Download as pdf or txt
Download as pdf or txt
You are on page 1of 11

Current fluctuations in nonequilibrium discontinuous phase transitions

C. E. Fiore,1, ∗ Pedro E. Harunari,1, 2, † C. E. Fernández Noa,1 and Gabriel T. Landi1, ‡


1
Instituto de Fı́sica da Universidade de São Paulo, 05314-970 São Paulo, Brazil
2
Complex Systems and Statistical Mechanics, Physics and Materials Science Research Unit,
University of Luxembourg, Luxembourg L-1511 G.D. Luxembourg
(Dated: September 2, 2021)
Discontinuous phase transitions out of equilibrium can be characterized by the behavior of macroscopic
stochastic currents. But while much is known about the the average current, the situation is much less understood
for higher statistics. In this paper, we address the consequences of the diverging metastability lifetime – a
arXiv:2109.00385v1 [cond-mat.stat-mech] 1 Sep 2021

hallmark of discontinuous transitions – in the fluctuations of arbitrary thermodynamic currents, including the
entropy production. In particular, we center our discussion on the conditional statistics, given which phase the
system is in. We highlight the interplay between integration window and metastability lifetime, which is not
manifested in the average current, but strongly influences the fluctuations. We introduce conditional currents
and find, among other predictions, their connection to average and scaled variance through a finite-time version
of Large Deviation Theory and a minimal model. Our results are then further verified in two paradigmatic
models of discontinuous transitions: Schlögl’s model of chemical reactions, and a 12-states Potts model subject
to two baths at different temperatures.

I. INTRODUCTION Conversely, the behavior of higher order statistics, such as the


variance, is much less understood.
In microscopic systems, currents of heat, work and entropy Cumulants of thermodynamic currents are usually assessed
production must be treated as random variables, which fluc- via numerical approaches, such as Monte Carlo simula-
tuate over different runs of an experiment [1, 2]. This rep- tions [40], or large deviation theory (LDT) [7, 44–48]. In
resents a paradigm shift in thermodynamics, and has already both cases, cumulants are computed from long-time sample
led to fundamental advancements in the field, such as fluc- averages, integrated over a time window τ. Ultimately, one
tuation theorems [3–8] and, more recently, the discovery of is interested in taking τ → ∞, at least in principle. But in
thermodynamic uncertainty relations [9–13]. It also entails systems presenting discontinuous transitions this can become
practical consequences, e.g. in the design of Brownian en- an issue, since the phase coexistence is characterized by states
gines [14–17], molecular motors [18–21], information-driven with very long metastability lifetimes τm . In fact, τm increases
devices [22, 23], and bacterial baths [24]. In these systems, exponentially with the system volume V. As a consequence,
both the output power [25, 26] and the efficiency [27–30] may the order of the limits τ → ∞ and V → ∞ becomes non-
fluctuate significantly, leading to possible violations of macro- trivial [49].
scopic predictions, such as the Carnot limit. [14]. In this paper we approach this issue by introducing the idea
A scenario of particular interest is that of non-equilibrium of conditional currents, given which phase the system is in.
steady-states (NESSs), which occur when a system is placed We focus, in particular, on the diffusion coefficient (scaled
in contact with multiple reservoirs at different temperatures variance). We formulate a finite-time large deviation theory,
T i and/or chemical potentials µi . NESSs are characterized by which neatly highlights the non-trivial interplay between τ
finite currents of energy and matter, and thus also a finite en- and τm . This is then specialized to a minimal 2-state model,
tropy production rate σt [1, 31–34]. At the stochastic level, that is able to capture the key features of the problem and
these become fluctuating quantities, associated to a probabil- also provides useful predictions. These are then tested on two
ity distribution. Understanding the behavior of said distribu- paradigmatic examples of discontinuous transitions: Schlögl’s
tions constitutes a major area of research, as they form the model of chemical kinetics, and a 12-states Potts model sub-
basis for extending the laws of the thermodynamics towards ject to two baths at different temperatures.
the microscale, providing insights in non-trivial properties of This paper is organized as follows: Sec. II presents the main
non-equilibrium physics. Of particular interest is their behav- concepts and assumptions considered. The conditional large
ior across non-equilibrium phase transitions [35]. Most of our deviation theory is developed in Sec. III and then specialized
understanding, however, is centered on the average current. to a minimal model in Sec. IV. Applications are then consid-
For instance, the average entropy production rate has been ered in Sec. V and our conclusions are summarized in Sec. VI.
found to be always finite around the transition point, with the
first derivative either diverging, in continuous transitions [36–
42], or presenting a jump in discontinuous ones [39, 40, 43].
II. MODELS AND ASSUMPTIONS

We consider a stochastic system X(t) undergoing Marko-


∗ fiorecarlos.cf@gmail.com vian evolution. For simplicity, we assume continuous-time
† pedroharunari@gmail.com and a discrete (possibly infinite) set of states X(t) ∈ S. The
‡ gtlandi@gmail.com system probability p x (t) is assumed to evolve according to the
2

master equation [50] The main feature we introduce in this paper is the notion of
X conditional currents, given which phase i = 0, 1 the system is
X
ṗ x (t) = W xy py − Wyx p x := W xy py , (1) in. Inserting the identity 1 = (1 − It ) + It inside the integral (2)
y y allows us to define the current when the system is in phase 1
as
where W xy ≡ PWy→x denotes the transition rates from y to x
and W xx ≡ − y,x Wyx . The dynamics is taken to be ergodic, Zτ X
and such that W xy > 0 whenever Wyx > 0, ensuring the system Jτ|1 = dt It+ dyz δX(t− ),y δX(t+ ),z . (4)
will relax to a unique steady state p∗x . In general, p∗x will be a 0
y,z

non-equilibrium steady-state (NESS).


