This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

Properties of the dissipation functions for passive and active systems

Harsh Soni School of Physical Sciences, IIT Mandi, Kamand, Mandi, HP 175005 India
Abstract

The dissipation function for a system is defined as the natural logarithm of the ratio between probabilities of a trajectory and its time-reversed trajectory, and its probability distribution follows a well-known relation called the fluctuation theorem. Using the generic Langevin equations, we derive the expressions of the dissipation function for passive and active systems. For passive systems, the dissipation function depends only on the initial and the final values of the dynamical variables of the system, not on the trajectory of the system. Furthermore, it does not depend explicitly on the reactive or dissipative coupling coefficients of the generic Langevin equations. In addition, we study a 1D case numerically to verify the fluctuation theorem with the form of the dissipation function we obtained. For active systems, we define the work done by active forces along a trajectory. If the probability distribution of the dynamical variables is symmetric under time reversal, in both cases, the average rate of change of the dissipation function with trajectory duration is nothing but the average entropy production rate of the system and reservoir.

I Introduction

Irreversibility of a system can be quantified by using the dissipation function which is defined as the natural logarithm of the ratio of the probability density of a trajectory to that of its time-reversed trajectory. The probability distribution function of the dissipation function exhibits an interesting symmetry relation known as the fluctuation theorem [1, 2, 3]. The fluctuation theorem has been substantially explored using theory [4, 5, 6, 7, 8, 9, 10, 11, 12] and experiment [13, 14, 15, 16, 17, 18, 19]. For stochastic processes, it has been investigated mainly for the single-particle or single-variable case [5, 6, 20]. Moreover, little attention has been paid to the systems described by the Langevin equations with multiplicative noise, except for a few studies [20, 12].

This paper discusses the fluctuation relations for a wide class of systems described by the generic Langevin equations [21, 22]. Assuming that the slow variables of a system vary much slower than its microscopic degrees of freedom, one can consider that the system is always in local thermodynamic equilibrium at temperature TT. The dynamics of such systems is well-explained by the generic Langevin equations. We consider the active as well as passive systems. We use the path integral approach to calculate the probability density of a trajectory of the system with α\alpha-discritization [23, 20, 24].

Our main results are as follows. We first show that the generic Langevin equations describe a passive system, irrespective of the value of α\alpha. We then derive the expression of the dissipation function for passive systems relaxing towards thermodynamic equilibrium. Interestingly, the dissipation function is independent of the trajectory followed by the system; it only depends on the initial and the final values of the dynamical variables of the system. Moreover, it is not an explicit function of the coefficients appearing in the generic Langevin equation. Using Brownian dynamics simulation, we also verify the fluctuation theorem for a 1D single-particle problem with state-dependent diffusion. Finally, we construct an expression of the dissipation function for the active systems, and we define the work done by the active forces. For both active and passive systems, the average rate of change of the dissipation function with the time duration is the same as the rate of change of the entropy of the system and reservoir, assuming that the probability distribution of the dynamical variables is invariant under time reversal.

In the next section, we will discuss passive systems, and in section III, we will explore active systems.

II Passive systems

This section is arranged as follows. In the next subsection, we summarize the generic Langevin equations. We then calculate the ratio of the probabilities of a trajectory and its time-reversed trajectory (see subsection II.2). In subsection II.3, fluctuation relations and the dissipation function for the passive systems are presented. In subsection II.4, we talk about the quenched systems, along with an example of a system of a single colloidal particle.

II.1 Generic Langevin equations

Here we consider a passive system relaxing towards equilibrium. Its macroscopic dynamics is described by a set of nn number of slow variables 𝑨{A1,A2,.An}\bm{A}\equiv\{A_{1},A_{2},....A_{n}\}. Let AisiAiA_{i}\to s_{i}A_{i} under time reversal, where si=1s_{i}=1 and si=1s_{i}=-1 if AiA_{i} is even and odd under time reversal, respectively; e.g., si=1s_{i}=1 for position and si=1s_{i}=-1 for momentum. The generic Langevin equations for the system at temperature TT can be written in the following form [21, 22]:

dAidt=ΓijAj+kBTΓijAj+ηi(t).\dfrac{dA_{i}}{dt}=-\Gamma_{ij}\dfrac{\partial\mathcal{H}}{\partial A_{j}}+k_{\text{B}}T\dfrac{\partial\Gamma_{ij}}{\partial A_{j}}+\eta_{i}(t). (1)

where \mathcal{H} is the coarse-grained or effective Hamiltonian of the system and the coefficients Γij\Gamma_{ij} satisfy the following property:

Γij=sisjΓji.\Gamma_{ij}=s_{i}s_{j}\Gamma_{ji}. (2)

In Eq. (1), the terms with sisj=1s_{i}s_{j}=-1 are the Poisson bracket or reactive terms, whereas the terms with sisj=1s_{i}s_{j}=1 are the dissipative terms [25]. The last term ηi(t)\eta_{i}(t) represents the rapid fluctuations due to the dynamics of the microscopic degrees of freedom of the system. We assume that ηi(t)\eta_{i}(t) is white Gaussian noise, and its autocorrelation function is given by

ηi(t)ηj(t)=2kBTΓijsδ(tt),\left\langle\eta_{i}(t)\eta_{j}(t^{\prime})\right\rangle=2k_{\text{B}}T\Gamma^{\text{s}}_{ij}\delta(t-t^{\prime}), (3)

where Γijs(Γij+Γji)/2\Gamma^{\text{s}}_{ij}\equiv(\Gamma_{ij}+\Gamma_{ji})/2 is the symmetric part of Γij\Gamma_{ij}. From Eq.(2), Γijs\Gamma^{\text{s}}_{ij} shows the following symmetry property:

Γijs=sisjΓijs.\Gamma^{\text{s}}_{ij}=s_{i}s_{j}\Gamma^{\text{s}}_{ij}. (4)

Further, it is assumed to be invertible. Here in Eq. (1), we use Einstein notation, which will be carried through the rest of the paper, unless otherwise stated. Writing ηi(t)\eta_{i}(t) as the linear combination of time series of the white Gaussian noise ξj(t)\xi_{j}(t) having no correlation with each other i.e. ξi(t)ξj(t)=δijδ(tt)\left\langle\xi_{i}(t)\xi_{j}(t)\right\rangle=\delta_{ij}\delta(t-t^{\prime}):

ηi(t)=Nijξj(t),\eta_{i}(t)=N_{ij}\xi_{j}(t), (5)

where, from Eq. (3), NijN_{ij} is given by the solution of the equations:

NikNjk=2kBTΓijs.N_{ik}N_{jk}=2k_{\text{B}}T\Gamma^{\text{s}}_{ij}. (6)

Since NikN_{ik} must be real, Γijs\Gamma^{\text{s}}_{ij} must have positive eigenvalues [26]. As Γijs\Gamma^{\text{s}}_{ij} is considered to be invertible, NikN_{ik} is invertible as well. It should be noted that NjkN_{jk} is not uniquely defined by the above equation. However, NjkN_{jk} is just a dummy matrix which does not appear anywhere in the final results. Substituting (5) into Eq. (1):

dAidt=ΓijAj+kBTΓijAj+Nijξj(t).\dfrac{dA_{i}}{dt}=-\Gamma_{ij}\dfrac{\partial\mathcal{H}}{\partial A_{j}}+k_{\text{B}}T\dfrac{\partial\Gamma_{ij}}{\partial A_{j}}+N_{ij}\xi_{j}(t). (7)

The above stochastic equations have no ambiguity when NijN_{ij} does not depend explicitly on 𝑨\bm{A}. However, NijN_{ij} is the function of 𝑨\bm{A} for many systems; in such cases, the above equations are not well-defined unless their discrete scheme is specified. We here use α\alpha-discretization method to discretize the above equations [20, 23, 24], which leads to a drift of αNljNij/Al\alpha N_{lj}\partial N_{ij}/\partial A_{l} to the value of AiA_{i} due to the noise term [27, 24]. On the contrary, the noise terms in the generic Langevin equations represent the thermal fluctuations and do not contribute to the average dynamics of the slow variables. One can eliminate the noise-induced drift by adding a correction term αNljNij/Al-\alpha N_{lj}\partial N_{ij}/\partial A_{l} to Eq. (7). Therefore, the generic Langevin equations can be completely described as follows:

dAidt\displaystyle\dfrac{dA_{i}}{dt} =\displaystyle= ΓijAj+kBTΓijAjαNljNijAl+Nijξj\displaystyle-\Gamma_{ij}\dfrac{\partial\mathcal{H}}{\partial A_{j}}+k_{\text{B}}T\dfrac{\partial\Gamma_{ij}}{\partial A_{j}}-\alpha N_{lj}\dfrac{\partial N_{ij}}{\partial A_{l}}+N_{ij}\xi_{j} (8)
\displaystyle\equiv i+Nijξj\displaystyle\mathcal{F}_{i}+N_{ij}\xi_{j} (9)

with their discrete form

dAi(l)=ϵi(𝑨¯lf)+ϵNij(𝑨¯lf)ξjl,dA_{i}(l)=\epsilon\mathcal{F}_{i}(\bar{\bm{A}}^{\text{f}}_{l})+\sqrt{\epsilon}N_{ij}(\bar{\bm{A}}^{\text{f}}_{l})\xi_{j}^{l}, (10)

where ϵ\epsilon is the time step, dAi(l)Ai(ϵl)Ai(ϵ(l1))dA_{i}(l)\equiv A_{i}(\epsilon l)-A_{i}(\epsilon(l-1)), 𝑨¯lfα𝑨(ϵl)+(1α)𝑨(ϵ(l1))\bar{\bm{A}}^{\text{f}}_{l}\equiv\alpha\bm{A}(\epsilon l)+(1-\alpha)\bm{A}(\epsilon(l-1)),

iΓijAj+kBTΓijAjαNljNijAl,\mathcal{F}_{i}\equiv-\Gamma_{ij}\dfrac{\partial\mathcal{H}}{\partial A_{j}}+k_{\text{B}}T\dfrac{\partial\Gamma_{ij}}{\partial A_{j}}-\alpha N_{lj}\dfrac{\partial N_{ij}}{\partial A_{l}}, (11)

and,

ξjl1ϵ(l1)ϵlϵξj(t)𝑑t.\xi_{j}^{l}\equiv\dfrac{1}{\sqrt{\epsilon}}\int^{l\epsilon}_{(l-1)\epsilon}\xi_{j}(t)dt. (12)

are the series of random numbers having normal distribution with standard deviation 1 and mean 0. The parameter α\alpha can take any “absolute constant” between 0 and 1; α=0\alpha=0 and α=1/2\alpha=1/2 cases are referred to as Itô and Stratonovich methods, respectively. As we have another parameter α\alpha in the problem now, one of the questions we ask here is, do different values of α\alpha correspond to different systems? If yes, do all the values of α\alpha belong to passive systems?

Based on the behavior under time reversal, dividing i\mathcal{F}_{i} into the following three parts is\mathcal{F}^{\text{s}}_{i}, ia\mathcal{F}^{\text{a}}_{i} and iN\mathcal{F}^{\text{N}}_{i}:

is(𝑨)=ΓijsAj+kBTΓijsAj\displaystyle\mathcal{F}^{\text{s}}_{i}(\bm{A})=-\Gamma^{\text{s}}_{ij}\dfrac{\partial\mathcal{H}}{\partial A_{j}}+k_{\text{B}}T\dfrac{\partial\Gamma^{\text{s}}_{ij}}{\partial A_{j}} (13)
ia(𝑨)=ΓijaAj+kBTΓijaAj\displaystyle\mathcal{F}^{a}_{i}(\bm{A})=-\Gamma^{a}_{ij}\dfrac{\partial\mathcal{H}}{\partial A_{j}}+k_{\text{B}}T\dfrac{\partial\Gamma^{\text{a}}_{ij}}{\partial A_{j}} (14)
iN=αNkjNijAk\displaystyle\mathcal{F}^{\text{N}}_{i}=-\alpha N_{kj}\dfrac{\partial N_{ij}}{\partial A_{k}} (15)

where Γija(ΓijΓji)/2\Gamma^{\text{a}}_{ij}\equiv(\Gamma_{ij}-\Gamma_{ji})/2 is the antisymmetric part of Γij\Gamma_{ij}. From Eq.(2), Γija\Gamma^{\text{a}}_{ij} exhibits the following symmetry property (not in Einstein notation):

Γija=sisjΓija.\Gamma^{\text{a}}_{ij}=-s_{i}s_{j}\Gamma^{\text{a}}_{ij}. (16)

Since (𝒔𝑨)=(𝑨)\mathcal{H}(\bm{s}\circ\bm{A})=\mathcal{H}(\bm{A}) and Γijs(𝒔𝑨)=Γijs(𝑨)\Gamma^{\text{s}}_{ij}(\bm{s}\circ\bm{A})=\Gamma^{\text{s}}_{ij}(\bm{A}) [21], from Eqs. (4) & (16), under time reversal,

𝓕s(𝑨)𝓕s(𝒔𝑨)=𝒔𝓕s(𝑨),\displaystyle\bm{\mathcal{F}}^{\text{s}}(\bm{A})\to\bm{\mathcal{F}}^{\text{s}}(\bm{s}\circ\bm{A})=\bm{s}\circ\bm{\mathcal{F}}^{\text{s}}(\bm{A}), (17)
𝓕a(𝑨)𝓕a(𝒔𝑨)=𝒔𝓕a(𝑨).\displaystyle\bm{\mathcal{F}}^{\text{a}}(\bm{A})\to\bm{\mathcal{F}}^{\text{a}}(\bm{s}\circ\bm{A})=-\bm{s}\circ\bm{\mathcal{F}}^{\text{a}}(\bm{A}). (18)

where \circ stands for Hadamard product i.e. 𝒔𝑨{s1A1,s2A2,.snAn}\bm{s}\circ\bm{A}\equiv\{s_{1}A_{1},s_{2}A_{2},....s_{n}A_{n}\}. For given Γijs\Gamma^{\text{s}}_{ij}, is(𝑨)\mathcal{F}^{\text{s}}_{i}(\bm{A}) and ia(𝑨)\mathcal{F}^{\text{a}}_{i}(\bm{A}) do not depend on NijN_{ij}. In general, iN\mathcal{F}^{\text{N}}_{i} does not follow any of the above time reversal symmetries.

II.2 The ratio between the probability densities of a trajectory and its time-reversed trajectory

Let p0(𝑨)p_{0}(\bm{A}) be the probability distribution function of 𝑨\bm{A} at t=0t=0. In ϵ0\epsilon\to 0 limit, the probability density of a trajectory of the system (𝑨0,𝑨1,𝑨2,..𝑨N)(\bm{A}_{0},\bm{A}_{1},\bm{A}_{2},.....\bm{A}_{N}) (here 𝑨l𝑨(lϵ)\bm{A}_{l}\equiv\bm{A}(l\epsilon)) between t=0t=0 and t=τNϵt=\tau\equiv N\epsilon is given by [12] (see Appendix A)

P\displaystyle P \displaystyle\simeq p0(𝑨0)l=1N{(2kBT)1/2(2πϵ)n/2exp[14ϵkBT[dAi(l)ϵis(𝑨¯lf)ϵia(𝑨¯lf)](Γs1)ij(𝑨¯lf)[dAj(l)ϵjs(𝑨¯lf)ϵja(𝑨¯lf)].\displaystyle p_{0}(\bm{A}_{0})\prod^{N}_{l=1}\left\{\dfrac{(2k_{\text{B}}T)^{-1/2}}{(2\pi\epsilon)^{n/2}}\exp\Biggl{[}-\dfrac{1}{4\epsilon k_{\text{B}}T}\left[dA_{i}(l)-\epsilon\mathcal{F}^{\text{s}}_{i}(\bar{\bm{A}}^{\text{f}}_{l})-\epsilon\mathcal{F}^{\text{a}}_{i}(\bar{\bm{A}}^{\text{f}}_{l})\right]{{{\color[rgb]{0,0,0}(\Gamma^{s}}}^{-1})_{ij}}(\bar{\bm{A}}^{\text{f}}_{l})\left[dA_{j}(l)-\epsilon\mathcal{F}^{\text{s}}_{j}(\bar{\bm{A}}^{\text{f}}_{l})-\epsilon\mathcal{F}^{\text{a}}_{j}(\bar{\bm{A}}^{\text{f}}_{l})\right]\Biggr{.}\right. (19)
α[(dAi(l)ϵis(𝑨¯lf)ϵia(𝑨¯lf))[(Γs1)ijΓjksAk]𝑨¯lf+ϵ[isAi+iaAi]𝑨¯lf]]det(𝚪s(𝑨¯lf))1/2\displaystyle\left.-\alpha\left[\left(dA_{i}(l)-\epsilon\mathcal{F}^{\text{s}}_{i}(\bar{\bm{A}}^{\text{f}}_{l})-\epsilon\mathcal{F}^{\text{a}}_{i}(\bar{\bm{A}}^{\text{f}}_{l})\right)\left[({\Gamma^{s}}^{-1})_{ij}\dfrac{\partial\Gamma^{s}_{jk}}{\partial A_{k}}\right]_{\bar{\bm{A}}^{\text{f}}_{l}}+\epsilon\left[\dfrac{\partial\mathcal{F}^{\text{s}}_{i}}{\partial A_{i}}+\dfrac{\partial\mathcal{F}^{\text{a}}_{i}}{\partial A_{i}}\right]_{\bar{\bm{A}}^{\text{f}}_{l}}\right]\right]\text{det}\left(\bm{\Gamma^{\text{s}}}(\bar{\bm{A}}^{\text{f}}_{l})\right)^{-1/2}
×exp[α2ϵkBT[2ΓijsAiAjΓiksAk(Γs1)ijΓjpsAp]𝑨¯lf]}.\displaystyle\times\left.\exp\left[\alpha^{2}\epsilon k_{\text{B}}T\left[\dfrac{\partial^{2}\Gamma^{s}_{ij}}{\partial A_{i}\partial A_{j}}-\dfrac{\partial\Gamma^{s}_{ik}}{\partial A_{k}}({\Gamma^{s}}^{-1})_{ij}\dfrac{\partial\Gamma^{s}_{jp}}{\partial A_{p}}\right]_{\bar{\bm{A}}^{\text{f}}_{l}}\right]\right\}.

