Skip to main content
Log in

Thermodynamics of Currents in Nonequilibrium Diffusive Systems: Theory and Simulation

  • Published:
Journal of Statistical Physics Aims and scope Submit manuscript

Abstract

Understanding the physics of nonequilibrium systems remains as one of the major challenges of modern theoretical physics. We believe nowadays that this problem can be cracked in part by investigating the macroscopic fluctuations of the currents characterizing nonequilibrium behavior, their statistics, associated structures and microscopic origin. This fundamental line of research has been severely hampered by the overwhelming complexity of this problem. However, during the last years two new powerful and general methods have appeared to investigate fluctuating behavior that are changing radically our understanding of nonequilibrium physics: a powerful macroscopic fluctuation theory (MFT) and a set of advanced computational techniques to measure rare events. In this work we study the statistics of current fluctuations in nonequilibrium diffusive systems, using macroscopic fluctuation theory as theoretical framework, and advanced Monte Carlo simulations of several stochastic lattice gases as a laboratory to test the emerging picture. Our quest will bring us from (1) the confirmation of an additivity conjecture in one and two dimensions, which considerably simplifies the MFT complex variational problem to compute the thermodynamics of currents, to (2) the discovery of novel isometric fluctuation relations, which opens an unexplored route toward a deeper understanding of nonequilibrium physics by bringing symmetry principles to the realm of fluctuations, and to (3) the observation of coherent structures in fluctuations, which appear via dynamic phase transitions involving a spontaneous symmetry breaking event at the fluctuating level. The clear-cut observation, measurement and characterization of these unexpected phenomena, well described by MFT, strongly support this theoretical scheme as the natural theory to understand the thermodynamics of currents in nonequilibrium diffusive media, opening new avenues of research in nonequilibrium physics.

This is a preview of subscription content, log in via an institution to check access.

Access this article

Price excludes VAT (USA)
Tax calculation will be finalised during checkout.

Instant access to the full article PDF.

Fig. 1
Fig. 2
Fig. 3
Fig. 4
Fig. 5
Fig. 6
Fig. 7
Fig. 8
Fig. 9

Similar content being viewed by others

Notes

  1. For this reason, LDFs are known as rate functions in the mathematical literature [16, 17].

  2. Note that the path integral formalism here described is based on a discretized Langevin equation of Ito-type [25].

  3. We explicitly exclude in this description the asymmetric and totally-asymmetric exclusion processes (ASEP and TASEP, respectively), as these models are not diffusive. See [41, 42] for a review on these interesting models of nonequilibrium behavior.

  4. Note that these rates converge for large N to the standard ones found in literature, namely \(\frac{1}{2}(1\pm\frac{E}{N})\), but avoid problems with negative rates for small N. In any case, the hydrodynamic descriptions of both variants of the model are identical.

  5. Note that the macroscopic current J is related to this microscopic current q through J=q N. Hence the microscopic large deviation function \(\mathcal{F}(\mathbf{q})\) scales with the system size as \(\mathcal{F}(\mathbf{q})=N^{d-2} G(\mathbf{J}=\mathbf{q} N)\), where G(J) is now the current LDF appearing in the diffusively-scaled macroscopic fluctuation theory of Sect. 2, see Eq. (12). This can be proved by noting that in the microscopic case we have \(P(\mathbf{q})=\mathrm{P}(\mathbf{Q}_{t},t)\sim\exp[+t \mathcal{F}(\mathbf{q})]\) while in macroscopic limit we have P(J)∼exp[+τN d G(J)]. Thus, by writing the latter probability in terms of the diffusive-scaled time variable τ=t/N 2, we get that P(J=q N)∼exp[+tN d−2 G(J=q N)] which compared to the microscopic probability gives us the scaling mentioned above. In a similar manner, if \(\theta(\boldsymbol{\lambda})=\max_{\mathbf{q}}[{\mathcal{F}}(\mathbf{q}) + \boldsymbol{\lambda}\cdot\mathbf{q}]\) and μ(λ)=max J [G(J)+λJ] are the Legendre transforms of \(\mathcal{F}(\mathbf{q})\) and G(J), respectively, they are related via the simple scaling relation μ(λ)=N 2−d θ(λ N d−1).

