Skip to main content
Log in

Local Thermal Equilibrium for Certain Stochastic Models of Heat Transport

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

Abstract

This paper is about nonequilibrium steady states (NESS) of a class of stochastic models in which particles exchange energy with their “local environments” rather than directly with one another. The physical domain of the system can be a bounded region of \(\mathbb R^d\) for any \(d \ge 1\). We assume that the temperature at the boundary of the domain is prescribed and is nonconstant, so that the system is forced out of equilibrium. Our main result is local thermal equilibrium in the infinite volume limit. In the Hamiltonian context, this would mean that at any location x in the domain, local marginal distributions of NESS tend to a probability with density \(\frac{1}{Z} e^{-\beta (x) H}\), permitting one to define the local temperature at x to be \(\beta (x)^{-1}\). We prove also that in the infinite volume limit, the mean energy profile of NESS satisfies Laplace’s equation for the prescribed boundary condition. Our method of proof is duality: by reversing the sample paths of particle movements, we convert the problem of studying local marginal energy distributions at x to that of joint hitting distributions of certain random walks starting from x, and prove that the walks in question become increasingly independent as system size tends to infinity.

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.

Similar content being viewed by others

References

  1. Billingsley, P.: Convergence of Probability Measures. Wiley, New York (1968)

    MATH  Google Scholar 

  2. Bernardin, C., Olla, S.: Fourier’s law for a microscopic model of heat conduction. J. Stat. Phys. 121, 271–289 (2005)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  3. Bricmont, J., Kupiainen, A.: Towards a derivation of Fouriers law for coupled anharmonic oscillators. Commun. Math. Phys. 274(3), 555–626 (2007)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  4. Bonetto, F., Lebowitz, J.L., Lukkarinen, J.: Fourier’s law for a harmonic crystal with self-consistent stochastic reservoirs. J. Stat. Phys. 116(1–4), 783–813 (2004)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  5. Bonetto, F., Lebowitz, J.L., Rey-Bellet, L.: Fourier’s law: a challenge to theorists. In: Proceedings ICMP-2000, Imperial College Press, London, pp. 128–150 (2000)

  6. Carinci, G., Giardinà, C., Giberti, C., Redig, F.: Duality for stochastic models of transport. J. Stat. Phys. 152, 657–697 (2013)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  7. Carinci, G., Giardinà, C., Redig, F., Sasamoto, T.: Asymetric stochastic transport models with \({\cal U}_q({\mathfrak{s}}{\mathfrak{u}} (1,1))\) symmetry, preprint arXiv:1507.01478 (2015)

  8. Chung, K.L., Hunt, C.A.: On the zeros of \(\sum _1^n \pm 1\). Ann. Math. 50, 385–400 (1949)

    Article  MathSciNet  Google Scholar 

  9. de Finetti, B.: Funzione caratteristica di un fenomeno aleatorio. In: Atti della R. Academia Nazionale dei Lincei, Serie 6. Memorie, Classe di Scienze Fisiche, Mathematice e Naturale, vol. 4, pp. 251–299 (1931)

  10. Dolgopyat, D., Liverani, C.: Energy transfer in a fast-slow Hamiltonian system. Commun. Math. Phys. 308(1), 201–225 (2011)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  11. Dolgopyat, D., Nándori, P., Non equilibrium density profiles in Lorentz tubes with thermostated boundaries preprint (2014)

  12. Erdős, P., Taylor, S.J.: Some problems concerning the structure of random walk paths. Acta Math. Acad. Sci. Hungar. 11, 137–162 (1960)

    Article  MathSciNet  MATH  Google Scholar 

  13. Eckmann, J.-P., Young, L.-S.: Nonequilibrium energy profiles for a class of 1-D models. Commun. Math. Phys. 262(1), 237–267 (2006)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  14. Eyink, G., Lebowitz, J.L., Spohn, H.: Hydrodynamics of stationary nonequilibrium states for some lattice gas models. Commun. Math. Phys. 132, 252–283 (1990)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  15. Eyink, G., Lebowitz, J.L., Spohn, H.: Lattice gas models in contact with stochastic reservoirs: local equilibrium and relaxation to the steady state. Commun. Math. Phys. 140, 119–131 (1991)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  16. Gaspard, P., Gilbert, T.: Heat conduction and Fourier’s law in a class of many particle dispersing billiards. New J. Phys. 10(10), 103004 (2008)