This NESS is also assumed to undergo a discontinuous tran- The current Jτ|0 is defined similarly, but with 1 − It instead.
sition by changing a certain control parameter λ to a threshold There is an ambiguity here as to whether we use It− or It+ . But
value λc . This means that in the vicinity of λc , there will exist a this only affects those jumps in which It− = 0(1) and It+ =
bistable region characterized by configurations with very long 1(0), which are extremely rare compared to all others. The
lifetimes. The two phases are labeled as 0 (for λ < λc ) and 1 total current (2) is then recovered as
(for λ > λc ). We monitor the phases by defining an indicator Jτ = Jτ|0 + Jτ|1 , (5)
random variable It = 0, 1 (henceforth called the phase indi-
cator), which specifies in which phase the system is at time t. an identity which holds at the stochastic level.
This can always be done by partitioning the set of states S into The conditional first moments are defined as
two subsets, S0 and S1 , representing each phase. The criteria
E(Jτ|i )
for doing so is model dependent, and will be discussed further µi = , (6)
below. The probability of finding the system in phase 1, in the τqi
NESS, is then q ≡ E(It ) = Pr(It = 1). We will also use the where the factor of qi in the denominator is placed to compen-
notation q1 = q and q0 = 1 − q, when convenient. sate for the varying times the system spends in each phase.
The crucial aspect of discontinuous transitions is that, when The average current is thus decomposed as
the volume V is large, transitions between coexisting phases
become extremely rare. The system will thus be governed J = (1 − q)µ0 + qµ1 . (7)
by two very distinct timescales, one describing fast relax-
ation within each phase and another describing seldom tran- As with J, the conditional averages µi will be shown below to
sitions between the phases. The latter will be referred to as also be independent of τ.
the metastability lifetime τm , and usually grows exponentially Similarly, we define conditional diffusion coefficients
with V [51]. 2
E(Jτ|i ) − E(Jτ|i )2
We consider the consequences of this type of scaling to Dτ|i = , (8)
the behavior of a generic integrated thermodynamic current. 2τqi
Given a certain time integration window τ, such a current may
which represent the fluctuations of the system within each
be defined as [11]
phase. From Eq. (5), we therefore see that the diffusion co-
Zτ efficient Dτ in Eq. (3) is split in three terms
X
Jτ = dt dyz δX(t− ),y δX(t+ ),z , (2) 1
Dτ = (1−q)Dτ|0 +qDτ|1 +Cτ , cov Jτ|0 , Jτ|1 , (9)

y,z Cτ :=
0 τ
where δi j is the Kronecker delta, X(t± ) is the state of the sys- where cov(A, B) = E(AB) − E(A)E(B) is the covariance be-
tem immediately before and after a transition and dyz is a func- tween conditional currents A and B, and is expected to be sig-
tion satisfying dyz = −dzy , which defines the current in ques- nificant only in the vicinity of the transition point.
tion. In the limit τ → ∞, such a current will behave according
to a large deviation principle [47]. But due to the sensitive
interplay between τ and τm , we will not assume τ → ∞, as is III. LARGE DEVIATION THEORY
customary. Instead, we will analyze the behavior of Jτ as a
function of τ. More specifically, our interest is in the regime To shed light on the behaviour of conditional currents, we
where τ is large compared to the “within-phase” timescales, consider here a finite-time version of large deviation the-
but not necessarily larger than the metastability lifetime τm . ory [47]. We being with the unconditional quantities, and then
We will also focus on both the average Jτ , and diffusion coef- adapt our results to the conditional case. Let Gτ (η) = E(eηJτ )
ficient (scaled variance) Dτ , defined as denote the moment generating function (MGF) P associated to
  the currentP(2). Decomposing it as Gτ (η) = x E(eηJτ |Xτ =
Jτ = E(Jτ )/τ, Dτ = E(Jτ2 ) − E(Jτ )2 /(2τ). (3) x)p x (τ) = x G x (η), we find that the entries G x (η) will evolve
according to equation
It turns out that Jτ ≡ J is independent of τ, irrespective of
whether τ is large or not [47]. Conversely, for Dτ , this will be dG x (η) X
= L xy (η)Gy (η), (10)
the case iff τ  τm . dτ y
3

where the tilted operator L(η) depends on both the transition on the interplay between τ and all eigenvalues λi of W. If
matrix W in Eq. (1), and the type of current in question, ac- τ  1/|λi |, for all eigenvalues λi , 0, then the term eλi τ − 1
cording to may be neglected, leading to the widely used expression from
large deviation
L(η) xy = eηdxy W xy , (11)

where, recall, d xx = 0. To evaluate J and Dτ , we only require Dτ = h1|L2 |pi − h1|L1 W+ L1 |pi, (16)
the series expansion of L(η), which we write as L(η) = W +
ηL1 + η2 L2 , for matrices L1(2) given by
where W+ = i,0 λ−1
P
i |xi ihyi | is the Moore-Penrose pseudoin-
verse of W (see Appendix A for more details). Close to the
(L1 ) xy = W xy d xy , (L2 ) xy = W xy d2xy /2. (12)
transition point, there will appear a clear separation of time
scales in the eigenvalues λi . At least one eigenvalue will be
A. Unconditional cumulants very small, of the order λi ∼ −1/τm , while all others will be
much larger (describing the within-phase dynamics). If τ is
large compared to these time scales, but not with respect to
We denote by |pi the column vector whose entries are the τm , then the approximation taking Eq. (15) to (16) will not
steady-state distribution p∗x , and h1| the row vector with all hold true. And since τm scales exponentially with the volume,
entries equal 1. Then, as discussed further in Appendix A, the as we approach the thermodynamic limit, larger and larger
first moment can be written, for arbitrary τ, as values of τ have to be considered. This is a direct illustration
of the non-commutativity of the limits τ → ∞ and V → ∞.
Jτ ≡ J = h1|L1 |pi, (13)
B. Conditional cumulants
which is independent of τ, as expected. Conversely, the diffu-
sion coefficient is written as Eqs. (13) and (14) also apply to the conditional currents (4).
Zτ Zτ 0 One simply has to modify accordingly the tilted operator L(η)
1 0 00 J2τ or, what is equivalent, the matrices L1 and L2 in Eq. (12). For
Dτ = h1|L2 |pi + dτ 0
dτ00 h1|L1 eW(τ −τ ) L1 |pi − .(14)
τ 2 each conditional currentPJτ|i , we define a projection operator
0 0 Πi such that Π1xy = δ x,y z∈S i δy,z ; i.e., which projects onto the
We can also obtain a more explicit expression if we assume states Si associated to phase i = 0, 1. The corresponding tilted
that W is diagonalizable, with eigenvalues λi , right eigenvec- operator will then be defined similarly, but with a current of
tors W|xi i = λi |xi i and left eigenvectors hyi |W = hyi |λi . Since the form dixy = d xy Πiyy , which means one should use instead
the steady-state is unique, one eigenvalue must be zero, say matrices L1 Πi and L2 Πi .
λ0 = 0. The corresponding eigenvectors are then |x0 i = |pi Eq. (13) then yields, taking also into account the factor qi
and hy0 | = h1|. Carrying out the integrals one then finds that in the denominator,
 λτ
 e i − 1 − λi τ 
X 
Dτ = h1|L2 |pi + h1|L1 |xi ihyi |L1 |pi   , (15) 1
i,0
λ2i τ µi = h1|L1 Πi |pi. (17)
qi
where we used the orthogonality relation h1|xi i = 0, for i , 0.
This expression makes it clear that Dτ will depend sensibly Proceeding similarly with Eq. (14), we find

Zτ Zτ0
h1|L2 Πi |pi 1 0 00 µ2i qi τ
Dτ|i = + dτ 0
dτ00 h1|L1 Πi eW(τ −τ ) L1 Πi |pi − . (18)
qi τqi 2
0 0