  6. This simulation scheme is well-suited for discrete-time Markov chains. A slightly different though equivalent version of the algorithm exists for continuous-time stochastic lattice gases [19].

  7. Note that this is the original formulation of the additivity principle for the integrated current stated by Bodineau and Derrida in [9]. In [39], Bertini et al. state an additivity principle for the density field which involves either a maximization or a minimization depending on the convexity of the rate functional for the considered model.

  8. Note however that the additivity hypothesis has been recently confirmed in Hamiltonian models with anomalous, non-diffusive transport properties [49]. These results considerably broaden the range of applicability of the additivity conjecture.

  9. See inset of Fig. 11 of Ref. [24] where J versus λ is displayed.

  10. For equilibrium dynamics, T L =T R , this symmetry is obvious as the system has x↔1−x symmetry.

  11. In fact, it suffices to determine the current distribution along an arbitrary direction, say e.g. \(\mathbf{J}=|\mathbf{J}| \hat{\mathbf{x}}\), to reconstruct the whole distribution P τ (J) using the IFR.

  12. The optimal traveling wave profile solution of Eq. (98) must be symmetric because this differential equation remains invariant under the transformation x→1−x. Periodicity implies that profile extrema, if any, come in pairs (maximum and minimum). Moreover, the profile has at most a single pair of extrema because, when we make \(w'_{J}(x)=0\) in Eq. (98), the resulting equation is third order for the KMP model and WASEP. This, together with the symmetry of the profile, implies in turn that the extrema ω 1=ω J (x 1) and ω 0=ω J (x 0) are such that |x 0x 1|=1/2.

  13. This results from the microreversibility of the model at hand: while we need a large size limit in order to verify the predictions derived from MFT (which is a macroscopic theory), no finite-size corrections affect the fluctuation theorem, whose validity can be used to ascertain the range of applicability of the cloning algorithm used to sample large deviations [28].