The ϵ3/2\epsilon^{3/2}- and higher-order terms have been neglected here. It is apparent from the above expression that, for given Γijs\Gamma^{\text{s}}_{ij}, PP is independent of NijN_{ij}. So no statistical property of the system has a dependence upon the choice of NikN_{ik}. Therefore, as mentioned earlier, NikN_{ik} is merely a dummy matrix. The probability density PP depends on α\alpha; thus, the different values of α\alpha correspond to different systems. Later in this subsection, we will see that Eq. (8) provides the dynamics of a passive system for any α\alpha. The time-reversed trajectory of the trajectory (𝑨0,𝑨1,𝑨2,..𝑨N)(\bm{A}_{0},\bm{A}_{1},\bm{A}_{2},.....\bm{A}_{N}) would be (𝒔𝑨N,𝒔𝑨N1,..𝒔𝑨1)(\bm{s}\circ\bm{A}_{N},\bm{s}\circ\bm{A}_{N-1},.....\bm{s}\circ\bm{A}_{1}), so its probability density can be calculated by replacing 𝑨i\bm{A}_{i} by 𝒔𝑨Ni\bm{s}\circ\bm{A}_{N-i} in the above equation; that is, (see Appendix B)

Pr\displaystyle P_{r} \displaystyle\simeq l=1N{(2kBT)1/2(2πϵ)n/2exp[14ϵkBT[dAi(l)ϵis(𝑨¯lr)+ϵia(𝑨¯lr)](Γs1)ij(𝑨¯lr)[dAj(l)ϵjs(𝑨¯lr)+ϵja(𝑨¯lr)].\displaystyle\prod^{N}_{l=1}\left\{\dfrac{(2k_{\text{B}}T)^{-1/2}}{(2\pi\epsilon)^{n/2}}\exp\Biggl{[}-\dfrac{1}{4\epsilon k_{\text{B}}T}\left[-dA_{i}(l)-\epsilon\mathcal{F}^{\text{s}}_{i}(\bar{\bm{A}}^{\text{r}}_{l})+\epsilon\mathcal{F}^{\text{a}}_{i}(\bar{\bm{A}}^{\text{r}}_{l})\right]{{{\color[rgb]{0,0,0}(\Gamma^{s}}}^{-1})_{ij}}(\bar{\bm{A}}^{\text{r}}_{l})\left[-dA_{j}(l)-\epsilon\mathcal{F}^{\text{s}}_{j}(\bar{\bm{A}}^{\text{r}}_{l})+\epsilon\mathcal{F}^{\text{a}}_{j}(\bar{\bm{A}}^{\text{r}}_{l})\right]\Biggr{.}\right. (20)
α[(dAi(l)ϵis(𝑨¯lr)+ϵia(𝑨¯lr))[(Γs1)ijΓjksAk]𝑨¯lr+ϵ[isAiiaAi]𝑨¯lr]]det(𝚪s(𝑨¯lr))1/2\displaystyle\left.-\alpha\left[\left(-dA_{i}(l)-\epsilon\mathcal{F}^{\text{s}}_{i}(\bar{\bm{A}}^{\text{r}}_{l})+\epsilon\mathcal{F}^{\text{a}}_{i}(\bar{\bm{A}}^{\text{r}}_{l})\right)\left[({\Gamma^{s}}^{-1})_{ij}\dfrac{\partial\Gamma^{s}_{jk}}{\partial A_{k}}\right]_{\bar{\bm{A}}^{\text{r}}_{l}}+\epsilon\left[\dfrac{\partial\mathcal{F}^{\text{s}}_{i}}{\partial A_{i}}-\dfrac{\partial\mathcal{F}^{\text{a}}_{i}}{\partial A_{i}}\right]_{\bar{\bm{A}}^{\text{r}}_{l}}\right]\right]\text{det}\left(\bm{\Gamma^{\text{s}}}(\bar{\bm{A}}^{\text{r}}_{l})\right)^{-1/2}
×exp[α2ϵkBT[2ΓijsAiAjΓiksAk(Γs1)ijΓjpsAp]𝑨¯lr]}p0(𝒔𝑨N).\displaystyle\times\left.\exp\left[\alpha^{2}\epsilon k_{\text{B}}T\left[\dfrac{\partial^{2}\Gamma^{s}_{ij}}{\partial A_{i}\partial A_{j}}-\dfrac{\partial\Gamma^{s}_{ik}}{\partial A_{k}}({\Gamma^{s}}^{-1})_{ij}\dfrac{\partial\Gamma^{s}_{jp}}{\partial A_{p}}\right]_{\bar{\bm{A}}^{\text{r}}_{l}}\right]\right\}p_{0}(\bm{s}\circ\bm{A}_{N}).

In ϵ0\epsilon\to 0 limit, the ratio P/PrP/P_{r} takes the following form (see Appendix 84):

PPr=p0(𝑨(0))p0(𝒔𝑨(τ))exp[1kBT((𝑨(τ))(𝑨(0)))]\dfrac{P}{P_{r}}=\dfrac{p_{0}(\bm{A}(0))}{p_{0}(\bm{s}\circ\bm{A}(\tau))}\exp\left[-\dfrac{1}{k_{\text{B}}T}\left(\mathcal{H}(\bm{A}(\tau))-\mathcal{H}(\bm{A}(0))\right)\right] (21)

In the stationary state, p0(𝑨)=exp((𝑨)/kBT)/𝒵p_{0}(\bm{A})=\exp(-\mathcal{H}(\bm{A})/k_{\text{B}}T)/\mathcal{Z} (see Appendix H), the above equation then yields

P=Pr.P=P_{r}. (22)

It implies that any system whose dynamics is given by Eq. (8) has the time reversal symmetry in its stationary state, regardless of the value of α\alpha. Hence, Eq. (8) describes a passive system for any value of α\alpha between 0 and 1.

II.3 Fluctuation relations and the dissipation function for the passive systems

One can readily show that [12] the dissipation function τ\mathcal{R}_{\tau} for the trajectory (𝑨0,𝑨1,𝑨2,..𝑨N)(\bm{A}_{0},\bm{A}_{1},\bm{A}_{2},.....\bm{A}_{N}) defined as

τ=ln[PPr]\mathcal{R_{\tau}}=\ln\left[\dfrac{P}{P_{r}}\right] (23)

satisfies the relation

𝒫(τ=X)𝒫(τ=X)=exp(X),{{\color[rgb]{0,0,0}\dfrac{\mathcal{P}(\mathcal{R}_{\tau}=X)}{\mathcal{P}(\mathcal{R}_{\tau}=-X)}=\exp(X),}} (24)

where 𝒫\mathcal{P} is the probability distribution of τ\mathcal{R}_{\tau}. The above relation is known as the fluctuation theorem [2, 5]. From Eq. (21),

τ=lnp0(𝑨(0))p0(𝒔𝑨(τ))1kBT[(𝑨(τ))(𝑨(0))].\mathcal{R}_{\tau}=\ln\dfrac{p_{0}(\bm{A}(0))}{p_{0}(\bm{s}\circ\bm{A}(\tau))}-\dfrac{1}{k_{\text{B}}T}\left[\mathcal{H}(\bm{A}(\tau))-\mathcal{H}(\bm{A}(0))\right]. (25)

Intriguingly, τ\mathcal{R}_{\tau} does not depend explicitly on Γij\Gamma_{ij}. Moreover, it depends only the initial and final values of 𝑨\bm{A}, not on the trajectory followed by 𝑨\bm{A}. Note that the ratio P/PrP/P_{r} in Eq. (21) also has the same functional properties.

One can also define the dissipation function for the system as follows [1]:

τ=lnp(𝑨(0),𝑨(τ);τ)p(𝒔𝑨(τ),𝒔𝑨(0);τ),\mathcal{R}^{\prime}_{\tau}=\ln\dfrac{p(\bm{A}(0),\bm{A}(\tau);\tau)}{p(\bm{s}\circ\bm{A}(\tau),\bm{s}\circ\bm{A}(0);\tau)}, (26)

where p(𝑨(0),𝑨(τ);τ)p(\bm{A}(0),\bm{A}(\tau);\tau) is the net probability that the system goes from 𝑨(0)\bm{A}(0) to 𝑨(τ)\bm{A}(\tau) in time τ\tau:

p(𝑨(0),𝑨(τ);τ)=P,p(\bm{A}(0),\bm{A}(\tau);\tau)=\sum P, (27)

where the summation is performed over all the trajectories between 𝑨(0)\bm{A}(0) and 𝑨(τ)\bm{A}(\tau). Since the ratio P/PrP/P_{r} is independent of the trajectory between 𝑨(0)\bm{A}(0) and 𝑨(τ)\bm{A}(\tau), τ=τ\mathcal{R^{\prime}}_{\tau}=\mathcal{R}_{\tau}.

The integrated form of the relation (24) is

exp(τ)=1,\left\langle\exp(-\mathcal{R}_{\tau})\right\rangle=1, (28)

where angular bracket stands for the ensmeble average. Using this relation, one can show that [28, 5]

τ0.\left\langle\mathcal{R}_{\tau}\right\rangle\geq 0. (29)

In equilibrium, P=PrP=P_{r}, thus τ=τ=0\left\langle\mathcal{R}_{\tau}\right\rangle=\mathcal{R}_{\tau}=0. So τ\left\langle\mathcal{R}_{\tau}\right\rangle behaves like the change in the entropy of system and it can be used to evaluate that how far the system is from equilibrium. Since the solution of the Fokker-Planck equation corresponding to Eq. (8) does not depend on α\alpha, τ\left\langle\mathcal{R}_{\tau}\right\rangle is constant in α\alpha (see Appendix I).

Generalizing the expression of the dissipation function given in Eq. (25) for the trajectories starting at arbitrary time tt:

τ(t)\displaystyle\mathcal{R}_{\tau}(t) =\displaystyle= lnpt(𝑨(t))pt(𝒔𝑨(t+τ))\displaystyle\ln\dfrac{p_{t}(\bm{A}(t))}{p_{t}(\bm{s}\circ\bm{A}(t+\tau))} (30)
1kBT[(𝑨(t+τ))(𝑨(t))].\displaystyle-\dfrac{1}{k_{\text{B}}T}\left[\mathcal{H}(\bm{A}(t+\tau))-\mathcal{H}(\bm{A}(t))\right].

The instantaneous irreversibility can be evaluated by calculating τ(t)\mathcal{R}_{\tau}(t) in τ0\tau\to 0 limit, that is,

τ(t)s˙(t)τ+lnpt(𝑨(t+τ))pt(𝒔𝑨(t+τ)),{{\color[rgb]{0,0,0}\mathcal{R}_{\tau}(t)\simeq\dot{{s}}(t)\tau+\ln\dfrac{p_{t}(\bm{A}(t+\tau))}{p_{t}(\bm{s}\circ\bm{A}(t+\tau))},}} (31)

where

s˙(t)=ddt(lnpt(𝑨(t))+1kBT(𝑨(t)))|t=t.\dot{{s}}(t)=-\left.\dfrac{d}{dt^{\prime}}\left(\ln p_{t}(\bm{A}(t^{\prime}))+\dfrac{1}{k_{\text{B}}T}\mathcal{H}(\bm{A}(t^{\prime}))\right)\right|_{t^{\prime}=t}. (32)

As discussed in Appendix E, kBs˙(t)k_{\text{B}}\left\langle\dot{{s}}(t)\right\rangle is nothing but the rate of change of total entropy of the system and the reservoir (see Eq. (107)). According to Eq. (31), the irreversible behavior of the system results from entropy production and from the asymmetric behavior of pt(𝑨)p_{t}(\bm{A}) under time reversal. If pt(𝒔𝑨)pt(𝑨)p_{t}(\bm{s}\circ\bm{A})\neq p_{t}(\bm{A}), the system is instantaneously irreversible since

0(t)=lnpt(𝑨(t))pt(𝒔𝑨(t))0.\mathcal{R}_{0}(t)=\ln\dfrac{p_{t}(\bm{A}(t))}{p_{t}(\bm{s}\circ\bm{A}(t))}\neq 0. (33)

Since (𝒔𝑨)=(𝑨)\mathcal{H}(\bm{s}\circ\bm{A})=\mathcal{H}(\bm{A}), the equilibrium probability distribution peq(𝑨)exp((𝑨)/kBT)/𝒵p_{\text{eq}}(\bm{A})\equiv\exp(-\mathcal{H}(\bm{A})/k_{\text{B}}T)/\mathcal{Z} always follows the symmetry property peq(𝒔𝑨)=peq(𝑨)p_{\text{eq}}(\bm{s}\circ\bm{A})=p_{\text{eq}}(\bm{A}); it is a fundamental property of peq(𝑨)p_{\text{eq}}(\bm{A}). The nonzero value of 0(t)\left\langle\mathcal{R}_{0}(t)\right\rangle for an out-of-equilibrium system signifies that the system violates this symmetry. Note that 0(t)0\left\langle\mathcal{R}_{0}(t)\right\rangle\geq 0. An example of such systems is as follows: consider a colloidal particle moving with a nonzero average velocity 𝒗0\bm{v}_{0} and having the probability distribution p0(𝒗)=Cexp((𝒗𝒗0)2/2)p_{0}(\bm{v})=C\exp(-(\bm{v}-\bm{v}_{0})^{2}/2), at t=0t=0. Under time reversal 𝒗𝒗\bm{v}\to-\bm{v}, so 𝒔={1,1,1}\bm{s}=\{-1,-1,-1\}. Hence p0(𝒔𝒗)p0(𝒗)p_{0}(\bm{s}\circ\bm{v})\neq p_{0}(\bm{v}).

For pt(𝒔𝑨)=pt(𝑨)p_{t}(\bm{s}\circ\bm{A})=p_{t}(\bm{A}) case, 0(t)=0\mathcal{R}_{0}(t)=0, so from Eq. (31),

s˙(t)=limτ0τ(t)τ.\dot{{s}}(t)=\lim_{\tau\to 0}\dfrac{\mathcal{R}_{\tau}(t)}{\tau}. (34)

Thus, the average rate of change of τ(t)\mathcal{R}_{\tau}(t) with τ\tau is the same as the rate of the total entropy production of the system and the reservoir; from Eq. (29), the second law of thermodynamics is evident, s˙(t)>0\left\langle\dot{{s}}(t)\right\rangle>0. In the next subsection, we discuss a broad class of passive systems with pt(𝒔𝑨)=pt(𝑨)p_{t}(\bm{s}\circ\bm{A})=p_{t}(\bm{A}).

The form of the dissipation function used by Seifert et al. [5] is briefly discussed in Appendix F.

II.4 The dissipation function for quenched systems

Here we consider that the system is initially in a thermodynamic equilibrium state and the state variables of the system 𝜷{β1,β2,,βn}\bm{\beta}\equiv\{\beta_{1},\beta_{2},...,\beta_{n}\} are abruptly changed at t=0t=0. Then the system will start evolving towards the equilibrium state corresponding to the modified values of 𝜷\bm{\beta}. Writing the coarse-grained Hamiltonian of the system as the function of 𝜷\bm{\beta}: (𝑨;𝜷)\mathcal{H}\equiv\mathcal{H}(\bm{A};\bm{\beta}). Let 𝜷=𝜷I\bm{\beta}=\bm{\beta}_{\text{I}} at t=0t=0 then

p0(𝑨)=1𝒵(𝜷I)exp[(𝑨;𝜷I)kBT]p_{0}(\bm{A})=\dfrac{1}{\mathcal{Z}(\bm{\beta}_{\text{I}})}\exp\left[-\dfrac{\mathcal{H}(\bm{A};\bm{\beta}_{\text{I}})}{k_{\text{B}}T}\right] (35)

From Eq.(25), for a quench from 𝜷=𝜷I\bm{\beta}=\bm{\beta}_{\text{I}} to 𝜷=𝜷F\bm{\beta}=\bm{\beta}_{\text{F}} at t=0t=0, the dissipation function for the system takes the following form

τ=1kBT[(𝑨(0);𝜷F)(𝑨(0);𝜷I)\displaystyle\mathcal{R}_{\tau}=\dfrac{1}{k_{\text{B}}T}\left[\mathcal{H}(\bm{A}(0);\bm{\beta}_{\text{F}})-\mathcal{H}(\bm{A}(0);\bm{\beta}_{\text{I}})\right.
((𝑨(τ);𝜷F)(𝑨(τ);𝜷I))].\displaystyle-\left.(\mathcal{H}(\bm{A}(\tau);\bm{\beta}_{\text{F}})-\mathcal{H}(\bm{A}(\tau);\bm{\beta}_{\text{I}}))\right]. (36)

We will now discuss an example of quenched systems.

II.4.1 Colloidal particle in a harmonic potential well

Consider a colloidal particle trapped in a harmonic potential U=k𝐫2/2U=k\mathbf{r}^{2}/2, where kk is the stiffness of the potential. Imagine that initially the particle is in thermodynamic equilibrium with k=k0k=k_{0} and the value of kk is instantaneously changed from k0k_{0} to k1k_{1} at t=0t=0 [15]. Ignoring the kinetic energy, the coarse-grained Hamiltonian for this system would be simply =U\mathcal{H}=U. Then, from Eq. (II.4), the dissipation function for a trajectory between 𝐫=𝐫0\mathbf{r}=\mathbf{r}_{0} and 𝐫=𝐫τ\mathbf{r}=\mathbf{r}_{\tau} in time τ\tau is given by

Rτ=12(k0k1)(𝐫τ2𝐫02)R_{\tau}=\dfrac{1}{2}(k_{0}-k_{1})(\mathbf{r}_{\tau}^{2}-\mathbf{r}^{2}_{0}) (37)

The above expression was derived by Carberry et al. [29, 15] for spatially uniform diffusion constant. As we have considered the dependence of Γij\Gamma_{ij} on 𝑨\bm{A} in the derivation of τ\mathcal{R}_{\tau}, the above expression of the dissipation function is more general; it is valid for the systems having state dependent diffusion as well. In order to verify our prediction, we numerically solve the Langevin equation for a colloidal particle with the diffusion coefficient varying with position. For simplicity, we consider the 1D case. From Eq. (8), the overdamped Langevin equation for the colloidal particle reads

dxdt=1kBTD(x)kx+(1α)dD(x)dx+2D(x)η(t),\dfrac{dx}{dt}=-\dfrac{1}{k_{\text{B}}T}D(x)kx+(1-\alpha)\dfrac{dD(x)}{dx}+\sqrt{2D(x)}\eta(t), (38)

with its discreate form (see Eq. (10))

x(t+dt)=x(t)1kBTD(x¯f)kx¯f\displaystyle x(t+dt)=x(t)-\dfrac{1}{k_{\text{B}}T}D(\bar{x}^{\text{f}})k\bar{x}^{\text{f}} +\displaystyle+ (1α)[dD(x)dx]x¯f\displaystyle(1-\alpha)\left[\dfrac{dD(x)}{dx}\right]_{\bar{x}^{\text{f}}} (39)
+\displaystyle+ 2D(x¯f)dtηt,\displaystyle\sqrt{2D(\bar{x}^{\text{f}})dt}\eta_{t},

where x¯f=αx(t+dt)+(1α)x(t)\bar{x}^{\text{f}}=\alpha x(t+dt)+(1-\alpha)x(t) and D(x)D(x) is the state-dependent diffusion. The above equation is a self-consistent equation of x(t+dt)x(t+dt) for given x(t)x(t). There are many examples of the systems having state-dependent diffusion; e.g., a colloidal particle near a wall [23]. We here consider a hypothetical system having Gaussian profile of the diffusion coefficient:

D(x)=D0exp(x2L2).D(x)=D_{0}\exp\left(-\dfrac{x^{2}}{L^{2}}\right). (40)
Refer to caption
Figure 1: ln[𝒫(X)/𝒫(X)]\ln[\mathcal{P}(X)/\mathcal{P}(-X)] vs XX for a 1D colloidal particle in a potential well U=kx2/2U=kx^{2}/2 with the diffusion coefficient D(x)=D0exp(x2/L2)D(x)=D_{0}\exp(-x^{2}/L^{2}), where 𝒫\mathcal{P} is the probability distribution function for the dissipation function given by Eq. (37). The value of kk is suddenly changed from k=k0k=k_{0} to k=k1k=k_{1} at t=0t=0. Here L/k1/kBT=1L/\sqrt{k_{1}/k_{\text{B}}T}=1, k1/k0=4k_{1}/k_{0}=4, the time duration of the trajectory τ=3kBT/D0k1\tau=3k_{\text{B}}T/D_{0}k_{1}, and 8,00,000 no. of trajectories are used for the statistics.

