Sunteți pe pagina 1din 20

Entropy production in field theories without time reversal symmetry:

Quantifying the non-equilibrium character of active matter


Cesare Nardini,1, 2, 3 Étienne Fodor,1, 4 Elsen Tjhung,1 Frédéric van Wijland,4 Julien Tailleur,4 and Michael E. Cates1
1
DAMTP, Centre for Mathematical Sciences, University of Cambridge, Wilberforce Road, Cambridge CB3 0WA, UK
2
SUPA, School of Physics and Astronomy, University of Edinburgh,
Peter Guthrie Tait Road, Edinburgh EH9 3FD, UK
3
Service de Physique de l’État Condensé, CNRS UMR 3680, CEA-Saclay, 91191 Gif-sur-Yvette, France
4
Laboratoire Matière et Systèmes Complexes, UMR 7057 CNRS/P7, Université Paris Diderot,
10 rue Alice Domon et Léonie Duquet, 75205 Paris cedex 13, France
arXiv:1610.06112v2 [cond-mat.stat-mech] 23 Apr 2017

Active matter systems operate far from equilibrium due to the continuous energy injection at the
scale of constituent particles. At larger scales, described by coarse-grained models, the global en-
tropy production rate S quantifies the probability ratio of forward and reversed dynamics and hence
the importance of irreversibility at such scales: it vanishes whenever the coarse-grained dynamics
of the active system reduces to that of an effective equilibrium model. We evaluate S for a class of
scalar stochastic field theories describing the coarse-grained density of self-propelled particles with-
out alignment interactions, capturing such key phenomena as motility-induced phase separation.
We show how the entropy production can be decomposed locally (in real space) or spectrally (in
Fourier space), allowing detailed examination of the spatial structure and correlations that underly
departures from equilibrium. For phase-separated systems, the local entropy production is concen-
trated mainly on interfaces with a bulk contribution that tends to zero in the weak-noise limit. In
homogeneous states, we find a generalized Harada-Sasa relation that directly expresses the entropy
production in terms of the wavevector-dependent deviation from the fluctuation-dissipation relation
between response functions and correlators. We discuss extensions to the case where the particle
density is coupled to a momentum-conserving solvent, and to situations where the particle current,
rather than the density, should be chosen as the dynamical field. We expect the new conceptual
tools developed here to be broadly useful in the context of active matter, allowing one to distinguish
when and where activity plays an essential role in the dynamics.

PACS numbers: 05.40.-a; 05.70.Ce; 82.70.Dd; 87.18.Gh

Active matter consists of systems where energy is in- tures have been measured experimentally, for instance in
jected at the level of each constituent particle, for in- living systems from the dynamics of injected tracers [27–
stance to power a self-propelled motion, before being dis- 31]. Generically, however, there is no direct connection
sipated locally [1]. Interacting assemblies of such par- between the value of the effective temperature and the
ticles exhibit a rich phenomenology, ranging from the non-equilibrium nature of the dynamics encoded in the
transition to collective motion [2–6] to the emergence of breakdown of time-reversal symmetry (TRS). Another
spatio-temporal chaos [7] and large-scale vortices of self- characteristic feature of non-equilibrium systems is the
propelled composite structures [8, 9]. The sustained in- emergence of steady-state currents, whose study has long
jection and dissipation of energy at the microscopic level been of interest [32–37]. On the other hand, TRS break-
drives the dynamics out of equilibrium. Despite this, down in steady state is in general quantified by the global
it is sometimes difficult to pinpoint a truly nonequilib- entropy production rate S [38]. This can be found, even
rium signature in the emergent collective properties: the far from equilibrium, directly from the probability ratio
strong microscopic departure from equilibrium does not of each realization of the dynamics to its time reversed
necessarily survive in the large scale physics. This is counterpart [38–41].
particularly striking in the emergence of cohesive mat- The possibility that temporal or spatial coarse-graining
ter in the absence of cohesive forces through the mecha- of a nonequilibrium system can restore or partially re-
nism of motility-induced phase separation [10–12]. While store TRS, creating an effective equilibrium dynamics,
clearly non-thermal, the large-scale dynamics does not has been theoretically addressed in a number of different
lead to steady mass currents and closely resembles equi- studies [42–47]. So far however, only a few of these ad-
librium phase separation. Accordingly, many attempts dress active matter directly [10, 48–52]. Our goal in the
have been made to connect this phenomenology to equi- present work is to understand the connection between
librium physics [10, 13–21]. emergent phenomena, such as phase separation, and the
The departure from equilibrium in active systems has existence of irreversibility at coarse-grained scales. We
often been studied by introducing effective temperatures address this question by studying the entropy produc-
as defined by the ratio of response functions to correlators tion of stochastic field theories, which describe at coarse-
[22]. These reduce to the true temperature in the equi- grained level active systems undergoing motility-induced
librium limit because the fluctuation-dissipation theorem phase separation. Importantly, we progress beyond the
then holds [16, 23–26]. In particular, effective tempera- evaluation of the global entropy production rate, which
in steady state is a single number S, to address the more at mean-field level and is broken by fluctuations, as we
detailed question of how this is built up of contributions show, both directly, and by constructing a spectral de-
from different regions of real space or reciprocal space. composition of the entropy production. This decomposi-
This allows us to develop tools for quantifying deviations tion generalizes to active field theories the Harada-Sasa
from equilibrium in a locally or spectrally resolved fash- relation (HSR) [47] which relates S to violations of the
ion, leading to new insight. fluctuation-dissipation relation. The experimental feasi-
To capture the collective physics of self-propelled par- bility to measure entropy production from HRS has al-
ticles, various continuum theories have been proposed ready been proven [75–77] while its application to many-
based either on coarse-graining procedures or on sym- body active matter systems is still elusive. We finally
metry arguments [1]. A prototypical example is “Active extend our approach to more general systems by con-
Model B” [53] which describes diffusive phase separation sidering (in Section III) a diffusive active density field
between two isotropic phases of active matter at differ- coupled to a momentum-conserving fluid (Active Model
ent particle densities. It captures in stylized form the H), and (in Section IV) a diffusive model in which the
coarse-grained many-body dynamics of active particles mean current J has nonzero curl (Active Model B+). In
with either discrete [10, 21] or continuous [54, 55] an- the latter case we can choose whether to calculate S from
gular relaxation of their self-propulsion direction. (Such φ(r, t) as before, or from J(r, t) which now contains more
particles can phase separate even when their interactions information. The results differ: this shows that the en-
are purely repulsive [12].) Within a gradient expansion of tropy production depends on what variables one chooses
the particle density field φ(r), at zeroth order a bulk free to retain in a coarse-grained description. In Section V
energy density f (φ) can be constructed, thus providing we give our conclusions.
a mapping to equilibrium [10]. But at square-gradient
level the mapping to equilibrium is lost and TRS is bro-
ken [53, 54]. Active Model B thus adds minimal TRS- I. ACTIVE MODEL B
breaking gradient terms to Model B, which is an equi-
librium square-gradient theory widely employed in the We consider a conserved, scalar phase-ordering system
theory of critical phenomena [56, 57]. This resembles at the fluctuating hydrodynamics level. Based on sym-
how the KPZ equation [58] was constructed, by extension metry grounds, the simplest dynamics of its order pa-
of the Edwards-Wilkinson model [59], as a prototypical rameter field φ(r, t) satisfies
non-equilibrium model of interfacial growth.
The methods used to quantify entropy production rely φ̇ = −∇ · J, J = −∇µ + Λ, (1)
on path integral representations of stochastic PDEs and
weak-noise large deviation theory [60–64]. These have where J is a fluctuating current and Λ a spatio-temporal
recently been used to address the macroscopic behaviour Gaussian white noise field satisfying
of diffusive systems [35, 65]. In that context, several

stochastic field theories associated via suitable coarse Λ= 2DΓ ; hΓα (r, t)Γβ (r0 , t0 )i = δαβ δ(r − r0 )δ(t − t0 ).
graining [66–69] with lattice [35, 63, 70] and continuum (2)
models [64] have been considered, and large-deviation re- Here the noise strength D is the ratio of the collective
sults for both density [35, 64] and current [32, 35, 65] diffusivity to the collective mobility; the latter has been
have been obtained. Other results on stochastic thermo- set to unity in (1). For systems en route to equilibrium,
dynamics and fluctuation theorems for field theories have the deterministic part of the current takes the form
been offered [71–74].
We give in Section I a brief review of Active Model δF[φ]
B. Then we define in Section II the global entropy pro- Jd ≡ −∇µ, µ = µE ≡ . (3)
δφ
duction rate S from the probability of observing a real-
ization of the fluctuating density field φ(r, t). A success This is Model B [56, 57]. The chemical potential µE de-
of S is to capture in a single number the importance rives from a free energy F[φ] which is conveniently chosen
of non-equilibrium effects in steady state. However, a of the φ4 -type
nonzero value does not allow straightforward physical in- Z h
sight into where or how TRS is broken in the coarse- κ i
F[φ] = f (φ) + |∇φ|2 dr, (4)
grained dynamics. To rectify this we introduce a spa- 2
tial decomposition of the entropy production that allows a2 φ2 a4 φ4
us to locally resolve the nonequilibrium effects of activ- f (φ) = + . (5)
2 4
ity. For a phase-separated state of Active Model B, we
find that the main contribution arises from the interfaces. Phase separation then arises, at mean-field level, when-
On this basis, one might be tempted to assume that in ever a2 < 0; here a4 and κ are both positive.
homogeneous phases the nonequilibrium nature of the For a class of phase-separating active matter models,
dynamics becomes unimportant, yielding effective equi- it has been argued that the main effect of the activity can
librium behavior [10, 55]. But in fact this holds only be captured at the fluctuating hydrodynamic level by an

2
additional contribution to the chemical potential given, to the entropy production rate S; the initial condition is
in its simplest form, by: then irrelevant. All terms of equilibrium form (µE ) con-
tribute only to this transient as we will see below. Im-
µ = µE + µA , µA ≡ λ|∇φ|2 . (6) portantly, although individual paths can have negative
Eqs. (1–6) define Active Model B [53, 54], which is the entropy production, S τ as defined in (7) cannot be neg-
simplest coarse-grained description of phase-separating ative. Using standard field-theoretical methods [69, 80–
active systems. Note that the explicit coarse-graining 82], the trajectory weight can be written as
of self-propelled particles interacting via a density- P[φ] = exp [−A] , (8)
dependent propulsion speed indeed leads to a closely
related dynamics, albeit with more complex density- where A[φ(r, t)] is the dynamical action.
dependence of the various coefficients [21, 54]. For Active Model B as studied here, we show below
The defining property of µA is that it cannot be writ- that the global entropy production can be written as
ten as the functional derivative of any F[φ]. It rep- Z
resents a nonequilibrium chemical potential contribu- S = hσ̂(φ, ∇φ, . . . )i dr, (9)
tion, which violates TRS by undermining the free-energy
structure of the steady state. Interestingly, while the
where σ̂ is a local function of the field φ and its deriva-
break-down of TRS has often been modelled at the micro-
tives. We then interpret the integrand σ(r) ≡ hσ̂i as a
scopic level with nonequilibrium noise terms that directly
steady-state local entropy production density. Yet, such
break fluctuation-dissipation relations, it is the determin-
an interpretation carries two caveats. First, there are
istic part the current J which deviates from equilibrium
several different possible expressions for σ(r) which dif-
form in Active Model B. This differs from recent studies
fer by transient and/or total derivative terms, all giving
of the impact of colored noises on Models A and B [78].
the same integral S; we return to this issue below. Sec-
Note also that the decomposition of µ into its equilib-
ond, the existence of a local entropy production density
rium and nonequilibrium parts is not unique; since the
seemingly implies additivity of S over subsystems. This
only defining property of µA is that it does not derive
is however a subtle point which is discussed in Appendix
from a free energy, an arbitrary equilibrium contribution
A.
can be moved into it from µE . This does not change any
The dynamical action for Active Model B is2 [80]
of the results of the next Section, but we will have to
be more careful in Section III where activity affects the 1
Z
stress tensor as well as the chemical potential. A[φ] = A(r, t)drdt,
4D (11)
A = −(φ̇ + ∇ · Jd )∇−2 (φ̇ + ∇ · Jd ).
II. ENTROPY PRODUCTION
Here the stochastic integral, like all those below, is de-
fined in the Stratonovich sense [83]. In Eq. (11), we have
To quantify the breaking of TRS, the main quantity silently omitted a time-symmetric contribution stemming
of interest is the noise-averaged, global, steady-state en- from the Stratonovich time-discretisation (see Appen-
tropy production rate S. According to the precepts of dices B and C for details). The integral operator ∇−2
stochastic thermodynamics [38, 79], it is defined as [39] is the functional inverse to the Laplacian (a Coulomb
1