References

  1. Bertini, L., De Sole, A., Gabrielli, D., Jona-Lasinio, G., Landim, C.: Phys. Rev. Lett. 87, 040601 (2001)

    Article  ADS  MathSciNet  Google Scholar 

  2. Bertini, L., De Sole, A., Gabrielli, D., Jona-Lasinio, G., Landim, C.: J. Stat. Phys. 107, 635 (2002)

    Article  ADS  MATH  Google Scholar 

  3. Bertini, L., De Sole, A., Gabrielli, D., Jona-Lasinio, G., Landim, C.: Phys. Rev. Lett. 94, 030601 (2005)

    Article  ADS  Google Scholar 

  4. Bertini, L., De Sole, A., Gabrielli, D., Jona-Lasinio, G., Landim, C.: J. Stat. Phys. 123, 237 (2006)

    Article  ADS  MATH  MathSciNet  Google Scholar 

  5. Bertini, L., De Sole, A., Gabrielli, D., Jona-Lasinio, G., Landim, C.: J. Stat. Mech. P07014 (2007)

  6. Bertini, L., De Sole, A., Gabrielli, D., Jona-Lasinio, G., Landim, C.: J. Stat. Phys. 135, 857 (2009)

    Article  ADS  MATH  MathSciNet  Google Scholar 

  7. Derrida, B.: J. Stat. Mech. P07023 (2007)

  8. Seifert, U.: Rep. Prog. Phys. 75, 126001 (2012)

    Article  ADS  Google Scholar 

  9. Bodineau, T., Derrida, B.: Phys. Rev. Lett. 92, 180601 (2004)

    Article  ADS  Google Scholar 

  10. Hurtado, P.I., Garrido, P.L.: Phys. Rev. Lett. 102, 250601 (2009)

    Article  ADS  Google Scholar 

  11. Gallavotti, G., Cohen, E.D.G.: Phys. Rev. Lett. 74, 2694 (1995)

    Article  ADS  Google Scholar 

  12. Evans, D.J., Cohen, E.G.D., Morriss, G.P.: Phys. Rev. Lett. 71, 2401 (1993)

    Article  ADS  MATH  Google Scholar 

  13. Evans, D.J., Searles, D.J.: Phys. Rev. E 50, 1645 (1994)

    Article  ADS  Google Scholar 

  14. Lebowitz, J.L., Spohn, H.: J. Stat. Phys. 95, 333 (1999)

    Article  ADS  MATH  MathSciNet  Google Scholar 

  15. Kurchan, J.: J. Phys. A 31, 3719 (1998)

    Article  ADS  MATH  MathSciNet  Google Scholar 

  16. Ellis, R.S.: Entropy, Large Deviations and Statistical Mechanics. Springer, New York (1985)

    Book  MATH  Google Scholar 

  17. Touchette, H.: Phys. Rep. 478, 1 (2009)

    Article  ADS  MathSciNet  Google Scholar 

  18. Giardinà, C., Kurchan, J., Peliti, L.: Phys. Rev. Lett. 96, 120603 (2006)

    Article  ADS  Google Scholar 

  19. Lecomte, V., Tailleur, J.: J. Stat. Mech. P03004 (2007)

  20. Lecomte, V., Tailleur, J.: AIP Conf. Proc. 1091, 212 (2009)

    ADS  Google Scholar 

  21. Giardinà, C., Kurchan, J., Lecomte, V., Tailleur, J.: J. Stat. Phys. 145, 787 (2011)

    Article  ADS  MATH  MathSciNet  Google Scholar 

  22. Spohn, H.: Large Scale Dynamics of Interacting Particles. Springer, Berlin (1991)

    Book  MATH  Google Scholar 

  23. Binney, J.J., Dowrick, N.J., Fisher, A.J., Newman, M.E.J.: The Theory of Critical Phenomena: An Introduction to the Renormalization Group. Oxford University Press, Oxford (1998)

