations (145) are associated to the following Fokker-Planck
in a more appropriate form. To this end we proceed as follows. equation:
We multiply (141) by σim and sum in i to get a rst equation. ∂P  ∂ 1  ∂
Next we multiply (141) by Rim and sum in i to get a second =− (vi P ) − (Fi P )
∂t i
∂xi m i ∂vi
equation. From these two equations we get an equation for the
right-hand side of (142) from which we reach the following  αi ∂ 1  ∂2
expression: + (vi P ) + 2 (Bij P ), (148)
i
m ∂vi m ij ∂vi ∂vj
β  ̂∗ e m  
Lm = Wij Pj σi − σjm σi − σj equation that gives the time evolution of the probability
2 ij
distribution P (x,v,t) of x and v at time t. It is convenient
1  ∗ e to write down the Fokker-Planck equation in the following
− Wij Pj (Ri − Rj  )(Rim − Rj m ). (143) form:
2β ij
∂P ( ∂Ji
)
=− Ki + , (149)
From this expression it follows that ∂t ∂vi
i
Lm = Lm , (144) where Ki and Ji are given by
which is the Onsager reciprocal relation [53]. ∂P Fi ∂P
Ki = vi + (150)
∂xi m ∂vi
IV. FOKKER-PLANCK EQUATION and
A. Langevin equations αi 1  ∂
Ji = − vi P − 2 (Bij P ). (151)
In this section we are concerned with systems that follow m m j ∂vj
a continuous time Markovian process in the continuous state
space, the phase space. We consider a system of particles that Let us consider now the time variation of entropy S, given
follows a dynamics described by the following set of Langevin by
equations, interpreted according to Itô, 
S(t) = −kB P (x,v,t) ln P (x,v,t)dxdv. (152)
dvi
m = Fi (x) − αi vi + Fi (t), (145)
dt The derivative of S gives
where m is the mass of each particle, vi = dxi /dt and xi is  ( )
dS ∂P
the position of the i-th particle, and x denotes the vector x = = −kB ln P dxdv. (153)
dt ∂t
(x1 , . . . ,xn ). We will also use the notation v = (v1 , . . . ,vn ).
The quantity Fi (x) is the force acting on the i-th particle, and After replacing (149) into this equation and performing
Fi (t) is a stochastic variable with the properties appropriate integrations by parts we reach the following
expression:
Fi (t) = 0, (146)
dS   Ji ( ∂P )
= −kB dxdv. (154)
dt P ∂vi
Fi (t)Fj (t  ) = 2Bij δ(t − t  ), (147) i

The terms corresponding to Ki vanish, that is,


where Bij may depend on x and v.
Notice that we are considering the so-called underdamped 
−kB Ki ln P dxdv = 0. (155)
systems, for which the state of a particle is dened by its
i
position and velocity [38], in opposition to the overdamped
case, for which the state of a particle is dened only by its We are assuming that P and its derivatives vanish at the
position [29]. boundary of integration.
The quantities Fi (t) are random forces acting on the
particles including the ones that describe the contact of the B. Microcanonical ensemble
system with the environment. We will treat two cases: one Here we treat the case of an isolated system, with no contact
in which the system is isolated (microcanonical ensemble) with the environment so the energy is strictly conserved.
and the other in which the system is in contact with a heat We thus assume that the force Fi are conservative so Fi =
reservoir (canonical ensemble). In the rst case the forces Fi −∂V /∂xi , which allows us to dene the energy function
are conservative and the stochastic forces are set up in such a E(v,x) as
way that the energy is conserved in any stochastic trajectory. m
In thermodynamic equilibrium they will lead to the Gibbs E(x,v) = vi2 + V (x). (156)
microcanonical probability distribution. In the second case the i
2

042140-12
STOCHASTIC APPROACH TO EQUILIBRIUM AND . . . PHYSICAL REVIEW E 91, 042140 (2015)