To obtain the trajectory of the particle, at each time step, we solve the Eq. (39) for x(t+dt)x(t+dt) with fixed point iteration method with the accuracy of 10410^{-4}. In Fig. 1, we show ln[𝒫(X)/𝒫(X)]\ln[\mathcal{P}(X)/\mathcal{P}(-X)] vs XX: clearly, the dissipation function given by Eq. (37) obeys the fluctuation relation, for all the values of α\alpha.

III The dissipation function for active systems

In this section, we consider the active systems [30] whose dynamics is governed by the equations of motion having the following form:

dAidt=i+𝒳i+Nijξj(t),\dfrac{dA_{i}}{dt}=\mathcal{F}_{i}+\mathcal{X}_{i}+N_{ij}\xi_{j}(t), (41)

where the addition term 𝒳i\mathcal{X}_{i} represents the active driving forces. Due to the presence of the active forces, the active systems are always away from equilibrium. However, they can achieve a nonequilibrium steady state. Writing 𝒳i{\mathcal{X}}_{i} as the sum of two terms 𝒳is{\mathcal{X}}^{\text{s}}_{i} and 𝒳ia{\mathcal{X}}^{\text{a}}_{i} such that 𝓧s(𝒔𝑨)=𝒔𝓧s(𝑨)\bm{\mathcal{X}}^{\text{s}}(\bm{s}\circ\bm{A})=\bm{s}\circ\bm{\mathcal{X}}^{\text{s}}(\bm{A}) and 𝓧a(𝒔𝑨)=𝒔𝓧a(𝑨)\bm{\mathcal{X}}^{\text{a}}(\bm{s}\circ\bm{A})=-\bm{s}\circ\bm{\mathcal{X}}^{\text{a}}(\bm{A}) (see Eqs. (115) & (116), Appendix G). It should be noted that NijN_{ij} serves as a dummy matrix here as well because the form of PP will be the same as that in Eq. (19), except that is\mathcal{F}^{\text{s}}_{i} and ia\mathcal{F}^{\text{a}}_{i} will have additional active components 𝒳is{\mathcal{X}}^{\text{s}}_{i} and 𝒳ia{\mathcal{X}}^{\text{a}}_{i}, respectively. Following the approach used in subsection II.2, we obtain the following expression of the dissipation function:

τ(t)=lnpt(𝑨(t))pt(𝒔𝑨(t+τ))1kBT[(𝑨(t+τ))\displaystyle\mathcal{R}_{\tau}(t)=\ln\dfrac{p_{t}(\bm{A}(t))}{p_{t}(\bm{s}\circ\bm{A}(t+\tau))}-\dfrac{1}{k_{\text{B}}T}\left[\mathcal{H}(\bm{A}(t+\tau))\right.
(𝑨(t))]+1kBTt+τtw(t)dt,\displaystyle\left.-\mathcal{H}(\bm{A}(t))\right]+\dfrac{1}{k_{\text{B}}T}\int^{t+\tau}_{t}w(t^{\prime})dt^{\prime}, (42)

where

w(t)\displaystyle w(t) =\displaystyle= (𝑨(t))Ai𝒳ia(𝑨(t))kBT𝒳ia(𝑨(t))Ai\displaystyle\dfrac{\partial\mathcal{H}(\bm{A}(t))}{\partial A_{i}}\mathcal{X}^{\text{a}}_{i}(\bm{A}(t))-k_{\text{B}}T\dfrac{\partial\mathcal{X}^{\text{a}}_{i}(\bm{A}(t))}{\partial A_{i}} (43)
+\displaystyle+ (Γs1)ij(𝑨(t))𝒳js(𝑨(t))[dAidt𝒴ia(𝑨(t))]\displaystyle{{\color[rgb]{0,0,0}({\Gamma^{\text{s}}}^{-1})_{ij}}}(\bm{A}(t))\mathcal{X}^{\text{s}}_{j}(\bm{A}(t))\left[\dfrac{dA_{i}}{dt}-\mathcal{Y}^{\text{a}}_{i}(\bm{A}(t))\right]

and

𝒴ia=𝒳ia+AkΓkiakBTΓkiaAk.\mathcal{Y}^{\text{a}}_{i}=\mathcal{X}^{\text{a}}_{i}+\dfrac{\partial\mathcal{H}}{\partial A_{k}}\Gamma^{\text{a}}_{ki}-k_{\text{B}}T\dfrac{\partial\Gamma^{\text{a}}_{ki}}{\partial A_{k}}. (44)

The integration in Eq. (43) is performed using midpoint rule. In contrast to passive systems, τ(t)\mathcal{R}_{\tau}(t) here depends on Γij\Gamma_{ij}, though not on α\alpha. Moreover, τ(t)\mathcal{R}_{\tau}(t) is trajectory-dependent, so τ(t)\left\langle\mathcal{R}_{\tau}(t)\right\rangle is a function of α\alpha because the probability density PP of a trajectory depends on α\alpha (as in passive systems, see Eq. (19)). The ensemble average of w(t)w(t) is given by (with an assumption, see Appendix G)

w(t)\displaystyle\left\langle w(t)\right\rangle =\displaystyle= (dAidtAkΓkia+kBTΓkiaAk)(Γs1)ij𝒳js\displaystyle\left\langle\left(\dfrac{dA_{i}}{dt}-\dfrac{\partial\mathcal{H}}{\partial A_{k}}\Gamma^{\text{a}}_{ki}+k_{\text{B}}T\dfrac{\partial\Gamma^{\text{a}}_{ki}}{\partial A_{k}}\right){{\color[rgb]{0,0,0}({\Gamma^{\text{s}}}^{-1})_{ij}}}\mathcal{X}^{\text{s}}_{j}\right\rangle (45)
(dAidt𝒴ia)(Γs1)ij𝒳ja,\displaystyle-\left\langle\left(\dfrac{dA_{i}}{dt}-\mathcal{Y}^{\text{a}}_{i}\right){{\color[rgb]{0,0,0}({\Gamma^{\text{s}}}^{-1})_{ij}}}\mathcal{X}^{\text{a}}_{j}\right\rangle,

where the first term is the average rate of work done by active force (Γs1)ij𝒳js{{\color[rgb]{0,0,0}({\Gamma^{\text{s}}}^{-1})_{ij}}}\mathcal{X}^{\text{s}}_{j} and the second term is by the active force (Γs1)ij𝒳ja{{\color[rgb]{0,0,0}({\Gamma^{\text{s}}}^{-1})_{ij}}}\mathcal{X}^{\text{a}}_{j}. So w(t)w(t) can be interpreted as the rate of work done by active forces along the trajectory at time tt. For pt(𝒔𝑨)=pt(𝑨)p_{t}(\bm{s}\circ\bm{A})=p_{t}(\bm{A}), the rate of change of dissipation function with τ\tau (see Eqs. (34) and (III)) is given by

s˙(t)=1kBTw(t)ddt(lnpt(𝑨(t))+1kBT(𝑨(t)))|t=t.\displaystyle\dot{{s}}(t)=\dfrac{1}{k_{\text{B}}T}w(t)-\left.\dfrac{d}{dt^{\prime}}\left(\ln p_{t}(\bm{A}(t^{\prime}))+\dfrac{1}{k_{\text{B}}T}\mathcal{H}(\bm{A}(t^{\prime}))\right)\right|_{t^{\prime}=t}.
(46)

and its average reads (see Appendix G)

s˙(t)=1kBT(Ji(𝑨,t)pt(𝑨)𝒴ia(𝑨))(Γs1)ij(𝑨)\displaystyle\left\langle\dot{{s}}(t)\right\rangle=\dfrac{1}{k_{\text{B}}T}\left\langle\left(\dfrac{J_{i}(\bm{A},t)}{p_{t}(\bm{A})}-\mathcal{Y}^{\text{a}}_{i}(\bm{A})\right){{\color[rgb]{0,0,0}({\Gamma^{\text{s}}}^{-1})_{ij}}}(\bm{A})\right.
×(Jj(𝑨,t)pt(𝑨)𝒴ja(𝑨)),\displaystyle\times\left.\left(\dfrac{J_{j}(\bm{A},t)}{p_{t}(\bm{A})}-\mathcal{Y}^{\text{a}}_{j}(\bm{A})\right)\right\rangle, (47)

where Ji(𝑨,t)J_{i}(\bm{A},t) is the probability current for 𝑨\bm{A} [23]:

Ji(𝑨,t)\displaystyle J_{i}(\bm{A},t) =\displaystyle= (Γsij(𝑨)(𝑨)Aj+𝒴ia(𝑨)+𝒳js(𝑨))pt(𝑨)\displaystyle\left(-{\Gamma^{\text{s}}}_{ij}(\bm{A})\dfrac{\partial\mathcal{H}(\bm{A})}{\partial A_{j}}+\mathcal{Y}^{\text{a}}_{i}(\bm{A})+\mathcal{X}^{\text{s}}_{j}(\bm{A})\right)p_{t}(\bm{A}) (48)
kBTΓijs(𝑨)pt(𝑨)dAj.\displaystyle\qquad\qquad\qquad\qquad-k_{\text{B}}T\Gamma^{\text{s}}_{ij}(\bm{A})\dfrac{\partial p_{t}(\bm{A})}{dA_{j}}.

Using Eq (6), it is easy to show that s˙(t)>0\left\langle\dot{{s}}(t)\right\rangle>0, as expected from the integrated fluctuation theorem (28). Again, kBs˙(t)k_{\text{B}}\left\langle\dot{{s}}(t)\right\rangle is nothing but the rate of the total entropy production of the system and the reservoir (see Appendix E).

As discussed for the passive systems in subsection II.3, for pt(𝒔𝑨)pt(𝑨)p_{t}(\bm{s}\circ\bm{A})\neq p_{t}(\bm{A}) case, time reversal asymmetry in pt(𝑨)p_{t}(\bm{A}) also contributes to irreversibility. This contribution can also be observed in the stationary states of many active systems. Active systems with polar alignment [31, 32, 33, 34] are examples of this type of system; as the velocities of the particles are globally aligned, the velocity distribution is not an even function for these systems. The passive systems, being in equilibrium in their stationary states, cannot demonstrate this irreversibility.

A few special cases for the active systems are as follows: (a) if initially the system is in equilibrium (that is, 𝒳i=0\mathcal{X}_{i}=0) and the active term 𝒳i\mathcal{X}_{i} is switched on at t=0t=0 then p0(𝑨)=exp((𝑨)/kBT)/𝒵p_{0}(\bm{A})=\exp(-\mathcal{H}(\bm{A})/k_{\text{B}}T)/\mathcal{Z}. In this case, τ\mathcal{R}_{\tau} is just the net work done by the active forces during the trajectory (see Eq. (III)):

τ=1kBT0τw(t)𝑑t.\mathcal{R}_{\tau}={{\color[rgb]{0,0,0}\dfrac{1}{k_{\text{B}}T}\int^{\tau}_{0}}}w(t^{\prime})dt^{\prime}. (49)

(b) In the stationary state (i.e. in tt\to\infty limit), the time averaged work done by the active forces,

wav=limτ1τtt+τ𝑑tw(t),w_{\text{av}}=\lim_{\tau\to\infty}\dfrac{1}{\tau}\int^{t+\tau}_{t}dt^{\prime}w(t^{\prime}), (50)

is independent of tt. So, in τ\tau\to\infty limit, in Eq. (III), the last term is proportional to τ\tau and we can ignore the first two terms. Then,

τ\displaystyle\mathcal{R}_{\tau} \displaystyle\simeq 1kBTtt+τw(t)𝑑t.\displaystyle\dfrac{1}{k_{\text{B}}T}\int^{t+\tau}_{t}w(t^{\prime})dt^{\prime}. (51)
\displaystyle\simeq 1kBTwavτ.\displaystyle\dfrac{1}{k_{\text{B}}T}w_{\text{av}}\tau. (52)

From Eq. (24), the probability distribution 𝒫w(wav)\mathcal{P}_{\text{w}}(w_{\text{av}}) of wavw_{\text{av}} satisfies the relation

X=limτkBT1τln𝒫w(wav=X)𝒫w(wav=X),X=\lim_{\tau\to\infty}k_{\text{B}}T\dfrac{1}{\tau}\ln\dfrac{\mathcal{P}_{\text{w}}(w_{\text{av}}=X)}{\mathcal{P}_{\text{w}}(w_{\text{av}}=-X)}, (53)

This is called the steady-state fluctuation theorem [28].

IV Conclusion

Starting with the generic Langevin equations, using path integral approach, we first calculated the ratio of the probability densities of a trajectory and its time-reversed trajectory for passive systems using α\alpha-discritization: it is independent of the value of α\alpha. Irrespective of the value of α\alpha, the stationary solutions of generic Langevin equations have time reversal symmetry, so the generic Langevin equations with any value of α\alpha describes a passive system. Next we calculated the dissipation function for the passive systems which is found to be independent of the trajectory of the system, it depends only on the intial and the final values of the dynamical variables of the system. Furthermore, it is not an explicit function of coefficients of the generic Langevin equations. We also verify the fluctuation theorem for a 1D particle trapped in a potential well whose stiffness is suddenly changed, with the state-dependent diffusion. Finally, we obtained the expression of the dissipation function for active systems and defined the work done by the active forces. For both passive and active systems, the average of the rate of change of dissipation function with the duration of the trajectory is just the entropy production rate of the system and the reservoir.

Appendix A The probability density of a trajectory for passive systems

The generic Langevin equations for passive systems in discrete form (see Eq. (10)):

dAi(l)=ϵi(𝑨¯lf)+ϵNij(𝑨¯lf)ξjl,dA_{i}(l)=\epsilon\mathcal{F}_{i}(\bar{\bm{A}}^{\text{f}}_{l})+\sqrt{\epsilon}N_{ij}(\bar{\bm{A}}^{\text{f}}_{l})\xi_{j}^{l}, (54)

where dAi(l)Ai(ϵl)Ai(ϵ(l1))dA_{i}(l)\equiv A_{i}(\epsilon l)-A_{i}(\epsilon(l-1)), 𝑨¯lfα𝑨(ϵl)+(1α)𝑨(ϵ(l1))\bar{\bm{A}}^{\text{f}}_{l}\equiv\alpha\bm{A}(\epsilon l)+(1-\alpha)\bm{A}(\epsilon(l-1)), and

iΓijAj+kBTΓijAjαNljNijAl.\mathcal{F}_{i}\equiv-\Gamma_{ij}\dfrac{\partial\mathcal{H}}{\partial A_{j}}+k_{\text{B}}T\dfrac{\partial\Gamma_{ij}}{\partial A_{j}}-\alpha N_{lj}\dfrac{\partial N_{ij}}{\partial A_{l}}. (55)

Solving the above equations for ξil\xi_{i}^{l}, we obtain

ξil=1ϵ(N1)ij(𝑨¯lf)(dAj(l)ϵj(𝑨¯lf)).\xi_{i}^{l}=\dfrac{1}{\sqrt{\epsilon}}{{\color[rgb]{0,0,0}(N^{-1})_{ij}}}(\bar{\bm{A}}^{\text{f}}_{l})(dA_{j}(l)-\epsilon\mathcal{F}_{j}(\bar{\bm{A}}^{\text{f}}_{l})). (56)

Since ξil\xi_{i}^{l} are the uncorrelated series of random numbers having normal distribution with zero mean and variance one, the probability density function of a trajectory of the system (𝑨0,𝑨1,𝑨2,..𝑨N)(\bm{A}_{0},\bm{A}_{1},\bm{A}_{2},.....\bm{A}_{N}) (here 𝑨l𝑨(lϵ)\bm{A}_{l}\equiv\bm{A}(l\epsilon)) between t=0t=0 and t=τNϵt=\tau\equiv N\epsilon is given by [12]

P=p0(𝑨0)|𝒥|l=1N1(2π)n/2exp[12ϵ[dAi(l)ϵi(𝑨¯lf)](N1)ki(𝑨¯lf)(N1)kj(𝑨¯lf)[dAj(l)ϵj(𝑨¯lf)]],P=p_{0}(\bm{A}_{0})\left|\mathcal{J}\right|\prod^{N}_{l=1}\dfrac{1}{(2\pi)^{n/2}}\exp\left[-\dfrac{1}{2\epsilon}\left[dA_{i}(l)-\epsilon\mathcal{F}_{i}(\bar{\bm{A}}^{\text{f}}_{l})\right]{{\color[rgb]{0,0,0}(N^{-1})_{ki}}}(\bar{\bm{A}}^{\text{f}}_{l}){{\color[rgb]{0,0,0}(N^{-1})_{kj}}}(\bar{\bm{A}}^{\text{f}}_{l})\left[dA_{j}(l)-\epsilon\mathcal{F}_{j}(\bar{\bm{A}}^{\text{f}}_{l})\right]\right], (57)

where p0(𝑨)p_{0}(\bm{A}) is the probability distribution of 𝑨\bm{A} at t=0t=0 and 𝒥\mathcal{J} is the Jacobean determinant for the transformation of the variables of the probability density function from the ξil\xi_{i}^{l} to Aj(ϵm)A_{j}(\epsilon m). From Eq. (56), the Nn×NnNn\times Nn Jacobean matrix for the variable transformation is given by

𝒥jmil\displaystyle\mathcal{J}^{il}_{jm} =\displaystyle= ξilAj(ϵm)\displaystyle\dfrac{\partial\xi_{i}^{l}}{\partial A_{j}(\epsilon m)} (58)
=\displaystyle= 1ϵ1/2[[[(N1)ikAj]𝑨¯lfdAk(l)ϵ[(N1)ikAjk+kAj(N1)ik]𝑨¯lf](αδlm+(1α)δ(l1)m)\displaystyle\dfrac{1}{\epsilon^{1/2}}\left[\left[\left[\dfrac{\partial{{\color[rgb]{0,0,0}(N^{-1})_{ik}}}}{\partial A_{j}}\right]_{\bar{\bm{A}}^{\text{f}}_{l}}dA_{k}(l)-\epsilon\left[\dfrac{\partial{{\color[rgb]{0,0,0}(N^{-1})_{ik}}}}{\partial A_{j}}\mathcal{F}_{k}+\dfrac{\partial\mathcal{F}_{k}}{\partial A_{j}}{{\color[rgb]{0,0,0}(N^{-1})_{ik}}}\right]_{\bar{\bm{A}}^{\text{f}}_{l}}\right](\alpha\delta_{lm}+(1-\alpha)\delta_{(l-1)m})\right.
.+(N1)ij(𝑨¯lf)(δlmδ(l1)m)]\displaystyle\Bigg{.}+{{\color[rgb]{0,0,0}(N^{-1})_{ij}}}(\bar{\bm{A}}^{\text{f}}_{l})(\delta_{lm}-\delta_{(l-1)m})\Bigg{]}
=\displaystyle= 1ϵ1/2[[[(N1)ikAj]𝑨¯lf(dAk(l)ϵk(𝑨¯lf))ϵ[kAj(N1)ik]𝑨¯lf](αδlm+(1α)δ(l1)m)\displaystyle\dfrac{1}{\epsilon^{1/2}}\left[\left[\left[\dfrac{\partial{{\color[rgb]{0,0,0}(N^{-1})_{ik}}}}{\partial A_{j}}\right]_{\bar{\bm{A}}^{\text{f}}_{l}}\left(dA_{k}(l)-\epsilon\mathcal{F}_{k}(\bar{\bm{A}}^{\text{f}}_{l})\right)-\epsilon\left[\dfrac{\partial\mathcal{F}_{k}}{\partial A_{j}}{{\color[rgb]{0,0,0}(N^{-1})_{ik}}}\right]_{\bar{\bm{A}}^{\text{f}}_{l}}\right](\alpha\delta_{lm}+(1-\alpha)\delta_{(l-1)m})\right.
.+(N1)ij(𝑨¯lf)(δlmδ(l1)m)].\displaystyle\Bigg{.}+{{\color[rgb]{0,0,0}(N^{-1})_{ij}}}(\bar{\bm{A}}^{\text{f}}_{l})(\delta_{lm}-\delta_{(l-1)m})\Bigg{]}.