    Google Scholar 

  24. Hurtado, P.I., Garrido, P.L.: Phys. Rev. E 81, 041102 (2010)

    Article  ADS  Google Scholar 

  25. Zinn-Justin, J.: Quantum Field Theory and Critical Phenomena vol. 142. Clarendon Press, Oxford (2002)

    Book  Google Scholar 

  26. Pérez-Espigares, C., del Pozo, J.J., Garrido, P.L., Hurtado, P.I.: AIP Conf. Proc. 1332, 204 (2011)

    Article  ADS  Google Scholar 

  27. Jona-Lasinio, G.: Prog. Theor. Phys. 124, 731 (2010)

    Article  ADS  MATH  Google Scholar 

  28. Hurtado, P.I., Garrido, P.L.: J. Stat. Mech. P02032 (2009)

  29. Hurtado, P.I., Pérez-Espigares, C., del Pozo, J.J., Garrido, P.L.: Proc. Natl. Acad. Sci. USA 108, 7704 (2011)

    Article  ADS  MATH  MathSciNet  Google Scholar 

  30. Prados, A., Lasanta, A., Hurtado, P.I.: Phys. Rev. Lett. 107, 140601 (2011)

    Article  ADS  Google Scholar 

  31. Prados, A., Lasanta, A., Hurtado, P.I.: Phys. Rev. E 86, 031134 (2012)

    Article  ADS  Google Scholar 

  32. Hurtado, P.I., Lasanta, A., Prados, A.: Phys. Rev. E 88, 022110 (2013)

    Article  ADS  Google Scholar 

  33. Hurtado, P.I., Krapivsky, P.L.: Phys. Rev. E 85, 060103(R) (2012)

    Article  ADS  Google Scholar 

  34. Bodineau, T., Lagouge, M.: J. Stat. Phys. 139, 201 (2010)

    Article  ADS  MATH  MathSciNet  Google Scholar 

  35. Bodineau, T., Lagouge, M.: Ann. Appl. Probab. 22, 2282 (2012)

    Article  MATH  MathSciNet  Google Scholar 

  36. Jona-Lasinio, G.: Prog. Theor. Phys. Suppl. 184, 262 (2010)

    Article  ADS  MATH  Google Scholar 

  37. Kipnis, C., Marchioro, C., Presutti, E.: J. Stat. Phys. 27, 65 (1982)

    Article  ADS  MathSciNet  Google Scholar 

  38. Spohn, H.: J. Phys. A 16, 4275 (1983)

    Article  ADS  MathSciNet  Google Scholar 

  39. Bertini, L., Gabrielli, D., Lebowitz, J.L.: J. Stat. Phys. 121, 843 (2005)

    Article  ADS  MATH  MathSciNet  Google Scholar 

  40. Pöschel, T., Luding, S. (eds.): Granular Gases. Lecture Notes in Physics, vol. 564. Springer, Berlin (2001)

    Google Scholar 

  41. Derrida, B.: Phys. Rep. 301, 65 (1998)

    Article  ADS  MathSciNet  Google Scholar 

  42. Chou, T., Mallick, K., Zia, R.K.P.: Rep. Prog. Phys. 74, 116601 (2011)

    Article  ADS  MathSciNet  Google Scholar 

  43. Derrida, B., Lebowitz, J.L.: Phys. Rev. Lett. 80, 209 (1998)

    Article  ADS  MATH  MathSciNet  Google Scholar 

  44. Rákos, A., Harris, R.J.: J. Stat. Mech. P05005(2008)

  45. Hurtado, P.I., Garrido, P.L.: Phys. Rev. Lett. 107, 180601 (2011)

    Article  ADS  Google Scholar 

  46. Bodineau, T., Derrida, B.: Phys. Rev. E 72, 066110 (2005)

    Article  ADS  MathSciNet  Google Scholar 

  47. Harris, R.J., Schütz, G.M.: J. Stat. Mech. P07020 (2007)

  48. Schütz, G.M.: In: Domb, C., Lebowitz, J.L. (eds.) Phase Transitions and Critical Phenomena, vol. 19. Academic, London (2001)

    Google Scholar 

  49. Saito, K., Dhar, A.: Phys. Rev. Lett. 107, 250601 (2011)

    Article  ADS  Google Scholar 

  50. Espigares, C.P.: Ph.D. Thesis (2012)

  51. Espigares, C.P., Garrido, P.L., Hurtado, P.I.: Phys. Rev. E 87, 032115 (2013)

    Article  ADS  Google Scholar 

  52. Espigares, C.P., Garrido, P.L., Hurtado, P.I.: (2013, submitted)

  53. Villavicencio-Sanchez, R., Touchette, H., Harris, R.J.: arXiv:1308.2595 (2013)

  54. Andrieux, D., Gaspard, P.: J. Stat. Mech. P02006 (2007)

  55. Gorissen, M., Vanderzande, C.: Phys. Rev. E 86, 051114 (2012)

    Article  ADS  Google Scholar 

  56. Tailleur, J., Kurchan, J.: Nat. Phys. 3, 203 (2007)

    Article  Google Scholar 

  57. Lam, K.-D.N.T., Kurchan, J., Levine, D.: J. Stat. Phys. 137, 1079 (2009)

    Article  ADS  MATH  MathSciNet  Google Scholar 

Download references

Acknowledgements

We thank T. Bodineau, B. Derrida, A. Lasanta, J.L. Lebowitz, V. Lecomte, A. Prados and J. Tailleur for illuminating discussions.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Pedro L. Garrido.

Additional information

Dedicated to Herbert Spohn on the occasion of his 65th birthday.

This work has been supported by Spanish MICINN project FIS2009-08451, University of Granada, and Junta de Andalucía projects P07-FQM02725 and P09-FQM4682.

Appendices

Appendix A: Additivity Predictions for Current Fluctuations in the 2D KMP Model

In this appendix we solve the simplified variational problem posed by MFT after the additivity conjecture of Sect. 5 for the particular case of the two-dimensional Kipnis-Marchioro-Presutti model defined in Sect. 3, when subject to a temperature gradient along the x-direction (i.e. in contact with thermal baths at the left and right boundaries at temperatures T L and T R , respectively) and with periodic boundary conditions along the y-direction. Due to the symmetry of the problem, the calculations here described are also useful for the 1D KMP model subject to the same temperature gradient, for which predictions follow by simply making zero all y-components.