P[φ]
 integral). The action A measures the logarithm of the
τ τ
S = lim S , S = ln R , (7) probability that a solution to (1) is arbitrarily close to a
τ →∞ τ P [φ]
given path {φ(r, t)}0≤t≤τ .
where P is the probability of a path {φ(r, t)}0≤t≤τ , and We now introduce the time-reversal operation
P R is the probability of its time-reversed realization. The
average h·i in (7) is taken with respect to noise realiza- t → τ − t,
(12)
tions. With suitable ergodicity assumptions,1 which we φ(r, t) → φR (r, t) = φ(r, τ − t).
make throughout this paper, averaging over one long tra-
jectory (τ → ∞) with a single noise realization is equiv- This defines φR (r, t) as the path found by running a movie
alent to the noise-average in (7). In this case the angular of φ(r, t) backwards in time. The quantity P R then repre-
brackets h·i in (7) can be dropped, and below we silently sents the probability of observing the trajectory φR (r, t)
do this whenever it suits our purposes. For general initial
conditions φ(r, 0), this single-trajectory time-average will
also include a transient contribution that scales sublin- 2 An equivalent expression for A[φ] is
early with its duration τ , and hence does not contribute
A = |∇−1 (φ̇ + ∇ · Jd )|2 (10)
with the definition ∇−1 X ≡ ∇∇−2 X. The latter operation
maps a scalar field X to a vector field Y such that ∇.Y = X
1 For phase-separated systems, ergodicity is assumed subject to with a gauge choice ∇ ∧ Y = 0. The two forms for A in (11) and
specified positions for the interfaces between phases. (10) are then related by integration by parts.

3
under the “forward” dynamics (1). Since in these dy- find that
namics the deterministic part of the current Jd is a func- Z
tional of φ, we have Jd (r, t) → JRd (r, t) = Jd (r, τ − t), 1
S = − lim µ(r, t)φ̇(r, t)drdt
τ →∞ Dτ
as found by substituting (12) into (1). Thus the deter- Z  (16)
ministic current is not reversed, even though the actual 1
= − lim µA (r, t)φ̇(r, t)drdt + ∆F .
current J = Jd + Λ is equal and opposite in the forward τ →∞ Dτ
and backward paths, because φ̇R = −φ̇. The forward and
backward trajectories are thus likely to require very dif- This result is central to all that follows. Here ∆F =
ferent realizations of the noise. Put differently, the most F(τ ) − F(0) is the difference between the final and ini-
probable forward trajectories have small noise contribu- tial values of the free energy functional F. This is, of
tions so that J ' Jd . The total current in the time- course, the only term present in the passive limit; the
reversed trajectory, JR , is opposed to the deterministic total free energy change then fixes τ S τ = −∆F, so that
one JR ' −JRd so that a very atypical noise realization S = 0. ∆F remains bounded even in active systems be-
may be required, such that the probability of observing cause F(τ ) is bounded below and F(0) bounded above
the reversed trajectories is very small. for sensible initial conditions, in each case by terms linear
Note that the notion of time-reversed trajectories con- in the system volume. Hence this transient contribution
tains a degree of ambiguity. Here, we simply measure the always vanishes as τ → ∞. Assuming ergodicity, the
probability of the noise realization required to make the steady-state entropy production then obeys
time-reversal of a trajectory φ(r, t) a solution of the for- Z
ward dynamics. This should not be confused with various 1
S=− hµA φ̇idr, (17)
“conjugacy” operations that instead map the forward dy- D
namics of one system onto the backward dynamics of an-
other [38]. Such mappings typically involve treating some where the average h·i is now taken over the stationary
of the parameters in the model as odd under time rever- measure. Importantly (see footnote 1) since we are inter-
sal. This happens in magnetism, where S can be found ested in phase separation and similar situations of broken
by comparing a trajectory in an external magnetic field translational symmetry, we should understand the angle
B with the time-reversed trajectory in an external mag- brackets in this expression as being noise-averages within
netic field −B [83]. Such an extension of TRS is there- an ensemble where interfaces between phases are held sta-
fore mandatory to ensure the absence of entropy produc- tionary in time. We can then define a local steady-state
tion in conservative magnetic systems. In the present entropy production density
case, the external parameters (such as λ in Active Model 1
B) must not be chosen odd under time reversal, as this σφ (r) ≡ − hµA φ̇i(r), (18)
D
would compare forward trajectories in one system with
backward trajectories in one with different dynamics, and whose integral is the entropy production:
indeed different phase equilibria [53]. Therefore, such ex- Z
tensions of TRS are not pertinent to the present paper. S = Sφ ≡ σφ (r)dr. (19)
Returning to the main issue, we now observe that the
only anti-symmetric part in the dynamical action is φ̇.
The probability for a forward trajectory to lie arbitrarily Here and in (18) we have added a subscript φ to dis-
close to the time reversed trajectory {φR (r, t)}0≤t≤τ , as tinguish these forms from expressions appearing later on
a functional of {φ(r, t)}0≤t≤τ , is P R = exp [−AR ] with (Section IV) in which the current J rather than the den-
Z sity field φ is used to define trajectories.
1 As already mentioned following (9) above, there are
AR [φ] = AR (r, t)drdt,
4D (13) several other local quantities which have the same inte-
AR = −(φ̇ − ∇ · Jd )∇−2 (φ̇ − ∇ · Jd ), gral, S, and hence have equal claim to be called the local
entropy production density. Indeed,
where we have again omitted a time-symmetric contribu-
tion arising from the time discretization. Relation (7) for − σφ D ≡ hµφ̇i , hµA φ̇i , hJ.∇µA i , −hJ.Jd i (20)
the steady-state entropy production can now be written
are all equivalent for the purposes of computing S. The
AR − A
Z
1 first two forms differ by a transient contribution ∆F/τ →
S = lim = lim (AR − A)drdt. (14)
τ →∞ τ τ →∞ 4Dτ 0, as do the last two. The latter pair are found from
Using (11,13), we get the former by partial integration, differing from them by
  terms of the form ∇.Υd,A where Υd = Jµ and ΥA = JµA .
δF Our numerical studies suggest that these alternative can-
AR − A = 4(∇−1 φ̇)∇ + µA . (15) didates for local entropy production are practically indis-
δφ
tinguishable in the case of Active Model B. A more com-
Performing the spacial integral in (14) by parts and not- plex situation arises for Active Model B+, as described
ing that in the Stratonovich convention φ̇δF/δφ = Ḟ, we in Section IV.

4
A Spatial decomposition of entropy production

1 0.004

0.5 0.002

<φ(x)>

σφ(x)
0 0

<φ(x)>
-0.5 D=0.001 -0.002
D=0.002
D=0.02
-1 D=0.04 -0.004

0 40 80 120 160 200 240


x

FIG. 1. Left: Density map of a fluctuating phase-separated droplet in Active Model B in 2D. Center: Local contribution
to the entropy production σ(r) showing a strong contribution at the interfaces. Right: Density and entropy production
for a 1D system comprising a single domain wall, for various noise levels D  a22 /4a4 . The entropy production is strongly
inhomogeneous, attaining a finite value as D → 0 at the interface between dense and dilute regions and converging to zero in
the bulk in this limit. Values of the parameters used are a2 = −0.125, a4 = 0.125, κ = 8, and λ = 2.

A. Spatial decomposition of entropy production of φ0 and φ1 reduce to:


φ̇0 = −∇ · Jd (φ0 ) (23)
We now turn to the numerical evaluation of σφ (r) in  
2 δF0
Active Model B. We defer to future work any study of φ̇1 = ∇ + 2λ∇φ0 · ∇φ1 + ∇ · Γ (24)
the critical region of the model, focusing instead on the δφ1
weak noise limit (small D) where sharp interfaces form where Γ is a standard Gaussian white noise as in (2) and
between high and low density phases, respectively de-
noted by φh > 0 and φl < 0. Note that φh 6= −φl unless φ2 |∇φ1 |2
Z  
λ = 0, since activity breaks ±φ symmetry in F [53]. To F0 = (a2 + 3a4 φ20 ) 1 + κ dr. (25)
2 2
study phase separated states in one dimension we con-
sider a single domain wall in the centre of the system This shows, as expected, that the statistics of φ0,1 are
and impose ∇φ = 0 on the distant boundaries. We use independent of D at leading order.
finite difference methods with mid-point spatial discreti- We first consider the case where the mean-field dynam-
sation to integrate (1) via a fully explicit first order Euler ics for φ0 has relaxed to a constant profile. In this case,
algorithm. Importantly, the discretised system exactly it follows from (18) that
respects detailed balance whenever λ = 0, as shown in √
Appendix B where further numerical details are given. σφ (r) = −λ Dh|∇φ1 |2 φ̇1 i + O(D). (26)
For numerical purposes we evaluate the entropy produc-
tion density as Inspection of Eq. (24) shows that the only TRS-breaking
term is proportional to ∇φ0 so that, in a homogeneous
1
Z state, φ1 has an equilibrium dynamics controlled by the
σφ (r) = − lim µA (r, t)∇2 µ(r, t)dt, (21) √ F0 . The latter is even in φ1 so that the term
free-energy
t→∞ Dt
of order D in the expansion (26) of σφ vanishes by
symmetry. We conclude that in bulk phases σφ ∝ D. We
whose equivalence with (18) is established, within our have checked this by simulations in single-phase systems
discretization scheme, in Appendix D. where the total entropy production is indeed shown to
Our numerics in both one and two dimensions show scale as S ∝ D; see Fig. 2.
that the local entropy production density σφ (r) is In contrast however, Fig. 2 also shows that the total
strongly inhomogeneous, being large at the interface be- entropy production S for a phase-separated system re-
tween phases, but small within these phases (Fig. 1). To mains finite as D → 0. This D0 term dominates at small
quantitatively explain this, we consider the weak noise enough D and is caused by entropy production localised
expansion of the density near the interfaces between phases. This is graphically
confirmed in the spatial map of the local entropy produc-
√ tion density σφ (r) shown in Fig. 1 which shows that, as
φ = φ0 + Dφ1 + Dφ2 + O(D3/2 ). (22) D → 0, σ vanishes aways from interfaces but has a finite
contribution in their vicinity. This can be understood
In the weak noise limit, standard field-theoretical meth- from the expansion (22) where φ0 (r) is now a (steady)
ods, which we outline in Appendix E, show the dynamics phase-separated solution of the mean-field dynamics (23),