And to obtain the covariance in Eq. (9), we simply subtract the combination (1 − q)Dτ|0 + qDτ|1 from Dτ in Eq. (14). Recalling
that Π0 + Π1 = 1, this then yields
Zτ Zτ0 Zτ Zτ0
1 0 00 1 0 00
Cτ = dτ0 dτ00 h1|L1 Π0 eW(τ −τ )
L1 Π1 |pi + dτ0 dτ00 h1|L1 Π1 eW(τ −τ ) L1 Π0 |pi − q(1 − q)µ0 µ1 τ. (19)
τ τ
0 0 0 0

Concerning the timescales of the discontinuous transition, we notice that all diffusion coefficients, Dτ , Dτ|i and Cτ , are sub-
4

1
ject to a similar dependence, which is ultimately associated 1 4
0 (a) ( b)
with the matrix eW(τ−τ ) . Thus, one expects that all quantities
should scale similarly with τ.

q( 1- q)
1 1

q
2 8

C. Conditioning on the dynamics


0 λ 0 λ
λc λc
There is a subtle, but crucial difference between condition-
105 (d)
ing the currents and conditioning the dynamics. Eq. (4) is an 0.8 (c)
instance of the former: the current is conditioned on which
0.6 104
phase the system is in, but X(t) is still free to jump from one τ ≪ τm
phase to the other. Alternatively, one could define a condi- 0.4 1000 τ ≫ τm
tional dynamics, where the system is forced to remain only 0.2
within a certain phase. This could be accomplished, for in- 100
stance, by splitting the transition matrix W in Eq. (1) in blocks 0 5 10 15 20 0 5 10 15 20
of the form
τ/τm V
 
W00 W01 
W =   , (20) FIG. 1. Predictions of the minimal model of discontinuous transi-
W10 W11 tions. (a) The probability q = (1 + e−cV∆λ )−1 of finding the system in
phase 1, for increasing volumes (depicted by the arrow). (b) q(1 − q),
which is non-negligible only in the vicinity of the transition point.
referring to the two subsets S0 and S1 of each phase. A con-
(c) The quantity (e−τ/τm − 1 + τ/τm )/(τ/τm ) appearing in Eq. (29). It
ditional dynamics, given phase i, is one that is governed by tends to unity when τ  τm . (d) Prototypical behavior of the dif-
the restricted matrix Wii (with appropriate adjustments at the fusion coefficient (29) as a function of volume, for a fixed τ. When
boundaries to ensure that it remains a proper transition ma- V is such that τ  τm , the diffusion coefficient grows exponentially
trix). with V. But for a fixed τ, as V is increased, one must eventually cross
One can similarly adapt Eqs. (13) and (14) to this case. Let the point τ ∼ τm , after which the scaling becomes at most polyno-
|pi i denote the steady-state of Wii . For large volumes, since mial (due to the possible dependences of µi , Di on V). Parameters:
the two phases will be well separated, this will be quite similar c0 = ca = cb = λc = 1, µ0 = V/2, µ1 = 2V, γ0 = γ1 = V.
to q1i Πi |pi. Applying Eq. (13) will then yields exactly the same
first moment µi in Eq. (17). Hence, as far as the first moments
The corresponding second moment will thus be
are concerned, the distinction between conditional currents
and conditional dynamics is thus irrelevant. Zτ Zτ
However, for the diffusion coefficients this is absolutely 2
E(Jτ|1 ) = dt dt0 E(It It0 Zt Zt0 ). (22)
crucial. The reason is associated with the matrix exponen- 0 0
0 00
tial eW(τ −τ ) in Eq. (14). Conditioning on the dynamics
0 00
would lead instead to a matrix eWii (τ −τ ) . Since Wii is es- It hence depends, among other things, on the correlations be-
sentially Π WΠ (except for small modifications at the bound-
i i tween It and It0 , which decays very slowly around the transi-
aries), we therefore see that the problem amounts to the differ- tion point. For instance, in the simplest case where one can
0 00
ence between Πi eW(τ −τ ) Πi (conditioning on the currents) and assume a Markovian 2-state evolution for It (as will in fact be
eΠ WΠ (τ −τ ) (conditioning on the dynamics). The two objects
i i 0 00
considered further in Sec. IV), one has
are drastically different. The diffusion coefficients obtained 0

by conditioning the dynamics, which we shall henceforth refer C(t − t0 ) = cov(It , It0 ) = q(1 − q)e−(t−t )/τm , (23)
to as γτ|i , will thus fundamentally different from the diffusion
which will thus decay very slowly in time. This means that
coefficients Dτ|i in Eq. (8).
Dτ|i in Eq. (8) will depend very sensibly on the interplay be-
An intuitive argument as to why this is the case goes as tween τ and τm . Conversely, the diffusion coefficients γi , for
follows. The currents (4) are integrated over a certain time the conditional dynamics, will not. And hence, even for mod-
interval τ. Hence, its diffusion coefficient will depend on erately large τ, one expects it to be τ-independent.
correlations between different instants of time, and these are
dramatically affected by the long timescale τm introduced
by the discontinuous transition. In fact, let us define Zt = IV. MINIMAL MODEL
y,z dyz δX(t− ),y δX(t+ ),z , so that Eq. (4) can be written as
P

Many discontinuous non-equilibrium transitions can be ap-


Zτ proximated, for large volumes V, by a 2-state model [51].
Jτ|1 = dt It Zt . (21) That is, one reduces the dynamics essentially to the monitor-
0 ing of the phase indicator It . In general, the dynamics of It will
5

be non-Markovian, as this would represent a hidden Markov where γ = (1 − q)γ0 + qγ1 is independent of τ and
chain. Instead, a minimal model is one where the dynamics
of It can be assumed to be Markovian, which is justified when f (t) = (e−t − 1 + t)/t. (30)
V is sufficiently large. In this case, instead of the full master
equation (1), we may restrict the dynamics to The interesting part is the last term in Eq. (29). First, it de-
pends on q(1 − q), which is non-negligible only in the vicinity
of the transition point (Fig. 1(b)). Second, it depends on the
 
d X −a b 
qi = Wi j q j , W =   . (24) interplay between τ and τm through the function f , which is
dt a −b
j=0,1 shown in Fig. 1(c).
When τ  τm we get f (τ/τm ) ' τ/2τm , so that Eq. (29) can
Here a and b represent the rates for the system to jump from be approximated to
phase 0 → 1 and 1 → 0. The steady-state yields q ≡ q1 =
E(It ) = a/(a + b). Moreover, the metastability lifetime in this Dτ ' γ + q(1 − q)(µ1 − µ0 )2 τ/2, τ  τm , (31)
case reads τm = 1/(a + b). Finally, from (24) one can compute
the two-time correlation function, which is given in Eq. (23). which is thus linear in τ. Conversely, when τ  τm , we get
And since It can take on only two values, once C(t − t0 ) is
known we can reconstruct the full joint distribution Pr(It = Dτ ' γ + q(1 − q)(µ1 − µ0 )2 τm , τ  τm , (32)
i, It0 = i0 ), for arbitrary times t, t0 :
which is independent of τ, but linear in τm . Hence, when V
q2 + C(t − t0 ) i = i0 = 1,



 is large, this will become exponentially dominant. As a con-


 sequence, the large volume diffusion coefficient will actually
Pr(It = i, It0 = i0 ) = (1 − q)2 + C(t − t0 ) i = i0 = 0, (25)

become independent of the γi , and will instead be governed


essentially by the mismatch in conditional averages (µ1 −µ0 )2 ,



q(1 − q) − C(t − t0 ) i , i0 .


