Skip to main content
Log in

Hydrodynamics of the Generalized N-urn Ehrenfest Model

  • Published:
Potential Analysis Aims and scope Submit manuscript

Abstract

In this paper we are concerned with a generalized N-urn Ehrenfest model, where balls perform independent random walks between N boxes uniformly laid on [0,1]. After a proper scaling of the transition rates function of the aforesaid random walk, we derive the hydrodynamic limit of the model, i.e., the law of large numbers which the empirical measure of the model follows, under an assumption where the initial number of balls in each box independently follows a Poisson distribution. We show that the empirical measure of the model converges weakly to a deterministic measure with density driven by an integral equation. Furthermore, we derive non-equilibrium fluctuation of the model, i.e, the central limit theorem from the above hydrodynamic limit. We show that the non-equilibrium fluctuation of the model is driven by a time-inhomogeneous generalized O-U process on the dual of C[0,1]. At last, we prove a large deviation principle from the hydrodynamic limit under an assumption where the transition rates function from [0,1] × [0,1] to \([0, +\infty )\) of the aforesaid random walk is a product of two marginal functions from [0,1] to \([0, +\infty )\).

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.

We’re sorry, something doesn't seem to be working properly.

Please try refreshing the page. If that doesn't work, please contact support so we can address the problem.

Data availability statements

Data sharing not applicable to this article as no datasets were generated or analysed during the current study.

References

  1. Andjel, E., Cocozza-Thivent, C., Roussignol, M.: Quelques compléments sur le processus des misanthropes et le processus ≪zéro-range≫. Ann. l’Inst. Henri Poincaré Probab. Stat. 21, 363–382 (1985)

    MATH  Google Scholar 

  2. Bahadoran, C., Guiol, H., Ravishankar, K., Saada, E.: Strong hydrodynamic limit for attractive particle systems on \(\mathbb {Z}\). Electron. J. Probab. 15, 1–43 (2010)

    Article  MathSciNet  MATH  Google Scholar 

  3. Balázs, M., Seppäläinen, T.: Exact connections between current fluctuations and the second class particle in a class of deposition models. J. Stat. Phys. 127, 431–455 (2007)

    Article  MathSciNet  MATH  Google Scholar 

  4. Blom, G.: Mean transition times for the Ehrenfest urn model. Adv. Appl. Probab. 21, 479–480 (1989)

    Article  MathSciNet  MATH  Google Scholar 

  5. Chen, Y. -P., Goldstein, I. H., Lathrop, E. D., Nelsen, R. B.: Computing an expected hitting time for the 3-urn Ehrenfest model via electric networks. Stat. Probab. Lett. 127, 42–48 (2017)

    Article  MathSciNet  MATH  Google Scholar 

  6. Cheng, X., Zhao, M., Yao, Q., Cui, E.: On the distribution of the hitting time for the N-urn Ehrenfest model. Stat. Probab. Lett. 157, 108625, 11 (2020)

  7. Cocozza-Thivent, C.: Processus des misanthropes. Z. Wahrscheinlichkeitstheorie Verwandte Gebiete 70, 509–523 (1985)

    Article  MathSciNet  MATH  Google Scholar 

  8. De Masi, A., Ferrari, P.: A remark on the hydrodynamics of the zero-range processes. J. Stat. Phys. 36, 81–87 (1984)

    Article  MathSciNet  MATH  Google Scholar 

  9. Erhard, D., Franco, T., Gonçalves, P., Neumann, A., Tavares, M.: Non-equilibrium fluctuations for the SSEP with a slow bond. Ann. l’Inst. Henri Poincaré Probab. Stat. 56, 1099–1128 (2020)

    MathSciNet  MATH  Google Scholar 

  10. Eymard, R., Roussignol, M., Tordeux, A.: Convergence of a misanthrope process to the entropy solution of 1D problems. Stoch. Process. Appl. 122, 3648–3679 (2012)