The equation for the optimal density field under the additivity conjecture reads

$$ D[\rho_{\mathbf{J}}]^2 (\boldsymbol{\nabla}\rho_{\mathbf{J}})^2= \mathbf{J}^2 \bigl(1+2\sigma[\rho _{\mathbf{J}}]K\bigl(\mathbf{J}^2 \bigr) \bigr) $$
(111)

where K(J 2) is a constant of integration which guarantees the correct boundary conditions for ρ J (r), which in this particular case are

$$ \rho_{\mathbf{J}}(x=0,y)=T_L ; \quad\quad\rho_{\mathbf{J}}(x=1,y)=T_R ; \quad\quad\rho_{\mathbf{J}}(x,y=0)=\rho_{\mathbf{J}}(x,y=1) . $$
(112)

Moreover, for the KMP model recall that the diffusivity and mobility transport coefficients entering MFT are D[ρ]=1/2 and σ[ρ]=ρ 2, respectively. Interestingly, the optimal field solution of Eq. (111) depends exclusively on the modulus of the current vector J=|J|, that is ρ J (r)=ρ J (r). In addition, the symmetry of the problem suggests that ρ J (r) depends exclusively on x, with no structure in the y-direction, compatible with the presence of an external gradient along the x-direction, i.e. ρ J (r)≡ρ J (x). In the last expression we have simplified our notation in order not to clutter formulas below. Under these considerations, the differential equation (111) for the optimal profile in the 2D KMP model becomes

$$ \biggl( \frac{d\rho_{J}(x)}{dx} \biggr)^2=4J^2 \bigl(1+2K(J) \rho _{J}^2(x) \bigr) $$
(113)

Here two different scenarios appear. On one hand, for large enough K(J) the r.h.s. of Eq. (113) does not vanish ∀x∈[0,1] and the resulting profile is monotone. In this case, and assuming T L >T R henceforth without loss of generality,

$$ \frac{d \rho_{J}(x)}{dx} = - 2J \sqrt{1+2\rho_{J}^2(x) K(J)} . $$
(114)

On the other hand, for K(J)<0 the r.h.s. of Eq. (113) may vanish at some points, resulting in a ρ J (x) that is non-monotone and takes an unique value \(\rho_{J}^{*}\equiv\sqrt{-1/2K(J)}\) in the extrema. Notice that the r.h.s. of Eq. (113) may be written in this case as \(4J^{2}[1-(\rho_{J}(x)/\rho_{J}^{*})^{2}]\). It is then clear that, if non-monotone, the profile ρ J (x) can only have a single maximum because: (i) \(\rho_{J}(x)\le\rho_{J}^{*}\)x∈[0,1] for the profile to be a real function, and (ii) several maxima are not possible because they should be separated by a minimum, which is not allowed because of (i). Hence for the non-monotone case (recall T L >T R )

$$\begin{aligned} \frac{d \rho_{J}(x)}{dx} = \left \{ \begin{array}{l@{\quad}l} +2J \sqrt{1- (\frac{\rho_{J}(x)}{\rho _J^*} )^2} ,& { x<x^* } \\ - 2J \sqrt{1- (\frac{\rho_{J}(x)}{\rho _J^*} )^2} ,& { x>x^* } \end{array} \right . \end{aligned}$$
(115)

where x locates the profile maximum. This leaves us with two separated regimes for current fluctuations, with the crossover happening for \(J_{0}\equiv\frac{T_{L}}{2} [\frac{\pi}{2} - \sin ^{-1} (\frac{T_{R}}{T_{L}} ) ]\). This crossover current can be obtained from Eq. (121) below by letting \(\rho_{J}^{*}\rightarrow T_{L}\)

1.1 A.1 Region I: Monotonous Regime (J<J 0)

In this case, using Eq. (114) to change variables in Eq. (38) we have

$$ G(\mathbf{{J}})=\int_{T_L}^{T_R} d\rho_{J} \frac{1}{4J\rho_{J}^2\sqrt {1+2K(J)\rho_{J}^2}} \Bigl[ \Bigl(J_x-J\sqrt{1+2K(J) \rho_{J}^2} \Bigr)^2+J_y^2 \Bigr] , $$
(116)

with J α the α component of vector J. This results in

$$ G(\mathbf{{J}})=\frac{J_x}{2} \biggl(\frac{1}{T_R}-\frac{1}{T_L} \biggr)-J^2K(J)+\frac{J}{2} \biggl[\frac{\sqrt{1+2K(J)T_L^2}}{T_L}- \frac {\sqrt{1+2K(J)T_R^2}}{T_R} \biggr] $$
(117)