in agreement with previous studies on Schlögl’s model [44].
This offers another explicit illustration of the order of limits
The key feature of discontinuous transitions is the fact that
issue, which we depict graphically in Fig. 1(d): For a given τ,
transitions between phases are seldom when V is large. Close
as we increase V the diffusion coefficient will at first increase
to λc , the transition rates a and b will usually behave, up to
exponentially according to Eq. (32). But if τ is fixed, then a
polynomial corrections, as
point will always be reached around which τ ∼ τm . And be-
a ∼ e−V(c0 −ca ∆λ) , b ∼ e−V(c0 +cb ∆λ) , (26) yond this point, the scaling will be given by Eq. (31), which
is at most polynomial in V (due to a potential polynomial vol-
where c0 , ca , cb > 0 are constants and ∆λ = λ − λc . Note ume dependence of µi , γi ).
how the rates are exponentially decreasing with V. Transi- Even though these results were developed for a 2-level
tions hence become rare when V is large. From (26) we also model, they are still expected to hold for a broad class of
get τm ∼ ec0 V , which is the aforementioned exponential de- discontinuous transitions. The reason is that, as discussed in
pendence. Finally, q = (1 + e−cV∆λ )−1 , where c = ca + cb > 0; Ref. [53], the eigenvalues and eigenvectors of the two-level
hence q changes abruptly from 0 to 1 as λ crosses λc , as illus- transition matrix (24) are connected to some of the eigenval-
trated in Fig. 1(a). Since the conditional averages are weakly ues and eigenvectors of the full matrix W in Eq. (1). But, in
dependent on ∆λ, from Eq. (7) we therefore see that J should addition, the full W will also have several other eigenvalues
also change abruptly around λc , interpolating from µ0 to µ1 . associated to the within-phase dynamics. Thus, the step from
Eq. (15) to (29) only assumes that τ is much larger than all
other λi , so that within-phase terms can be neglected.
A. Unconditional diffusion coefficient

B. Conditional diffusion coefficients


As shown in [52], in this two-level model the tilted operator
can be written, up to order λ2 , as
We can also use this minimal model to relate the diffusion
−a + λµ0 + λ2 γ0 coefficients Dτ|i in Eq. (8) with the parameters µi , γi . To do so,
 
b
L(λ) = 

 (27) we use Eq. (18) with W now replaced by the two-state matrix
a −b + λµ1 + λ2 γ1 W in Eq. (24). As a result, we find
:= W + λL1 + λ2 L2 . (28) Dτ|1 = γ1 + µ21 (1 − q)τm f (τ/τm ), (33)
where γi are the diffusion coefficients conditioned on the dy- Dτ|0 = γ0 + µ20 qτm f (τ/τm ), (34)
namics, not the currents (as introduced in Sec. III C).
For the matrix W defined in Eq. (24) we have λ1 = −1/τm , Cτ = −2q(1 − q)µ0 µ1 τm f (τ/τm ), (35)
|pi = (1 − q, q), |x1 i = (−1, 1) and |y1 i = (−q, 1 − q). Hence,
using the explicit forms of L1 and L2 in Eq. (28), we get which can be combined together in the form (9), to yield
Eq. (29). All conditional quantities are thus found to scale
Dτ = γ + q(1 − q)(µ1 − µ0 )2 τm f (τ/τm ), (29) similarly with τ, according to the function f in Eq. (30). This
6

allows us to conclude that even the conditional diffusion coef- The bistable region is defined as the interval in the control pa-
ficients will be dominated by jumps between phases, and will rameters for which this equation has three real roots, x0 , x∗ , x1 .
be negligibly affected by the internal fluctuations within each The first and last represent stable states for the most likely
phase. We find this result both relevant and non-trivial. density within each phase, whereas x∗ is unstable and serves
It is also interesting to notice how the sign of the covari- as the phase separator. Hence, we define the phase-indicator
ance (35) depends only on the signs of µ0 and µ1 . A positively in the Schlögl’s model as a random variable It such that It = 1
correlated covariance means that fluctuations above (below) when X(t) > V x∗ and 0 otherwise.]
average in one phase tend to lead to fluctuations above (be- For concreteness, we choose as thermodynamic current the
low) the average in the other; and vice-versa for C < 0. We entropy production Jτ = στ . Whenever there is a transition,
see in Eq. (35) that the covariance will be negative whenever the net current (2) changes by an increment δστ defined ac-
µ0 , µ1 have the same sign. cording to the following rules:
k1
2X + A −−−−→ 3X δστ = µA ,
V. APPLICATIONS
k−1
3X −−−−−→ 2X + A δστ = −µA ,
Next we shall exemplify our main findings in two represen- (41)
k−2
tative systems displaying discontinuous phase transitions: The X −−−−−→ B δστ = µB ,
second Schlögl [54] and a 12-state Potts models connected to
k2
two baths at different temperatures. The former was recently B −−−−→ X δστ = −µB ,
analyzed in Ref. [44]. It represents an ideal laboratory for test-
ing our main prescriptions, since it presents an exact solution. where µA = ln ak1 /k−1 and µB = ln k−2 /bk2 .
The Potts model, on the other hand, is defined in a regular The model was simulated using the Gillespie algorithm.
lattice and exhibits a nonequilibrium phase transition under a We fix ak1 = k−2 = 1, bk2 = 0.2, and take as control pa-
different mechanism. Despite the absence of an exact solu- rameter the chemical potential gradient ∆µ = µB − µA =
tion, all main features about the phase transition and statistics ln [(k−2 ak1 )/(k−1 bk2 )]. For these parameters, the phase co-
about entropy production fluctuations are present. existence point in the thermodynamic limit occurs at ∆µ0 ∼
3.047 [44]. Figs. 2(a) and (b) present a basic characterization
of the steady-state. First, in Fig. 2(a) we show the numerically
A. Schlögl’s model computed metastability timescale τm , as a function of the vol-
ume V, confirming the exponential dependence with V. This
The second Schlögl model [54] describes a system with 3 is obtained by collecting the mean first passage time T xi →x∗
chemical species, A, B and X, supporting two types of chemi- for the system to go from each stable point x0(1) to the unsta-
cal reactions: ble point, x∗ . The rates a and b in Eq. (24) are then given by
k1 k2 a = (2T x0 →x∗ )−1 and b = (2T x1 →x∗ )−1 [56], from which we de-
−*
2X + A −
)−− 3X, −*
)
B−−− X. (36) termine τm = 1/(a + b). Second, Fig. 2(b) characterizes the
k−1 k−2
probability q of finding the system in phase 1, as a function of
Here k±1 , k±2 are kinetic constants that account, respectively, ∆µ − ∆µ0 , for different values of V, where markers are simu-
for catalytic, spontaneous creation and spontaneous annihi- lation data and the curves are a fit of q = (1 + e−cV(∆µ−∆µ0 ) )−1 ;
lation of X. The concentrations of A and B are fixed at a both agree very well for large volumes and/or small ∆µ − ∆µ0 .
and b due to the presence of chemostats. The dynamics of This is expected, since Schlögl’s model is known to have a
pn (t) = P(X(t) = n), for n = 0, 1, 2, . . ., is then described by well defined 2-state limit [53, 55] when V is large.
the master equation [53, 55] Sample stochastic trajectories of the current Jτ [Eq. (2)] as
a function of τ are shown in Fig. 2(c), for fixed ∆µ = 3.35 and
ṗn = fn−1 pn−1 + gn+1 pn+1 − ( fn + gn )pn , (37) V = 10. Red and blue curves represent the situations where
the system start in phases 1 and 0 respectively. For short τ the
where curves tend to remain well separated, so that Jτ behaves as
ak1 n(n − 1) either Jτ|1 or Jτ|0 . The corresponding statistics of Jτ , shown
fn B + bk2 V, (38) in the inset, would thus be a prototypical bimodal distribu-
V
tion. Conversely, when τ  τm ∼ 40, transitions between the
k−1 n(n − 1)(n − 2) phases begin to occur, which cause the corresponding distri-
gn B + k−2 n. (39)
V2 bution to change to unimodal.
The conditional mean current and diffusion coefficients are
The concentration x(t) = X(t)/V presents a bistable behavior
shown in Fig. 2(d)-(g). For concreteness, we focus on the
for large V [53], which is determined by the roots of the differ-
special point q = 1/2; i.e., where the two phases are equally
ential equation governing the deterministic behavior of x for
likely. As this depends on V, for each volume we first fix ∆µ
large volumes
as the point where q = 1/2. This reduces the free parameters
dx to V and τ only. The conditional averages µ0(1) as a function
= ak1 x2 + bk2 − k−1 x3 − k−2 x = 0. (40) of the volume are shown in Fig. 2(a). They are both found to
dt
7