5
A Spatial decomposition of entropy production

0.3 10
one-phase ξ = √2
two-phase ξ = 2√2
ξ = 4√2

0.2 1


0.1
0.1

0.01
0 0.1 1 10 100
0 0.01 0.02 0.03 0.04 0.05 0.06 2 2
λ /ξ
D

FIG. 2. Total entropy production Sφ for Active Model B FIG. 3. Scaling of the total entropy production as a func-
in one dimension for the case where the system consists of showing that Sφ ∝ λ2 , as predicted analytically.
tion of λ, q
a single uniform phase (a2 > 0, a4 > 0, red) and for the Here ξ = − 2aκ2 is the passive interface width. Values of
case where it shows coexistence between a high-density and
a low density phase (a2 < 0, a4 > 0, green). (Parameter the parameters used are (a2 = −0.5, a4 = 0.5, κ = 2, D =
values are a2 = ±0.25 for the uniform (+) and phase-separates 0.001), (−0.25, 0.25, 4, 0.001) and (−0.125, 0.125, 8, 0.001) for
(−) states, with a4 = 0.25, κ = 4, λ = 1.) In the two-phase red, green and blue points respectively.
system,
√ any putative sub-extensive (interfacial) term scaling
as D would be swamped by the extensive D1 contribution
from bulk phases. by activity plays its largest role at interfaces [21, 53].
More generally, the results exemplify how the local en-
R tropy production density σ(r) can provide a good quanti-
with a prescribed value of φ(r)dr. Inserting (22) in the tative tool to understand the breaking of TRS when the
definition (18) of σφ gives system is inhomogeneous. This is likely to be important
in situations more complicated than phase separation,
λ including those where activity parameters such as swim-
σφ (r) = − √ |∇φ0 |2 hφ̇1 i − λ|∇φ0 |2 hφ̇2 i
D (27) speed vary in space (see [84] and also Section IV). In such
√ cases analytical progress could be more difficult, but the
− 2λ∇φ0 · h∇φ1 φ̇1 i + O( D).
numerical evaluation of σφ (r) should remain tractable.
On the other hand, this approach gives little information
In the steady state, hφ̇i i = ∂t hφi i = 0 so that
about the character of TRS breaking in homogeneous
√ phases whose translational invariance forces σφ (r) to be
σφ (r) = −2λ∇φ0 · hφ̇1 ∇φ1 i + O( D) (28) simply flat. In such cases spatial structure is captured
by correlation functions rather than mean values; we ad-
where the average is taken over the stationary measure
dress this next.
of the dynamics (24) for φ1 . Given that the statistics of
φ1 are, by construction, D-independent, it follows that
σφ (r) = O(D0 ) wherever the gradient of the determinis-
tic solution is finite, i.e., at interfaces. B. Spectral decomposition of entropy production
In Fig. 3, we show the dependence of the total en- from fluctuation and response
tropy production Sφ on λ; this scales as Sφ ∝ λ2 . No
linear term is possible, because it would mean that of The fluctuation-dissipation theorem (FDT) is a funda-
two different Active Model B systems, with other param- mental property of equilibrium dynamics [85]. It states
eters the same but with opposite signs of λ, one would that the response to an external perturbation is entirely
have negative global entropy production in steady state. determined by spontaneous fluctuations in absence of
Accordingly in (28), the term hφ̇1 ∇φ1 i must itself be of the perturbation. For active field theories, and non-
order λ in general. This explains the quadratic scaling, equilibrium systems more widely, there is no general re-
but only to leading order in small λ (and D). In practice lation between correlation and response. Thus, violation
we find this scaling over a wide range of λ at small D of the FDT in a coarse grained description at some scale
(Fig. 3), and also at larger D (not shown) but we have is a proof that activity matters dynamically at this scale,
no explanation for these results at present. since no passive model could ever give rise to such a vio-
In summary, we have found above that σφ ∼ D0 where lation. In equilibrium, the ratio of correlators to response
h∇φi 6= 0 whereas σφ ∼ D in bulk phases (Fig. 2). This functions is set by the bath temperature. For nonequilib-
confirms quantitatively that the TRS breakdown induced rium dynamics, an effective temperature is sometimes in-

6
B Spectral decomposition of entropy production from fluctuation and response

troduced by analogy with the FDT by looking for asymp- equilibrated modes, as enforced by the FDT: ωC(k, ω) =
totic regimes (in time or frequency) where constant val- 2DR(k, ω). The spectral decomposition is particularly
ues arise for the correlator/response ratio [25, 86]. One interesting for uniform systems, where the real-space lo-
drawback of such a definition of the effective tempera- cal density σφ (r) defined in Eq. (18) is obliged to be
ture is that it generically is different for every perturba- constant by translational invariance. It is natural in
tion and observable under scrutiny, and therefore bears such systems to instead consider entropy production as
no universal meaning. a function of wavevector (and perhaps also frequency),
In this Section, we provide a quantitative relation be- and (30) provides the appropriate tool for doing so. Per-
tween the rate of entropy production S and the FDT vi- forming the ω integral allows one to quantitatively ex-
olation for the stochastic field dynamics governed by (1). plore whether activity matters dynamically at a given
We then present detailed numerical results for the case spatial scale. A further asset of the generalized HSR is
of Active Model B. For an overdamped single particle that it allows one to evaluate S without complete infor-
driven out of equilibrium by an external non-gradient mation about the equations of motion. In particular, as
force, such a connection was discovered by Harada and detailed below, we assumed these to lie in the broad class
Sasa [47]. They showed that the violation of the FDT defined by (1) but made no assumption about the form
relating the correlation of the particle position and the of the nonequilibrium chemical potential µA . This is in
response to a constant force provides a direct access to contrast to the real-space methods presented in Sec. II A.
the entropy production rate. Therefore, one important We now give the derivation of the HSR result. In terms
success of the Harada-Sasa relation (HSR) was to iden- of the dynamic action Ah of the perturbed dynamics,
tify, among all possible violations of the FDT in a non- using the property δP = −PδAh , the response can be
equilibrium system, the particular observable and per- written as
turbation that quantitatively determine S. The HSR
δAh
 
was later generalised to other types of equilibrium dy- R(r1 , r2 , t) = − φ(r1 , t) . (32)
namics, including the under-damped case and that with δh(r2 , 0) h=0
temporally correlated noise [87]. HSRs have also been
developed for systems with a time-scale separation be- We need Ah only to first order in h:
tween fast and slow degrees of freedom [43], and for the Z
1
density field in a system of particles driven by a time- δAh = − h ∂t φ − ∇2 µ drdt + O h2 ,
 
(33)
independent external force field [88]. In active matter, 2D
some of us recently announced a generalisation of the
so that the spatially diagonal response obeys
HSR in a non-equilibrium microscopic model of particles
propelled by correlated noises [52]. 1

φ(r, t) ∂t φ − ∇2 µ (r, 0) .

We consider the correlation function C and the re- R(r, r, t) = (34)
2D
sponse R to a change of the chemical potential µ → µ−h:
The time-antisymmetric part follows as
C(r1 , r2 , t − s) ≡ hφ(r1 , t)φ(r2 , s)i ,

(29) 1
δhφ(r1 , t)i R(r, r, t) − R(r, r, −t) = − ∂t C(r, r, t)−
R(r1 , r2 , t − s) ≡ . D
δh(r2 , s) h=0 1 2
− ∇ hφ(r, t)µ(r, 0) − φ(r, 0)µ(r, t)i ,
Before outlining the derivation of our generalized HSR, 2D (2)
we give the result: (35)
where we have used C(r, r, t) = C(r, r, −t) via (29). Here
∇2(2) denotes the Laplacian operator acting on the second
Z
S = σφ (k, ω)dkdω,
spatial variable of the function on which it acts. Compar-
(30) ing with (17) and ignoring subdominant boundary terms,
ωk −2
σφ (k, ω) ≡ [ωC(k, ω) − 2DR(k, ω)] , we deduce
(2π)d+1 D
Z
1
where S = lim ∂t ∇−2 (2) {D [R(r, r, t) − R(r, r, −t)]
Z D t→0 (36)
C(k, ω) ≡ C(r1 , r2 , t)eik·(r1 −r2 )+iωt dr1,2 dt, +∂t C(r, r, t)} drdt.
Z (31) This is our generalised Harada-Sasa relation in real space
R(k, ω) ≡ R(r1 , r2 , t)eik·(r1 −r2 ) sin(ωt)dr1,2 dt. and time. Fourier transforming then yields (30).
We have computed numerically the spectral decompo-
The spectral decomposition of the entropy production sition of the entropy production rate for Active Model B,
rate (30) expresses S as an integral over Fourier modes using periodic boundary conditions. We extracted the re-
of a spectral density σφ (k, ω). This density can for- sponse using the algorithm of [89] where two systems, one
mally be seen as the contribution to S of the modes with unperturbed and the other with perturbed chemical
in [k, k + dk] × [ω, ω + dω]. It vanishes for thermally potential, are driven by the same noise realisation. This

7
B Spectral decomposition of entropy production from fluctuation and response

mismatch to σφ (k, ω) makes the latter highly sensitive


to both high-frequency and low-wavenumber informa-
tion. RWe have checked that the total entropy production
S = σφ (k, ω)dkdω nonetheless approaches numerically
the same quantity calculated in real space via (21). The
spectral density plot in Fig. 5(a) does therefore include
most of the relevant frequency and wavenumber range.
Deferring fuller investigations to future work, we can con-
clude that our generalized HSR for Active Model B yields
a nontrivial spectral decomposition of the entropy pro-
duction, including features not captured by the effective
temperature paradigm.
Note that measuring the probability of trajectories and
their time-reversed counterparts is difficult to do in sim-
ulations and even more so in experiments. On the other
hands, physicists have an established expertise in mea-
suring FDT both in simulations and experiments. Eq.
(30) provides a direct link for such measurements to en-
tropy production that waits exploration for active matter
systems.

III. ACTIVE MODEL H

We now turn attention to Active Model H in which the


diffusive dynamics of φ(r, t) is coupled to a momentum-
conserving fluid with velocity v(r, t). This model was
first introduced in [90] to address the role of fluid motion
in active phase separation. In that work, it was shown
that for so-called contractile activity (where particles pull
fluid inward along a polar axis, determined in effect by
∇φ, and expel it equatorially) domain growth can cease
at a length scale where diffusive coarsening is balanced by
FIG. 4. (a) Correlation ωC(k, ω) and (b) response the active stretching of interfaces. As a generalisation to
2DR(k, ω) for Active Model B in one dimension, found by active systems of the well known Model H [56], this model
numerical simulations as detailed in the main text. Parame- offers an interesting arena in which to explore coupling
ter values: −a2 = a4 = 0.25, k = 4, D = 0.05 and λ = 8. of an active scalar to a second field whose dynamics is
dissipative, but which would obey TRS without coupling
to the active field.
allows a very precise measurement of the response, even Diffusive dynamics now takes place in the frame of the
without averaging over noise realizations. Nonetheless, a moving fluid so that (1) acquires an advective term,
separate simulation for each value of k and ω is needed
in order to compute the response R(k, ω). In contrast, a φ̇ + v · ∇φ = −∇ · J, (37)
single run (long enough to ensure good statistical aver-
where we retain from Active Model B the equations for
aging) is sufficient to measure the correlator C(k, ω) for
the diffusive current
all k and ω simultaneously. This is why our statistical
accuracy is better for C(k, ω) than for R(k, ω). J = Jd + Λ, Jd = −∇(µE + µA ). (38)
In Fig. 4, we plot both quantities in a single-phase
system for a wide range of wavelengths and frequencies. We assume the fluid is incompressible and of unit mass
Their mismatch spectrally quantifies the breakdown of density. The thermal Navier-Stokes equation for momen-
TRS via the entropy production, as shown in Fig. 5(a). tum conservation then reads
In Fig. 5(b), we present an ‘effective temperature’, de-
fined through D(k, ω)/D = ωC(k, ω)/2DR(k, ω). This (∂t + v · ∇)v = η∇2 v − ∇p + ∇ · (Σ + Γ̃), (39)
is a popular measure of FDT violations (see above) whose with the noise stress specified by [57]
shortcomings are apparent here. In particular, while the
effective temperature is peaked at large wavenumbers,
D E
Γ̃αβ (r, t)Γ̃γν (r0 , t0 ) =
this is true neither of the FDT mismatch nor of σφ (k, ω) (40)
itself. The additional factor of ω/k 2 relating the FDT 2ηD (δαγ δβν + δαν δβγ ) δ (r − r0 ) δ (t − t0 ) .