Notice that, for ρ J (x) to be monotone, \(1+2K(J)\rho_{J}^{2}>0\). Thus, \(K(J)>-(2T_{L}^{2})^{-1}\). Integrating now Eq. (114) we obtain the following implicit equation for ρ J (x) in this regime

$$\begin{aligned} 2xJ = \left \{ \begin{array}{l@{\quad}l} {\frac{1}{\sqrt{2K(J)}} \ln [ \frac{ T_L + \sqrt {T_L^2 + \frac{1}{2K(J)}}}{ \rho_{J}(x) + \sqrt{\rho_{J}(x)^2 + \frac{1}{2K(J)}}} ] }, & { K(J)>0 } \\ \frac{\sin^{-1} [T_L\sqrt{-2K(J)} ] - \sin ^{-1} [\rho_{J}(x)\sqrt{-2K(J)} ]}{\sqrt{-2K(J)}} , & {-\frac{1}{2T_L^2}<K(J)<0 } \end{array} \right . \end{aligned}$$
(118)

Making x=1 and ρ J (x=1)=T R in the previous equation, we obtain the implicit expression for the constant K(J). To get a feeling on how it depends on J, note that in the limit \(K(J)\rightarrow(-1/2T_{L}^{2})\), the current JJ 0, while for K(J)→∞ one gets J→0. In addition, from Eq. (118) we see that for K(J)→0 we find J=(T L T R )/2=〈J〉.

Sometimes it is interesting to work with the Legendre transform of the current LDF [17, 10, 24, 26], μ(λ)=max J [G(J)+λJ], where λ is a vector parameter conjugate to the current. Using the previous results for G(J) is easy to show [26] that μ(λ)=−K(λ)J (λ)2, where J (λ) is the current associated with a given λ, and the constant K(λ)=K(|J (λ)|). The expression for J (λ) can be obtained from Eq. (118) above in the limit x→1. Finally, the optimal profile for a given λ is just \(\rho_{\boldsymbol{\lambda}}(x)=\rho_{|\mathbf{J}^{*}(\boldsymbol{\lambda})|}(x)\).

1.2 A.2 Region II: Non-monotonous Regime (J>J 0)

In this case the optimal profile has a single maximum \(\rho_{J}^{*}\equiv \rho_{J}(x=x^{*})\) with \(\rho_{J}^{*}=1/\sqrt{-2K(J)}\) and \(-1/2T_{L}^{2}<K(J)<0\). Splitting the integral in Eq. (38) at x , and using now Eq. (115) to change variables, we arrive at

$$\begin{aligned} G(\mathbf{{J}}) = & \frac{J_x}{2} \biggl(\frac{1}{T_R}- \frac{1}{T_L} \biggr) - \frac{J}{2} \biggl[\frac{1}{T_L}\sqrt {1- \biggl(\frac{T_L}{\rho _J^*} \biggr)^2} + \frac{1}{T_R} \sqrt{1- \biggl(\frac{T_R}{\rho_J^*} \biggr)^2} \\ &{}- \frac{1}{2\rho_J^*} \biggl(\pi-\sin^{-1} \biggl( \frac{T_L}{\rho _J^*} \biggr)-\sin^{-1} \biggl(\frac{T_R}{\rho_J^*} \biggr) \biggr) \biggr] . \end{aligned}$$
(119)

Integrating Eq. (115) one gets an implicit equation for the non-monotone optimal profile

$$\begin{aligned} 2xJ = \left \{ \begin{array}{l@{\quad}l} {\rho_J^* [ \sin^{-1} ( \frac{\rho_{J}(x)}{\rho _J^*} ) - \sin^{-1} ( \frac{T_L}{\rho_J^*} ) ] } & \text{for } 0\le x<x^* \\ {2J + \rho_J^* [ \sin^{-1} ( \frac{T_R}{\rho _J^*} ) - \sin^{-1} ( \frac{\rho_{J}(x)}{\rho_J^*} ) ] } & \text{for } x^*<x\le1 \end{array} \right . \end{aligned}$$
(120)