The above matrix is a block triangular matrix of n×nn\times n submatrices with fixed (l,ml,m), so its determinant will be the multiplication of all the diagonal submatrices (i.e. with l=ml=m):

𝒥=l=1N1ϵn/2det(𝑴(l)),\mathcal{J}=\prod^{N}_{l=1}\dfrac{1}{\epsilon^{n/2}}{{\color[rgb]{0,0,0}\text{det}\left(\bm{M}(l)\right)}}, (59)

where

Mij(l)\displaystyle M_{ij}(l) =\displaystyle= (N1)ij(𝑨¯lf)+α[(N1)ikAj]𝑨¯lf(dAk(l)ϵk(𝑨¯lf))ϵα[kAj(N1)ik]𝑨¯lf\displaystyle{{\color[rgb]{0,0,0}(N^{-1})_{ij}}}(\bar{\bm{A}}^{\text{f}}_{l})+\alpha\left[\dfrac{\partial{{\color[rgb]{0,0,0}(N^{-1})_{ik}}}}{\partial A_{j}}\right]_{\bar{\bm{A}}^{\text{f}}_{l}}\left(dA_{k}(l)-\epsilon\mathcal{F}_{k}(\bar{\bm{A}}^{\text{f}}_{l})\right)-\epsilon\alpha\left[\dfrac{\partial\mathcal{F}_{k}}{\partial A_{j}}{{\color[rgb]{0,0,0}(N^{-1})_{ik}}}\right]_{\bar{\bm{A}}^{\text{f}}_{l}} (60)
=\displaystyle= (N1)ip(𝑨¯lf)[δpj+α[Npq(N1)qkAj]𝑨¯lf(dAk(l)ϵk(𝑨¯lf))αϵ[pAj]𝑨¯lf].\displaystyle{{\color[rgb]{0,0,0}(N^{-1})_{ip}}}(\bar{\bm{A}}^{\text{f}}_{l})\left[\delta_{pj}+\alpha\left[N_{pq}\dfrac{\partial{{\color[rgb]{0,0,0}(N^{-1})_{qk}}}}{\partial A_{j}}\right]_{\bar{\bm{A}}^{\text{f}}_{l}}\left(dA_{k}(l)-\epsilon\mathcal{F}_{k}(\bar{\bm{A}}^{\text{f}}_{l})\right)-\alpha\epsilon\left[\dfrac{\partial\mathcal{F}_{p}}{\partial A_{j}}\right]_{\bar{\bm{A}}^{\text{f}}_{l}}\right].

Using the power series expression of ln(det(𝑰+δ𝓑))\ln({{\color[rgb]{0,0,0}\text{det}\left(\bm{I}+\delta\bm{\mathcal{B}}\right)}}) for any matrix 𝓑\bm{\mathcal{B}} in δ0\delta\to 0 limit (such that δ𝓑<1||\delta\bm{\mathcal{B}}||<1), that is,

ln(det(𝑰+δ𝓑))=[Tr[𝓑]δ12Tr[𝓑𝓑]δ2+𝒪(δ3)],\ln({{\color[rgb]{0,0,0}\text{det}\left(\bm{I}+\delta\bm{\mathcal{B}}\right)}})=\left[\text{Tr}[\bm{\mathcal{B}}]\delta-\dfrac{1}{2}\text{Tr}[\bm{\mathcal{B}}\cdot\bm{\mathcal{B}}]\delta^{2}+\mathcal{O}(\delta^{3})\right], (61)

the determinant of 𝑴(l)\bm{M}(l) can be written as

det(𝑴(l))\displaystyle{{\color[rgb]{0,0,0}\text{det}\left(\bm{M}(l)\right)}} =\displaystyle= det(𝑵1(𝑨¯lf))exp[α[[Njq(N1)qkAj]𝑨¯lf(dAk(l)ϵk(𝑨¯lf))ϵ[jAj]𝑨¯lf]\displaystyle{{\color[rgb]{0,0,0}\text{det}\left(\bm{N}^{-1}(\bar{\bm{A}}^{\text{f}}_{l})\right)}}\exp\left[\alpha\left[\left[N_{jq}\dfrac{\partial{{\color[rgb]{0,0,0}(N^{-1})_{qk}}}}{\partial A_{j}}\right]_{\bar{\bm{A}}^{\text{f}}_{l}}\left(dA_{k}(l)-\epsilon\mathcal{F}_{k}(\bar{\bm{A}}^{\text{f}}_{l})\right)-\epsilon\left[\dfrac{\partial\mathcal{F}_{j}}{\partial A_{j}}\right]_{\bar{\bm{A}}^{\text{f}}_{l}}\right]\right. (62)
12α2[Npq(N1)qkAjNjr(N1)riAp]𝑨¯lfdAk(l)dAi(l)+𝒪(ϵ3/2)],\displaystyle\left.{{\color[rgb]{0,0,0}-\dfrac{1}{2}\alpha^{2}\left[N_{pq}\dfrac{\partial(N^{-1})_{qk}}{\partial A_{j}}N_{jr}\dfrac{\partial(N^{-1})_{ri}}{\partial A_{p}}\right]_{\bar{\bm{A}}^{\text{f}}_{l}}dA_{k}(l)dA_{i}(l)+\mathcal{O}(\epsilon^{3/2})}}\right],
=\displaystyle= det(𝑵1(𝑨¯lf))exp[α[[(N1)qkNjqAj]𝑨¯lf(dAk(l)ϵk(𝑨¯lf))+ϵ[jAj]𝑨¯lf]\displaystyle{{\color[rgb]{0,0,0}\text{det}\left(\bm{N}^{-1}(\bar{\bm{A}}^{\text{f}}_{l})\right)}}\exp\left[-\alpha\left[\left[{{\color[rgb]{0,0,0}(N^{-1})_{qk}}}\dfrac{\partial N_{jq}}{\partial A_{j}}\right]_{\bar{\bm{A}}^{\text{f}}_{l}}\left(dA_{k}(l)-\epsilon\mathcal{F}_{k}(\bar{\bm{A}}^{\text{f}}_{l})\right)+\epsilon\left[\dfrac{\partial\mathcal{F}_{j}}{\partial A_{j}}\right]_{\bar{\bm{A}}^{\text{f}}_{l}}\right]\right.
12α2ϵ[NpmAjNjmAp]𝑨¯lf+𝒪(ϵ3/2)].\displaystyle\left.{{\color[rgb]{0,0,0}-\dfrac{1}{2}\alpha^{2}\epsilon\left[\dfrac{\partial N_{pm}}{\partial A_{j}}\dfrac{\partial N_{jm}}{\partial A_{p}}\right]_{\bar{\bm{A}}^{\text{f}}_{l}}+\mathcal{O}(\epsilon^{3/2})}}\right].

Note that dAk(l)dA_{k}(l) has a ϵ1/2\epsilon^{1/2}-term, so dAi(l)dAk(l)dA_{i}(l)dA_{k}(l) is of the order of ϵ\epsilon. Relations (85) & (90) (see Appendix D) have been used to get the last term of the above equation. Eq. (60) then reads

𝒥\displaystyle\mathcal{J} =\displaystyle= l=1N1ϵn/2det(𝑵1(𝑨¯lf))exp[α[[(N1)qkNjqAj]𝑨¯lf(dAk(l)ϵk(𝑨¯lf))+ϵ[jAj]𝑨¯lf]\displaystyle\prod^{N}_{l=1}\dfrac{1}{\epsilon^{n/2}}{{\color[rgb]{0,0,0}\text{det}\left(\bm{N}^{-1}(\bar{\bm{A}}^{\text{f}}_{l})\right)}}\exp\left[-\alpha\left[\left[{{\color[rgb]{0,0,0}(N^{-1})_{qk}}}\dfrac{\partial N_{jq}}{\partial A_{j}}\right]_{\bar{\bm{A}}^{\text{f}}_{l}}\left(dA_{k}(l)-\epsilon\mathcal{F}_{k}(\bar{\bm{A}}^{\text{f}}_{l})\right)+\epsilon\left[\dfrac{\partial\mathcal{F}_{j}}{\partial A_{j}}\right]_{\bar{\bm{A}}^{\text{f}}_{l}}\right]\right. (63)
12α2ϵ[NpmAjNjmAp]𝑨¯lf+𝒪(ϵ3/2)].\displaystyle\left.{{\color[rgb]{0,0,0}-\dfrac{1}{2}\alpha^{2}\epsilon\left[\dfrac{\partial N_{pm}}{\partial A_{j}}\dfrac{\partial N_{jm}}{\partial A_{p}}\right]_{\bar{\bm{A}}^{\text{f}}_{l}}+\mathcal{O}(\epsilon^{3/2})}}\right].

Substituting the above expression of 𝒥\mathcal{J} into Eq.(57):

P\displaystyle P =\displaystyle= p0(𝑨0)l=1N{1(2πϵ)n/2|det(𝑵1(𝑨¯lf))|exp[12ϵ[dAi(l)ϵi(𝑨¯lf)](N1)ki(𝑨¯lf)(N1)kj(𝑨¯lf)[dAj(l)ϵj(𝑨¯lf)]]\displaystyle p_{0}(\bm{A}_{0})\prod^{N}_{l=1}\left\{\dfrac{1}{(2\pi\epsilon)^{n/2}}{{\color[rgb]{0,0,0}\left|\text{det}\left(\bm{N}^{-1}(\bar{\bm{A}}^{\text{f}}_{l})\right)\right|}}\exp\left[-\dfrac{1}{2\epsilon}\left[dA_{i}(l)-\epsilon\mathcal{F}_{i}(\bar{\bm{A}}^{\text{f}}_{l})\right]{{\color[rgb]{0,0,0}(N^{-1})_{ki}}}(\bar{\bm{A}}^{\text{f}}_{l}){{\color[rgb]{0,0,0}(N^{-1})_{kj}}}(\bar{\bm{A}}^{\text{f}}_{l})\left[dA_{j}(l)-\epsilon\mathcal{F}_{j}(\bar{\bm{A}}^{\text{f}}_{l})\right]\right]\right. (64)
×exp[α[(dAi(l)ϵi(𝑨¯lf))[(N1)jiNkjAk]𝑨¯lf+ϵ[iAi]𝑨¯lf]12α2ϵ[NpmAjNjmAp]𝑨¯lf+𝒪(ϵ3/2)]}.\displaystyle\left.\times\exp\left[-\alpha\left[\left(dA_{i}(l)-\epsilon\mathcal{F}_{i}(\bar{\bm{A}}^{\text{f}}_{l})\right)\left[{{\color[rgb]{0,0,0}(N^{-1})_{ji}}}\dfrac{\partial N_{kj}}{\partial A_{k}}\right]_{\bar{\bm{A}}^{\text{f}}_{l}}+\epsilon\left[\dfrac{\partial\mathcal{F}_{i}}{\partial A_{i}}\right]_{\bar{\bm{A}}^{\text{f}}_{l}}\right]{{\color[rgb]{0,0,0}-\dfrac{1}{2}\alpha^{2}\epsilon\left[\dfrac{\partial N_{pm}}{\partial A_{j}}\dfrac{\partial N_{jm}}{\partial A_{p}}\right]_{\bar{\bm{A}}^{\text{f}}_{l}}+\mathcal{O}(\epsilon^{3/2})}}\right]\right\}.

From Eq. (2), (N1)ki(N1)kj=(Γs1)ij/2kBT{{\color[rgb]{0,0,0}(N^{-1})_{ki}}}{{\color[rgb]{0,0,0}(N^{-1})_{kj}}}={{\color[rgb]{0,0,0}(\Gamma^{s}}}^{-1})_{ij}/2k_{\text{B}}T, so

P\displaystyle P =\displaystyle= p0(𝑨0)l=1N{(2kBT)1/2(2πϵ)n/2det(𝚪s(𝑨¯lf))1/2exp[14ϵkBT[dAi(l)ϵi(𝑨¯lf)](Γs1)ij(𝑨¯lf)[dAj(l)ϵj(𝑨¯lf)]]\displaystyle p_{0}(\bm{A}_{0})\prod^{N}_{l=1}\left\{\dfrac{{{\color[rgb]{0,0,0}(2k_{\text{B}}T)^{-1/2}}}}{(2\pi\epsilon)^{n/2}}{{\color[rgb]{0,0,0}\text{det}\left(\bm{\Gamma^{\text{s}}}(\bar{\bm{A}}^{\text{f}}_{l})\right)^{-1/2}}}\exp\left[-\dfrac{1}{4\epsilon k_{\text{B}}T}\left[dA_{i}(l)-\epsilon\mathcal{F}_{i}(\bar{\bm{A}}^{\text{f}}_{l})\right]{{{\color[rgb]{0,0,0}(\Gamma^{s}}}^{-1})_{ij}}(\bar{\bm{A}}^{\text{f}}_{l})\left[dA_{j}(l)-\epsilon\mathcal{F}_{j}(\bar{\bm{A}}^{\text{f}}_{l})\right]\right]\right.
×exp[α[(dAi(l)ϵi(𝑨¯lf))[(N1)jiNkjAk]𝑨¯lf+ϵ[iAi]𝑨¯lf]12α2ϵ[NpmAjNjmAp]𝑨¯lf+𝒪(ϵ3/2)]}.\displaystyle\left.\times\exp\left[-\alpha\left[\left(dA_{i}(l)-\epsilon\mathcal{F}_{i}(\bar{\bm{A}}^{\text{f}}_{l})\right)\left[{{\color[rgb]{0,0,0}(N^{-1})_{ji}}}\dfrac{\partial N_{kj}}{\partial A_{k}}\right]_{\bar{\bm{A}}^{\text{f}}_{l}}+\epsilon\left[\dfrac{\partial\mathcal{F}_{i}}{\partial A_{i}}\right]_{\bar{\bm{A}}^{\text{f}}_{l}}\right]{{\color[rgb]{0,0,0}-\dfrac{1}{2}\alpha^{2}\epsilon\left[\dfrac{\partial N_{pm}}{\partial A_{j}}\dfrac{\partial N_{jm}}{\partial A_{p}}\right]_{\bar{\bm{A}}^{\text{f}}_{l}}+\mathcal{O}(\epsilon^{3/2})}}\right]\right\}.

Now let us break i\mathcal{F}_{i} into two terms,

i0=ΓijAj+kBTΓijAj\mathcal{F}^{0}_{i}=-\Gamma_{ij}\dfrac{\partial\mathcal{H}}{\partial A_{j}}+k_{\text{B}}T\dfrac{\partial\Gamma_{ij}}{\partial A_{j}} (66)

and

iN=αNljNijAl.\mathcal{F}^{\text{N}}_{i}=-\alpha N_{lj}\dfrac{\partial N_{ij}}{\partial A_{l}}. (67)

Replacing i\mathcal{F}_{i} by i0+iN\mathcal{F}^{0}_{i}+\mathcal{F}^{\text{N}}_{i} in Eq. (A):

P\displaystyle P =\displaystyle= p0(𝑨0)l=1N{(2kBT)1/2(2πϵ)n/2det(𝚪s(𝑨¯lf))1/2exp[14ϵkBT[dAi(l)ϵi0(𝑨¯lf)](Γs1)ij(𝑨¯lf)[dAj(l)ϵj0(𝑨¯lf)]]\displaystyle p_{0}(\bm{A}_{0})\prod^{N}_{l=1}\left\{\dfrac{(2k_{\text{B}}T)^{-1/2}}{(2\pi\epsilon)^{n/2}}\text{det}\left(\bm{\Gamma^{\text{s}}}(\bar{\bm{A}}^{\text{f}}_{l})\right)^{-1/2}\exp\left[-\dfrac{1}{4\epsilon k_{\text{B}}T}\left[dA_{i}(l)-\epsilon\mathcal{F}^{0}_{i}(\bar{\bm{A}}^{\text{f}}_{l})\right]{{(\Gamma^{s}}^{-1})_{ij}}(\bar{\bm{A}}^{\text{f}}_{l})\left[dA_{j}(l)-\epsilon\mathcal{F}^{0}_{j}(\bar{\bm{A}}^{\text{f}}_{l})\right]\right]\right. (68)
×exp[α[(dAi(l)ϵi0(𝑨¯lf))[(N1)jiNkjAk]𝑨¯lf+ϵ[i0Ai]𝑨¯lf]]\displaystyle\times\exp\left[-\alpha\left[\left(dA_{i}(l)-\epsilon\mathcal{F}^{0}_{i}(\bar{\bm{A}}^{\text{f}}_{l})\right)\left[(N^{-1})_{ji}\dfrac{\partial N_{kj}}{\partial A_{k}}\right]_{\bar{\bm{A}}^{\text{f}}_{l}}+\epsilon\left[\dfrac{\partial\mathcal{F}^{0}_{i}}{\partial A_{i}}\right]_{\bar{\bm{A}}^{\text{f}}_{l}}\right]\right]
×exp[12kBTiN(𝑨¯lf)(Γs1)ij(𝑨¯lf)[dAj(l)ϵj0(𝑨¯lf)]ϵ4kBT[iN(Γs1)ijjN]𝑨¯lf]\displaystyle\times\exp\left[\dfrac{1}{2k_{\text{B}}T}\mathcal{F}^{\text{N}}_{i}(\bar{\bm{A}}^{\text{f}}_{l}){{(\Gamma^{s}}^{-1})_{ij}}(\bar{\bm{A}}^{\text{f}}_{l})\left[dA_{j}(l)-\epsilon\mathcal{F}^{0}_{j}(\bar{\bm{A}}^{\text{f}}_{l})\right]-\dfrac{\epsilon}{4k_{\text{B}}T}\left[\mathcal{F}^{\text{N}}_{i}{{(\Gamma^{s}}^{-1})_{ij}}\mathcal{F}^{\text{N}}_{j}\right]_{\bar{\bm{A}}^{\text{f}}_{l}}\right]
×exp[αϵ[iN(N1)jiNkjAkiNAi]𝑨¯lf12α2ϵ[NpmAjNjmAp]𝑨¯lf+𝒪(ϵ3/2)]}.\displaystyle\left.\times\exp\left[\alpha\epsilon\left[\mathcal{F}^{\text{N}}_{i}(N^{-1})_{ji}\dfrac{\partial N_{kj}}{\partial A_{k}}-\dfrac{\partial\mathcal{F}^{\text{N}}_{i}}{\partial A_{i}}\right]_{\bar{\bm{A}}^{\text{f}}_{l}}-\dfrac{1}{2}\alpha^{2}\epsilon\left[\dfrac{\partial N_{pm}}{\partial A_{j}}\dfrac{\partial N_{jm}}{\partial A_{p}}\right]_{\bar{\bm{A}}^{\text{f}}_{l}}+\mathcal{O}(\epsilon^{3/2})\right]\right\}.