8
form (as arises in MIPS to zeroth order in gradients), this
does not lead to a relation ∇ · ΣA = −φ∇µA as one might
expect from (41) [90]. This means it is no longer possible,
as it was in Active Model B, to shift some arbitrary part
of µE into µA without changing the results. Conversely,
one cannot declare all terms in µ that are of equilibrium
form to be part of µE as we did for Active Model B in
Section I; we must now be careful to include in µA any
chemical potential contribution that does not contribute
to the stress via (41). For this reason, when an external
perturbation µ → µ − h(r) is used to calculate response
functions below, we will need to specify whether it is µA
or µE that is being perturbed. Note also that applying a
perturbation h(r) to the coarse grained theory need not
be equivalent to applying a similar one at microscopic
level; the latter would cause more complicated shifts in
both µE and µA [21].
Active Model H comprises the equations (37-42).
In [90], it was further assumed that F[φ] is given by (4),
from which the deviatoric equilibrium stress ΣE takes a
dyadic gradient form. This form turns out to be shared
by ΣA to leading order in gradients:

ΣE = −κ(∇φ)(∇φ), ΣA = −ζ(∇φ)(∇φ). (43)

Here ζ is a mechanical activity parameter which is posi-


tive for extensile and negative for contractile swimmers.
Active Model H thus breaks TRS via two different chan-
nels; first through the non-equilibrium chemical potential
µA in (38), and second because, for ζ 6= 0, the stress ten-
sor and the free energy are not related by (41). Our
analysis covers the case where (4,43) hold, as assumed
in [90], but in fact we need no such restrictions on the
form of F[φ], µA [φ], and ΣA [φ], so far as the latter is a
FIG. 5. (a) Correlator-response mismatch ωC − 2DR̃
∝ σφ (k, ω)k2 /ω and (b) normalised effective temperature
symmetric tensor.
D(k, ω)/D as a function of k and ω for Active Model B in Using the methods outlined for Active Model B in Sec-
one dimension. Parameter values as in Fig. 4. (Control sim- tion I, we can now write the dynamical action density of
ulations for the passive limit λ = 0 recover a statistically Active Model H as
insignificant signal for both plots, not shown, confirming that Z
1
our numerical algorithm respects TRS when present.) A= (Aφ + Aη )drdt,
4D
2
Aφ = ∇−1 [(∂t + v · ∇)φ + ∇ · Jd ] ,

(Recall that we set the mobility of the φ field to unity
1  2
so that D = kB T .) In (39), the pressure field p(r, t) Aη = ∇−1 (∂t + v·∇)vα − η∇2 vα + ∂α p − ∂β Σαβ

subsumes all isotropic stress contributions and enforces η
fluid incompressibility (∇ · v = 0). The deviatoric stress 1 h −2  i2
− ∇ ∂α vγ ∂γ vα − ∂α ∂γ ΣAαγ + ∇2 p ,
tensor Σ is traceless and symmetric. 2η
In passive systems, Σ = ΣE can be derived from a (44)
free-energy functional F[φ] by standard procedures [57]; as shown in Appendix F. The pressure p appears in the
restoring isotropic pressure contributions one finds equations of motion as a Lagrange multiplier for fluid
incompressibility; accordingly A is as given above for
∇ · ΣE = −φ∇µE . (41) those trajectories {φ(r, t), v(r, t), p(r, t)}0≤t≤τ that have
For an active system, the relation between the deviatoric ∇ · v = 0, but is infinite for all others. We next con-
stress tensor and the free energy breaks down, taking struct a time reversal transformation by supplementing
instead the more general form (12) with
Σ = ΣE + ΣA , (42) ρR (r, t) = ρ(r, τ − t),
where ΣE obeys (41). Note that even if activity creates vαR (r, t) = −vα (r, τ − t), (45)
a chemical potential contribution that is of equilibrium pR (r, t) = p(r, τ − t).

9
We note that both µA and ΣA must be treated even under where
time reversal, in line with the discussion already given in Z
Section II.3 It is then straightforward to compute the R(k, ω) ≡ R(r1 , r2 , t)eik·(r1 −r2 ) cos(ωt)dr1,2 dt,
action for the time-reversed dynamics and construct the Z (51)
entropy production via an obvious extension of (7)
R(k, ω) ≡ R(r1 , r2 , t)eik·(r1 −r2 ) cos(ωt)dr1,2 dt.
 
τ τ 1 P[φ, v]
Sφ,v = lim S , S = ln R , (46)
τ →∞ τ P [φ, v] The coupling of the scalar field with an external veloc-
ity field thus requires us to consider two different FDT
with the following result: violations when evaluating S: one associated with the φ
Z dynamics, and the other with the Navier-Stokes sector.
1  
Sφ,v = − µA (∂t + v · ∇)φ + vαβ ΣAαβ dr. (47) While a pair of such FDTs are satisfied independently
D in the passive case, the nature of the perturbations and
Here vαβ = (∂α vβ + ∂β vα )/2 is the symmetrised velocity observables entering the HSR could not have been an-
gradient tensor. This result reduces to (18) for Active ticipated a priori. In particular, the chemical potential
Model B when v = 0, and applies under the same broad perturbation δµ = −h(r, t) must act in the diffusive sec-
conditions. In particular, it is a steady-state quantity tor alone and not in the thermodynamic stress. This
that excludes transient contributions. (The latter would requires h to be considered part of µA not of µE . In con-
be the only terms present in the phase ordering of passive sequence (50) does not connect with standard expressions
Model H without boundary driving [91].) for the (transient) entropy production in the passive limit
The Harada-Sasa relation can be further generalised to in the way one might have expected. This finding is rel-
the case of Active Model H. To this end, we consider per- evant to the wider agenda of generalising the HSR to
turbations of the active parts of the chemical potential more complicated active field theories, for instance with
µA → µA − h and of the stress tensor Σ → Σ − ε, with ε orientational order [1]: since care is required, a relatively
symmetric and traceless. We introduce the correspond- formal approach is advisable. Our result demonstrates
ing response functions as how the HSR framework singles out, given a set of equa-
tions of motion, exactly which FDT violations should be
probed to quantify TRS breakdown.

δ h(∂t φ + vα ∂α φ)(r1 , t)i
R(r1 , r2 , t − s) = ,
δh(r2 , s)
h=0
(48)
δ hvαβ (r1 , t)i
R(r1 , r2 , t − s) = . IV. CIRCULATING CURRENTS
δεαβ (r2 , s) ε=0

We define the associated autocorrelation functions: Active Model B, while abandoning the free energy
structure of its passive counterpart, retains in (1) the gra-
C (r1 , r2 , t) = h(∂t φ + vα ∂α φ)(r1 , t)(∂t φ + vα ∂α φ)(r2 , 0)i , dient structure of the deterministic current: Jd = −∇µ.
C(r1 , r2 , t) = hvαβ (r1 , t)vαβ (r2 , 0)i . This implies ∇ ∧ Jd ≡ 0 so that, at zeroth order in noise,
(49) there can be no circulating currents in steady state. In
We demontrate in Appendix F that the entropy produc- this Section we outline how our approach generalizes to
tion rate can be expressed as scalar field theories that allow currents of nonzero curl,
Z as would be needed to capture (for instance) the self-
assembled many-body ratched behavior reported for ac-
Sφ,v = [σφ (k, ω) + σv (k, ω)]dkdω,
tive Brownian particles in [84].
k −2 One such model, which will be addressed in more detail
σφ (k, ω) = (C − 2DR) (k, ω), (50) in a separate publication [92], has φ̇ = −∇·(Jd +Λ) with
D(2π)d+1
2η Jd = −∇µ̃ + [(κ0 + κ1 φ)∇2 φ + ν|∇φ|2 ]∇φ (52)
σv (k, ω) = (C − 2DR) (k, ω),
D(2π)d+1 µ̃ = µB + (λ0 + λ1 φ)|∇φ|2
µB = −h + a2 φ + a3 φ2 + a4 φ3 − k∇2 φ.

3 To understand this for Active Model H, consider the stress tensor This model, which we call Active Model B+, is written
defined via (43). At first sight the above choice of time reversal here in the presence of a static free energy contribution
appears to contradict the statement in [90] that ζ is odd under −h(r)φ(r). In Jd it includes, for completeness, all active
time reversal. However, this statement refers to the fact that terms to order (φ3 , ∇3 ). Of these, only κ1 and ν can
if one reverses the direction of the fluid flow caused by a swim-
create nonzero curl. Indeed only ν can do this in one
mer (or set of swimmers) as in (45), this has the same effect as
reversing the sign of ζ, which interchanges the contractile and dimension, in the sense of creating a nonzero net current
extensile cases. So, rather than strict time reversal, this is a con- around a loop closed via periodic boundary conditions.
jugacy operation [38] comparing the probabilities of forward and (This is because φ∇2 φ is reduces to gradient form in one
reverse paths in two different systems. dimension.) By Helmholtz decomposition, Jd can always