104 300
(a) ( b) (c) τ = 15
5000 0.85
250

0.65 200 τ = 950


1000

τ /τ
500
τm

150

q
0.45 V=5 μ1

V = 15 100
100 Jτ
0.25 V = 25
50 50
V = 40 μ0
0.05 0
0 10 20 30 40 50 60 70 - 0.1 0 0.1 1 5 10 50 100 500 1000
V Δμ-Δμ0 τ
12 19.
15 (d ) 549. 17 (e) (f ) 1.3 ( g) τ = 10 1.3
16 10 17.4
τ = 200
15 8 15.9 1.2 1.2
10 366.

ln( Dτ 0)

ln( Dτ 1)
τ = 1000
ln( Dτ)

14
μ0

μ1

r
6 14.4
13 1.1 1.1
5 183. 4 12.8
12
11 2 11.3
1. 1.
0 0. 10 0 9.77 0 10 20 30 40 50 60 70
0 10 20 30 40 50 60 70 0 10 20 30 40 50 60 70 0 10 20 30 40 50 60 70
V V V V

FIG. 2. Conditional currents for Schlögl’s model, computed using the Gillespie algorithm. (a) τm vs.V. (b) q vs. ∆µ − ∆µ0 for different values
of V. Solid lines are a fit of q = (1 + e−cV(∆µ−∆µ0 ) )−1 . (c) Stochastic trajectories of Jτ /τ vs. τ, starting either in phase 1 (red) or phase 0 (blue).
The insets show the corresponding histograms at different times τ. (d)-(g) Mean and diffusion coefficient as a function of V, with τ = 103 and
∆µ fixed by setting q = 1/2. (d) Conditional means µ0(1) [Eq. (6)]. (e) Diffusion coefficient Dτ [Eq. (3)]. (f) Conditional diffusion coefficients
Dτ|i [Eq. (8)]. (g) The ratio r in Eq. (42), as a function of V, for different values of τ. Other parameters: a = k1 = k2 = k−2 = 1 and b = 0.2.

H(s) = − Vi=1 zδ=1 δ si ,si+δ , where s = (s1 , . . . , sV ). The


P P
be extensive in V, as expected; moreover, the activity in phase
1 is generally much larger, causing µ1  µ0 . equilibrium properties of this model have been studied ex-
Conversely, the diffusion coefficient Dτ (Fig. 2(e)) and their tensively in [57–61]. Here, we consider a non-equilibrium
conditional counterparts Dτ|i (Fig. 2(f)) are both exponential version where the even and odd sites of the lattice are cou-
in V, in line with previous studies [44]). For large volumes, pled to thermal baths at temperatures T 1 and T 1 + ∆T respec-
these are also well described by the third term in Eq. (29) tively, forming a checkerboard pattern. For concreteness, we
(or (33)-(34)). We confirm this by plotting in Fig. 2(g) the fix ∆T = 0.9. This temperature gradient ensures a steady heat
ratio flux from one bath to the other, and hence a non-vanishing
production of entropy [36, 62].

r= , (42) The model is simulated using standard Monte Carlo meth-
q(1 − q)(µ1 − µ0 )2 τm f (τ/τm ) ods. The dynamics is assumed to be governed by Markovian
single-site transitions si → s0i , occurring with rate ω si ,s0i =
where all quantities in the rhs are computed independently min{1, exp[−∆Ei /T i ]}, where ∆Ei = H(s0 ) − H(s) and T i is
from the simulations. One can also consider similar defini- the temperature of site i. For the current (2), we once again
tions for r0(1) . Since the γi are at most polynomial in V, if this focus on the net entropy production rate to the environment
ratio tends to r → 1 when V is large, it serves as a confir- which is characterized by increments ∆Ei /T i [39, 62].
mation that, for large V, the model effectively behaves as the As in the equilibrium version, the phase transition is ex-
2-state minimal model of Sec. IV. As is clear in Fig. 2(g), this pected to be discontinuous for Q > 4. Moreover, for Q = 12,
is indeed the case. the discontinuity is expected to become very sharp for suf-
ficient large V, since it involves Q distinct ordered phases
coexisting with a single disordered one. The nonequilib-
B. Q = 12-states Potts model rium phase transition can be quantified by the order-parameter
φ = Q[(Nmax /V) − 1]/(Q − 1), where Nmax = max{N1 , ..., NQ }
As a second application, we study a Q = 12 states Potts is the maximum number of spins among all Q configurations
model coupled to two thermal baths at different temperatures. [60, 63]. Fig. 3(a) shows results for φ as a function of T 1 ,
The model is defined in a regular 2D lattice with V sites, for different lattice sizes V. The emergence of a discontin-
where each site i assumes one of Q = 12 values si = 1, . . . , Q uous transition as V increases is clearly visible. The inset
and interacts with its z = 4 nearest neighbors, with energy in Fig. 3(a) shows the metastability lifetime, which is again
8