The strict conservation of energy means to say that E(x,v) is given by


should be a constant along any stochastic trajectory in phase 
space. This condition is fullled by the following set of Ji = Jij vj , (163)
Langevin equations, understood according the Stratonovich j (=i)

interpretation: ( )
1 ∂P ∂P
 Jij = 2 λij vi − vj , (164)
dvi m ∂vj ∂vi
m = Fi (x) + ξij vj , (157)
dt j (=i) Let us determine now the time derivative of entropy, which
is given by Eq. (154). After replacing (163) into Eq. (154)
where ξij are stochastic variables with the antisymmetric
and performing appropriate integration by parts we reach the
property ξj i = −ξij . The multiplicative noise at the right-hand
following expression:
side changes the velocities of the particles while keeping the
 ( )
kinetic energy invariant and can be interpreted as random dS kB  1 ∂P ∂P 2
elastic collisions of the particles with themselves or with = 2 λij vj − vi dxdv. (165)
dt m i<j P ∂vi ∂vj
immobile scatters. A similar noise has been used do describe
a particle that moves at constant speed but changes direction We are assuming that P and its derivatives vanish at the
at random times [94,95]. boundary of integration. The right-hand side of this equation is
Multiplying (157) by vi and summing in i we may conclude, clearly non-negative and is therefore identied as the entropy
after using the antisymmetric relation ξj i = −ξij , that E(v,x) production rate,
is strictly conserved along any stochastic path x(t), v(t).  ( )
Therefore, the equation of motion (157) describes a system kB  1 ∂P ∂P 2
= 2 λij vj − vi dxdv, (166)
of particles evolving in time in such a way that the energy m i<j P ∂vi ∂vj
is strictly constant. In analogy with equilibrium statistical
mechanics, this denes a microcanonical ensemble. which can also be written in the form
The stochastic variables ξij (t) are dened by the relations  m2  Jij2
ξij (t) = 0 (158)  = kB dxdv, (167)
λ
i<j ij
P
and
where the summation is over ij such that λij = 0, so
ξij (t)ξij (t  ) = 2λij δ(t − t  ), (159)
dS
= . (168)
where λij > 0 is a parameter that gives the strength of the dt
stochastic noise. Using the Stratonovich interpretation, and In the present case there is no entropy ux,
taking into account the antisymmetric property ξj i = −ξij of
the noise, we may write down the associate Fokker-Planck  = 0, (169)
equation, given by
which is consistent with our interpretation that Eqs. (157)
∂P  ∂ 1  ∂ describe an isolated system. Taking into account that  > 0 it
=− (vi P ) − (Fi P ) follows at once that dS/dt > 0 for an isolated system.
∂t i
∂xi m i ∂vi
( ) In the stationary state, which is a thermodynamic equi-
1  ∂ ∂P ∂ ∂P librium, the probability distribution P e (x,v) depends on x
+ 2 λij vj vj − vj vi ,
m ij ∂vi ∂vi ∂vi ∂vj and v only through E(x,v), that is, P e (x,v) is a function
of E(x,v). This statement can be checked by substitution
(160) on the right-hand side of the Fokker-Planck equation (160).
an equation that gives the time evolution of the probability Since E(x,y) is invariant along any path in phase space and
distribution P (x,v,t) of x and v at time t. The last summation supposing that initially its value is U , it follows that
extends over i = j and we recall that λj i = λij > 0. 1
It is worth mentioning that Eq. (157), understood in the P e (x,v) = δ[U − E(x,v)], (170)

Stratonovich sense, is equivalent to the following equation,
where  is a normalization constant that depends on U .
interpreted according to Itô:
We remark that in this case , given by (166), vanishes, as
dvi  expected.
m = Fi (x) − αi vi + ξij vj , (161)
dt j (=i)
C. Canonical ensemble
where
 Now we consider the case of a system in contact with
αi = λij . (162) a heat reservoir. In fact, we will consider the more general
j (=i) case in which each particle i is in contact with a reservoir at
temperature Ti . The appropriate set of Langevin equations that
Of course, this equation leads to the same Fokker-Planck
describes this situation is given by
equation (160).
It is convenient to write down the Fokker-Planck equation dvi
m = Fi (x) − αi vi + ζi (t), (171)
in the form given by (149) where Ki is given by (150) and Ji dt