At x=x both branches of the above equation must coincide, and this condition provides simple equations for both x and \(\rho_{J}^{*}\)

$$ J=\frac{\rho_J^*}{2} \biggl[\pi-\sin^{-1} \biggl(\frac{T_L}{\rho _J^*} \biggr)-\sin^{-1} \biggl(\frac{T_R}{\rho_J^*} \biggr) \biggr] ;\quad\quad x^*= \frac{\frac{\pi}{2}-\sin^{-1} (\frac{T_L}{\rho_J^*} ) }{\pi-\sin^{-1} (\frac{T_L}{\rho_J^*} )-\sin^{-1} (\frac{T_R}{\rho_J^*} )} . $$
(121)

Finally, as in the monotone regime, we can compute the Legendre transform of the current LDF, obtaining as before μ(λ)=−K(λ)J (λ)2, where J (λ) and the constant K(λ)=K(|J (λ)|) can be easily computed from the previous expressions [24]. Note that, in λ-space, monotone profiles are expected for \(|\boldsymbol{\lambda}+\boldsymbol{\epsilon}| \leq\frac{1}{2 T_{R}}\sqrt{1- (\frac{T_{R}}{T_{L}} )^{2}}\), where \(\boldsymbol{\epsilon}=(\frac{1}{2} (T_{L}^{-1}-T_{R}^{-1} ),0)\) is the driving force in this case, while non-monotone profiles appear for \(\frac{1}{2 T_{R}}\sqrt{1- (\frac{T_{R}}{T_{L}} )^{2}} \leq|\boldsymbol{\lambda}+\boldsymbol{\epsilon}| \leq\frac{1}{2} (\frac{1}{T_{L}}+\frac{1}{T_{R}} )\).

Figure 10 shows the predicted G(J) (left panel) and its Legendre transform (right panel) for the 2D-KMP model. Notice that the LDF is zero for J=〈J〉=((T L T R )/2,0) and negative elsewhere. For small current fluctuations, J≈〈J〉, G(J) obeys the following quadratic form

$$ G(\mathbf{J})\approx-\frac{1}{2} \biggl(\frac{(J_x-(T_L-T_R)/2)^2}{\sigma _x^2}+\frac{J_y^2}{\sigma_y^2} \biggr) , $$
(122)

with \(\sigma_{x}^{2}=(T_{L}^{2}+T_{L}T_{R}+T_{R}^{2})/3\) and \(\sigma_{y}^{2}=T_{L} T_{R}\), resulting in Gaussian statistics for currents near the average as expected from the central limit theorem. A similar expansion for the Legendre transform yields

$$ \mu(\boldsymbol{\lambda})\approx\frac{\lambda_x}{2} \bigl[(T_L-T_R)+ \sigma_x^2\lambda _x \bigr]+ \frac{\sigma_y^2}{2}\lambda_y^2. $$
(123)

Notice that beyond this restricted Gaussian regime, current statistics is in general non-Gaussian. In particular, for large enough current deviations, G(J) decays linearly, meaning that the probability of such fluctuations is exponentially small in J (rather than J 2).

Fig. 10
figure 10

Left panel: G(J) for the 2D-KMP model for T L =2 and T R =1. The blue circle signals the crossover from monotone (J<J 0π/3) to non-monotone (J>π/3) optimal profiles. The green surface corresponds to the Gaussian approximation for small current fluctuations. Right panel: Legendre transform of both G(J) (brown) and G(J→〈J〉) (green), together with the projection in λ-space of the crossover between monotonous and non-monotonous regime (Color figure online)

Despite the complex structure of G(J) in both regime I and II above, it can be easily checked that for any pair of isometric current vectors J and J′, such that |J|=|J′|, the current LDF obeys

$$ G(\mathbf{J})-G\bigl(\mathbf{J}'\bigr)=\boldsymbol{\epsilon}\cdot\bigl(\mathbf{J}- \mathbf{J}'\bigr) . $$
(124)

This relation, known as Isometric Fluctuation Relation, is not particular for the 2D KMP model but a general results for time-reversible diffusive systems, see Sect. 6.