0,9 V=900
V=1600
V=2500 equal area 1e+12 1e+12 (b)
(a) V=3600 0,068 (b) q=1/2 Dτ
0,6 (a)
Dτ|0
φ 1e+09 6
T1V τ=10 1e+08
0,006 6 Dτ|1
1e+06 Pφ τ=5.10
0,066 -Cτ
τm 1e+06 τ=10
7
0,3
10000
10000
0 0 2000 4000 0 2000 4000
0 0,5 1 V=900 V V
0
V 5000 0,064 φ V=1600 3e+09 Dτ|0
0 V=2116
Dτ|1
0,062 0,064 0,066 0,068 0,07 0 0,0005 -1 0,001 1e+09 2e+09
T1 V (c) -Cτ

1e+09 (d)
(c) (d)
0,8 0,3 0 0
0 5e+04 1e+05 0 5e+04 1e+05
τ τ=1000
τ
1,1 (f)
J τ=5000 1,01
q (e) τ=10000
0,4 0,2 τ=20000
1 1,1

r|1
r

r|0
0,99 1
1
0,1 1500 3000
0 0,98
0,062 0,064 0,066 0,068 0,07 0,062 0,064 0,066 0,068 0,07 1000 2000 3000 1000 2000 V 3000
T1
V
T1

FIG. 3. Characterization of the Q = 12 Potts model in contact FIG. 4. Unconditional and conditional diffusion coefficients for the
with two thermal baths of temperatures T 1 and T 1 + ∆T (with fixed Q = 12 Potts model. (a) Dτ vs. V for different values of τ. (b) Dτ|i
∆T = 0.9). (a) Order parameter φ vs. T 1 for different volumes and Cτ vs. V with τ = 5 × 106 . (c) Dτ vs.τ for different V. (d)
V. Inset: metastability lifetime τm vs. V. (b) Finite-size analysis Dτ|i and Cτ vs. τ for V = 1600. Continuous lines in (c) and (d) are
of the transition point T 1V vs. V −1 , yielding the asymptotic value the theoretical predictions from Eq. (29). (e) The ratio (42) between
T 01 = 0.0651(1). Inset: distribution of φ at T 1V , for different vol- Dτ and the predictions of the minimal model, Eq. (29), which tends
umes. (c) Phase probability q vs. T 1 , again for different volumes. to unity for large volumes. Curves are for different values of τ. (f)
The continuous lines are fits of q = [1 + Qe−Vc(T1 −T10 ) ]−1 . (d) Average Same, but for r|0 (main plot) and r|1 (inset). In all curves, for each
entropy production rate current J [Eq. (3)], which follows closely the V, we fix T 1 as the value T 1V for which q = 1/2. Other details are as
behavior of q. in Fig. 3.

found to grow exponentially with V.


The sharp features of discontinuous phase transitions be- findings are observed.
come rounded at the vicinity of the coexistence point, due
to finite size effects. To locate the transition point, we re- The unconditional diffusion coefficient Dτ [Eq. (3)] is
sort to the finite size scaling theory [64], establishing that the shown in Fig. 4(a) for different values of τ. In agreement with
”pseudo-transition” point T 1V , in which both phases have the the predictions of Eq. (29), for each τ the diffusion coefficient
same weight (equal-area order-parameter probability) reaches initially grows exponentially with V. But for a sufficiently
its asymptotic value T 10 according to the relation T 1V − T 10 ∼ large V, τm becomes comparable to τ and Dτ bends down-
V −1 . This is shown in Fig. 3(b), from which we find T 10 = wards. This is exactly the behavior predicted by the minimal
0.0651(1). model [Fig. 1(d)]. The corresponding conditional diffusion
A histogram of the order parameter φ is shown in the inset coefficients are shown in Fig. 4(b). They follow a similar de-
of Fig. 3(b). It shows that there is a clear separation between pendence on V as Dτ , which is in agreement with the expec-
the two phases, allowing us to define a separator φ∗ = 1/2, tations of Eqs. (33)-(35).
such that the phase-indicator It assumes the value It = 1 when The dependence of Dτ , Dτ|i and Cτ as a function of τ, for
φ(t) > φ∗ . The resulting phase probability q is presented in different V, are shown in Figs. 4(c),(d). In all cases, when
Fig. 3(c). As in the other models, it presents a sharp transition τ is small the diffusion coefficients tend to be linear in τ, in
at T 10 = 0.0651(1), and is well described by the expression agreement with Eq. (31). If V is not too large, then when τ
q = [1 + Qe−Vc(T1 −T10 ) ]−1 . Contrarily to Schlögl’s model, how- becomes large one recovers instead a τ-independent behavior,
ever, the curves for different volumes do not cross at q = 1/2, as predicted by Eq. (32). For large V something similar is
but instead at q ' 1/13. The unconditional current J [Eq. (3)] expected to occur, although it may require unrealistically large
is presented in Fig. 3(d). As predicted by Eq. (3), it follows values of τ.
very closely the behavior of q [Fig. 3(c)], smoothly interpolat-
ing between µ0 and µ1 . Finally, we study the ratio (42), between the actual diffu-
We now turn to an analysis of the unconditional and con- sion coefficients and the predictions of the minimal model
ditional diffusion coefficients. The results are summarized in [Eq. (29)]. The results, for both unconditional and conditional
Fig. 4. To reduce the number of free parameters, we proceed quantities, is shown in Fig. 4(e),(f). In all cases, the plots
similarly to Schlögl’s model, and set, for each volume V, the clearly show that the ratio seems to tend to unity for suffi-
temperature to T 1V (i.e., so that q = 1/2). In Fig. 5 we repeat ciently large V. This strongly indicates that the Potts model
the same analysis, but fixing instead the temperature at T 10 will also behave as an effective 2-state minimal model in the
(the thermodynamic limit transition point) for all V. Similar thermodynamic limit.
9