042140-13
TÂNIA TOMÉ AND MÁRIO J. DE OLIVEIRA PHYSICAL REVIEW E 91, 042140 (2015)

where ζi (t) is a stochastic variable with the properties Let us assume that the forces are conservative, Fi =
−∂V /∂xi . In this case, we dene the energy of the system
ζi (t) = 0, (172) as
m 2
E(x,v) = v + V (x). (181)
ζi (t)ζj (t  ) = 2αi kB Ti δij δ(t − t  ), (173) 2 i i
where Ti and αi are parameters. The two last terms in Eq. (171) Using the Fokker-Planck equation in the form (149) we get
are interpreted as describing the contact of the i-th particle with the following expression for the time derivative of the average
the heat bath at a temperature Ti and αi is the strength of the energy U = E(x,v):
interaction with the heat reservoir.
To the set of Langevin equations (171) is associated the dU
= −u , (182)
Fokker-Planck equation dt
∂P  ∂ 1  ∂ where
=− (vi P ) − (Fi P )  
∂t ∂xi m i ∂vi
i u = − m vi Ji dxdv (183)
1  ∂ kB  ∂ 2P i
+ αi (vi P ) + 2 αi Ti 2 , (174)
m i ∂vi m i ∂vi is the ux of energy from the system to outside. To reach this
expression we have performed appropriate integration by parts
an equation that gives the time evolution of the probability and assumed that P and its derivatives vanish at the boundaries
distribution P (x,v,t) of x and v at time t. of integration. Using the denition of Ji , given by (175), we
The Fokker-Planck equation again can be written in the may write the energy ux as
form given by (149) where Ki is given by (150) and Ji is given  ( 〈 〉 k B Ti )
by u = αi vi2 − . (184)
i
m
αi vi αi kB Ti ∂P
Ji = − P− . (175)
m m2 ∂vi When all temperatures are the same Ti = T we have  =
u /T so
Again the derivative of entropy is given by (154). Replac-
ing (175) into (154) we get the following expression [38]: dS 1 dU
− = , (185)
dt T dt
  ( )
dS m2 Ji2 m from which it follows that the time variation of F = U − T S
= + vi Ji dxdv. (176) is given by dF /dt = −T  so dF /dt 6 0.
dt i
αi Ti P Ti
Thermodynamic equilibrium occurs when all temperatures
The summation in (176) extends only to the terms for which are the same, Ti = T , and the forces are conservative, Fi =
αi = 0 and Ti = 0. −∂V /∂xi . In this case, Ki = 0 and Ji = 0, which leads to the
The rst term on the right-hand side of equation (176) following result for the equilibrium probability distribution:
is non-negative and is identied as the entropy production
1 −E(x,v)/kB T
rate [38], P e (x,v) = e , (186)
Z
 m2  J 2
= i
dxdv. (177) where Z is a normalization constant.
i
αi T i P If we integrate Eq. (185) in time, from an initial time t0
until innity, when the system is in equilibrium, we get
Although this identication may seem to be arbitrary, as has
been argued [41], we will see in the next section that in fact 1
S − S0 − (U − U0 ) > 0. (187)
it is in accordance with the expression (3). It vanishes only T
when Ji = 0, which is the equilibrium condition. The second Let us suppose that the system is in contact with just
summation is thus the entropy ux one heat reservoir at temperature T and that is temperature
m is varying slowly so dT /dt = α is small. This is again the
=− vi Ji dxdv, (178) quasistatic process that we have already discussed. In this
i
Ti
case, the quantity Ji will be of the order α so  will be of
which can also be written as the order α 2 . On the other hand,  remains at the linear order
 1 ( 〈 〉 αi Ti ) in α and we may write from (185) dS/dt = (1/T )dU/dt.
= αi vi2 − . (179) It follows that the entropy and energy cannot be arbitrary
i
T i m but are connected by the relation T dS = dU so a system
performing a quasistatic process may be considered to be in
After replacing Ji , given by (175), into (178) and performing equilibrium. From the result (187) we see that the curve that
an integration by parts, the variation of the entropy of the connect U and S has the property of convexity. To perceive this
system becomes it sufces to imagine that at the initial time t0 the energy U0
dS and entropy S0 correspond to values of equilibrium at a certain
=  − . (180) temperature T0 .
dt

042140-14
STOCHASTIC APPROACH TO EQUILIBRIUM AND . . . PHYSICAL REVIEW E 91, 042140 (2015)