Using Eq. (6) (that is, NikNjk=2kBTΓijsN_{ik}N_{jk}=2k_{\text{B}}T\Gamma^{\text{s}}_{ij}), one can write

(N1)jiNkjAk\displaystyle(N^{-1})_{ji}\dfrac{\partial N_{kj}}{\partial A_{k}} =\displaystyle= δim(N1)jmNkjAk\displaystyle\delta_{im}(N^{-1})_{jm}\dfrac{\partial N_{kj}}{\partial A_{k}} (69)
=\displaystyle= (Γs1)ipΓspm(N1)jmNkjAk\displaystyle({\Gamma^{s}}^{-1})_{ip}{\Gamma^{s}}_{pm}(N^{-1})_{jm}\dfrac{\partial N_{kj}}{\partial A_{k}}
=\displaystyle= 12kBT(Γs1)ipNpjNkjAk\displaystyle\dfrac{1}{2k_{\text{B}}T}({\Gamma^{s}}^{-1})_{ip}N_{pj}\dfrac{\partial N_{kj}}{\partial A_{k}}
=\displaystyle= 12kBT(Γs1)ip((NkjNpj)AkNkjNpjAk)\displaystyle\dfrac{1}{2k_{\text{B}}T}({\Gamma^{s}}^{-1})_{ip}\left(\dfrac{\partial(N_{kj}N_{pj})}{\partial A_{k}}-N_{kj}\dfrac{\partial N_{pj}}{\partial A_{k}}\right)
=\displaystyle= (Γs1)ipΓspkAk+12kBTα(Γs1)ippN.\displaystyle({\Gamma^{s}}^{-1})_{ip}\dfrac{\partial{\Gamma^{s}}_{pk}}{\partial A_{k}}+\dfrac{1}{2k_{\text{B}}T\alpha}({\Gamma^{s}}^{-1})_{ip}\mathcal{F}^{\text{N}}_{p}.

With the above expression, Eq. (68) reduces to

P\displaystyle P =\displaystyle= p0(𝑨0)l=1N{(2kBT)1/2(2πϵ)n/2det(𝚪s(𝑨¯lf))1/2exp[14ϵkBT[dAi(l)ϵi0(𝑨¯lf)](Γs1)ij(𝑨¯lf)[dAj(l)ϵj0(𝑨¯lf)].\displaystyle p_{0}(\bm{A}_{0})\prod^{N}_{l=1}\left\{\dfrac{(2k_{\text{B}}T)^{-1/2}}{(2\pi\epsilon)^{n/2}}\text{det}\left(\bm{\Gamma^{\text{s}}}(\bar{\bm{A}}^{\text{f}}_{l})\right)^{-1/2}\exp\Biggl{[}-\dfrac{1}{4\epsilon k_{\text{B}}T}\left[dA_{i}(l)-\epsilon\mathcal{F}^{0}_{i}(\bar{\bm{A}}^{\text{f}}_{l})\right]{{{\color[rgb]{0,0,0}(\Gamma^{s}}}^{-1})_{ij}}(\bar{\bm{A}}^{\text{f}}_{l})\left[dA_{j}(l)-\epsilon\mathcal{F}^{0}_{j}(\bar{\bm{A}}^{\text{f}}_{l})\right]\Biggr{.}\right.
α[(dAi(l)ϵi0(𝑨¯lf))[(Γs1)ijΓjksAk]𝑨¯lf+ϵ[i0Ai]𝑨¯lf]]\displaystyle\left.-\alpha\left[\left(dA_{i}(l)-\epsilon\mathcal{F}^{0}_{i}(\bar{\bm{A}}^{\text{f}}_{l})\right)\left[({\Gamma^{s}}^{-1})_{ij}\dfrac{\partial\Gamma^{s}_{jk}}{\partial A_{k}}\right]_{\bar{\bm{A}}^{\text{f}}_{l}}+\epsilon\left[\dfrac{\partial\mathcal{F}^{0}_{i}}{\partial A_{i}}\right]_{\bar{\bm{A}}^{\text{f}}_{l}}\right]\right]
×exp[ϵ4kBT[iN(Γs1)ij(jN+4αkBTΓjksAk)4αkBTiNAi2kBTα2NpmAjNjmAp]𝑨¯lf+𝒪(ϵ3/2)]}.\displaystyle\times\left.\exp\left[\dfrac{\epsilon}{4k_{\text{B}}T}\left[\mathcal{F}^{\text{N}}_{i}({\Gamma^{s}}^{-1})_{ij}\left(\mathcal{F}^{\text{N}}_{j}+4\alpha k_{\text{B}}T\dfrac{\partial\Gamma^{s}_{jk}}{\partial A_{k}}\right)-4\alpha k_{\text{B}}T\dfrac{\partial\mathcal{F}^{\text{N}}_{i}}{\partial A_{i}}-2k_{\text{B}}T\alpha^{2}\dfrac{\partial N_{pm}}{\partial A_{j}}\dfrac{\partial N_{jm}}{\partial A_{p}}\right]_{\bar{\bm{A}}^{\text{f}}_{l}}+\mathcal{O}(\epsilon^{3/2})\right]\right\}.

Using the relation NikNjk=2kBTΓijsN_{ik}N_{jk}=2k_{\text{B}}T\Gamma^{\text{s}}_{ij}, further simplifying the last term of the above equation yields

P\displaystyle P =\displaystyle= p0(𝑨0)l=1N{(2kBT)1/2(2πϵ)n/2det(𝚪s(𝑨¯lf))1/2exp[14ϵkBT[dAi(l)ϵi0(𝑨¯lf)](Γs1)ij(𝑨¯lf)[dAj(l)ϵj0(𝑨¯lf)].\displaystyle p_{0}(\bm{A}_{0})\prod^{N}_{l=1}\left\{\dfrac{(2k_{\text{B}}T)^{-1/2}}{(2\pi\epsilon)^{n/2}}\text{det}\left(\bm{\Gamma^{\text{s}}}(\bar{\bm{A}}^{\text{f}}_{l})\right)^{-1/2}\exp\Biggl{[}-\dfrac{1}{4\epsilon k_{\text{B}}T}\left[dA_{i}(l)-\epsilon\mathcal{F}^{0}_{i}(\bar{\bm{A}}^{\text{f}}_{l})\right]{{{\color[rgb]{0,0,0}(\Gamma^{s}}}^{-1})_{ij}}(\bar{\bm{A}}^{\text{f}}_{l})\left[dA_{j}(l)-\epsilon\mathcal{F}^{0}_{j}(\bar{\bm{A}}^{\text{f}}_{l})\right]\Biggr{.}\right. (71)
α[(dAi(l)ϵi0(𝑨¯lf))[(Γs1)ijΓjksAk]𝑨¯lf+ϵ[i0Ai]𝑨¯lf]]\displaystyle\left.-\alpha\left[\left(dA_{i}(l)-\epsilon\mathcal{F}^{0}_{i}(\bar{\bm{A}}^{\text{f}}_{l})\right)\left[({\Gamma^{s}}^{-1})_{ij}\dfrac{\partial\Gamma^{s}_{jk}}{\partial A_{k}}\right]_{\bar{\bm{A}}^{\text{f}}_{l}}+\epsilon\left[\dfrac{\partial\mathcal{F}^{0}_{i}}{\partial A_{i}}\right]_{\bar{\bm{A}}^{\text{f}}_{l}}\right]\right]
×exp[α2ϵkBT[2ΓijsAiAjΓiksAk(Γs1)ijΓjpsAp]𝑨¯lf+𝒪(ϵ3/2)]}.\displaystyle\times\left.\exp\left[\alpha^{2}\epsilon k_{\text{B}}T\left[\dfrac{\partial^{2}\Gamma^{s}_{ij}}{\partial A_{i}\partial A_{j}}-\dfrac{\partial\Gamma^{s}_{ik}}{\partial A_{k}}({\Gamma^{s}}^{-1})_{ij}\dfrac{\partial\Gamma^{s}_{jp}}{\partial A_{p}}\right]_{\bar{\bm{A}}^{\text{f}}_{l}}+\mathcal{O}(\epsilon^{3/2})\right]\right\}.

We further split i0\mathcal{F}^{0}_{i} into the two terms,

is(𝑨)=ΓijsAj+kBTΓijsAj\mathcal{F}^{\text{s}}_{i}(\bm{A})=-\Gamma^{\text{s}}_{ij}\dfrac{\partial\mathcal{H}}{\partial A_{j}}+k_{\text{B}}T\dfrac{\partial\Gamma^{\text{s}}_{ij}}{\partial A_{j}} (72)

and

ia(𝑨)=ΓijaAj+kBTΓijaAj,\mathcal{F}^{a}_{i}(\bm{A})=-\Gamma^{a}_{ij}\dfrac{\partial\mathcal{H}}{\partial A_{j}}+k_{\text{B}}T\dfrac{\partial\Gamma^{\text{a}}_{ij}}{\partial A_{j}}, (73)

such that, under time reversal (see  (17) and  (18)),

𝓕s(𝑨)𝓕s(𝒔𝑨)=𝒔𝓕s(𝑨),\displaystyle\bm{\mathcal{F}}^{\text{s}}(\bm{A})\to\bm{\mathcal{F}}^{\text{s}}(\bm{s}\circ\bm{A})=\bm{s}\circ\bm{\mathcal{F}}^{\text{s}}(\bm{A}), (74)
𝓕a(𝑨)𝓕a(𝒔𝑨)=𝒔𝓕a(𝑨).\displaystyle\bm{\mathcal{F}}^{\text{a}}(\bm{A})\to\bm{\mathcal{F}}^{\text{a}}(\bm{s}\circ\bm{A})=-\bm{s}\circ\bm{\mathcal{F}}^{\text{a}}(\bm{A}). (75)

Eq. (A) then becomes

P\displaystyle P =\displaystyle= p0(𝑨0)l=1N{(2kBT)1/2(2πϵ)n/2exp[14ϵkBT[dAi(l)ϵis(𝑨¯lf)ϵia(𝑨¯lf)](Γs1)ij(𝑨¯lf)[dAj(l)ϵjs(𝑨¯lf)ϵja(𝑨¯lf)].\displaystyle p_{0}(\bm{A}_{0})\prod^{N}_{l=1}\left\{\dfrac{(2k_{\text{B}}T)^{-1/2}}{(2\pi\epsilon)^{n/2}}\exp\Biggl{[}-\dfrac{1}{4\epsilon k_{\text{B}}T}\left[dA_{i}(l)-\epsilon\mathcal{F}^{\text{s}}_{i}(\bar{\bm{A}}^{\text{f}}_{l})-\epsilon\mathcal{F}^{\text{a}}_{i}(\bar{\bm{A}}^{\text{f}}_{l})\right]{{{\color[rgb]{0,0,0}(\Gamma^{s}}}^{-1})_{ij}}(\bar{\bm{A}}^{\text{f}}_{l})\left[dA_{j}(l)-\epsilon\mathcal{F}^{\text{s}}_{j}(\bar{\bm{A}}^{\text{f}}_{l})-\epsilon\mathcal{F}^{\text{a}}_{j}(\bar{\bm{A}}^{\text{f}}_{l})\right]\Biggr{.}\right. (76)
α[(dAi(l)ϵis(𝑨¯lf)ϵia(𝑨¯lf))[(Γs1)ijΓjksAk]𝑨¯lf+ϵ[isAi+iaAi]𝑨¯lf]]det(𝚪s(𝑨¯lf))1/2\displaystyle\left.-\alpha\left[\left(dA_{i}(l)-\epsilon\mathcal{F}^{\text{s}}_{i}(\bar{\bm{A}}^{\text{f}}_{l})-\epsilon\mathcal{F}^{\text{a}}_{i}(\bar{\bm{A}}^{\text{f}}_{l})\right)\left[({\Gamma^{s}}^{-1})_{ij}\dfrac{\partial\Gamma^{s}_{jk}}{\partial A_{k}}\right]_{\bar{\bm{A}}^{\text{f}}_{l}}+\epsilon\left[\dfrac{\partial\mathcal{F}^{\text{s}}_{i}}{\partial A_{i}}+\dfrac{\partial\mathcal{F}^{\text{a}}_{i}}{\partial A_{i}}\right]_{\bar{\bm{A}}^{\text{f}}_{l}}\right]\right]\text{det}\left(\bm{\Gamma^{\text{s}}}(\bar{\bm{A}}^{\text{f}}_{l})\right)^{-1/2}
×exp[α2ϵkBT[2ΓijsAiAjΓiksAk(Γs1)ijΓjpsAp]𝑨¯lf+𝒪(ϵ3/2)]}.\displaystyle\times\left.\exp\left[\alpha^{2}\epsilon k_{\text{B}}T\left[\dfrac{\partial^{2}\Gamma^{s}_{ij}}{\partial A_{i}\partial A_{j}}-\dfrac{\partial\Gamma^{s}_{ik}}{\partial A_{k}}({\Gamma^{s}}^{-1})_{ij}\dfrac{\partial\Gamma^{s}_{jp}}{\partial A_{p}}\right]_{\bar{\bm{A}}^{\text{f}}_{l}}+\mathcal{O}(\epsilon^{3/2})\right]\right\}.

Clearly, for given Γijs\Gamma^{\text{s}}_{ij}, PP is independent of the choice of NijN_{ij}.

Appendix B The probability density for the time-reversed trajectory

As the time-reversed trajectory of the trajectory (𝑨0,𝑨1,𝑨2,..𝑨N)(\bm{A}_{0},\bm{A}_{1},\bm{A}_{2},.....\bm{A}_{N}) is (𝒔𝑨N,𝒔𝑨N1,..𝒔𝑨1)(\bm{s}\circ\bm{A}_{N},\bm{s}\circ\bm{A}_{N-1},.....\bm{s}\circ\bm{A}_{1}), under time reversal, 𝑨l𝒔𝑨Nl\bm{A}_{l}\to\bm{s}\circ\bm{A}_{N-l} and therefore,

dAi(l)\displaystyle dA_{i}(l) =\displaystyle= Ai(ϵl)Ai(ϵ(l1))\displaystyle A_{i}(\epsilon l)-A_{i}(\epsilon(l-1)) (77)
\displaystyle\to si(Ai(ϵ(Nl))Ai(ϵ(Nl+1))\displaystyle s_{i}(A_{i}(\epsilon(N-l))-A_{i}(\epsilon(N-l+1))
\displaystyle\to sidAi(Nl+1),\displaystyle-s_{i}dA_{i}(N-l+1),

(Einstein’s convention is not used here) and

𝑨¯lf\displaystyle\bar{\bm{A}}^{\text{f}}_{l} =\displaystyle= α𝑨l+(1α)𝑨l1\displaystyle\alpha\bm{A}_{l}+(1-\alpha)\bm{A}_{l-1} (78)
\displaystyle\to (α𝒔𝑨Nl+(1α)𝒔𝑨Nl+1)\displaystyle(\alpha\bm{s}\circ\bm{A}_{N-l}+(1-\alpha)\bm{s}\circ\bm{A}_{N-l+1})
\displaystyle\to 𝒔𝑨¯Nl+1r,\displaystyle\bm{s}\circ\bar{\bm{A}}^{\text{r}}_{N-l+1},

where 𝑨¯lr(1α)𝑨(l)+α𝑨(l1)\bar{\bm{A}}^{\text{r}}_{l}\equiv(1-\alpha)\bm{A}(l)+\alpha\bm{A}(l-1). With the above transformations, using the relation Γijs=sisjΓijs\Gamma^{\text{s}}_{ij}=s_{i}s_{j}\Gamma^{\text{s}}_{ij} and Eqs. (74),  (75) & (76), we obtain the following expression of the probability density of the time-reversed trajectory:

Pr\displaystyle P_{r} =\displaystyle= l=1N{(2kBT)1/2(2πϵ)n/2exp[14ϵkBT[dAi(l)ϵis(𝑨¯lr)+ϵia(𝑨¯lr)](Γs1)ij(𝑨¯lr)[dAj(l)ϵjs(𝑨¯lr)+ϵja(𝑨¯lr)].\displaystyle\prod^{N}_{l=1}\left\{\dfrac{(2k_{\text{B}}T)^{-1/2}}{(2\pi\epsilon)^{n/2}}\exp\Biggl{[}-\dfrac{1}{4\epsilon k_{\text{B}}T}\left[-dA_{i}(l^{\prime})-\epsilon\mathcal{F}^{\text{s}}_{i}(\bar{\bm{A}}^{\text{r}}_{l^{\prime}})+\epsilon\mathcal{F}^{\text{a}}_{i}(\bar{\bm{A}}^{\text{r}}_{l^{\prime}})\right]{{{\color[rgb]{0,0,0}(\Gamma^{s}}}^{-1})_{ij}}(\bar{\bm{A}}^{\text{r}}_{l^{\prime}})\left[-dA_{j}(l^{\prime})-\epsilon\mathcal{F}^{\text{s}}_{j}(\bar{\bm{A}}^{\text{r}}_{l^{\prime}})+\epsilon\mathcal{F}^{\text{a}}_{j}(\bar{\bm{A}}^{\text{r}}_{l^{\prime}})\right]\Biggr{.}\right. (79)
α[(dAi(l)ϵis(𝑨¯lr)+ϵia(𝑨¯lr))[(Γs1)ijΓjksAk]𝑨¯lr+ϵ[isAiiaAi]𝑨¯lr]]det(𝚪s(𝑨¯lr))1/2\displaystyle\left.-\alpha\left[\left(-dA_{i}(l^{\prime})-\epsilon\mathcal{F}^{\text{s}}_{i}(\bar{\bm{A}}^{\text{r}}_{l^{\prime}})+\epsilon\mathcal{F}^{\text{a}}_{i}(\bar{\bm{A}}^{\text{r}}_{l^{\prime}})\right)\left[({\Gamma^{s}}^{-1})_{ij}\dfrac{\partial\Gamma^{s}_{jk}}{\partial A_{k}}\right]_{\bar{\bm{A}}^{\text{r}}_{l^{\prime}}}+\epsilon\left[\dfrac{\partial\mathcal{F}^{\text{s}}_{i}}{\partial A_{i}}-\dfrac{\partial\mathcal{F}^{\text{a}}_{i}}{\partial A_{i}}\right]_{\bar{\bm{A}}^{\text{r}}_{l^{\prime}}}\right]\right]\text{det}\left(\bm{\Gamma^{\text{s}}}(\bar{\bm{A}}^{\text{r}}_{l^{\prime}})\right)^{-1/2}
×exp[α2ϵkBT[2ΓijsAiAjΓiksAk(Γs1)ijΓjpsAp]𝑨¯lr+𝒪(ϵ3/2)]}p0(𝒔𝑨N),\displaystyle\times\left.\exp\left[\alpha^{2}\epsilon k_{\text{B}}T\left[\dfrac{\partial^{2}\Gamma^{s}_{ij}}{\partial A_{i}\partial A_{j}}-\dfrac{\partial\Gamma^{s}_{ik}}{\partial A_{k}}({\Gamma^{s}}^{-1})_{ij}\dfrac{\partial\Gamma^{s}_{jp}}{\partial A_{p}}\right]_{\bar{\bm{A}}^{\text{r}}_{l^{\prime}}}+\mathcal{O}(\epsilon^{3/2})\right]\right\}p_{0}(\bm{s}\circ\bm{A}_{N}),

where l=Nl+1l^{\prime}=N-l+1. In the above equation, the index ll^{\prime} runs from NN to 1 so we can replace l=1N\prod^{N}_{l=1} by l=N1l=1N\prod^{1}_{l^{\prime}=N}\equiv\prod^{N}_{l^{\prime}=1}. Hence

Pr\displaystyle P_{r} =\displaystyle= l=1N{(2kBT)1/2(2πϵ)n/2exp[14ϵkBT[dAi(l)ϵis(𝑨¯lr)+ϵia(𝑨¯lr)](Γs1)ij(𝑨¯lr)[dAj(l)ϵjs(𝑨¯lr)+ϵja(𝑨¯lr)].\displaystyle\prod^{N}_{l=1}\left\{\dfrac{(2k_{\text{B}}T)^{-1/2}}{(2\pi\epsilon)^{n/2}}\exp\Biggl{[}-\dfrac{1}{4\epsilon k_{\text{B}}T}\left[-dA_{i}(l)-\epsilon\mathcal{F}^{\text{s}}_{i}(\bar{\bm{A}}^{\text{r}}_{l})+\epsilon\mathcal{F}^{\text{a}}_{i}(\bar{\bm{A}}^{\text{r}}_{l})\right]{{{\color[rgb]{0,0,0}(\Gamma^{s}}}^{-1})_{ij}}(\bar{\bm{A}}^{\text{r}}_{l})\left[-dA_{j}(l)-\epsilon\mathcal{F}^{\text{s}}_{j}(\bar{\bm{A}}^{\text{r}}_{l})+\epsilon\mathcal{F}^{\text{a}}_{j}(\bar{\bm{A}}^{\text{r}}_{l})\right]\Biggr{.}\right. (80)
α[(dAi(l)ϵis(𝑨¯lr)+ϵia(𝑨¯lr))[(Γs1)ijΓjksAk]𝑨¯lr+ϵ[isAiiaAi]𝑨¯lr]]det(𝚪s(𝑨¯lr))1/2\displaystyle\left.-\alpha\left[\left(-dA_{i}(l)-\epsilon\mathcal{F}^{\text{s}}_{i}(\bar{\bm{A}}^{\text{r}}_{l})+\epsilon\mathcal{F}^{\text{a}}_{i}(\bar{\bm{A}}^{\text{r}}_{l})\right)\left[({\Gamma^{s}}^{-1})_{ij}\dfrac{\partial\Gamma^{s}_{jk}}{\partial A_{k}}\right]_{\bar{\bm{A}}^{\text{r}}_{l}}+\epsilon\left[\dfrac{\partial\mathcal{F}^{\text{s}}_{i}}{\partial A_{i}}-\dfrac{\partial\mathcal{F}^{\text{a}}_{i}}{\partial A_{i}}\right]_{\bar{\bm{A}}^{\text{r}}_{l}}\right]\right]\text{det}\left(\bm{\Gamma^{\text{s}}}(\bar{\bm{A}}^{\text{r}}_{l})\right)^{-1/2}
×exp[α2ϵkBT[2ΓijsAiAjΓiksAk(Γs1)ijΓjpsAp]𝑨¯lr+𝒪(ϵ3/2)]}p0(𝒔𝑨N).\displaystyle\times\left.\exp\left[\alpha^{2}\epsilon k_{\text{B}}T\left[\dfrac{\partial^{2}\Gamma^{s}_{ij}}{\partial A_{i}\partial A_{j}}-\dfrac{\partial\Gamma^{s}_{ik}}{\partial A_{k}}({\Gamma^{s}}^{-1})_{ij}\dfrac{\partial\Gamma^{s}_{jp}}{\partial A_{p}}\right]_{\bar{\bm{A}}^{\text{r}}_{l}}+\mathcal{O}(\epsilon^{3/2})\right]\right\}p_{0}(\bm{s}\circ\bm{A}_{N}).

Appendix C Calculation of the ratio between the probability densities of a trajectory and its time-reversed trajectory

Using relations (86) & (87), expanding Nij(𝑨¯lf)N_{ij}(\bar{\bm{A}}^{\text{f}}_{l}) and Nij(𝑨¯lr)N_{ij}(\bar{\bm{A}}^{\text{r}}_{l}) around 𝑨=𝑨¯l(𝑨l+𝑨l1)/2\bm{A}=\bar{\bm{A}}_{l}\equiv(\bm{A}_{l}+\bm{A}_{l-1})/2:

Nij(𝑨¯lf)\displaystyle N_{ij}(\bar{\bm{A}}^{\text{f}}_{l}) =\displaystyle= Nij(𝑨¯l)+2α12[NijAk]𝑨¯ldAk(l)+12(2α12)2[2NijAkAm]𝑨¯ldAk(l)dAm(l)+𝒪(ϵ3/2),\displaystyle N_{ij}(\bar{\bm{A}}_{l})+\dfrac{2\alpha-1}{2}\left[\dfrac{\partial N_{ij}}{\partial A_{k}}\right]_{\bar{\bm{A}}_{l}}dA_{k}(l)+\dfrac{1}{2}\left(\dfrac{2\alpha-1}{2}\right)^{2}\left[\dfrac{\partial^{2}N_{ij}}{\partial A_{k}\partial A_{m}}\right]_{\bar{\bm{A}}_{l}}dA_{k}(l)dA_{m}(l)+\mathcal{O}(\epsilon^{3/2}), (81)
Nij(𝑨¯lr)\displaystyle N_{ij}(\bar{\bm{A}}^{\text{r}}_{l}) =\displaystyle= Nij(𝑨¯l)2α12[NijAk]𝑨¯ldAk(l)+12(2α12)2[2NijAkAm]𝑨¯ldAk(l)dAm(l)+𝒪(ϵ3/2).\displaystyle N_{ij}(\bar{\bm{A}}_{l})-\dfrac{2\alpha-1}{2}\left[\dfrac{\partial N_{ij}}{\partial A_{k}}\right]_{\bar{\bm{A}}_{l}}dA_{k}(l)+\dfrac{1}{2}\left(\dfrac{2\alpha-1}{2}\right)^{2}\left[\dfrac{\partial^{2}N_{ij}}{\partial A_{k}\partial A_{m}}\right]_{\bar{\bm{A}}_{l}}dA_{k}(l)dA_{m}(l)+\mathcal{O}(\epsilon^{3/2}). (82)

Then, using the relation NikNjk=2kBTΓijsN_{ik}N_{jk}=2k_{\text{B}}T\Gamma^{\text{s}}_{ij} and Eq. (61), we obtain

det(𝚪s(𝑨¯lf))1/2det(𝚪s(𝑨¯lr))1/2\displaystyle\dfrac{\text{det}\left(\bm{\Gamma^{\text{s}}}(\bar{\bm{A}}^{\text{f}}_{l})\right)^{-1/2}}{\text{det}\left(\bm{\Gamma^{\text{s}}}(\bar{\bm{A}}^{\text{r}}_{l})\right)^{-1/2}} =\displaystyle= |det(𝑵1(𝑨¯lf))||det(𝑵1(𝑨¯lr))|\displaystyle\dfrac{\left|\text{det}\left(\bm{N}^{-1}(\bar{\bm{A}}^{\text{f}}_{l})\right)\right|\ }{\left|\text{det}\left(\bm{N}^{-1}(\bar{\bm{A}}^{\text{r}}_{l})\right)\right|} (83)
=\displaystyle= exp[(2α1)[(N1)jmNmjAk]𝑨¯ldAk(l)+𝒪(ϵ3/2)]\displaystyle\exp\left[-(2\alpha-1)\left[{{\color[rgb]{0,0,0}(N^{-1})_{jm}}}\dfrac{\partial N_{mj}}{\partial A_{k}}\right]_{\bar{\bm{A}}_{l}}dA_{k}(l)+\mathcal{O}(\epsilon^{3/2})\right]
=\displaystyle= exp[2α12[(Γs1)jm(Γs)mjAk]𝑨¯ldAk(l)+𝒪(ϵ3/2)].\displaystyle\exp\left[-\dfrac{2\alpha-1}{2}\left[({\Gamma^{s}}^{-1})_{jm}\dfrac{\partial(\Gamma^{s})_{mj}}{\partial A_{k}}\right]_{\bar{\bm{A}}_{l}}dA_{k}(l)+\mathcal{O}(\epsilon^{3/2})\right].

In ϵ0\epsilon\to 0 limit, dividing  (76) by  (80) first, and using the above equation and the relations given in Appendix D, we get

PPr\displaystyle\dfrac{P}{P_{r}} =\displaystyle= p0(𝑨0)p0(𝒔𝑨N)exp[1kBTl=1N((𝑨l)(𝑨l1))+𝒪(ϵ3/2)]\displaystyle\dfrac{p_{0}(\bm{A}_{0})}{p_{0}(\bm{s}\circ\bm{A}_{N})}\exp\left[-\dfrac{1}{k_{\text{B}}T}\sum^{N}_{l=1}\left(\mathcal{H}(\bm{A}_{l})-\mathcal{H}(\bm{A}_{l-1})\right)+\mathcal{O}(\epsilon^{3/2})\right] (84)
=\displaystyle= p0(𝑨(0))p0(𝒔𝑨(τ))exp[1kBT((𝑨(τ))(𝑨(0)))],\displaystyle\dfrac{p_{0}(\bm{A}(0))}{p_{0}(\bm{s}\circ\bm{A}(\tau))}\exp\left[-\dfrac{1}{k_{\text{B}}T}\left(\mathcal{H}(\bm{A}(\tau))-\mathcal{H}(\bm{A}(0))\right)\right],

where 𝑨(0)𝑨0\bm{A}(0)\equiv\bm{A}_{0} and 𝑨(τ)𝑨N\bm{A}(\tau)\equiv\bm{A}_{N}.

Appendix D Various relations needed for the calculation in subsection II.2

Recalling Eq. (10) of the main text

dAi(l)=ϵi(𝑨¯lf)+ϵNij(𝑨¯lf)ξjl.dA_{i}(l)=\epsilon\mathcal{F}_{i}(\bar{\bm{A}}^{\text{f}}_{l})+\sqrt{\epsilon}N_{ij}(\bar{\bm{A}}^{\text{f}}_{l})\xi_{j}^{l}. (85)

Note that the lowest order term in dAi(l)dA_{i}(l) is a ϵ1/2\epsilon^{1/2}-term. Let us consider a function G(𝑨)G(\bm{A}); expanding G(𝑨¯lf)G(\bar{\bm{A}}^{\text{f}}_{l}) and G(𝑨¯lr)G(\bar{\bm{A}}^{\text{r}}_{l}) around 𝑨=𝑨¯l(𝑨l+𝑨l1)/2\bm{A}=\bar{\bm{A}}_{l}\equiv(\bm{A}_{l}+\bm{A}_{l-1})/2:

G(𝑨¯lf)\displaystyle G(\bar{\bm{A}}^{\text{f}}_{l}) =\displaystyle= G(𝑨¯l+2α12d𝑨l)\displaystyle G(\bar{\bm{A}}_{l}+\dfrac{2\alpha-1}{2}d\bm{A}_{l}) (86)
=\displaystyle= G(𝑨¯l)+2α12[GAk]𝑨¯ldAk(l)+12(2α12)2[2GAkAm]𝑨¯ldAk(l)dAm(l)+𝒪(ϵ3/2)\displaystyle G(\bar{\bm{A}}_{l})+\dfrac{2\alpha-1}{2}\left[\dfrac{\partial G}{\partial A_{k}}\right]_{\bar{\bm{A}}_{l}}dA_{k}(l)+\dfrac{1}{2}\left(\dfrac{2\alpha-1}{2}\right)^{2}\left[\dfrac{\partial^{2}G}{\partial A_{k}\partial A_{m}}\right]_{\bar{\bm{A}}_{l}}dA_{k}(l)dA_{m}(l)+\mathcal{O}(\epsilon^{3/2})

and

G(𝑨¯lr)\displaystyle G(\bar{\bm{A}}^{\text{r}}_{l}) =\displaystyle= G(𝑨¯l2α12d𝑨l)\displaystyle G(\bar{\bm{A}}_{l}-\dfrac{2\alpha-1}{2}d\bm{A}_{l}) (87)
=\displaystyle= G(𝑨¯l)2α12[GAk]𝑨¯ldAk(l)+12(2α12)2[2GAkAm]𝑨¯ldAk(l)dAm(l)+𝒪(ϵ3/2),\displaystyle G(\bar{\bm{A}}_{l})-\dfrac{2\alpha-1}{2}\left[\dfrac{\partial G}{\partial A_{k}}\right]_{\bar{\bm{A}}_{l}}dA_{k}(l)+\dfrac{1}{2}\left(\dfrac{2\alpha-1}{2}\right)^{2}\left[\dfrac{\partial^{2}G}{\partial A_{k}\partial A_{m}}\right]_{\bar{\bm{A}}_{l}}dA_{k}(l)dA_{m}(l)+\mathcal{O}(\epsilon^{3/2}),

where d𝑨l=𝑨l𝑨l1d\bm{A}_{l}=\bm{A}_{l}-\bm{A}_{l-1}. Then

dAi(l)dAj(l)G(𝑨¯lf)\displaystyle dA_{i}(l)dA_{j}(l)G(\bar{\bm{A}}^{\text{f}}_{l}) =\displaystyle= dAi(l)dAj(l)G(𝑨¯l)+2α12[GAk]𝑨¯ldAi(l)dAj(l)dAk(l)\displaystyle dA_{i}(l)dA_{j}(l)G(\bar{\bm{A}}_{l})+\dfrac{2\alpha-1}{2}\left[\dfrac{\partial G}{\partial A_{k}}\right]_{\bar{\bm{A}}_{l}}dA_{i}(l)dA_{j}(l)dA_{k}(l) (88)
+12(2α12)2[2GAkAm]𝑨¯ldAi(l)dAj(l)dAk(l)dAm(l)+𝒪(ϵ5/2).\displaystyle+\dfrac{1}{2}\left(\dfrac{2\alpha-1}{2}\right)^{2}\left[\dfrac{\partial^{2}G}{\partial A_{k}\partial A_{m}}\right]_{\bar{\bm{A}}_{l}}dA_{i}(l)dA_{j}(l)dA_{k}(l)dA_{m}(l)+\mathcal{O}(\epsilon^{5/2}).