10
be written as This is a classical form, familiar from the case of external
driving, such as a current of charged particles in a circular
Jd = Jgd
+ Jrd
g
wire driven by a tangential electromotive force.
Jd = −∇µ ; Jrd = ∇ ∧ A, (53) Interestingly, the generalised Harada-Sasa relation (30)
for some choice of scalar field µ and vector field A, with for Sφ is insensitive to the additional terms in (52) and
superscripts g, r standing for gradient and rotational. By follows from (18) directly as before, without modifica-
similarly decomposing the noise terms one can also for- tion. Indeed, since the circulating currents have no effect
mally define gradient and circulating parts of the total on the dynamics of φ, the FDT mismatch of its correla-
current J = Jg + Jr . tors can give no information about their contribution to
Since Jr is divergence-free, it plays no part in the dy- the entropy production via the last term in (55). A con-
namics of φ. Accordingly (7), which defines Sφ , is obliv- nection between Harada-Sasa theory and this term might
ious to any entropy production arising from circulating be made, by considering the fluctuations of the curly cur-
currents. To capture this additional source of irreversibil- rents themselves and their response to perturbations that
ity, we now promote J(r, t) into an explicit dynamical couple to them. We leave this for future investigations.
variable. This leaves φ dynamically redundant,
Rt as it can
be reconstructed via φ(r, t) = φ(r, 0) − 0 ∇ · J(r, s)ds.
From the point of view of comparing forward and back- V. CONCLUSIONS
ward trajectories, this corresponds to comparing movies
in which the current is recorded (e.g., using tracer parti- We have studied the violation of time reversal symme-
cles) rather than just density fluctuations. try (TRS) in a class of scalar field theories relevant for the
If we fix a current trajectory {J(x, t)}0≤t≤τ , the prob- description of active matter such as self-propelled parti-
ability that J is arbitrarily close to it has again an expo- cles without alignment interactions. We have presented
nential form with action given by general tools for addressing the question of whether ac-
1
Z tivity matters dynamically—meaning, whether it con-
2
AJ = |J − Jd | (r, t)drdt. (54) tributes distinctly to TRS-breaking physics, or whether
4D
the dynamics at large scales could be reproduced by some
A straightforward computation shows that the entropy (possibly complicated) equilibrium model with TRS.
production at the level of currents is Studying the entropy production allows one to quantify
Z TRS breakdown directly at coarse-grained level: it un-
1
SJ = Sφ + hJr .Jrd i(r)dr . (55) ambiguously assesses the extent to which dynamics at
D this scale is genuinely nonequilibrium.
For comparison, note that using (20) Sφ can be written The main models we considered were Active Models
Z B and H [53, 90]. For Active Model B, we defined a lo-
1 cal entropy production density and confirmed this to be
Sφ = hJg .Jgd i(r)dr. (56)
D strongly localized at the interface between phases, with
There are no cross terms between rotational and gra- a bulk contribution that is much weaker at low noise lev-
dient contributions because Jg .(∇ ∧ A)dr vanishes on
R els, by constructing spatial maps to quantify where ac-
partial integration. (This allows either of Jr or Jg to tivity plays a role. We also offered a generalisation of the
be replaced by the full current J in (55,56).) The result Harada-Sasa relation (HSR) to field theories, quantita-
(56) is compatible with the forms (18,21) found above for tively relating the entropy production rate to the viola-
σφ (r), although µA is now a nonlocal functional 4 of φ. tion of the fluctuation-dissipation theorem (FDT). This
We listed in (20) several, potentially inequivalent, allows entropy production to be spectrally decomposed
forms for the local entropy production σφ differing by across spatial and temporal Fourier modes, even in trans-
transient contributions and/or formal entropy currents. lationally invariant cases where real-space maps convey
Further options arise for the correspondingly defined lo- no useful information.
cal quantity σJ ; we explore these elsewhere [92]. Mean- Assigning wave-number dependence to an entropy pro-
while, the simplest case of an additional entropy pro- duction allows one to give quantitative underpinning to
duction arising via (55) is when the current contains a a statement like “this active system is effectively passive,
deterministic rotational part Jrd that is non-vanishing as but only at scales larger than the correlation length”.
D → 0. The resulting contribution to SJ is The quantitative link between FDT violation and TRS
Z Z breakdown reinforces the fact that TRS is, first, a scale-
1 r 2 1 dependent phenomenon, and second, one that depends
|Jd | dr = |∇ ∧ A|2 dr. (57)
D D on which variables are retained in a coarse-grained de-
scription. Absence of FDT violation ensures zero entropy
production, but only if the FDT is tested across the full
4 Its calculation involves first constructing µ as a Coulomb integral subset of dynamical fields that we wish to describe. As
of ∇ · J, and then identifying as µA the part of it that cannot be we showed for Active Model H, the requirements are not
represented as a free energy derivative. always obvious. It might therefore be wise to focus first

11
on the entropy production, and only when the form of Appendix A: Additivity and locality of entropy
this is clear turn to Harada-Sasa type constructions to production
make an explicit connection to FDT violations. This
should ensure that no important terms are overlooked. For Active Model B and its relatives we established re-
An open task is to construct a HSR capable of detecting lations such as (9,17) relating the global steady state en-
the entropy production from circulating currents that are tropy production rate S to the integral of a local quantity
invisible in the dynamics of the density field φ(r, t), as σ(r). This shows that the entropy production is additive
exemplified by Active Model B+. over regions of space within a larger system. However,
In scalar active matter the micro-dynamics is usually if we were to isolate one of these regions as a subsys-
very far from equilibrium, despite which several approx- tem, its entropy production rate would depend on what
imate mappings to equilibrium. Coarse-grained approxi- boundary conditions were imposed. Only in the ther-
mations for the dynamics have been proposed in a num- modynamic limit where all subsystems are much larger
ber of systems [93–96], establishing at times a surprising than the correlation length, do the contributions from
connection to equilibrium [10, 97]. Note the question such internal boundaries become unimportant.
of whether a given non-TRS dynamics can be approx- In this Appendix we consider the case where one spec-
imated, at coarse grained level, by one with TRS is a ifies fixed information about the time evolution of φ and
distinct issue from whether the original microscopic dy- its gradients on each internal boundary which is then
namics is close to equilibrium [44]. In this article we pro- shared by the subsystems on either side of that bound-
vide the tools to decided whether TRS breaking survives ary. We examine the additivity of the entropy production
coarse-graining in active matter systems. in this case, and discuss conditions under which it makes
In some active matter systems, mesoscopic or macro- sense to view σ(r) as a local entropy production density.
scopic violations of TRS are obvious – for instance when We consider a region of space Ω1 , with Ω2 its comple-
fluid vortices are visible at the scale of interest. Our ma- ment; ∂Ω is the boundary separating the two regions. We
chinery can quantify such violations, but is likely to be restrict attention to local field theories, defined as those
more useful in cases where they are harder to detect. For in which the trajectories within Ω1 have no further de-
example, it would be an interesting follow up of this work pendence on those in Ω2 once the boundary information
to apply the theoretical results developed here to Model χ is given. The probability for the system to follow a
B coupled to a logistic birth/death term which causes given trajectory ψ which equals ψ 1 (r, t) inside Ω1 , and
arrest of phase separation at a particular scale [98]. Par- ψ 2 (r, t) in Ω2 , can then be written as
ticles (representing bacteria) divide in dilute regions and
die off in dense ones, manifestly violating TRS at the P[ψ] = P1 [ψ 1 |χ] P2 [ψ 2 |χ] P∂Ω [χ] , (A1)
level of currents. It would be interesting to know if such
a model still has finite entropy production when the den- where Pi [ψ i |χ] is the conditional probability of finding
sity field alone is monitored. ψ i in Ωi , given that these fields and their derivatives
satisfy some boundary information χ(t) on ∂Ω, whose
Finally, there is clearly a connection between the
unconditional probability is P∂Ω [χ].
idea of scale-dependent entropy production and the
For Active Model B, Eq. (A1) is obviously true if one
renormalization-group concept of “relevance”. However,
works with the current (ψ = J) of which the action A[J]
these are not interchangeable, since activity could be rel-
is a local functional. Less obviously, this is true also
evant (changing a universality class) without itself sur-
working with the density φ, so long as the boundary data
viving coarse-graining (the new class might restore TRS
χ contains enough derivatives of φ both to specify (φ̇ +
at its fixed point). We hope our work will inform future
∇ · Jd ) on ∂Ω and to ensure that the Poisson equation
RG studies both of scalar active field theories and the
more complex vector and tensor theories of active mat- ∇2 f = (φ̇ + ∇ · Jd ) has an unique solution inside Ωi . In
ter [1, 99–102], for which RG studies of are starting to such cases one has
systematically uncover new classes. Wi [ψ i |χ]
Pi [ψ i |χ] = (A2)
Zi [χ]

where Wi is a statistical weight for the conditioned trajec-


ACKNOWLEDGMENTS
tories in Ωi , and Zi [χ] its integral over such trajectories.
For any local action A, we have Wi = exp −Ai [ψ i ] for
We thank Hugues Chaté, Romain Mari, Fred MacKin- trajectories respecting the boundary data and Wi = 0
tosh, Davide Marenduzzo, Thomas Speck, David Tong, otherwise; here Ai is an action integrated only over the
Paolo Visco and Raphael Wittkowski for illuminating dis- domain Ωi . From this follows the statistical weight of
cussions. C. Nardini acknowledges the hospitality pro- the boundary data itself, W∂Ω [χ] = Z1 [χ]Z2 [χ], and its
vided by DAMTP, University of Cambridge while this normalized probability density
work was being done. This research has been supported
by EPSRC grant EP/J007404. MEC holds a Royal So- Z1 [χ]Z2 [χ]
P∂Ω [χ] = R . (A3)
ciety Research Professorship. Z1 [χ]Z2 [χ] D[χ]

12
Substituting (A2,A3) in (A1) for the forward probability In (A8-A11), Si , P and W are now functionals of config-
P[ψ], and then repeating exactly the same calculation urations and not trajectories.
for the backward probability P R [ψ], gives These arguments for qualified additivity support the
interpretation of σ(r) = hσ̂i in (9) as a local entropy
S = hS1 [χ]iP∂Ω + hS2 [χ]iP∂Ω + Sχ (A4) density, on the basis that σ̂(ψ, ∇ψ, ...) stands conceptu-
ally in relation to the global entropy production S just
where the entropy production of the boundary process is as the equilibrium entropy density S(ψ, ∇ψ, ...) stands in
written relation to the global equilibrium entropy S. Note that in
1

P∂Ω
 the main text we have mainly focused on the weak noise
Sχ = lim ln R . (A5) limit, for which the integral in (9) is performed at a suffi-
τ →∞ τ P∂Ω P∂Ω
ciently coarse-grained level that long-wavelength fluctua-
tion contributions are unimportant. In equilibrium mod-
In (A4), the first two terms sum the entropy productions
els this would correspond to a treatment of S(ψ, ∇ψ, ...)
in our two subsystems; each of these is first calculated
at mean-field or density-functional level.
with quenched boundary data χ as
Having established qualified additivity, and given the
1

Pi
 non-negativity of S, a natural question is whether the
τ τ
Si [χ] = lim Si [χ] Si [χ] = ln R entropy production density σ(r) is itself non-negative.
τ →∞ τ Pi Pi [ψ |χ]
i In partial answer to this, observe first that the various
(A6) contributions in (A4) are non-negative since the relevant
and then averaged over χ (notated as h·iP∂Ω in (A4)). integral fluctuation theorems [38] apply equally to con-
From (A2) we obtain ditional and full probabilities. We thus have:
Si [χ] = 1

P∂Ω

τ
Z 
1 ZiR [χ] Si [χ] ≥ 0, ln R ≥ 0. (A12)
= σ̂(ψ, ∇ψ, ...)dr + lim ln τ P∂Ω P∂Ω
τ →∞ τ Zi [χ]
Ωi Pi [ψ i |χ]
(A7) This might seems to imply the positivity of σ(ψ, ∇ψ, ...)
with σ̂(ψ, ∇ψ, ...) a local function of ψ and its gradients. when the continuum limit is taken. However, due to the
The first term represents the bulk contribution to S while presence of the second term of (A7), this is actually not a
the second only depends on the boundary data. This correct inference unless the statistics of the boundary in-
does not vanish in general, but it does so whenever the formation χ itself exhibits TRS, so that the second term
boundary information χ itself exhibits TRS. vanishes. This suggests (as is physically reasonable) that
Thus the entropy production rate S for two systems any local region of negative σ must be driven, through its
with a shared boundary is strictly additive only if a movie boundary, by a larger, positive entropy production hap-
of events on the boundary itself looks statistically the pening elsewhere. We defer to future work the question
same when shown in reverse. Generically this cannot be of whether such situations can arise in practice for the
expected, so that there is indeed a finite (sub-extensive) active field theories studied here.
boundary contribution, which becomes negligible only
if the subsystems are large compared to the correlation
length. Appendix B: Discretised dynamics
This type of ‘qualified additivity’ is familiar from the
study of systems in equilibrium. Indeed, using arguments In this Appendix, we detail how to perform spatial dis-
that exactly parallel those above, one can show that in cretisation such that detailed balance is always recovered
equilibrium the global entropy defined as in the equilibrium limit for Active Model B. This is the
Z discretisation we used to numerically integrate the model,
S = − P[ψ] ln(P[ψ])D[ψ] (A8) giving the results in Sec. I.
We consider here 1D Model B (µA = 0), although it is
obeys (analogous to (A4)) easy to extend the results of this Appendix to higher
dimensions. We consider a system of finite width L
S = hS1 [χ]iP∂Ω + hS2 [χ]iP∂Ω + Sχ (A9) such that x ∈ [0, L] with periodic boundary conditions.
We discretise x into N lattice points with equal lattice
with the boundary contribution to the entropy spacing ∆ so that N ∆ = L, and the density field as
Z φ(x, t) → φi (t), where i = 1, 2, . . . N ; φi is the value of φ
at x = i∆. The dynamics then becomes:
Sχ = − P∂Ω [χ] ln(P∂Ω [χ]) D[χ] . (A10)
∂F √
Moreover, in analogy with (A7) we find ∂ t φ i = ∇2 + 2D∇ηi , (B1)
∂φi
Si [χ] = − hln W [ψ i |χ]iPi [ψ |χ] + ln Zi [χ] (A11) with hηi (t)ηj (t0 )i = δij δ(t − t0 ). We now choose the fol-
i