D. Nonequilibrium stationary state V. MASTER EQUATION REPRESENTATION OF THE


Let us take a look at the the energy variation per unit time, FOKKER-PLANCK EQUATION
or power, Pi , associated to the i-th particle, given by A. Microcanonical ensemble
m d 〈 2〉 It is possible to represent the Fokker-Planck in terms of a
Pi = vi Fi  + v , (188) master equation. This can be done by a discretization of the
2 dt i
phase space in a such a way that the continuum limit will
where the rst term is associated to the dissipation due to the
reduce the master equation to the Fokker-Planck equation.
force Fi acting on the particle and the second the time variation
From the representation we can easily identify the transition
of its kinetic energy. Using the Fokker-Planck equation it is
probabilities from which we can obtain, for instance, the
straightforward to show that
entropy production rate.
〈 〉 αi Ti To set up the discrete stochastic dynamics we imagine a
Pi = αi vi2 − . (189)
m representative point in the phase space following a stochastic
trajectory. We consider two types of transitions from a given
Therefore,
 point in the phase space. The rst type is dened by the
u = Pi (190) transitions determined by the Hamiltonian ow. This type of
i transition is dened by
and (x,v) → (Hi+ x,Hi+ v), (198)
 Pi
= , (191)
Ti where Hi+ x and Hi+ v are vectors with the same components
i
of the vectors x and v except the i-th components xi and vi
and we recall that the summation is over i such that αi = 0 which are transformed to xi and vi , where xi is given by
and Ti = 0.
Let us consider the contact of the system with two reservoirs xi = xi + bvi , (199)
A and B at temperatures T1 and T2 , respectively. The heat ux
from reservoir A to the system is given by and vi is determined in such a way that the energy is conserved,
 that is,
J = Pi , (192)
i∈A E(Hi+ x,Hi+ v) = E(x,v), (200)
where the summation is over the particles that are in contact
with reservoir A. A similar expression holds, where b > 0 is a parameter. Each transition occurs with rate
 1/b. If b is sufcient small, vi is given by
J = Pi , (193)
i∈B Fi
vi = vi + b . (201)

for the heat ux J from reservoir B to the system. In m
the stationary state,  = , and taking into account the
Notice that the Hamiltonian transition dened above has no
expression (191) for , we get
reverse in the sense that from a point (Hi+ x,Hi+ v) we cannot
J J reach the point (x,v) with this type of transition.
= + . (194) The transitions of the second type changes only the
T1 T2
velocities and preserves the kinetic energy. This type is dened
But in the stationary state u = J + J  = 0 so by
 = XJ , (195)
(x,v) → (x,Mij v), (202)
where
1 1 where Mij v is a vector with the same components of the vector
X= − . (196) v except the components i and j which are vi and vj given by
T1 T2
Let us assume that X is small so T = T2 − T1 is small. vi = vi cos θ − vj sin θ, vj = vi sin θ + vj cos θ, (203)
In this case J = LT , where L is the thermal coefcient.
Writing the probability distribution as P (x,v) = P e (x,v)[1 −
where θ > 0, so (vi )2 + (vj )2 = vi2 + vj2 and the kinetic energy
T a(x,v)], where P e (x,v) is the equilibrium distribution
is preserved. Another possible transition is dened by
when the temperatures of the reservoir is T1 , we may calculate
J to get
(x,v) → (x,Mj i v). (204)
 
L= αi vi2 a(x,v)P e (x,v)dxdv, (197) Each of these transition occurs with rate equal to λij /m2 θ 2 .
i∈A
Notice that this second type of transition has a reverse since
which may be understood as an average over the equilibrium from the point (x,Mij v) it is possible to reach the point (x,v).
distribution. It is sufcient to observe that Mj i (Mij v) = v.

042140-15
TÂNIA TOMÉ AND MÁRIO J. DE OLIVEIRA PHYSICAL REVIEW E 91, 042140 (2015)