1e+12 1e+12 Appendix A: Large deviation theory results at arbitrary times


1e+10
(a) (b)
Dτ 1e+09 6 1e+08 Dτ | 0 In this appendix, we derive the expressions for the first
τ = 10
6 Dτ | 1
1e+06
τ = 5.10 1e+06
- Cτ
and second current moments from the large deviation theory.
0 2000 4000 0 2000 4000
Unlike standard treatments, the main difference here is that
4e+08
V V we focus on finite integration times τ. The starting point is
(c) Dτ|0 (d)
V = 900
V = 1600
Dτ|1 Eq. (10), describing the evolution of the entries G x (η) of the
1e+08
Dτ 2e+08 V = 2500 -Cτ moment generating function (MGF). Treating it as a vector
|G(η)i and from its series expansion in powers of η, we have
0
0 25000 50000
0 that
0 25000 50000
τ 1,05 τ
τ=5000 (e) rτ | 0
τ=5000
τ=10000
(f) |G(η)i = |pi + η|g1 i + η2 |g2 i + . . . , (A1)
1,2 τ=10000 τ=20000
τ=20000 1
r
0,95
where |pi is the steady-state of W. Combining this with the
1
series expansion of the tilted operator, L(η) = W + ηL1 +
500 1000 1500 2000 2500
0,9
500 1000 1500 2000 2500 η2 L2 , and collecting terms of the same order in η, we have the
V V
following system of equations
FIG. 5. Same as Fig. 4, but fix T 1 fixed at the thermodynamic limit
value T 10 = 0.0651(1).
d
|pi = W|pi, (A2)

d
|g1 i = L1 |pi + W|g1 i, (A3)

VI. CONCLUSIONS
d
|g2 i = L2 |pi + L1 |g1 i + W|g2 i. (A4)

The statistics of thermodynamic currents is a fundamental
issue in nonequilibrium thermodynamics, which has recently From these, the first and second moments are promptly ob-
received significant interest. In this paper, we presented a tained as
simple and general description of the statistics of thermody- E(Jτ ) = h1|g1 i, E(Jτ2 ) = 2h1|g2 i, (A5)
namic currents for systems displaying discontinuous phase
transitions. We introduced the idea of conditional statistics, which follow from the definition of the MGF. Eq. (A2) is
accounting for the currents in each of the coexisting phases. automatically satisfied in the steady-state. The solution of
From large deviation theory, general relations for the uncon- Eq. (A3), with |g1 (τ = 0)i = 0, is given by
ditional and conditional cumulants of a generic current were

presented. We also proposed a minimal model, which cap- 0
tures all essential features of the problem. Our ideas were |g1 (τ)i = dτ0 eW(τ−τ ) L1 |pi. (A6)
illustrated in two representative systems: the exactly solvable 0
Schlögl’s model of chemical reactions, and a Q-states Potts
For concreteness, we assume W is diagonalizable as discussed
model subject to two baths at different temperatures. In both
above Eq. (15). We can then write
cases, the results were found to follow very well the theoreti-
X
cal predictions of the minimal model, illustrating not only its eWτ = |pih1| + eλi τ |xi ihyi |. (A7)
reliability but also the intricate role of distinct scaling times i,0
and the volume.
The eigenvectors satisfy h1|pi = hyi |xi i = 1 and h1|xi i =
As a final remark, we address some potential extensions of hyi |pi = 0. Thus, plugging (A7) in (A6), we find
our work. It would be interesting to extend such approach to
study the statistics of other quantities, such as the work. An- X eλi τ − 1
|g1 (τ)i = |pih1|L1 |pi τ + |xi ihyi |L1 |pi. (A8)
other interesting point to be investigated concerns the use of
i,0
λi
our framework for tackling statistics of efficiency of thermal
engines at the phase coexistence regimes. To obtain the first moment we take the inner product h1|g1 i;
the second term vanishes and we are left with
E(Jτ ) = h1|L1 |pi τ, (A9)
which yields Eq. (13).
ACKNOWLEDGMENTS Turning now to the second moment, the solution of
Eq. (A4) reads
The authors acknowledge the financial support of the Zτ
0
São Paulo Research Foundation FAPESP, under grants |g2 (τ)i = dτ0 eW(τ−τ ) (L2 |pi + L1 |g1 (τ0 )i). (A10)
2018/02405-1, 2017/24567-0, 2020/03708-8, 2019/14072-0. 0
10

We are only interested in h1|g2 i. Using Eq. (A7), together Hence,


with the fact that h1|xi i = 0, we are then left only with
E(Jτ2 ) − E(Jτ )2 = 2h1|L2 |pi τ (A14)
 eλi τ − 1 − λi τ 
X  
+2 h1|L1 |xi ihyi |L1 |pi   .
Zτ n o i,0
λ2i
h1|g2 (τ)i = dτ0 h1|L2 |pi + h1|L1 |g1 (τ0 )i . (A11)
Dividing by 2τ finally yields Eq. (15).
0
As a final comment, concerning now the computation of
Eq. (16), which is valid when τ  λi , it is convenient to ex-
press the solution in a way which is independent of the full
The first term is time-independent and hence will simply give eigendecomposition of W (and hence more convenient for nu-
a factor of τ. In the second term we use Eq. (A8), leading to merical computations). Let |Q1 i denote the solution of the
linear equation
 
W|Q1 i = 1 − |pih1| L1 |pi. (A15)
Zτ Zτ0
h1|g2 (τ)i = h1|L2 |pi τ + dτ 0 0 00
dτ00 h1|L1 eW(τ −τ ) L1 |pi. This equation actually has an infinite number of solutions,
which are of the form
0 0
(A12) |Q1 i = W+ L1 |pi + |pih1|wi, (A16)
This, combined with the first moment squared, yields Eq. (14).
for any vector |wi. Here, recall, W+ is the Moore-Penrose
To obtain the more explicit formula (15), we carry out the pseudo-inverse of W. Projecting out the contributions from
remaining integral, leading to the subspace |pih1|, we see that

1 − |pih1| |Q1 i = W+ L1 |pi.


 
(A17)
τ2 Hence, Eq. (16) can be rewritten as
h1|g2 (τ)i = h1|L2 |pi τ + h1|L1 |pi2 (A13)
2
 λτ Dτ = h1|L2 |pi − h1|L1 |Q1 i − h1|L1 |pih1|Q1 i. (A18)
 e i − 1 − λi τ 
X 
+ h1|L1 |xi ihyi |L1 |pi   .
i,0
λ2i This form of the diffusion coefficient is more familiar in the
LDT literature, as compared with Eq. (16). It has the advan-
tage that it requires solving a single linear equation (A15),
which is computationally much cheaper than fully diagonaliz-
The second term is identified as the first moment squared. ing W.