13
lowing discretisation for the gradient operator ∇: This is divergent as ∆ → 0, both in the part that de-
pends on f 00 (φi ) and in the second, constant contribu-
ψi+1 − ψi−1 tion. However one still has AS − ARS = 0, as promised,
∇ψi = , (B2)
2∆ which is all we need for the results of the main text. The
for any discrete fields ψi , which implies the discretisation use of any other convention and/or discretization would
for the Laplacian operator: give the same final result for S once the counterpart of
AS −ARS , which is no longer zero, is properly worked out.
ψi+2 − 2ψi + ψi−2
∇2 ψi = . (B3)
4∆2
Appendix D: Spatial discretization
To show detailed balance is satisfied, we substitute
Eqs. (B2,B3) into (B1) to obtain:
√ The result (21) follows from (18) only if hµA ∇.Λi = 0.
1 ∂F 2D This would hold automatically if the stochastic integrals
∂t φi = − 2 Aij + Bij ηj , (B4) were interpreted in the Ito convention in which the time
4∆ ∂φj 2∆
derivative is evaluated at the start of the timestep; how-
where the matrices A and B are given by: ever in this paper we use the (mid-timestep) Stratonovich
  convention, which anticipates part of the subsequent in-
2 0 −1 crement so that φ̇ cannot be simply replaced by ∇2 µ to
 0 2 0 −1  give (21) from (18). Here we address only the special case
 
 −1 0 2 0 −1  of Active Model B in one dimension. We show that (21)
 
A= .. , (B5) still holds, so long as we employ mid-point spatial dis-
 . 
  cretisation.
 . .. 
  Any stochastic integral in the Stratonovich convention
−1 0 2 can be transformed in one in the Ito convention. Sub-
tleties however arise when dealing with stochastic PDEs;
and these are closely linked to the AS term in the dynami-
0 1
  cal action discussed in Appendix C above. Let us first
consider a stochastic differential equation in the form
 −1 0 1
(xi ∈ R, x = (x1 , ..., xn ))


B= −1 0 −1 
, (B6)
 .. 
 .  ẋi = ai (x) + bij ηj , (D1)
−1 0
where hηi (t)ηj (s)i = δij δ(t − s). We want to consider the
with zero elements where not explicitly written. We ob- following Stratonovich integral
serve BB T = B T B = A, which is the condition for
this discretised dynamics to obey exactly detailed bal-
Z t

ance [83, 103]. Using the same discretization for the ac- Iil = fi (x(s))ηl (s)ds, (D2)
0
tive term µA , this ensures that TRS is respected fully in
the limit µA → 0. where x(s) satisfies (D1). Iil can be converted in an Ito
plus a non-stochastic (Riemann) integral as follows [108]
Appendix C: Time-symmetric contribution to A Z t
Iil = fi (x(s)) · ηl (s)ds + Iilconv , (D3)
0
The time symmetric contribution that was omitted
from (11) depends on the prescription used for defining where · denotes that the integral has to be understood in
stochastic integrals (Ito, Stratonovich, or intermediate the Ito sense, and
[104–107]) and also on the spatial discretisation used; in-
deed subscript S denotes our adopted Stratonovich con- Z t
1 ∂fi
vention. As discussed in Appendix B above, we use mid- Iilconv = (x(s))bjl ds. (D4)
2 0 ∂xj
point discretization to maintain exact TRS in the passive
limit. We therefore present here only the form of this This result, when formally generalised to the case of Ac-
term, notated AS in the following, for Active Model B in tive Model B, produces ill-defined formulae, involving
the Stratonovich convention with midpoint discretization the square of a Dirac delta. In order to progress, we
and the choice µA = λ|∇φ|2 . This is straightforwardly then midpoint-discretise the dynamics (1)-(4) as in Ap-
obtained from Eq. (5.15) in [104] as pendix B, with periodic boundary conditions. We have
1 X 00


 √
R
AS = AS = − 2 f (φi ) + . (C1) µi+2 − 2µi + µi−2 2D
2∆ i 4∆2 φ̇i = + Bij ηj , (D5)
4∆2 2∆

14
where B obeys (B6), that is Bij = δi+1,j − δi−1,j . Now pendix. We have
consider the part of the entropy production coming from 1
Z h i
the stochastic integral: A[φ] = − φ̇0 + ∇ · Jd [φ0 ]
4D
h i
∆ X
Z t ∇−2 φ̇0 + ∇ · Jd [φ0 ] drdt
I=− µA,i (ηi+1 − ηi−1 ). (D6)
2tD∆ i
Z h
0 1 i
− √ φ̇0 + ∇ · Jd [φ0 ] ∇−2
2 D
  
Applying (D4) and using the discrete form of µA 2 δF0
φ̇1 − ∇ + 2λ∇φ0 · ∇φ1 drdt
δφ1
 2 Z   
φi+1 − φi−1 1 δF0
µA,i = λ , (D7) − φ̇1 − ∇ 2
+ 2λ∇φ0 · ∇φ1 ∇−2
2∆ 4 δφ1
  
2 δF0
we have I conv = 0. We conclude that, with midpoint spa- φ̇1 − ∇ + 2λ∇φ0 · ∇φ1 drdt
δφ1
tial discretisation, the expression for the entropy produc- √
tion (18) can be equally interpreted in Ito or Stratonovich + O( D) ,
conventions. Then, using the non-anticipating property (E1)
of the Ito convention, (18) implies (21). where we have used the definition of F0 in (25). In the
small noise limit, the first term in (E1) must vanish to
This result requires the integrand µA,i in (18) to de-
avoid any divergence, yielding the mean-field equation:
pend only on i ± 1 and not on i ± 2. If we wanted to write
φ̇0 = −∇ · Jd (φ0 ). The second term thus also vanishes
the entropy production in the equivalent form
and the third one corresponds to the dynamics of φ1 given
Z by (24).
1
S=− hµφ̇i(r)dr, (D8)
D
Appendix F: Harada-Sasa for Active Model H
we would find (with midpoint spatial discretisation) that
the conversion from Stratonovich to Ito brings in a term We first prove that the action of Active Model H is
given by (44). Introducing the noise vector Υα = ∂β Γαβ ,
Z  its correlations read

1 ∆ X 3κ 00
− lim + f (φ i ) dt, (D9)
t→∞ t 2∆2 4∆2 hΥα (r, t)Υβ (r0 , t0 )i =
i (F1)
= −2ηD δαβ ∇2 + ∂α ∂β δ(r − r0 )δ(t − t0 ) .


which does not admit a continuous limit for ∆ → 0. How- The dynamic action corresponding to the fluid dynamics
ever this divergence cancels against another term is
Z
1
Υα (r, t)Ξαβ (r − r0 , t − t0 )Υβ (r0 , t0 ), (F2)
Z
1 Aη =
− hµ · ∇Λi(r)dr, (D10) 2 r,r0 ,t,t0
D
where
that appears in the Ito but not the Stratonovich inte- Z
gral for S. These fact have been checked in our numeri- Ξαγ (r − r00 , t − t00 ) hΥγ (r00 , t00 )Υβ (r0 , t0 )i =
r00 ,t00 (F3)
cal simulations of Active Model B. Let us finally observe
that (D9) is actually proportional to AS ; see (C1). This = δαβ δ(r − r0 )δ(t − t0 ).
is however non generic and due to the fact that the noise We get the explicit expression for Ξ as
in the dynamics is additive and BB T = B T B = A.
Ξαβ (r − r0 , t − t0 ) =
We conclude by noting that analogous apparent diver-  
gences would be found using other type of spatial dis- 1 1
=− ∇−2 δαβ − ∇−2 ∂α ∂β δ(r − r0 )δ(t − t0 ) .
cretisation or Fourier truncations in the numerical com- 2ηD 2
putation of the entropy production. (F4)
Using incompressibility, the contribution of the fluid dy-
namics to the total action follows as in (44).
We now consider the derivation of the generalised
Appendix E: Small noise expansion Harada-Sasa relation for Active Model H, stated in (50).
It is useful to start from an expression of the entropy
production equivalent to (47) up to boundary terms:
We expand the dynamic action with the D-expansion Z
of φ in (22). Since we use this action only for book- 1
keeping, we choose the simpler Itō convention in this ap- Sφ,v = − hµ(∂t + v · ∇)φ + vαβ Σαβ i dr. (F5)
D

15
This can be rewritten as dynamic action at linear order in ε reads
Z
1
δA = − ∂β εαβ ∇−2 [(∂t + vγ ∂γ ) vα
Z D 2ηD r,t
1
Sφ,v = − lim µ(r, 0) [(∂t + vα ∂α ) φ] (r, t) −η∇2 vα − ∂γ ΣAαγ + ∂α p

2D t→0 Z (F12)
+ µ(r, t) [(∂t + vα ∂α ) φ] (r, 0) (F6) 1
− ∂α ∂γ εαγ ∇−4 (∂β vµ ∂µ vβ
4ηD r,t
E
+ vαβ (r, 0)Σαβ (r, t) + vαβ (r, t)Σαβ (r, 0) dr
−∂β ∂µ ΣAβµ + ∇2 p + O(ε2 ) .


Now employing a formula analogous to (32), the response


The dynamics perturbed as µA → µA − h is given by follows as
1

R(r, r, t) = − vαβ (r, t) ∂β ∇−2 [(∂t + vγ ∂γ ) vα


2ηD
(∂t + vβ ∂β ) φ = −∂α Jα , Jα = −∂α (µ − h)+Γα . (F7)
−η∇2 vα − ∂γ ΣAαγ + ∂α p (r, 0)


+ vαβ (r, t) ∂α ∂β ∇−4 (∂γ vµ ∂µ vγ


The corresponding dynamical action shift at linear order 4ηD
in h reads −∂γ ∂µ ΣAγµ + ∇2 p (r, 0) .


(F13)
Z We now observe that we only need the integral over space:
1 2
µ drdt + O h2 .
  