The transitions above lead us to the following master thermodynamic equilibrium, is the Gibbs distribution. Indeed,
equation: the detailed balance of the master equation gives us the relation
∂ 1 P e (Hi− x,Hi− v) = P e (x,v), (210)
P (x,v) = {P (Hi− x,Hi− v) − P (x,v)}
∂t i
b e
which means that P (x,v) depends on (x,v) through E(x,v).
 λij Writing
+ {P (x,Mij v)+P (x,Mj i v)−2P (x,v)},
2ε2 1 −E(x,v)/kB T
ij P (x,v) = e , (211)
Z
(205)
we see that the other relation,
where Hi− is dened in a way similar to Hi+ except that the sign
A+ − e − − e
i (Ci v)P (x,Ci v) = Ai (v)P (x,v), (212)
in front of b in Eq. (199) is negative. It is straightforward to
show that expression (205) reduces to the Fokker-Planck (160) is fullled if we take into account that all temperatures are the
in the limit ε → 0 and b → 0. same, Ti = T .
Taking into account that all transitions preserve the energy
E(x,v), we see that in equilibrium the probability distribution C. Entropy production
P e (x,v) depends on (x,v) through E(x,v). If at initial time the
energy is equal to U , then P e (x,v) vanishes if E(x,v) = U and We have seen that the entropy production rate of a system
is a constant if E(x,v) = U , which is the Gibbs microcanonical described by a master equation is obtained by expression (3).
distribution. This expression is appropriate when the rates of the reversed
transitions are nonzero. This is the case of transitions dened
by (202). The entropy production rate M associated to these
B. Canonical ensemble transitions, according to (3), is given by
Next we set up a discrete stochastic dynamics, described by kB   λij
a master equation, whose continuous limit gives the Fokker- M = {P (x,Mij v) − P (x,v)}
Planck equation (174). The representative point in phase space 2 x,v ij 2m2 θ 2
(x,v) performs a stochastic trajectory. We consider again two P (x,Mij v)
types of transitions from a given point in the phase space. × ln . (213)
The rst type is the transition dened by the Hamiltonian ow P (x,v)
given by Eq. (198). The transition of the second type changes In the limit θ → 0, the right-hand side reduces to expression on
only the velocities but in general it does not preserve the kinetic the right-hand side of (165). The entropy ux M associated to
energy. This type of transition is dened by the transitions (202) is obtained by using (12), but it vanishes
identically,
(x,v) → (x,Ci± v), (206)
M = 0. (214)
where Ci± v is a vector with the same components of the vector
v except the i-th component vi which is given by Let us consider now the entropy production rate C
associated to the transitions dened by (206). According to
vi = vi ± a, (207) expression (3), it is given by
where a > 0 is a parameter and each transition occurs with kB   + −
rate C = {Ai (Ci v)P (x,Ci− v) − A−
i (v)P (x,v)}
2 x,v i
αi kB Ti ∓amvi /2kB Ti
A±i (v) = e . (208) A+ − −
i (Ci v)P (x,Ci v)
m2 a 2 × ln . (215)
A−i (v)P (x,v)
The master equation is written as
1 After taking the limit a → 0 this expression is reduced to the

P (x,v) = {P (Hi− x,Hi− v) − P (x,v)} result (177). The corresponding entropy ux C is obtained
∂t i
b from (12) and is given by

{A+ − − − kB   + −
= i (Ci v)P (x,Ci v) − Ai (v)P (x,v)} C = {Ai (Ci v)P (x,Ci− v) − A−i (v)P (x,v)}
i 2 x,v i

+ + +
+m {A−
i (Ci v)P (x,Ci v) − Ai (v)P (x,v)}. A+ −
i (Ci v)
× ln . (216)
i
A−i (v)
(209)
The limit a → 0 leads us to the result (178).
It is straighfoward to show that in the limit a → 0 and b → We now wish to consider the entropy production rate and
0, the master equation reduces to Eq. (174), and the master the ux of entropy coming from the parts of the stochastic
equation indeed can be understood as a representation of the trajectory associated to the Hamiltonian ow, given by the
Fokker-Planck equation (174). transitions dened by (200). We postulate that the entropy
The stationary solution of the master equation when ux associated to the Hamiltonion ow vanishes identically,
all temperatures are the same, which corresponds to the H = 0. Therefore, the entropy production rate associated

042140-16
STOCHASTIC APPROACH TO EQUILIBRIUM AND . . . PHYSICAL REVIEW E 91, 042140 (2015)