[1] U. Seifert, Reports on progress in physics 75, 126001 (2012). 012101 (2017).
[2] F. Ritort, Comptes Rendus Physique 8, 528 (2007), work, dissi- [13] P. Pietzonka and U. Seifert, Physical Review Letters 120,
pation, and fluctuations in nonequilibrium physics. 190602 (2017).
[3] D. J. Evans, E. G. D. Cohen, and G. P. Morriss, Physical Re- [14] I. A. Martı́nez, É. Roldán, L. Dinis, D. Petrov, J. M. Parrondo,
view Letters 71, 2401 (1993). and R. A. Rica, Nature physics 12, 67 (2016).
[4] G. Gallavotti and E. G. D. Cohen, Journal of Statistical Physics [15] V. Blickle and C. Bechinger, Nature Physics 8, 143 (2012).
80, 931 (1995). [16] K. Proesmans, Y. Dreher, M. Gavrilov, J. Bechhoefer, and
[5] C. Jarzynski, Physical Review Letters 78, 2690 (1997). C. Van den Broeck, Physical Review X 6, 041010 (2016).
[6] G. E. Crooks, Journal of Statistical Physics 90, 1481 (1998). [17] P. A. Quinto-Su, Nature communications 5, 1 (2014).
[7] M. Esposito, U. Harbola, and S. Mukamel, Reviews of Modern [18] K. Kinosita, R. Yasuda, H. Noji, and K. Adachi, Philosophical
Physics 81, 1665 (2009). Transactions of the Royal Society of London. Series B: Biolog-
[8] M. Campisi, P. Hänggi, and P. Talkner, Reviews of Modern ical Sciences 355, 473 (2000).
Physics 83, 771 (2011). [19] S. Liepelt and R. Lipowsky, Phys. Rev. Lett. 98, 258102 (2007).
[9] A. C. Barato and U. Seifert, Physical Review Letters 114, [20] S. Liepelt and R. Lipowsky, Phys. Rev. E 79, 011917 (2009).
158101 (2015). [21] A. Lau, D. Lacoste, and K. Mallick, Physical review letters 99,
[10] P. Pietzonka, A. C. Barato, and U. Seifert, Physical Review E 158102 (2007).
93, 052145 (2016). [22] J. V. Koski, V. F. Maisi, J. P. Pekola, and D. V. Averin, Proceed-
[11] T. R. Gingrich, J. M. Horowitz, N. Perunov, and J. L. England, ings of the National Academy of Sciences 111, 13786 (2014).
Physical Review Letters 116, 120601 (2016). [23] S. Toyabe, T. Sagawa, M. Ueda, E. Muneyuki, and M. Sano,
[12] P. Pietzonka, F. Ritort, and U. Seifert, Physical Review E 96, Nature physics 6, 988 (2010).
11

[24] S. Krishnamurthy, S. Ghosh, D. Chatterji, R. Ganapathy, and [44] B. Nguyen and U. Seifert, Physical Review E 102, 022101
A. Sood, Nature Physics 12, 1134 (2016). (2020).
[25] P. Pietzonka and U. Seifert, Phys. Rev. Lett. 120, 190602 [45] L. Levitov and G. Lesovik, JETP letters 58, 230 (1993).
(2018). [46] C. Flindt, C. Fricke, F. Hohls, T. Novotny, K. Netocny, T. Bran-
[26] T. Denzler and E. Lutz, arXiv preprint arXiv:2007.01034 des, and R. J. Haug, Proceedings of the National Academy of
(2020). Sciences 106, 10116 (2009).
[27] G. Verley, M. Esposito, T. Willaert, and C. Van den Broeck, [47] H. Touchette, Physics Reports 478, 1 (2009).
Nature communications 5, 4721 (2014). [48] Z. Koza, Journal of Physics A: Mathematical and General 32,
[28] G. Verley, T. Willaert, C. Van den Broeck, and M. Esposito, 7637 (1999).
Physical Review E 90, 052145 (2014). [49] F. Baras, M. M. Mansour, and J. E. Pearson, The Journal of
[29] M. Polettini, G. Verley, and M. Esposito, Physical review let- Chemical Physics 105, 8257 (1996).
ters 114, 050601 (2015). [50] N. G. van Kampen, Stochastic Processes in Physics and Chem-
[30] T. Denzler and E. Lutz, Phys. Rev. Research 2, 032062 (2020). istry (North-Holland Personal Library, 2007).
[31] M. Esposito and C. Van Den Broeck, Physical Review E - Sta- [51] P. Hanggi, H. Grabert, P. Talkner, and H. Thomas, Physical
tistical, Nonlinear, and Soft Matter Physics 82, 011143 (2010). Review A 29, 371 (1984).
[32] C. Van den Broeck and M. Esposito, Physical Review E 82, [52] P. Pietzonka, K. Kleinbeck, and U. Seifert, New Journal of
011144 (2010). Physics 18, 052001 (2016).
[33] T. Tomé and M. J. de Oliveira, Phys. Rev. E 82, 021120 (2010). [53] M. Vellela and H. Qian, Journal of the Royal Society Interface
[34] T. Tomé and M. J. de Oliveira, Phys. Rev. E 91, 042140 (2015). 6, 925 (2009).
[35] J. Marro and R. Dickman, Nonequilibrium Phase Transitions [54] F. Schlögl, Zeitschrift für Physik 253, 147 (1972).
in Lattice Models, 1st ed. (Cambridge University Press, Cam- [55] P. Hänggi, H. Grabert, P. Talkner, and H. Thomas, Physical
bridge, 1999). Review A 29, 371 (1984).
[36] T. Tomé and M. C. De Oliveira, Physical Review Letters 108, [56] D. T. Gillespie, The Journal of Chemical Physics 74, 5295
020601 (2012). (1981).
[37] P. S. Shim, H. M. Chun, and J. D. Noh, Physical Review E 93, [57] F. Y. Wu, Rev. Mod. Phys. 54, 235 (1982).
012113 (2016). [58] D. Kim, Physics Letters A 87, 127 (1981).
[38] L. Crochik and T. Tomé, Physical Review E - Statistical, Non- [59] R. Baxter, Journal of Physics A: Mathematical and General 15,
linear, and Soft Matter Physics 72, 057103 (2005). 3329 (1982).
[39] Y. Zhang and A. C. Barato, Journal of Statistical Mechanics: [60] C. E. Fiore and M. G. E. da Luz, The Jour-
Theory and Experiment 2016, 113207 (2016). nal of Chemical Physics 138, 014105 (2013),
[40] C. E. F. Noa, P. E. Harunari, M. J. de Oliveira, and C. E. Fiore, https://doi.org/10.1063/1.4772809.
Physical Review E 100, 012104 (2019). [61] C. E. Fiore and M. G. E. da Luz, Phys. Rev. Lett. 107, 230601
[41] T. Herpich and M. Esposito, Physical Review E 99, 022135 (2011).
(2019). [62] T. Martynec, S. H. L. Klapp, and S. A. M. Loos, New Journal
[42] B. O. Goes, C. E. Fiore, and G. T. Landi, Phys. Rev. Research of Physics 22, 093069 (2020).
2, 013136 (2020). [63] M. S. S. Challa, D. P. Landau, and K. Binder, Phys. Rev. B 34,
[43] T. Herpich, J. Thingna, and M. Esposito, Physical Review X 8, 1841 (1986).
31056 (2018). [64] M. M. de Oliveira, M. G. E. da Luz, and C. E. Fiore, Phys. Rev.
E 97, 060101 (2018).

You might also like