δA = − h (∂t + vβ ∂β ) φ − ∂αα Z Z
2D 1
R(r, r, t)dr = hvαβ (r, t)vαβ (r, 0)i dr
(F8) 2D
From this follows the response 1
Z
+ hvαβ (r, t)Σαβ (r, 0)i dr (F14)
4ηD
Z
1
R(r, r, t) = + hvα (r, t) (∂t + vγ ∂γ ) vα (r, 0)i dr,
4ηD
1
 2

[(∂t + vα ∂α ) φ] (r, t) (∂t + vβ ∂β ) φ − ∂ββ µ (r, 0) , where we have used the incompressibility condition and
2D
(F9) integration by parts to eliminate both the pressure term
and its symmetrized form and the contribution given by the third and the fourth
line of (F13). In the t → 0 limit, we have
Z
R(r, r, t) + R(r, r, −t) = lim hvα (r, t) (∂t + vγ ∂γ ) vα (r, 0)i dr =
t→0
1 (F15)
h[(∂t + vα ∂α ) φ] (r, t) [(∂t + vβ ∂β ) φ] (r, 0)i Z  
D 1
= ∂t vα2 (r, 0) + vα vγ ∂γ vα (r, 0) dr = 0.
1 2 (F10) 2
− ∇ h[(∂t + vα ∂α ) φ] (r, t)µ(r, 0)i
2D (2) Using the definition of C in (49) we obtain
1 2
− ∇ h[(∂t + vα ∂α ) φ] (r, 0)µ(r, t)i . Z Z
2D (2) lim [R(r, r, t) + R(r, r, −t)] dr =
1
lim C (r, r, t)dr
t→0 D t→0
Z
1
We see that the first two lines of (F6) can be written + lim hvαβ (r, t)Σαβ (r, 0)i dr
4ηD t→0
Z
1
+ lim hvαβ (r, 0)Σαβ (r, t)i dr.
1 4ηD t→0
h[(∂t + vα ∂α ) φ] (r, t)µ(r, 0)i + (F16)
2
1 Plugging Eqs. (F10) and (F16) in (F6), we finally get
h[(∂t + vα ∂α ) φ] (r, 0)µ(r, t)i
2 Sφ,v =
= ∇−2 −2
(2) C (r, r, t) − D∇(2) [R(r, r, t) + R(r, r, −t)] , 1
Z
(F11) lim ∇−2 (2) {D [R(r, r, t) + R(r, r, −t)] − C (r, r, t)} dr
D t→0
where we employed the definition (49) for the correlation. 2η
Z
+ lim {C(r, r, t) − D [R(r, r, t) + R(r, r, −t)]} dr.
To obtain the last line in the expression for the entropy D t→0
production (F6), we now consider that the response with (F17)
respect to the perturbation ΣA → ΣA − ε with εαβ sym- The generalized Harada-Sasa relation (50) is deduced by
metric in the exchange of α and β and traceless. The Fourier transforming in space and time.

16
[1] MC Marchetti, JF Joanny, S Ramaswamy, TB Liver- of state in suspensions of active colloids,” Physical
pool, J Prost, Madan Rao, and R Aditi Simha, “Hy- Review X 5, 011004 (2015).
drodynamics of soft active matter,” Reviews of Modern [18] Sho C Takatori and John F Brady, “Towards a ther-
Physics 85, 1143 (2013). modynamics of active matter,” Physical Review E 91,
[2] Tamás Vicsek, András Czirók, Eshel Ben-Jacob, Inon 032117 (2015).
Cohen, and Ofer Shochet, “Novel type of phase transi- [19] T F F Farage, P Krinninger, and Joseph M Brader,
tion in a system of self-driven particles,” Physical review “Effective interactions in active brownian suspensions,”
letters 75, 1226 (1995). Physical Review E 91, 042310 (2015).
[3] Guillaume Grégoire and Hugues Chaté, “Onset of col- [20] Umberto Marini Bettolo Marconi and Claudio Maggi,
lective and cohesive motion,” Physical review letters 92, “Towards a statistical mechanical theory of active flu-
025702 (2004). ids,” Soft matter 11, 8768–8781 (2015).
[4] Alexandre P Solon, Hugues Chaté, and Julien Tailleur, [21] Alexandre P. Solon, Joakim Stenhammar, Michael E.
“From phase to microphase separation in flocking mod- Cates, Yariv Kafri, and Julien Tailleur, “Generalized
els: The essential role of nonequilibrium fluctuations,” thermodynamics of phase equilibria in scalar active mat-
Physical review letters 114, 068101 (2015). ter,” arXiv preprint arXiv:1609.03483 (2016).
[5] Julien Deseigne, Olivier Dauchot, and Hugues Chaté, [22] Leticia F Cugliandolo, “The effective temperature,”
“Collective motion of vibrated polar disks,” Physical re- Journal of Physics A: Mathematical and Theoretical 44,
view letters 105, 098001 (2010). 483001 (2011).
[6] Antoine Bricard, Jean-Baptiste Caussin, Nicolas [23] Davide Loi, Stefano Mossa, and Leticia F Cugliandolo,
Desreumaux, Olivier Dauchot, and Denis Bartolo, “Effective temperature of active matter,” Physical Re-
“Emergence of macroscopic directed motion in popu- view E 77, 051111 (2008).
lations of motile colloids,” Nature 503, 95–98 (2013). [24] Davide Loi, Stefano Mossa, and Leticia F Cugliandolo,
[7] Henricus H Wensink, Jörn Dunkel, Sebastian Heiden- “Effective temperature of active complex matter,” Soft
reich, Knut Drescher, Raymond E Goldstein, Hartmut Matter 7, 3726–3729 (2011).
Löwen, and Julia M Yeomans, “Meso-scale turbulence [25] Demian Levis and Ludovic Berthier, “From single-
in living fluids,” Proceedings of the National Academy particle to collective effective temperatures in an active
of Sciences 109, 14308–14313 (2012). fluid of self-propelled particles,” EPL 111, 60006 (2015).
[8] Volker Schaller, Christoph Weber, Christine Semmrich, [26] Boris Lander, Jakob Mehl, Valentin Blickle, Clemens
Erwin Frey, and Andreas R Bausch, “Polar patterns of Bechinger, and Udo Seifert, “Noninvasive measurement
driven filaments,” Nature 467, 73–77 (2010). of dissipation in colloidal systems,” Physical Review E
[9] Yutaka Sumino, Ken H Nagai, Yuji Shitaka, Dan 86, 030401 (2012).
Tanaka, Kenichi Yoshikawa, Hugues Chaté, and [27] Daisuke Mizuno, Catherine Tardin, C. F. Schmidt, and
Kazuhiro Oiwa, “Large-scale vortex lattice emerging F. C. MacKintosh, “Nonequilibrium mechanics of active
from collectively moving microtubules,” Nature 483, cytoskeletal networks,” Science 315, 370–373 (2007).
448–452 (2012). [28] Claire Wilhelm, “Out-of-equilibrium microrheology in-
[10] J. Tailleur and M. E. Cates, “Statistical mechanics of side living cells,” Phys. Rev. Lett. 101, 028101 (2008).
interacting run-and-tumble bacteria,” Phys. Rev. Lett. [29] Francois Gallet, Delphine Arcizet, Pierre Bohec, and
100, 218103 (2008). Alain Richert, “Power spectrum of out-of-equilibrium
[11] Yaouen Fily and M Cristina Marchetti, “Athermal phase forces in living cells: amplitude and frequency depen-
separation of self-propelled particles with no alignment,” dence,” Soft Matter 5, 2947–2953 (2009).
Physical review letters 108, 235702 (2012). [30] É. Fodor, M. Guo, N. S. Gov, P. Visco, D. A. Weitz, and
[12] Michael E. Cates and Julien Tailleur, “Motility-induced F. van Wijland, “Activity-driven fluctuations in living
phase separation,” Annu. Rev. Condens. Matter Phys. cells,” EPL 110, 48005 (2015).
6, 219–244 (2015). [31] W. W. Ahmed, E. Fodor, M. Almonacid, M. Busson-
[13] Sho C Takatori, Wen Yan, and John F Brady, “Swim nier, M.-H. Verlhac, N. S. Gov, P. Visco, F. van Wij-
pressure: stress generation in active matter,” Physical land, and T. Betz, “Active mechanics reveal molecular-
review letters 113, 028103 (2014). scale force kinetics in living oocytes,” ArXiv e-prints
[14] Xingbo Yang, M Lisa Manning, and M Cristina arXiv:1510.08299.
Marchetti, “Aggregation and segregation of confined ac- [32] Thierry Bodineau and Bernard Derrida, “Cumulants
tive particles,” Soft matter 10, 6477–6484 (2014). and large deviations of the current through non-
[15] Thomas Speck, Julian Bialké, Andreas M Menzel, and equilibrium steady states,” Comptes Rendus Physique
Hartmut Löwen, “Effective cahn-hilliard equation for 8, 540–555 (2007).
the phase separation of active brownian particles,” [33] Thierry Bodineau and Bernard Derrida, “Current fluc-
Physical Review Letters 112, 218304 (2014). tuations in nonequilibrium diffusive systems: an ad-
[16] AP Solon, ME Cates, and J Tailleur, “Active brownian ditivity principle,” Physical review letters 92, 180601
particles and run-and-tumble particles: A comparative (2004).
study,” The European Physical Journal Special Topics [34] Lorenzo Bertini, Alberto De Sole, Davide Gabrielli, Gi-
224, 1231–1262 (2015). anni Jona-Lasinio, and Claudio Landim, “Current fluc-
[17] Félix Ginot, Isaac Theurkauff, Demian Levis, tuations in stochastic lattice gases,” Physical review let-
Christophe Ybert, Lydéric Bocquet, Ludovic Berthier, ters 94, 030601 (2005).
and Cécile Cottin-Bizonne, “Nonequilibrium equation [35] Lorenzo Bertini, Alberto De Sole, Davide Gabrielli, Gio-