    Article  ADS  Google Scholar 

  17. Groot, S.R., Mazur, P.: Non-equilibrium Thermodynamics. North-Holland, Amsterdam (1962)

    MATH  Google Scholar 

  18. Guo, M.Z., Papanicolau, G.C., Varadhan, S.R.S.: Nonlinear diffusion limit for a system with nearest neighbor interactions. Commun. Math. Phys. 118, 31–59 (1988)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  19. Jansen, S., Kurt, N.: On the notion(s) of duality for Markov processes. Probab. Surv. 11, 59–120 (2014)

    Article  MathSciNet  MATH  Google Scholar 

  20. Kipnis, C., Marchioro, C., Presutti, E.: Heat flow in an exactly solvable model. J. Stat. Phys. 27, 65–74 (1982)

    Article  ADS  MathSciNet  Google Scholar 

  21. Lawler, G.: Intersections of Random Walks. Birkhäuser, Boston (1991)

    Book  MATH  Google Scholar 

  22. Lin, K., Young, L.-S.: Nonequilibrium steady states for certain Hamiltonian models. J. Stat. Phys. 139(4), 630–657 (2010)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  23. Liverani, C., Olla, S.: Toward the Fourier law for a weakly interacting anharmonic crystal. J. Am. Math. Soc. 25, 555–583 (2012)

    Article  MathSciNet  MATH  Google Scholar 

  24. Larralde, H., Leyvraz, F., Mejia-Monasterio, C.: Transport properties of a modified Lorentz gas. J. Stat. Phys. 113(1), 197–231 (2003)

    Article  MathSciNet  MATH  Google Scholar 

  25. Petrov, V.V.: Sums of Independent Random Variables. Akademie-Verlag, Berlin (1975)

    Book  MATH  Google Scholar 

  26. Ravishankar, K., Young, L.-S.: Local thermodynamic equilibrium for some models of Hamiltonian origin. J. Stat. Phys. 128, 3 (2007)

    Article  MathSciNet  MATH  Google Scholar 

  27. Ruelle, D.: A mechanical model for Fourier’s law of heat conduction. Commun. Math. Phys. 311(3), 755–768 (2012)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  28. Zessin, H.: The method of moments for random measures. Z. Wahrscheinlichkeitstheorie verw. Gebiete 62, 395–409 (1983)

    Article  MathSciNet  MATH  Google Scholar 

Download references

Acknowledgments

LSY was supported in part by NSF Grant DMS-1363161.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Péter Nándori.

Appendix: Proof of Proposition 3

Appendix: Proof of Proposition 3

Our proof uses the continuous mapping theorem, which can be stated as follows (see e.g. Theorem 5.1 in [1]). As before, weak convergence is denoted by “\(\Rightarrow \)”, and for a mapping f and measure \(\mu \), \(f_*\mu \) is the measure given by \(f_*\mu (A)=\mu (f^{-1}(A))\).

Theorem 6

(Continuous mapping theorem) Let X and \(X'\) be separable metric spaces, and let P(X) denote the set of Borel probability measures on X. We consider a Borel measurable mapping \(f:X \rightarrow X'\) with discontinuity set \(D_f\), and let \(\mu _n, \mu \in P(X)\) be such that (i) \(\mu _n \Rightarrow \mu \) as \(n \rightarrow \infty \) and (ii) \(\mu (D_f) = 0\). Then \(f_*{\mu _n} \Rightarrow f_*\mu \) as \(n \rightarrow \infty \).

To fix some notation, for \(\mathfrak T \in \mathbb R^+\) we let

$$\begin{aligned} X = C([0, \mathfrak T], \mathbb R^d) \end{aligned}$$

be the set of continuous maps from \([0, \mathfrak T]\) to \(\mathbb R^d\) endowed with the sup norm, making it a separable metric space. For \(a \in \mathbb R^d\), we let \(B^a\) be the standard Brownian motion starting from \(a \in \mathbb R^d\) up to time \(\mathfrak T\), and with a slight abuse of notation, we use \(B^a\) to denote also the corresponding measure on X. Let \(W^a_n\) be the rescaled SSRW up to time \(\lfloor n\mathfrak T \rfloor \) starting from \(a \in \mathbb R^d\), i.e. if \(\hat{S}_k\) is a d-dimensional SSRW with \(\hat{S}_0=\langle a \sqrt{n} \rangle \), then \(W^a_n(k/n) = \frac{\hat{S}_k}{\sqrt{n}}\) for \(k = 0,1,\ldots , \lfloor n\mathfrak T \rfloor \), and \(W^a_n(t)\) is obtained by interpolating linearly between \( t = k/n\) and \(t = (k+1)/n\). As with \(B^a\), we use \(W^a_n\) to denote also the corresponding measure on X. By the invariance principle,