Appendix B: Constants of Motion

A sufficient condition for the Isometric Fluctuation Relation (IFR), Eqs. (58) or (124), to hold is that

$$ \frac{\delta\boldsymbol{\pi}_1[\rho(\mathbf{r})]}{\delta\rho(\mathbf{r}')} = 0 , $$
(125)

with the functional π 1[ρ(r)] defined in Eq. (40) above. We have shown that condition (125) follows from the time-reversibility of the dynamics, in the sense that the evolution operator in the Fokker-Planck formulation of Eq. (1) obeys a local detailed balance condition, see Eq. (55). Condition (125) implies that π 1[ρ(r)] is in fact a constant of motion, ϵ, independent of the profile ρ(r). Therefore we can use an arbitrary profile ρ(r), compatible with boundary conditions, to compute ϵ. We now choose boundary conditions to be gradient-like in the \(\hat {x}\)-direction, with densities ρ L and ρ R at the left and right reservoirs, respectively, and periodic boundary conditions in all other directions. Given these boundaries, we now select a linear profile

$$ \rho(\mathbf{r})=\rho_L + (\rho_R-\rho_L) x , $$
(126)

to compute ϵ, with x∈[0,1], and assume very general forms for the current and mobility functionals

$$\begin{aligned} \mathbf{Q} \bigl[\rho(\mathbf{r}) \bigr] \equiv& D_{0,0}[\rho] \boldsymbol{\nabla}\rho+ \sum _{n,m>0} D_{nm}[\rho] \bigl( \boldsymbol{\nabla}^m \rho \bigr)^{2n} \boldsymbol{\nabla}\rho , \\ \sigma \bigl[\rho(\mathbf{r}) \bigr] \equiv& \sigma_{0,0}[\rho] + \sum _{n,m>0} \sigma _{nm}[\rho] \bigl( \boldsymbol{\nabla}^m \rho \bigr)^{2n} , \end{aligned}$$

where as a convention we denote as F[ρ] a generic functional of the profile but not of its derivatives. It is now easy to show that \(\boldsymbol{\epsilon}=\varepsilon\hat{x}+\mathbf{E}\), with

$$ \varepsilon= \int_{\rho_L}^{\rho_R} d \rho \frac{D_{0,0}(\rho)+\sum_{n>0} D_{n1}(\rho)(\rho_R-\rho_L)^{2n}}{ \sigma_{0,0}(\rho)+\sum_{m>0} \sigma_{m1}(\rho)(\rho_R-\rho_L)^{2m}} , $$
(127)

and \(\hat{x}\) the unit vector along the gradient direction. In a similar way, if the following condition holds

$$ \frac{\delta}{\delta\rho(\mathbf{r}')} \int\mathbf{Q}\bigl[\rho(\mathbf{r})\bigr] d\mathbf{r}= 0 , $$
(128)

together with time-reversibility, Eq. (125), the system can be shown to obey an extended Isometric Fluctuation Relation which links any current fluctuation J in the presence of an external field E with any other isometric current fluctuation J′ in the presence of an arbitrarily-rotated external field E , and reduces to the standard IFR for E=E , see Eq. (79) in the main text. Condition (128) implies that ν≡∫Q[ρ(r)]d r is another constant of motion, which can be now written as \(\boldsymbol{\nu}=\nu\hat{x}\), with

$$ \nu= \int_{\rho_L}^{\rho_R} d \rho \biggl[D_{0,0}( \rho)+\sum_{n>0} D_{n1}(\rho) ( \rho_R-\rho_L)^{2n} \biggr] , $$
(129)

As an example, for a diffusive system Q[ρ(r)]=−D[ρ] ρ(r), with D[ρ] the diffusivity functional, and the above equations yield the familiar results

$$\begin{aligned} \varepsilon = & \int_{\rho_R}^{\rho_L} \frac{D(\rho)}{\sigma(\rho )}d\rho , \\ \nu = & \int_{\rho_R}^{\rho_L} D(\rho)d\rho , \end{aligned}$$

for a standard local mobility σ[ρ].

Rights and permissions

Reprints and permissions

About this article

Cite this article

Hurtado, P.I., Espigares, C.P., del Pozo, J.J. et al. Thermodynamics of Currents in Nonequilibrium Diffusive Systems: Theory and Simulation. J Stat Phys 154, 214–264 (2014). https://doi.org/10.1007/s10955-013-0894-6

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s10955-013-0894-6

Keywords

Navigation