to the Hamiltonian ow should be equal to part of dS/dt This required the introduction of the transition rates which thus
coming from the Hamiltonian ow. This part can be obtained play a fundamental role in the present approach, similar to the
by inserting the rst summation of the right-hand side of the Gibbs distribution in the case of equilibrium.
master equation (205) into Eq. (8). After doing this, we get the Using the property that the production of entropy is non-
following expression for the entropy production rate associated negative, which is understood as the dynamic formulation of
to the Hamiltonian ow: the second law of thermodynamics, we were able to show that
 1 P (x,v) in the quasi-static process, the representative point in the ther-
H = k B P (x,v) ln , (217) modynamic space approaches a surface and that this surface
x,v i
b P (H + x,H + v)
has the property of convexity. These statements are usually
which is similar to expression (10). introduced as postulates in equilibrium thermodynamics. We
Next we have to show that H > 0. To this end, we expand have also shown the bilinear form of entropy production,
each term in the summation in powers of b. Up to linear terms which is a sum of terms, each one being a product of a
in b the i element of the summation equals force and a ux. We remark that this is the macroscopic form
( ) ( ) used in nonequilibrium thermodynamics and should not be
∂P Fi ∂P b ∂P Fi ∂P 2 confused with the microscopic denition of entropy production
vi + + vi + . (218)
∂xi m ∂vi 2P ∂xi m ∂vi itself, which looks like a bilinear form. From the bilinear
form of entropy production, we have determined the Onsager
But the integral in x and v of the rst term vanishes so
coefcients and shown that they obey the reciprocal relations.
H > 0. In fact, it vanishes in the continuum limit b → 0.
The nonequilibrium steady states of a system with several
Therefore, in the continuum limit H = 0. From this result
chemical species and chemical reactions were studied by the
it follows that M is the total production of entropy for the
use of appropriate transition rates. From the denition of the
microcanonical case. Since in the continuous limit, it goes
entropy production rate it was possible to derive the bilinear
into (167), it follows that the expression given by (167) is
form, which in this case is written in terms of afnities and
indeed the entropy production rate, as we have assumed.
the rates of the extents of reaction. In equilibrium the afnities
Similarly, it follows that C is the total production rate for
vanish, which is the condition for chemical equilibrium.
the canonical case. In the continuous limit it is identied
Using appropriate transition rates or appropriate stochastic
with (177) so the expression given by (177) is indeed the
noise, in the case of the Fokker-Planck, it was possible to
entropy production rate, as assumed.
study several situations that were analogous to those related
to the microcanonical, canonical, and grand-canonical Gibbs
VI. CONCLUSION ensembles. For the microcanonical case in continuous state
We developed the stochastic approach to thermodynamics space we have introduced an energy-conserving stochastic
based on the stochastic dynamics. More specically, we used noise. For the canonical case we used the usual white Gaussian
the master equation, in the case of discrete state space, and noise. To make contact with the master equation, we have
the Fokker-Planck, in the case of continuous state space. used a master equation representation of the Fokker-Planck.
Our approach is founded on the use of a form for the Using this representation we conrmed the expression for the
production of entropy which is non-negative by denition production of entropy that was introduced by the splitting of
and vanishes in equilibrium. Based on these assumptions we the time derivative of entropy. In this case we postulated that
studied interacting systems with many degrees of freedom in a Hamiltonian transition induces no ux of entropy. Since the
equilibrium or out of thermodynamic equilibrium and how the entropy is constant along a Hamiltonian ow in continuous
macroscopic laws can be derived from the stochastic dynamics. space, this implies no production of entropy.

[1] R. Clausius, Ann. Phys. Chem. 176, 353 (1857). [12] J. S. Thomsen, Phys. Rev. 91, 1263 (1953).
[2] J. C. Maxwell, Philos. Mag. 19, 19 (1860); ,20, 21 (1860); ,20, [13] M. J. Klein and P. H. E. Meijer, Phys. Rev. 96, 250 (1954).
33 (1860). [14] M. J. Klein, Phys. Rev. 97, 1446 (1955).
[3] L. Boltzmann, Wien. Sitzungsber. 66, 275 (1872). [15] T. L. Hill and O. Kedem, J. Theor. Biol. 10, 399 (1966).
[4] L. Boltzmann, Wien. Sitzungsber. 76, 373 (1872). [16] D. A. McQuarrie, J. Appl. Prob. 4, 413 (1967).
[5] J. W. Gibbs, Elementary Principles in Statistical Mechanics [17] J. Schnakenberg, Rev. Mod. Phys. 48, 571 (1976).
(Scribner, New York, 1902). [18] L. Jiu-Li, C. Van den Broeck, and G. Nicolis, Z. Phys. B 56, 165
[6] R. C. Dunbar, J. Chem. Educ. 59, 22 (1982). (1984).
[7] A. Walstad, Am. J. Phys. 81, 555 (2013). [19] C. Y. Mou, J.-L. Luo, and G. Nicolis, J. Chem. Phys. 84, 7011
[8] H. G. Schuster, Deterministic Chaos (Physik Verlag, Weinheim, (1986).
1984). [20] D. T. Gillespie, Physica A 188, 404 (1992).
[9] G. Nicolis and C. Nicolis, Phys. Rev. A 38, 427 (1988). [21] A. Pérez-Madrid, J. R. Rubı́, and P. Mazur, Physica A 212, 231
[10] I. G. Sinai, Introduction ot Ergodic Theory (Princeton University (1994).
Press, Princeton, NJ, 1977). [22] B. Gaveau and L. S. Schulman, Phys. Lett. A 229, 347 (1997).
[11] B. J. Alder and T. E. Wainwright, J. Chem. Phys. 27, 1208 [23] K. Sekimoto, Prog. Theor. Phys. Suppl. 130, 17 (1998).
(1957). [24] T. Tomé and M. J. de Oliveira, Braz. J. Phys. 27, 525 (1997).