$$\begin{aligned} W^a_n \Rightarrow B^a \quad \text{ as } n \rightarrow \infty \end{aligned}$$
(34)

on any interval \([0, \mathfrak T]\). The convergence in (34) differs from that asserted in Proposition 3 in that the latter is for paths that terminate not at a fixed time but upon reaching \(\partial \mathcal D\).

Proof of Proposition 3

Let \(x \in \mathcal D\) and \(\varepsilon >0\) be given. We write

$$\begin{aligned} \mathcal D^1 = \{ y \in \mathbb R^d | \exists z \in \mathcal D, |y-z|<1\}, \end{aligned}$$

and let \(\mathfrak T = \mathfrak T (\varepsilon )\) be such that a Brownian motion starting from x reaches \(\partial \mathcal D^1\) before time \(\mathfrak T \) with probability at least \(1-\varepsilon /(2\Vert T\Vert _{\infty })\). For \(\omega \in X=C([0, \mathfrak T], \mathbb R^d)\), we define \(\tau (\omega ) = \min \{ t \in [0, \mathfrak T]: \omega (t) \in \partial \mathcal D \}\) if such a t exists, \(= \infty \) if it does not. Then we define \(f: X \rightarrow \mathbb R\) by

$$\begin{aligned} f( \omega ) \ = \ \left\{ \begin{array}{ll} T(\omega (\tau )) &{}\quad \text{ if } \tau \le \mathfrak T \\ \max _{y \in \partial \mathcal D} T(y) &{}\quad \text{ if } \tau = \infty . \end{array} \right. \end{aligned}$$

Lemma 18

\(B^x(D_f) = 0\)

We first finish the proof assuming the result of this lemma. Via a rigid translation, we may assume \(x =0\) (so that \(xL \in \mathbb Z^d\) for all L). Then with \(S_0=xL\), \(S_n/L\) in the proposition is \(W^0_{L^2} (n/L^2) \) in the notation above. Observe that we are now in the setting of the continuous mapping theorem: \(W^0_{L^2} \Rightarrow B^0\) as \(L \rightarrow \infty \) is condition (i) in Theorem 6, and the assertion in Lemma 18 is condition (ii). Thus the theorem applies, and its conclusion together with our choice of \(\mathfrak T\) gives exactly (5) in the case \(S_0=xL\).

To prove the full statement of Proposition 3, we observe that if the result was false, there would be a sequence \(x_k \in \mathbb R^d\) with \(x_k \rightarrow x\) and a sequence \(L_k \rightarrow \infty \) such that if \(S_n^{(k)}\) is the SSRW on \(\mathcal D_{L_k}\) with \(S^{(k)}_0 = x_k L_k\), then

$$\begin{aligned} \left| \mathbb E \left( T \left( \frac{S^{(k)}_{\tau }}{L_k} \right) \right) -u(x)\right| > \varepsilon \end{aligned}$$

where \(\tau \) is the smallest n such that \(S^{(k)}_n \in \mathcal B_{L_k}\). Such a scenario cannot occur: Since \(W^{x_k}_{L_k^2} = (x_k-x) + W^x_{L_k^2}\), it follows from (34) that \(W^{x_k}_{L_k^2} \Rightarrow B^x\) on \([0, \mathfrak T]\), and the argument in the last paragraph with \(W^{x_k}_{L_k^2}\) in the place of \(W^0_{L^2}\) gives the opposite inequality. \(\square \)

To complete the proof, it remains to show that the discontinuity set of f has zero Wiener measure.