17
vanni Jona-Lasinio, and Claudio Landim, “Macroscopic hammar, Rosalind J Allen, Davide Marenduzzo, and
fluctuation theory,” Reviews of Modern Physics 87, 593 Michael E Cates, “Scalar ϕ4 field theory for active-
(2015). particle phase separation,” Nature communications 5
[36] Sebastian Pilgram, Andrew N Jordan, E V Sukhorukov, (2014).
and M Büttiker, “Stochastic path integral formulation [54] Joakim Stenhammar, Adriano Tiribocchi, Rosalind J
of full counting statistics,” Physical review letters 90, Allen, Davide Marenduzzo, and Michael E Cates, “Con-
206801 (2003). tinuum theory of phase separation kinetics for active
[37] Christopher Battle, Chase P Broedersz, Nikta Fakhri, brownian particles,” Physical review letters 111, 145702
Veikko F Geyer, Jonathon Howard, Christoph F (2013).
Schmidt, and Fred C MacKintosh, “Broken detailed [55] M. E. Cates and J. Tailleur, “When are active brownian
balance at mesoscopic scales in active biological sys- particles and run-and-tumble particles equivalent? con-
tems,” Science 352, 604–607 (2016). sequences for motility-induced phase separation,” EPL
[38] Udo Seifert, “Stochastic thermodynamics, fluctuation 101, 20010 (2013).
theorems and molecular machines,” Reports on Progress [56] Pierre C Hohenberg and Bertrand I Halperin, “The-
in Physics 75, 126001 (2012). ory of dynamic critical phenomena,” Reviews of Modern
[39] Joel L Lebowitz and Herbert Spohn, “A gallavotti– Physics 49, 435 (1977).
cohen-type symmetry in the large deviation functional [57] Paul M Chaikin and Tom C Lubensky, Principles
for stochastic dynamics,” Journal of Statistical Physics of condensed matter physics, Vol. 1 (Cambridge Univ
95, 333–365 (1999). Press, 2000).
[40] Christian Maes, “The fluctuation theorem as a gibbs [58] Mehran Kardar, Giorgio Parisi, and Yi-Cheng Zhang,
property,” J. Stat. Phys. 95, 367–392 (1999). “Dynamic scaling of growing interfaces,” Physical Re-
[41] Jorge Kurchan, “Fluctuation theorem for stochastic dy- view Letters 56, 889 (1986).
namics,” Journal of Physics A: Mathematical and Gen- [59] Samuel F Edwards and DR Wilkinson, “The surface
eral 31, 3719 (1998). statistics of a granular aggregate,” in Proceedings of
[42] David A Egolf, “Equilibrium regained: From nonequi- the Royal Society of London A: Mathematical, Physical
librium chaos to statistical mechanics,” Science 287, and Engineering Sciences, Vol. 381 (The Royal Society,
101–104 (2000). 1982) pp. 17–31.
[43] Shou-Wen Wang, Kyogo Kawaguchi, Shin-ichi Sasa, [60] William G Faris and Giovanni Jona-Lasinio, “Large fluc-
and Lei-Han Tang, “Entropy production of nano tuations for a nonlinear heat equation with noise,” Jour-
systems with timescale separation,” arXiv preprint nal of Physics A: Mathematical and General 15, 3025
arXiv:1601.04463 (2016). (1982).
[44] Stefano Bo and Antonio Celani, “Entropy production [61] Hugo Touchette, “The large deviation approach to sta-
in stochastic systems with fast and slow time-scales,” tistical mechanics,” Physics Reports 478, 1–69 (2009).
Journal of Statistical Physics 154, 1325–1351 (2014). [62] Hugo Touchette and Rosemary J Harris, “Large devia-
[45] David A Egolf, “Equilibrium regained: From nonequi- tion approach to nonequilibrium systems,” Nonequilib-
librium chaos to statistical mechanics,” Science 287, rium Statistical Physics of Small Systems: Fluctuation
101–104 (2000). Relations and Beyond , 335–360 (2012).
[46] L Cerino and A Puglisi, “Entropy production for [63] Julien Tailleur, Jorge Kurchan, and Vivien Lecomte,
velocity-dependent macroscopic forces: The problem “Mapping out-of-equilibrium into equilibrium in one-
of dissipation without fluctuations,” EPL (Europhysics dimensional transport models,” Journal of Physics A:
Letters) 111, 40012 (2015). Mathematical and Theoretical 41, 505001 (2008).
[47] Takahiro Harada and Shin-ichi Sasa, “Equality con- [64] Freddy Bouchet, Krzysztof Gawedzki, and Cesare Nar-
necting energy dissipation with a violation of the dini, “Perturbative calculation of quasi-potential in non-
fluctuation-response relation,” Phys. Rev. Lett. 95, equilibrium diffusions: A mean-field example,” Journal
130602 (2005). of Statistical Physics 163, 1157 (2016).
[48] Pyoung-Seop Shim, Hyun-Myung Chun, and Jae Dong [65] Bernard Derrida, “Microscopic versus macroscopic ap-
Noh, “Macroscopic time-reversal symmetry breaking at proaches to non-equilibrium systems,” Journal of Statis-
a nonequilibrium phase transition,” Physical Review E tical Mechanics: Theory and Experiment 2011, P01030
93, 012113 (2016). (2011).
[49] Chulan Kwon, Joonhyun Yeo, Hyun Keun Lee, and [66] David S Dean, “Langevin equation for the density of a
Hyunggyu Park, “Anomalous entropy production in system of interacting langevin processes,” J. Phys. A:
the presence of momentum-dependent forces,” arXiv Math. Gen. 29, L613 (1996).
preprint arXiv:1506.02339 (2015). [67] Masao Doi, “Second quantization representation for
[50] Debasish Chaudhuri, “Active brownian particles: En- classical many-particle system,” Journal of Physics A:
tropy production and fluctuation response,” Physical Mathematical and General 9, 1465 (1976).
Review E 90, 022131 (2014). [68] L Peliti, “Path integral approach to birth-death pro-
[51] Debasish Chaudhuri, “Entropy production by active cesses on a lattice,” Journal de Physique 46, 1469–1483
particles: Coupling of odd and even functions of ve- (1985).
locity,” arXiv preprint arXiv:1605.00269 (2016). [69] Hans-Karl Janssen, “On a lagrangean for classical field
[52] Étienne Fodor, Cesare Nardini, Michael E. Cates, Julien dynamics and renormalization group calculations of dy-
Tailleur, Paolo Visco, and Frédéric van Wijland, “How namical critical properties,” Zeitschrift für Physik B
far from equilibrium is active matter?” Phys. Rev. Lett. Condensed Matter 23, 377–380 (1976).
117, 038103 (2016). [70] Julien Tailleur, Jorge Kurchan, and Vivien Lecomte,
[53] Raphael Wittkowski, Adriano Tiribocchi, Joakim Sten- “Mapping nonequilibrium onto equilibrium: the macro-

18
scopic fluctuations of simple transport models,” Physical [88] Takahiro Harada, “Macroscopic expression connecting
review letters 99, 150602 (2007). the rate of energy dissipation with the violation of the
[71] Kirone Mallick, Moshe Moshe, and Henri Orland, “A fluctuation response relation,” Physical Review E 79,
field-theoretic approach to non-equilibrium work iden- 030106R (2009).
tities,” Journal of Physics A: Mathematical and Theo- [89] D Villamaina, A Puglisi, and A Vulpiani, “The fluc-
retical 44, 095002 (2011). tuation dissipation relation in sub diffusive systems:
[72] T Leonard, B Lander, U Seifert, and T Speck, the case of granular single-file diffusion,” J. Stat. Mech.
“Stochastic thermodynamics of fluctuating density 2008, L10001 (2008).
fields: non-equilibrium free energy differences under [90] Adriano Tiribocchi, Raphael Wittkowski, Davide
coarse-graining,” The Journal of chemical physics 139, Marenduzzo, and Michael E Cates, “Active model h:
204109 (2013). Scalar active matter in a momentum-conserving fluid,”
[73] Pablo I Hurtado, Carlos Pérez-Espigares, Jesús J del Physical Review Letters 115, 188302 (2015).
Pozo, and Pedro L Garrido, “Symmetries in fluctua- [91] Alan J Bray, “Theory of phase-ordering kinetics,” Ad-
tions far from equilibrium,” Proceedings of the National vances in Physics 43, 357–459 (1994).
Academy of Sciences 108, 7704–7709 (2011). [92] Elsen Tjhung, Cesare Nardini, Raphael Wittkowski,
[74] Giacomo Gradenigo, Andrea Puglisi, and Alessan- and EM Cates, “Light-induced self-assembly of active
dro Sarracino, “Entropy production in non-equilibrium rectification devices,” in preparation.
fluctuating hydrodynamics,” The Journal of chemical [93] Eric Bertin, Michel Droz, and Guillaume Grégoire,
physics 137, 014509 (2012). “Boltzmann and hydrodynamic description for self-
[75] Shoichi Toyabe, Tetsuaki Okamoto, Takahiro propelled particles,” Physical Review E 74, 022101
Watanabe-Nakayama, Hiroshi Taketani, Seishi Kudo, (2006).
and Eiro Muneyuki, “Nonequilibrium energetics of a [94] Tanniemola B Liverpool and M Cristina Marchetti,
single f 1-atpase molecule,” Physical review letters 104, “Bridging the microscopic and the hydrodynamic in ac-
198103 (2010). tive filament solutions,” EPL (Europhysics Letters) 69,
[76] Shoichi Toyabe, Hong-Ren Jiang, Takenobu Nakamura, 846 (2005).
Yoshihiro Murayama, and Masaki Sano, “Experimental [95] Eric Bertin, Hugues Chaté, Francesco Ginelli, Shradha
test of a new equality: Measuring heat dissipation in an Mishra, Anton Peshkov, and Sriram Ramaswamy,
optically driven colloidal system,” Physical Review E “Mesoscopic theory for fluctuating active nematics,”
75, 011122 (2007). New Journal of Physics 15, 085032 (2013).
[77] Étienne Fodor, Wylie W Ahmed, Maria Almonacid, [96] Thomas Ihle, “Kinetic theory of flocking: Derivation
Matthias Bussonnier, Nir S Gov, M-H Verlhac, Timo of hydrodynamic equations,” Physical Review E 83,
Betz, Paolo Visco, and Frédéric van Wijland, “Nonequi- 030901 (2011).
librium dissipation in living oocytes,” EPL (Europhysics [97] AG Thompson, J Tailleur, ME Cates, and RA Blythe,
Letters) 116, 30008 (2016). “Lattice models of nonequilibrium bacterial dynamics,”
[78] Matteo Paoluzzi, Claudio Maggi, Umberto Marini Bet- Journal of Statistical Mechanics: Theory and Experi-
tolo Marconi, and Nicoletta Gnan, “Critical phenom- ment 2011, P02029 (2011).
ena in active matter,” arXiv preprint arXiv:1609.03483 [98] M. E. Cates, D. Marenduzzo, I. Pagonabarraga,
(2016). and J. Tailleur, “Arrested phase separation in
[79] Ken Sekimoto, Stochastic energetics, Vol. 799 (Springer, reproducing bacteria creates a generic route to
2010). pattern formation,” Proceedings of the National
[80] L. Onsager and S. Machlup, “Fluctuations and irre- Academy of Sciences 107, 11715–11720 (2010),
versible processes,” Phys. Rev. 91, 1505–1512 (1953). http://www.pnas.org/content/107/26/11715.full.pdf.
[81] P. C. Martin, E. D. Siggia, and H. A. Rose, “Statistical [99] Leiming Chen, John Toner, and Chiu Fan Lee, “Criti-
dynamics of classical systems,” Phys. Rev. A 8, 423–437 cal phenomenon of the order–disorder transition in in-
(1973). compressible active fluids,” New Journal of Physics 17,
[82] C. De Dominicis, “A lagrangian version of halperin- 042002 (2015).
hohenberg-ma models for the dynamics of critical phe- [100] Shradha Mishra, R Aditi Simha, and Sriram Ra-
nomena,” Lettere al Nuovo Cimento (1971-1985) 12, maswamy, “A dynamic renormalization group study of
567–574 (1975). active nematics,” Journal of Statistical Mechanics: The-
[83] Nicolaas Godfried Van Kampen, Stochastic processes in ory and Experiment 2010, P02003 (2010).
physics and chemistry, Vol. 1 (Elsevier, 1992). [101] John Toner and Yuhai Tu, “Long-range order in a two-
[84] Joakim Stenhammar, Raphael Wittkowski, Davide dimensional dynamical xy model: how birds fly to-
Marenduzzo, and Michael E. Cates, “Light-induced gether,” Physical Review Letters 75, 4326 (1995).
self-assembly of active rectification devices,” Sci- [102] Leiming Chen and John Toner, “Universality for moving
ence Advances 2 (2016), 10.1126/sciadv.1501850, stripes: A hydrodynamic theory of polar active smec-
http://advances.sciencemag.org/content/2/4/e1501850.full.pdf. tics,” Physical review letters 111, 088701 (2013).
[85] R Kubo, “The fluctuation-dissipation theorem,” Reports [103] Crispin W Gardiner et al., Handbook of stochastic meth-
on Progress in Physics 29, 255 (1966). ods, Vol. 3 (Springer Berlin, 1985).
[86] Ludovic Berthier and Jorge Kurchan, “Non-equilibrium [104] A. W. C. Lau and T. C. Lubensky, “State-dependent
glass transitions in driven and active matter,” Nat. diffusion: Thermodynamic consistency and its path in-
Phys. 9, 310–314 (2013). tegral formulation,” Phys. Rev. E 76, 011123 (2007).
[87] J M Deutsch and Onuttom Narayan, “Energy dissipa- [105] Camille Aron, Daniel G Barci, Leticia F Cugliandolo,
tion and fluctuation response for particles in fluids,” Zochil Gonzalez Arenas, and Gustavo S Lozano, “Dy-
Physical Review E 74, 026112 (2006). namical symmetries of markov processes with multi-

19
plicative white noise,” arXiv preprint arXiv:1412.7564 P11018 (2010).
(2014). [107] H.K. Janssen, “Lecture notes in physics in dynamical
[106] Camille Aron, Giulio Biroli, and Leticia F Cuglian- critical phenomena and related topics,” (1979).
dolo, “Symmetries of generating functionals of langevin [108] Peter Eris Kloeden, Eckhard Platen, and Henri Schurz,
processes with colored multiplicative noise,” Journal of Numerical solution of SDE through computer experi-
Statistical Mechanics: Theory and Experiment 2010, ments (Springer Science & Business Media, 2012).

20

S-ar putea să vă placă și