042140-17
TÂNIA TOMÉ AND MÁRIO J. DE OLIVEIRA PHYSICAL REVIEW E 91, 042140 (2015)

[25] P. Mazur, Physica A 274, 491 (1999). [63] M. Smoluchowski, Ann. Phys. 353, 1103 (1915).
[26] L. Crochik and T. Tomé, Phys. Rev. E 72, 057103 (2005). [64] M. Planck, Sitz. Knig. Preuss. Akad. Wiss. 324 (1917).
[27] R. K. P. Zia and B. Schmittmann, J. Phys. A: Math. Gen. 39, [65] L. S. Ornstein, KNAW, Proceedings, Amsterdam 21, 96 (1919).
L407 (2006). [66] H. A. Kramers, Physica 7, 284 (1940).
[28] D. Andrieux and P. Gaspard, Phys. Rev. E 74, 011906 (2006). [67] N. G. van Kampen, Stochastic Processes in Physics and
[29] T. Tomé, Braz. J. Phys. 36, 1285 (2006). Chemistry (North-Holland, Amsterdam, 1981).
[30] T. Schmiedl and U. Seifert, J. Chem. Phys. 126, 044101 (2007). [68] C. W. Gardiner, Handbook of Stochastic Methods for Physics,
[31] R. J. Harris and G. M. Schütz, J. Stat. Mech. (2007) P07020. Chemistry and Natural Sciences (Springer, Berlin, 1983).
[32] R. K. P. Zia and B. Schmittmann, J. Stat. Mech. (2007) P07012. [69] H. Risken, The Fokker-Planck Equation, Methods of Solution
[33] U. Seifert, Eur. Phys. J. B 64, 423 (2008). and Applications (Springer, Berlin, 1984).
[34] R. A. Blythe, Phys. Rev. Lett. 100, 010601 (2008). [70] T. Tomé and M. J. de Oliveira, Stochastic Dynamics and
[35] B. Gaveau, M. Moreau, and L. S. Schulman, Phys. Rev. E 79, Irreversibility (Springer, Heidelberg, 2015).
010102 (2009). [71] N. T. J. Bailey, The Mathematical Theory of Epidemics (Hafner,
[36] B. Gaveau and L. S. Schulman, Phys. Rev. E 79, 021112 (2009). New York, 1957).
[37] M. Esposito, K. Lindenberg, and C. Van den Broeck, Phys. Rev. [72] M. S. Bartlett, Stochastic Population Models in Ecology and
Lett. 102, 130602 (2009). Epidemiology (Methuen, London, 1960).
[38] T. Tomé and M. J. de Oliveira, Phys. Rev. E 82, 021120 (2010). [73] R. M. Nisbet and W. C. S. Gurney, Modelling Fluctuating
[39] C. Van den Broeck and M. Esposito, Phys. Rev. E 82, 011144 Populations (Wiley, New York, 1982).
(2010). [74] T. L. Hill, Free Energy Transduction and Biochemical Cycle
[40] T. Tomé and M. J. de Oliveira, Phys. Rev. Lett. 108, 020601 Kinetics (Springer, New York, 1989).
(2012). [75] P. Ao, Commun. Theor. Phys. 49, 1073 (2008).
[41] R. E. Spinney and I. J. Ford, Phys. Rev. E 85, 051113 (2012). [76] G. Lan, P. Sartori, S. Neumann, V. Sourjik, and Y. Tu, Nat. Phys.
[42] M. Esposito, Phys. Rev. E 85, 041125 (2012). 8, 422 (2012).
[43] F. Zhang, L. Xu, K. Zhang, E. Wang, and J. Wang, J. Chem. [77] H. Berry, Phys. Rev. E 67, 031907 (2003).
Phys. 137, 065102 (2012). [78] J. L. England, J. Chem. Phys. 139, 121923 (2013).