Proof of Lemma 18

First, we identify the discontinuity set \(D_f\). If \(\tau (\omega ) = \infty \), then the trajectory of \(\omega \) up to time \(\mathfrak T\) is bounded away from \(\partial \mathcal D\), hence f is continuous at \(\omega \). If \(\tau (\omega ) < \mathfrak T\), then \(\liminf _{\omega ' \rightarrow \omega } \tau (\omega ') \ge \tau (\omega )\) for the same reason, but the corresponding \(\limsup \) can be strictly greater than \(\tau (\omega )\) if the trajectory of \(\omega \) does not cross to the other side of \(\partial \mathcal D\) immediately following \(\tau (\omega )\). More precisely, we have deduced that \(D_f = \{\tau = \mathfrak T\} \cup E\) where

$$\begin{aligned} E = \{\omega : \tau (\omega ) < \mathfrak T \text{ and } \exists \eta =\eta (\omega ) >0 \text{ s.t. } \omega ((\tau , \tau +\eta )) \subset \bar{\mathcal D}\} \end{aligned}$$

where \(\bar{\mathcal D}\) is the closure of D.

Clearly, \(\{ \tau = \mathfrak T\}\) has measure 0, so it suffices to show \(B^x(E)=0\).

Since harmonic measure is absolutely continuous, the set of \(\omega \) for which \(\omega (\tau )\) lies at a point at which \(\partial \mathcal D\) is not \(C^2\) differentiable has measure zero. Let \(\omega \) be outside of this measure zero set, and fix an orthonormal basis \(\{e_1, \ldots , e_d\}\) of \(\mathbb R^d\) such that \(e_1\) is the outward normal to \(\partial \mathcal D\) at \(\omega (\tau )\). Then there exist \(K>0\) and a neighborhood U of \(\omega (\tau )\) in \(\mathbb R^d\) such that

$$\begin{aligned} y \in \bar{\mathcal D} \quad \text { implies } \quad y_1 < K \Vert (y_2, \ldots , y_d)\Vert ^2. \end{aligned}$$
(35)

Recall that by the strong Markov property, \(B^x\) starting from the stopping time \(\tau \) is a Brownian motion. In particular, by projecting this Brownian motion, which we call \(\hat{B}(t)\), to the line parallel to \(e_1\) and to the hyperplane spanned by \(e_2, \ldots , e_d\), we obtain two independent Brownian motions, \(\hat{B}_1(t)\) and \(\hat{B}_{d-1}(t)\). Since \(\hat{B}_1(t)/\sqrt{t}\) and \(\hat{B}_{d-1}(t)/\sqrt{t}\) have standard normal distributions,

$$\begin{aligned} \mathbb P(|\hat{B}_1(t)| < t^{2/3}) \rightarrow 0 \quad \text { and } \quad \mathbb P(\Vert \hat{B}_{d-1}(t)\Vert > t^{1/3}/\sqrt{K}) \rightarrow 0 \end{aligned}$$

as \(t \rightarrow 0\). Choosing \(t_n \downarrow 0\) so that

$$\begin{aligned} \sum _n \mathbb P(|\hat{B}_1(t_n)| < t_n^{2/3}), \ \ \sum _n \mathbb P(\Vert \hat{B}_{d-1}(t_n)\Vert > t_n^{1/3}/\sqrt{K}) \ < \ \infty , \end{aligned}$$

it follows from the Borel–Cantelli lemma that \(|\hat{B}_1(t_n)| > t_n^{2/3}\) and \(\Vert \hat{B}_{d-1}(t_n)\Vert < t_n^{1/3}/\sqrt{K}\) hold for all but finitely many n. For \(t_n\) for which these inequalities hold, we are guaranteed that \(\hat{B}(t_n) \not \in \bar{\mathcal D}\) if \(\hat{B}_1(t_n)>0\) and \(\hat{B}(t_n) \in U\).

Now it is a well known fact that on the time interval \([0, \eta ]\) for every \(\eta >0\), a 1D Brownian motion starting from 0 makes infinitely many excursions from 0, and each excursion is positive with probability 1 / 2 independently of other excursions. Applying this fact to \(\hat{B}_1(t)\), and assuming (as we may) that each \(t_n\) lies in a different excursion, it follows that with probability 1, \(\hat{B}_1(t_n)>0\) for infinitely many n. Since \(\mathbb P(\hat{B}(t) \in U, t \in [0, \eta ]) \rightarrow 1\) as \(\eta \rightarrow 0\), we have proved that \( B^x (E) = 0\). \(\square \)

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Li, Y., Nándori, P. & Young, LS. Local Thermal Equilibrium for Certain Stochastic Models of Heat Transport. J Stat Phys 163, 61–91 (2016). https://doi.org/10.1007/s10955-016-1466-3

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s10955-016-1466-3

Keywords

Navigation