From Eq. (85),

dAi(l)dAj(l)dAk(l)=ξplξqlξrlNipNjqNkrϵ3/2+(ξplξqlNipNjqk+ξplξrlNipNkrj+ξqlξrlNjqNkri)ϵ2+𝒪(ϵ5/2).dA_{i}(l)dA_{j}(l)dA_{k}(l)=\xi_{p}^{l}\xi_{q}^{l}\xi_{r}^{l}N_{ip}N_{jq}N_{kr}\epsilon^{3/2}+(\xi_{p}^{l}\xi_{q}^{l}N_{ip}N_{jq}\mathcal{F}_{k}+\xi_{p}^{l}\xi_{r}^{l}N_{ip}N_{kr}\mathcal{F}_{j}+\xi_{q}^{l}\xi_{r}^{l}N_{jq}N_{kr}\mathcal{F}_{i})\epsilon^{2}+\mathcal{O}(\epsilon^{5/2}). (89)

Our final expressions will be written in integral form, and since ξj(t)\xi_{j}(t) is the time derivative of a Wiener process, we can write:

ξplξql\displaystyle\xi_{p}^{l}\xi_{q}^{l} \displaystyle\equiv δpq\displaystyle\delta_{pq} (90)
ξplξqlξrl\displaystyle\xi_{p}^{l}\xi_{q}^{l}\xi_{r}^{l} \displaystyle\equiv δpqξrl+δprξql+δrqξpl\displaystyle\delta_{pq}\xi_{r}^{l}+\delta_{pr}\xi_{q}^{l}+\delta_{rq}\xi_{p}^{l} (91)
ξplξqlξrlξol\displaystyle\xi_{p}^{l}\xi_{q}^{l}\xi_{r}^{l}\xi_{o}^{l} \displaystyle\equiv δpqδro+δprδqo+δqrδpo.\displaystyle\delta_{pq}\delta_{ro}+\delta_{pr}\delta_{qo}+\delta_{qr}\delta_{po}. (92)

Using the above relations, Eq. (85) & Eq. (6) (NikNjk=2kBTΓijsN_{ik}N_{jk}=2k_{\text{B}}T\Gamma^{\text{s}}_{ij}), Eq. (89) can be written as

dAi(l)dAj(l)dAk(l)=2kBT(ΓijsdAk(l)+ΓjksdAi(l)+ΓkisdAj(l))ϵ+𝒪(ϵ5/2).dA_{i}(l)dA_{j}(l)dA_{k}(l)=2k_{\text{B}}T(\Gamma^{\text{s}}_{ij}dA_{k}(l)+\Gamma^{\text{s}}_{jk}dA_{i}(l)+\Gamma^{\text{s}}_{ki}dA_{j}(l))\epsilon+\mathcal{O}(\epsilon^{5/2}). (93)

Similarly,

dAi(l)dAj(l)dAk(l)dAm(l)\displaystyle dA_{i}(l)dA_{j}(l)dA_{k}(l)dA_{m}(l) =\displaystyle= ξplξqlξrlξolNipNjqNkrNmoϵ2+𝒪(ϵ5/2)\displaystyle\xi_{p}^{l}\xi_{q}^{l}\xi_{r}^{l}\xi_{o}^{l}N_{ip}N_{jq}N_{kr}N_{mo}\epsilon^{2}+\mathcal{O}(\epsilon^{5/2}) (94)
=\displaystyle= (2kBT)2(ΓijsΓkms+ΓiksΓjms++ΓimsΓjks)ϵ2+𝒪(ϵ5/2)\displaystyle(2k_{\text{B}}T)^{2}(\Gamma^{\text{s}}_{ij}\Gamma^{\text{s}}_{km}+\Gamma^{\text{s}}_{ik}\Gamma^{\text{s}}_{jm}++\Gamma^{\text{s}}_{im}\Gamma^{\text{s}}_{jk})\epsilon^{2}+\mathcal{O}(\epsilon^{5/2})

Substituting (93) and  (94) into Eq. (88), we obtain

dAi(l)dAj(l)G(𝑨¯lf)\displaystyle dA_{i}(l)dA_{j}(l)G(\bar{\bm{A}}^{\text{f}}_{l}) =\displaystyle= dAi(l)dAj(l)G(𝑨¯l)+(2α1)kBT[GAk]𝑨¯l(Γijs(𝑨¯l)dAk(l)+Γjks(𝑨¯l)dAi(l)+Γkis(𝑨¯l)dAj(l))ϵ\displaystyle dA_{i}(l)dA_{j}(l)G(\bar{\bm{A}}_{l})+(2\alpha-1)k_{\text{B}}T\left[\dfrac{\partial G}{\partial A_{k}}\right]_{\bar{\bm{A}}_{l}}(\Gamma^{\text{s}}_{ij}(\bar{\bm{A}}_{l})dA_{k}(l)+\Gamma^{\text{s}}_{jk}(\bar{\bm{A}}_{l})dA_{i}(l)+\Gamma^{\text{s}}_{ki}(\bar{\bm{A}}_{l})dA_{j}(l))\epsilon (95)
+12[(2α1)kBT]2[2GAkAm(ΓijsΓkms+ΓiksΓjms++ΓimsΓjks)]𝑨¯lϵ2+𝒪(ϵ5/2)\displaystyle+\dfrac{1}{2}\left[(2\alpha-1)k_{\text{B}}T\right]^{2}\left[\dfrac{\partial^{2}G}{\partial A_{k}\partial A_{m}}(\Gamma^{\text{s}}_{ij}\Gamma^{\text{s}}_{km}+\Gamma^{\text{s}}_{ik}\Gamma^{\text{s}}_{jm}++\Gamma^{\text{s}}_{im}\Gamma^{\text{s}}_{jk})\right]_{\bar{\bm{A}}_{l}}\epsilon^{2}+\mathcal{O}(\epsilon^{5/2})

Similarly, from Eq. (87), we readily obtain

dAi(l)dAj(l)G(𝑨¯lr)\displaystyle dA_{i}(l)dA_{j}(l)G(\bar{\bm{A}}^{\text{r}}_{l}) =\displaystyle= dAi(l)dAj(l)G(𝑨¯l)(2α1)kBT[GAk]𝑨¯l(ΓijsdAk(l)+ΓjksdAi(l)+ΓkisdAj(l))ϵ\displaystyle dA_{i}(l)dA_{j}(l)G(\bar{\bm{A}}_{l})-(2\alpha-1)k_{\text{B}}T\left[\dfrac{\partial G}{\partial A_{k}}\right]_{\bar{\bm{A}}_{l}}(\Gamma^{\text{s}}_{ij}dA_{k}(l)+\Gamma^{\text{s}}_{jk}dA_{i}(l)+\Gamma^{\text{s}}_{ki}dA_{j}(l))\epsilon (96)
+12[(2α1)kBT]2[2GAkAm]𝑨¯l(ΓijsΓkms+ΓiksΓjms++ΓimsΓjks)ϵ2+𝒪(ϵ5/2).\displaystyle+\dfrac{1}{2}\left[(2\alpha-1)k_{\text{B}}T\right]^{2}\left[\dfrac{\partial^{2}G}{\partial A_{k}\partial A_{m}}\right]_{\bar{\bm{A}}_{l}}(\Gamma^{\text{s}}_{ij}\Gamma^{\text{s}}_{km}+\Gamma^{\text{s}}_{ik}\Gamma^{\text{s}}_{jm}++\Gamma^{\text{s}}_{im}\Gamma^{\text{s}}_{jk})\epsilon^{2}+\mathcal{O}(\epsilon^{5/2}).

Likewise, we can easily derive the following relations:

dAi(l)G(𝑨¯lf)\displaystyle dA_{i}(l)G(\bar{\bm{A}}^{\text{f}}_{l}) =\displaystyle= dAi(l)G(𝑨¯l)+(2α1)kBTϵ[ΓijsGAj]𝑨¯l+𝒪(ϵ3/2)\displaystyle dA_{i}(l)G(\bar{\bm{A}}_{l})+(2\alpha-1)k_{\text{B}}T\epsilon\left[\Gamma^{s}_{ij}\dfrac{\partial G}{\partial A_{j}}\right]_{\bar{\bm{A}}_{l}}+\mathcal{O}(\epsilon^{3/2}) (97)
dAi(l)G(𝑨¯lr)\displaystyle dA_{i}(l)G(\bar{\bm{A}}^{\text{r}}_{l}) =\displaystyle= dAi(l)G(𝑨¯l)(2α1)kBTϵ[ΓijsGAj]𝑨¯l+𝒪(ϵ3/2)\displaystyle dA_{i}(l)G(\bar{\bm{A}}_{l})-(2\alpha-1)k_{\text{B}}T\epsilon\left[\Gamma^{s}_{ij}\dfrac{\partial G}{\partial A_{j}}\right]_{\bar{\bm{A}}_{l}}+\mathcal{O}(\epsilon^{3/2}) (98)
G(𝑨¯lf)\displaystyle G(\bar{\bm{A}}^{\text{f}}_{l}) =\displaystyle= G(𝑨¯l)+𝒪(ϵ1/2)\displaystyle G(\bar{\bm{A}}_{l})+\mathcal{O}(\epsilon^{1/2}) (99)
G(𝑨¯lr)\displaystyle G(\bar{\bm{A}}^{\text{r}}_{l}) =\displaystyle= G(𝑨¯l)+𝒪(ϵ1/2)\displaystyle G(\bar{\bm{A}}_{l})+\mathcal{O}(\epsilon^{1/2}) (100)
[GAi]𝑨¯ldAi(l)\displaystyle\left[\dfrac{\partial G}{\partial A_{i}}\right]_{\bar{\bm{A}}_{l}}dA_{i}(l) =\displaystyle= G(𝑨l)G(𝑨l1)+𝒪(ϵ3/2)\displaystyle G(\bm{A}_{l})-G(\bm{A}_{l-1})+\mathcal{O}(\epsilon^{3/2}) (101)

Appendix E The relation between entropy production rate and s˙\dot{s}

The free energy of the system at time tt would be

F(t)\displaystyle F(t) =\displaystyle= (𝑨)pt(𝑨)𝑑𝑨T[kBpt(𝑨)lnpt(𝑨)𝑑𝑨]\displaystyle\int\mathcal{H}(\bm{A})p_{t}(\bm{A})d\bm{A}-T\left[-k_{\text{B}}\int p_{t}(\bm{A})\ln p_{t}(\bm{A})d\bm{A}\right] (102)
=\displaystyle= [(𝑨)+kBTlnpt(𝑨)],\displaystyle\left\langle\left[\mathcal{H}(\bm{A})+k_{\text{B}}T\ln p_{t}(\bm{A})\right]\right\rangle,

where \left\langle\right\rangle stands for the ensemble average and pt(𝑨)p_{t}(\bm{A}) is the probability distribution of 𝑨\bm{A} at time tt. Let us define the free energy of a single trajectory of the system at time tt as

f(t)=(𝑨(t))+kBTlnpt(𝑨(t))f(t)=\mathcal{H}(\bm{A}(t))+k_{\text{B}}T\ln p_{t}(\bm{A}(t)) (103)

Then it is straight forward to show that

dF(t)dt=df(t)dt.\displaystyle\dfrac{d{F}(t)}{dt}=\left\langle\dfrac{df(t)}{dt}\right\rangle. (104)

E.1 For passive systems

From Eq (32), we readily get

df(t)dt=kBTs˙(t),\left\langle\dfrac{df(t)}{dt}\right\rangle=-k_{\text{B}}T\left\langle\dot{s}(t)\right\rangle, (105)

so from Eq. (104)

dF(t)dt=kBTs˙(t).\displaystyle\dfrac{d{F}(t)}{dt}=-k_{\text{B}}T\left\langle\dot{s}(t)\right\rangle. (106)

Assuming that the system is always in metastable thermal equilibrium with the reservoir, the rate of change total entropy of system and reservoir would be

dS(t)dt=1TdF(t)dt=kBs˙(t).\dfrac{dS(t)}{dt}=-\dfrac{1}{T}\dfrac{d{F}(t)}{dt}=k_{\text{B}}\left\langle\dot{{s}}(t)\right\rangle. (107)

E.2 For active systems

For active systems, from Eq. (46), one can trivially prove that

df(t)dt=kBTs˙(t)+w(t),\left\langle\dfrac{df(t)}{dt}\right\rangle=-k_{\text{B}}T\left\langle\dot{s}(t)\right\rangle+\left\langle w(t)\right\rangle, (108)

where w(t)\left\langle w(t)\right\rangle is average rate of the work perfomed by active forces. Hence, from Eq. (104),

dF(t)dt=kBTs˙(t)+w(t).\dfrac{d{F}(t)}{dt}=-k_{\text{B}}T\left\langle\dot{{s}}(t)\right\rangle+\left\langle w(t)\right\rangle. (109)

Therefore, the rate of total enetropy production of the system and the reservoir:

dS(t)dt\displaystyle\dfrac{dS(t)}{dt} =\displaystyle= 1TdF(t)dt+1Tw(t)\displaystyle-\dfrac{1}{T}\dfrac{d{F}(t)}{dt}+\dfrac{1}{T}\left\langle w(t)\right\rangle (110)
=\displaystyle= kBs˙(t).\displaystyle k_{\text{B}}\left\langle\dot{{s}}(t)\right\rangle.

Appendix F The dissipation function defined by Seifert et al. [5]

Seifert et al. [5] used the following form of the dissipation function:

𝒮τ=ln[PPr],\mathcal{S}_{\tau}=\ln\left[\dfrac{P}{P^{\prime}_{r}}\right], (111)

where PP is the probability density of a trajectory between t=0t=0 and t=τt=\tau which is given by Eq. (A), and PrP^{\prime}_{r} is the probability density of the time-reversed trajectory, considering that the time-reserved trajectory starts at t=τt=\tau, not at t=0t=0. Thus, the epxression of 𝒮τ\mathcal{S}_{\tau} is readily obtained by replacing p0(𝒔𝑨N)p_{0}(\bm{s}\circ\bm{A}_{N}) with pt(𝒔𝑨N)p_{t}(\bm{s}\circ\bm{A}_{N}) in the expression of τ\mathcal{R}_{\tau} (see Eq. (25)), that is

𝒮τ=lnp0(𝑨(0))pτ(𝒔𝑨(τ))1kBT[(𝑨(τ))(𝑨(0))]\mathcal{S}_{\tau}=\ln\dfrac{p_{0}(\bm{A}(0))}{p_{\tau}(\bm{s}\circ\bm{A}(\tau))}-\dfrac{1}{k_{\text{B}}T}\left[\mathcal{H}(\bm{A}(\tau))-\mathcal{H}(\bm{A}(0))\right] (112)

This dissipation function follows the fluctuation relation (24) in steady states only, not in general. However, as discussed by [5], it does always follow the integrated fluctuation relation (28) and thereofore 𝒮τ0\left\langle\mathcal{S}_{\tau}\right\rangle\geq 0.

Appendix G Calculation of s˙\dot{s} for the active systems

Recalling equations of motion for the active systems

dAidt=i+𝒳i+Nijξj(t),\dfrac{dA_{i}}{dt}=\mathcal{F}_{i}+\mathcal{X}_{i}+N_{ij}\xi_{j}(t), (113)

where

iΓijAj+kBTΓijAjαNljNijAl\mathcal{F}_{i}\equiv-\Gamma_{ij}\dfrac{\partial\mathcal{H}}{\partial A_{j}}+k_{\text{B}}T\dfrac{\partial\Gamma_{ij}}{\partial A_{j}}-\alpha N_{lj}\dfrac{\partial N_{ij}}{\partial A_{l}} (114)

and 𝒳i\mathcal{X}_{i} is the active term. Writing 𝓧\bm{\mathcal{X}} as 𝓧=𝓧s+𝓧a\bm{\mathcal{X}}=\bm{\mathcal{X}}^{\text{s}}+\bm{\mathcal{X}}^{\text{a}}, where

𝓧s(𝑨)\displaystyle\bm{\mathcal{X}}^{\text{s}}(\bm{A}) =\displaystyle= 12(𝓧(𝑨)+𝒔𝓧(𝒔𝑨)),\displaystyle\dfrac{1}{2}\left(\bm{\mathcal{X}}(\bm{A})+\bm{s}\circ\bm{\mathcal{X}}(\bm{s}\circ\bm{A})\right), (115)
𝓧a(𝑨)\displaystyle\bm{\mathcal{X}}^{\text{a}}(\bm{A}) =\displaystyle= 12(𝓧(𝑨)𝒔𝓧(𝒔𝑨))\displaystyle\dfrac{1}{2}\left(\bm{\mathcal{X}}(\bm{A})-\bm{s}\circ\bm{\mathcal{X}}(\bm{s}\circ\bm{A})\right) (116)

follow the properties 𝓧s(𝒔𝑨)=𝒔𝓧s(𝑨)\bm{\mathcal{X}}^{\text{s}}(\bm{s}\circ\bm{A})=\bm{s}\circ\bm{\mathcal{X}}^{\text{s}}(\bm{A}) and 𝓧a(𝒔𝑨)=𝒔𝓧a(𝑨)\bm{\mathcal{X}}^{\text{a}}(\bm{s}\circ\bm{A})=-\bm{s}\circ\bm{\mathcal{X}}^{\text{a}}(\bm{A}). The Fokker-Planck equation for the probability density pt(𝑨)p_{t}(\bm{A}) of 𝑨\bm{A} reads

pt(𝑨)t=Ji(𝑨,t)Ai,\dfrac{\partial p_{t}(\bm{A})}{\partial t}=-\dfrac{\partial J_{i}(\bm{A},t)}{\partial A_{i}}, (117)

where

Ji(𝑨,t)\displaystyle J_{i}(\bm{A},t) =\displaystyle= (Γsij(𝑨)(𝑨)Aj+𝒴ia(𝑨)+𝒳js(𝑨))pt(𝑨)\displaystyle\left(-{\Gamma^{\text{s}}}_{ij}(\bm{A})\dfrac{\partial\mathcal{H}(\bm{A})}{\partial A_{j}}+\mathcal{Y}^{\text{a}}_{i}(\bm{A})+\mathcal{X}^{\text{s}}_{j}(\bm{A})\right)p_{t}(\bm{A}) (118)
kBTΓijs(𝑨)pt(𝑨)dAj\displaystyle\qquad\qquad\qquad\qquad-k_{\text{B}}T\Gamma^{\text{s}}_{ij}(\bm{A})\dfrac{\partial p_{t}(\bm{A})}{dA_{j}}

is the probability current [23] and

𝒴ia=𝒳ia+AkΓkiakBTΓkiaAk.\mathcal{Y}^{\text{a}}_{i}=\mathcal{X}^{\text{a}}_{i}+\dfrac{\partial\mathcal{H}}{\partial A_{k}}\Gamma^{\text{a}}_{ki}-k_{\text{B}}T\dfrac{\partial\Gamma^{\text{a}}_{ki}}{\partial A_{k}}. (119)

Since 𝓧a(𝒔𝑨)=𝒔𝓧a(𝑨)\bm{\mathcal{X}}^{\text{a}}(\bm{s}\circ\bm{A})=-\bm{s}\circ\bm{\mathcal{X}}^{\text{a}}(\bm{A}), using the relation Γija=Γijasisj\Gamma^{\text{a}}_{ij}=-\Gamma^{\text{a}}_{ij}s_{i}s_{j} (Eq. (16)), it is easy to show that

𝓨a(𝒔𝑨)=𝒔𝓨a(𝑨).\bm{\mathcal{Y}}^{\text{a}}(\bm{s}\circ\bm{A})=-\bm{s}\circ\bm{\mathcal{Y}}^{\text{a}}(\bm{A}). (120)

As Eq. (117) has no term with NijN_{ij}, pt(𝑨)p_{t}(\bm{A}) would be independent of the choice of NijN_{ij}. Recalling Eq. (III)

s˙(t)=w(t)ddt(lnpt(𝑨(t))+1kBT(𝑨(t)))|t=t.\dot{s}(t)=w(t)-\left.\dfrac{d}{dt^{\prime}}\left(\ln p_{t}(\bm{A}(t^{\prime}))+\dfrac{1}{k_{\text{B}}T}\mathcal{H}(\bm{A}(t^{\prime}))\right)\right|_{t^{\prime}=t}. (121)

where

w(t)\displaystyle w(t) =(𝑨(t))Ai𝒳ia(𝑨(t))kBT𝒳ia(𝑨(t))Ai\displaystyle=\dfrac{\partial\mathcal{H}(\bm{A}(t))}{\partial A_{i}}\mathcal{X}^{\text{a}}_{i}(\bm{A}(t))-k_{\text{B}}T\dfrac{\partial\mathcal{X}^{\text{a}}_{i}(\bm{A}(t))}{\partial A_{i}} (122)
+\displaystyle+ (Γs1)ij(𝑨(t))𝒳js(𝑨(t))[dAidt𝒴ia(𝑨(t))].\displaystyle{{\color[rgb]{0,0,0}({\Gamma^{\text{s}}}^{-1})_{ij}}}(\bm{A}(t))\mathcal{X}^{\text{s}}_{j}(\bm{A}(t))\left[\dfrac{dA_{i}}{dt}-\mathcal{Y}^{\text{a}}_{i}(\bm{A}(t))\right].