[44] U. Seifert, Rep. Prog. Phys. 75, 126001 (2012). [79] R. J. Glauber, J. Math. Phys. 4, 294 (1963).
[45] X.-J. Zhang, H. Qian, and M. Qian, Phys. Rep. 510, 1 (2012). [80] T. E. Harris, Ann. Probab. 2, 969 (1974).
[46] H. Ge, M. Qian, and H. Qian, Phys. Rep. 510, 87 (2012). [81] T. M. Ligget, Interacting Particle Systems (Springer, New York,
[47] M. Santillan and H. Qian, Physica A 392, 123 (2013). 1985).
[48] D. Luposchainsky and H. Hinrichsen, J. Stat. Phys. 153, 828 [82] R. M. Ziff, E. Gulari, and Y. Barshad, Phys. Rev. Lett. 56, 2553
(2013). (1986).
[49] W. Wu and J. Wang, J. Chem. Phys. 141, 105104 (2014). [83] J. E. Satulovsky and T. Tomé, Phys. Rev. E 49, 5073 (1994).
[50] H. B. Callen, Thermodynamics (Wiley, New York, 1960). [84] M. J. de Oliveira, J. Stat. Phys. 66, 273 (1992).
[51] M. J. de Oliveira, Equilibrium Thermodynamics (Springer, [85] J. Marro and R. Dickman, Nonequilibrium Phase Transitions in
Heidelberg, 2013). Lattice Models (Cambridge University Press, Cambridge, UK,
[52] T. De Donder, L’Afnité (Lamertin, Bruxelles, 1927). 1999).
[53] L. Onsager, Phys. Rev. 37, 405 (1931); ,38, 2265 (1931). [86] M. J. de Oliveira, Phys. Rev. E 67, 066101 (2003).
[54] K. G. Denbigh, The Thermodynamics of the Steady State [87] R. Clausius, Ann. Phys. Chem. 201, 353 (1865).
(Methuen, London, 1951). [88] H. R. Künsch, Z. Wahrscheinlichkeit. 66, 407 (1984).
[55] I. Prigogine, Introduction to Thermodynamics of Irreversible [89] C. Godrèche, J. Stat. Mech. (2011) P04005.
Processes (Thomas, Springeld, 1955). [90] M. J. de Oliveira, J. Stat. Mech. (2011) P12012.
[56] S. R. de Groot and P. Mazur, Non-Equilibrium Thermodynamics [91] S. A. Arrhenius, Z. Physik. Chem. 4, 96 (1889); ,4, 226
(North-Holland, Amsterdam, 1962). (1889).
[57] P. Glansdorff and I. Prigogine, Thermodynamics of Structure, [92] W. J. Moore, Physical Chemistry (Longman, London, 1965).
Stability and Fluctuations (Wiley, New York, 1971). [93] G. N. Lewis and M. Randall, Thermodynamics and the Free
[58] G. Nicolis and I. Prigogine, Self-Organization in Nonequilib- Energy of Chemical Substances (McGraw-Hill, New York,
rium Systems (Wiley, New York, 1977). 1923).
[59] A. Einstein, Ann. Phys. 322, 549 (1905). [94] D. S. Lemons, An Introduction to Stochastic Processes in
[60] M. Smoluchowski, Ann. Phys. 326, 756 (1906). Physics (John Hopkins University Press, Baltimore, 2002).
[61] P. Langevin, Comp. Rend. 146, 530 (1908). [95] G. T. Landi and M. J. de Oliveira, Phys. Rev. E 89, 022105
[62] A. D. Fokker, Ann. Phys. 348, 810 (1914). (2014).

042140-18

Vous aimerez peut-être aussi