Using Eqs.  (117) & (118), one can write Eq. (III) in the following form:

s˙(t)\displaystyle\dot{{s}}(t) =\displaystyle= 1kBT(Ji(𝑨,t)pt(𝑨)𝒴ia(𝑨))(Γs1)ij(𝑨)\displaystyle\dfrac{1}{k_{\text{B}}T}\left(\dfrac{J_{i}(\bm{A},t)}{p_{t}(\bm{A})}-\mathcal{Y}^{\text{a}}_{i}(\bm{A})\right){{\color[rgb]{0,0,0}({\Gamma^{\text{s}}}^{-1})_{ij}}}(\bm{A}) (123)
×(Jj(𝑨,t)pt(𝑨)𝒴ja(𝑨))+Rs+R0,\displaystyle\times\left(\dfrac{J_{j}(\bm{A},t)}{p_{t}(\bm{A})}-\mathcal{Y}^{\text{a}}_{j}(\bm{A})\right)+R_{s}+R_{0},

where

Rs=1pt(𝑨)Ai(pt(𝑨)𝒴ia(𝑨))R_{s}=-\dfrac{1}{p_{t}(\bm{A})}\dfrac{\partial}{\partial A_{i}}\left(p_{t}(\bm{A})\mathcal{Y}^{\text{a}}_{i}(\bm{A})\right) (124)

and

R0\displaystyle R_{0} =(dAidtJi(𝑨,t)pt(𝑨))[1pt(𝑨)pt(𝑨)Ai\displaystyle=\left(\dfrac{dA_{i}}{dt}-\dfrac{J_{i}(\bm{A},t)}{p_{t}(\bm{A})}\right)\left[\dfrac{1}{p_{t}(\bm{A})}\dfrac{\partial p_{t}(\bm{A})}{\partial A_{i}}\right. (125)
1kBT((Γs1)ij(𝑨)𝒳js(𝑨)(𝑨)Aj)].\displaystyle\left.-\dfrac{1}{k_{\text{B}}T}\left({{\color[rgb]{0,0,0}({\Gamma^{\text{s}}}^{-1})_{ij}}}(\bm{A})\mathcal{X}^{\text{s}}_{j}(\bm{A})-\dfrac{\partial\mathcal{H}(\bm{A})}{\partial A_{j}}\right)\right].

The average of RsR_{s} reads

Rs\displaystyle\left\langle R_{s}\right\rangle =\displaystyle= 1pt(𝑨)Ai(pt(𝑨)𝒴ia(𝑨))pt(𝑨)dnA\displaystyle-\int\dfrac{1}{p_{t}(\bm{A})}\dfrac{\partial}{\partial A_{i}}\left(p_{t}(\bm{A})\mathcal{Y}^{\text{a}}_{i}(\bm{A})\right)p_{t}(\bm{A})d^{n}A (126)
=\displaystyle= Ai(pt(𝑨)𝒴ia(𝑨))dnA\displaystyle-\int\dfrac{\partial}{\partial A_{i}}\left(p_{t}(\bm{A})\mathcal{Y}^{\text{a}}_{i}(\bm{A})\right)d^{n}A

The above expression can be written as a surface integral with the integrand 𝑰s=pt(𝑨)𝓨a(𝑨)\bm{I}_{\text{s}}=-p_{t}(\bm{A})\bm{\mathcal{Y}}^{\text{a}}(\bm{A}). If the system is periodic in AiA_{i} (e.g., AiA_{i} is an angle), the surface integral is already zero. If AiA_{i} lies in the infinite interval (,)(-\infty,\infty), in Ai±A_{i}\to\pm\infty limit, pt(𝑨)0p_{t}(\bm{A})\to 0, given that 𝑨\bm{A} are physical variables. Assuming that 𝑰s||\bm{I}_{\text{s}}|| converges faster than 𝑨1n||\bm{A}||^{1-n}, the surface integral is again zero. For pt(𝒔𝑨)=pt(𝑨)p_{t}(\bm{s}\circ\bm{A})=p_{t}(\bm{A}) case, Rs\left\langle R_{s}\right\rangle is always zero as follows: setting 𝑨=𝒔𝑨\bm{A}=\bm{s}\circ\bm{A}^{\prime} gives dnA=dnAd^{n}A=d^{n}A^{\prime}, then using Eq. (120), we get

Rs\displaystyle\left\langle R_{s}\right\rangle =\displaystyle= Ai(pt(𝑨)𝒴ia(𝑨))dnA\displaystyle\int\dfrac{\partial}{\partial A^{\prime}_{i}}\left(p_{t}(\bm{A}^{\prime})\mathcal{Y}^{\text{a}}_{i}(\bm{A}^{\prime})\right)d^{n}A^{\prime} (127)
=\displaystyle= Rs,\displaystyle-\left\langle R_{s}\right\rangle,

so Rs=0\left\langle R_{s}\right\rangle=0. Since the ensemble average of dAi/dtdA_{i}/dt for given 𝑨\bm{A} and tt is just Ji(𝑨,t)/pt(𝑨)J_{i}(\bm{A},t)/p_{t}(\bm{A}), Ro=0\left\langle R_{o}\right\rangle=0. Therefore, the ensemble average of Eq. (123) is given by

s˙(t)\displaystyle\left\langle\dot{{s}}(t)\right\rangle =\displaystyle= 1kBT(Ji(𝑨,t)pt(𝑨)𝒴ia(𝑨))(Γs1)ij(𝑨)\displaystyle\dfrac{1}{k_{\text{B}}T}\left\langle\left(\dfrac{J_{i}(\bm{A},t)}{p_{t}(\bm{A})}-\mathcal{Y}^{\text{a}}_{i}(\bm{A})\right){{\color[rgb]{0,0,0}({\Gamma^{\text{s}}}^{-1})_{ij}}}(\bm{A})\right. (128)
×(Jj(𝑨,t)pt(𝑨)𝒴ja(𝑨)).\displaystyle\times\left.\left(\dfrac{J_{j}(\bm{A},t)}{p_{t}(\bm{A})}-\mathcal{Y}^{\text{a}}_{j}(\bm{A})\right)\right\rangle.

Similarly, w(t)\left\langle w(t)\right\rangle given by Eq. (122) can be written in the following form

w(t)\displaystyle\left\langle w(t)\right\rangle =\displaystyle= (dAidtAkΓkia+kBTΓkiaAk)(Γs1)ij𝒳js\displaystyle\left\langle\left(\dfrac{dA_{i}}{dt}-\dfrac{\partial\mathcal{H}}{\partial A_{k}}\Gamma^{\text{a}}_{ki}+k_{\text{B}}T\dfrac{\partial\Gamma^{\text{a}}_{ki}}{\partial A_{k}}\right){{\color[rgb]{0,0,0}({\Gamma^{\text{s}}}^{-1})_{ij}}}\mathcal{X}^{\text{s}}_{j}\right\rangle (129)
(dAidt𝒴ia)(Γs1)ij𝒳ja.\displaystyle-\left\langle\left(\dfrac{dA_{i}}{dt}-\mathcal{Y}^{\text{a}}_{i}\right){{\color[rgb]{0,0,0}({\Gamma^{\text{s}}}^{-1})_{ij}}}\mathcal{X}^{\text{a}}_{j}\right\rangle.

Here, if pt(𝒔𝑨)pt(𝑨)p_{t}(\bm{s}\circ\bm{A})\neq p_{t}(\bm{A}), we must assume that pt(𝑨)𝓧a(𝑨)||p_{t}(\bm{A})\bm{\mathcal{X}}^{\text{a}}(\bm{A})|| converges faster than 𝑨1n||\bm{A}||^{1-n}.

Appendix H Stationary solution of the Fokker-Planck equation associated with Eq. (8)

The Fokker-Planck equation for the probability distribution pt(𝑨)p_{t}(\bm{A}) of the solution of Eq (8) is given by

pt(𝑨)t=Ji(𝑨,t)Ai;\dfrac{\partial p_{t}(\bm{A})}{\partial t}=-\dfrac{\partial J_{i}(\bm{A},t)}{\partial A_{i}}; (130)

the expression of the probability current Ji(𝑨,t)J_{i}(\bm{A},t) reads [23]

Ji(𝑨,t)\displaystyle J_{i}(\bm{A},t) =\displaystyle= (Γsij(𝑨)(𝑨)Aj+AkΓkiakBTΓkiaAk)\displaystyle\left(-{\Gamma^{\text{s}}}_{ij}(\bm{A})\dfrac{\partial\mathcal{H}(\bm{A})}{\partial A_{j}}+\dfrac{\partial\mathcal{H}}{\partial A_{k}}\Gamma^{\text{a}}_{ki}-k_{\text{B}}T\dfrac{\partial\Gamma^{\text{a}}_{ki}}{\partial A_{k}}\right) (131)
×pt(𝑨)kBTΓijs(𝑨)pt(𝑨)dAj.\displaystyle\times p_{t}(\bm{A})-k_{\text{B}}T\Gamma^{\text{s}}_{ij}(\bm{A})\dfrac{\partial p_{t}(\bm{A})}{dA_{j}}.

Undoubtedly, the dynamics of pt(𝑨)p_{t}(\bm{A}) is independent of the choice of NijN_{ij} and α\alpha. Since

AkΓkiakBTΓkiaAk\dfrac{\partial\mathcal{H}}{\partial A_{k}}\Gamma^{\text{a}}_{ki}-k_{\text{B}}T\dfrac{\partial\Gamma^{\text{a}}_{ki}}{\partial A_{k}} (132)

is the Poisson bracket term, the stationary solution of the above equation is given by [25]

ps(𝑨)=1𝒵exp[(𝑨)kBT],p_{\text{s}}(\bm{A})=\dfrac{1}{\mathcal{Z}}\exp\left[-\dfrac{\mathcal{H}(\bm{A})}{k_{\text{B}}T}\right], (133)

where ZZ is the normalizing constant.

Appendix I Dependence of τ\left\langle\mathcal{R}_{\tau}\right\rangle on α\alpha for the passive systems

For the passive systems, τ\mathcal{R}_{\tau} depends only on the initial and final states of the system 𝑨0\bm{A}_{0} and 𝑨τ\bm{A}_{\tau}, so its average can be calculated using the formula [35]

τ=τp0(𝑨0)G(𝑨0,𝑨τ;τ)dnA0dnAτ,\left\langle\mathcal{R}_{\tau}\right\rangle=\int\mathcal{R}_{\tau}p_{0}(\bm{A}_{0})G(\bm{A}_{0},\bm{A}_{\tau};\tau)d^{n}A_{0}d^{n}A_{\tau}, (134)

where p0(𝑨)p_{0}(\bm{A}) is the probability distribution of 𝑨\bm{A} at t=0t=0, and G(𝑨,𝑨;τ)G(\bm{A},\bm{A}^{\prime};\tau) is the probability distribution of state 𝑨\bm{A}^{\prime} at t=τt=\tau given that the system was in the state 𝑨\bm{A} at t=0t=0; it is the solution of Eq. (130) with the initial condition G(𝑨,𝑨;τ=0)=𝜹(𝑨𝑨)G(\bm{A},\bm{A}^{\prime};\tau=0)=\bm{\delta}(\bm{A}-\bm{A}^{\prime}). As the solution of Eq. (130) is independent of α\alpha, τ\left\langle\mathcal{R}_{\tau}\right\rangle would be constant in α\alpha.

References

  • Evans et al. [1993] D. J. Evans, E. G. D. Cohen, and G. P. Morriss, Probability of second law violations in shearing steady states, Phys. Rev. Lett. 71, 2401 (1993).
  • Evans and Searles [2002] D. J. Evans and D. J. Searles, The fluctuation theorem, Advances in Physics 51, 1529 (2002).
  • Searles and Evans [2004] D. J. Searles and D. J. Evans, Fluctuations relations for nonequilibrium systems, Australian Journal of Chemistry 57, 1119 (2004).
  • Liphardt et al. [2002] J. Liphardt, S. Dumont, S. B. Smith, I. Tinoco, and C. Bustamante, Equilibrium information from nonequilibrium measurements in an experimental test of jarzynski’s equality,  296, 1832 (2002).
  • Seifert [2005] U. Seifert, Entropy production along a stochastic trajectory and an integral fluctuation theorem, Phys. Rev. Lett. 95, 040602 (2005).
  • Blickle et al. [2006] V. Blickle, T. Speck, L. Helden, U. Seifert, and C. Bechinger, Thermodynamics of a colloidal particle in a time-dependent nonharmonic potential, Phys. Rev. Lett. 96, 070603 (2006).
  • Seifert [2012] U. Seifert, Stochastic thermodynamics, fluctuation theorems and molecular machines, Reports on progress in physics 75, 126001 (2012).
  • Gallavotti [2008] G. Gallavotti, Heat and fluctuations from order to chaos, The European Physical Journal B 61, 1 (2008).
  • Jarzynski [2007] C. Jarzynski, Comparison of far-from-equilibrium work relations, Comptes Rendus Physique 8, 495 (2007).
  • Hurtado et al. [2011] P. I. Hurtado, C. Pèrez-Espigares, J. J. del Pozo, and P. L. Garrido, Symmetries in fluctuations far from equilibrium, Proceedings of the National Academy of Sciences  (2011).
  • Villavicencio-Sanchez et al. [2014] R. Villavicencio-Sanchez, R. J. Harris, and H. Touchette, Fluctuation relations for anisotropic systems, EPL 105, 30009 (2014).
  • Soni [2019] H. Soni, Flocks, Flow and Fluctuations in Inanimate Matter: Simulations and Theory, Ph.D. thesis (2019).
  • Wang et al. [2002] G. M. Wang, E. M. Sevick, E. Mittag, D. J. Searles, and D. J. Evans, Experimental demonstration of violations of the second law of thermodynamics for small systems and short time scales, Phys. Rev. Lett. 89, 050601 (2002).
  • Collin et al. [2005] D. Collin, F. Ritort, C. Jarzynski, S. B. Smith, I. Tinoco, and C. Bustamante, Verification of the crooks fluctuation theorem and recovery of rna folding free energies, Nature 437, 231 (2005).
  • Carberry et al. [2004] D. M. Carberry, J. C. Reid, G. M. Wang, E. M. Sevick, D. J. Searles, and D. J. Evans, Fluctuations and irreversibility: An experimental demonstration of a second-law-like theorem using a colloidal particle held in an optical trap, Phys. Rev. Lett. 92, 140601 (2004).
  • Evans and Searles [1994] D. J. Evans and D. J. Searles, Equilibrium microstates which generate second law violating steady states, Phys. Rev. E 50, 1645 (1994).
  • Khan and Sood [2010] M. Khan and A. Sood, Out-of-equilibrium microrheology using optical tweezers to probe directional viscoelastic properties under shear, EPL (Europhysics Letters) 92, 48001 (2010).
  • Kumar et al. [2011] N. Kumar, S. Ramaswamy, and A. K. Sood, Symmetry properties of the large-deviation function of the velocity of a self-propelled polar particle, Phys. Rev. Lett. 106, 118001 (2011).
  • Kumar et al. [2015] N. Kumar, H. Soni, S. Ramaswamy, and A. K. Sood, Anisotropic isometric fluctuation relations in experiment and theory on a self-propelled rod, Physical Review E 91, 030102 (2015).
  • Aron et al. [2016] C. Aron, D. G. Barci, L. F. Cugliandolo, Z. G. Arenas, and G. S. Lozano, Dynamical symmetries of markov processes with multiplicative white noise, Journal of Statistical Mechanics: Theory and Experiment 2016, 053207 (2016).
  • Dengler [2015] R. Dengler, Another derivation of generalized langevin equations, arXiv preprint arXiv:1506.02650  (2015).
  • Mazenko [2008] G. F. Mazenko, Nonequilibrium statistical mechanics (John Wiley & Sons, 2008).
  • Lau and Lubensky [2007] A. W. Lau and T. C. Lubensky, State-dependent diffusion: Thermodynamic consistency and its path integral formulation, Physical Review E 76, 011123 (2007).
  • Cugliandolo and Lecomte [2017] L. F. Cugliandolo and V. Lecomte, Rules of calculus in the path integral representation of white noise langevin equations: the onsager–machlup approach, Journal of Physics A: Mathematical and Theoretical 50, 345001 (2017).
  • Chaikin and Lubensky [1995] P. M. Chaikin and T. C. Lubensky, Principles of Condensed Matter Physics (Cambridge University Press, 1995) cambridge Books Online.
  • Onsager [1931] L. Onsager, Reciprocal relations in irreversible processes. i., Phys. Rev. 37, 405 (1931).
  • Gard [1988] T. C. Gard, Introduction to stochastic differential equations (New York : M. Dekker, 1988).
  • Sevick et al. [2008] E. Sevick, R. Prabhakar, S. R. Williams, and D. J. Searles, Fluctuation theorems, Annual Review of Physical Chemistry 59, 603 (2008), pMID: 18393680.
  • Reid et al. [2004] J. Reid, D. Carberry, G. Wang, E. M. Sevick, D. J. Evans, and D. J. Searles, Reversibility in nonequilibrium trajectories of an optically trapped particle, Physical Review E 70, 016111 (2004).
  • Marchetti et al. [2013] M. C. Marchetti, J. F. Joanny, S. Ramaswamy, T. B. Liverpool, J. Prost, M. Rao, and R. A. Simha, Hydrodynamics of soft active matter, Rev. Mod. Phys. 85, 1143 (2013).
  • Bricard et al. [2013] A. Bricard, J.-B. Caussin, N. Desreumaux, O. Dauchot, and D. Bartolo, Emergence of macroscopic directed motion in populations of motile colloids, Nature 503, 95 (2013).
  • Kumar et al. [2014] N. Kumar, H. Soni, S. Ramaswamy, and A. K. Sood, Flocking at a distance in active granular matter, Nat Commun 5 (2014).
  • Weber et al. [2013] C. A. Weber, T. Hanke, J. Deseigne, S. Léonard, O. Dauchot, E. Frey, and H. Chaté, Long-range ordering of vibrated polar disks, Phys. Rev. Lett. 110, 208001 (2013).
  • Chatterjee et al. [2021] R. Chatterjee, N. Rana, R. A. Simha, P. Perlekar, and S. Ramaswamy, Inertia drives a flocking phase transition in viscous active fluids, Physical Review X 11, 031063 (2021).
  • Doi et al. [1988] M. Doi, S. F. Edwards, and S. F. Edwards, The theory of polymer dynamics, Vol. 73 (oxford university press, 1988).