    Article  MathSciNet  MATH  Google Scholar 

  11. Fouque, J. P., Saada, E.: Totally asymmetric attractive particle systems on Z hydrodynamic limit for general initial profiles. Stoch. Process. Appl. 51, 9–23 (1994)

    Article  MathSciNet  MATH  Google Scholar 

  12. Funaki, T., Nishikawa, T.: Large deviations for the ginzburg ”CLandau ∇ϕ interface model. Probab. Theory Relat. Fields 120, 535–568 (2001)

    Article  MATH  Google Scholar 

  13. Gao, F. Q., Quastel, J.: Moderate deviations from the hydrodynamic limit of the symmetric exclusion process. Sci. China (Ser. A) 5, 577–592 (2003)

    Article  MathSciNet  MATH  Google Scholar 

  14. Gonċalves, P., Jara, M.: Nonlinear fluctuations of weakly asymmetric interacting particle systems. Arch. Ration. Mech. Anal. 212, 597–644 (2014)

    Article  MathSciNet  MATH  Google Scholar 

  15. Holley, R. A., Stroock, D. W.: Generalized Ornstein-Uhlenbeck processes and infinite particle branching brownian motions. Publ. Res. Inst. Math. Sci. 14, 741–788 (1978)

    Article  MathSciNet  MATH  Google Scholar 

  16. Jona-Lasinio, G., Landim, C., Vares, M. E.: Large deviations for a reaction diffusion model. Probab. Theory Relat. Fields 97, 339–361 (1993)

    Article  MathSciNet  MATH  Google Scholar 

  17. Jara, M., Menezes, O.: Non-equilibrium fluctuations for a reaction-diffusion model via relative entropy. arXiv:1810.03418 (2018)

  18. Kipnis, C., Landim, C.: Scaling limits of interacting particle systems. Springer, Berlin (1999)

  19. Kipnis, C., Olla, S., Varadhan, S. R. S.: Hydrodynamics and large deviation for simple exclusion processes. Commun. Pure Appl. Math. 42, 115–137 (1989)

    Article  MathSciNet  MATH  Google Scholar 

  20. Lang, S.: Undergraduate Analysis. Springer, New York (1983)

  21. Lathrop, E. D., Goldstein, I. H., Chen, Y. -P.: A note on a generalized Ehrenfest urn model: another look at the mean transition times. J. Appl. Probab. 53, 630–632 (2016)

    Article  MathSciNet  MATH  Google Scholar 

  22. Liggett, T.M.: Interacting Particle Systems. Springer, New York (1985)

  23. Meerson, B., Zilber, B.: Large deviations of a long-time average in the Ehrenfest urn model. J. Stat. Mech. Theory Exper. 5, 053202, 11, 82C05 (2018)

  24. Mitoma, I.: Tightness of probabilities on \(C(\lbrack 0, 1\rbrack ;{\mathscr{S}},')\) and \(D(\lbrack 0, 1\rbrack ;{\mathscr{S}}')\). Ann. Probab. 11, 989–999 (1983)

    Article  MathSciNet  Google Scholar 

  25. Palacios, J. L.: Another look at the Ehrenfest urn via electric networks. Adv. Appl. Probab. 26, 820–824 (1994)

    Article  MathSciNet  MATH  Google Scholar 

  26. Presutti, E., Spohn, H.: Hydrodynamics of the voter model. Ann. Probab. 11, 867–875 (1983)

    Article  MathSciNet  MATH  Google Scholar 

  27. Puhalskii, A.: The method of stochastic exponentials for large deviations. Stoch. Process. Appl. 54, 45–70 (1994)

    Article  MathSciNet  Google Scholar 

  28. Ren, L. R., Xue, X. F.: Moderate deviations of generalized N-urn Ehrenfest models. arXiv:2108.03040 (2021)

  29. Rezakhanlou, F.: Hydrodynamic limit for attractive particle systems on \(\mathbb {Z}\). Commun. Math. Phys. 140, 417–448 (1991)

    Article  MathSciNet  MATH  Google Scholar 

  30. Schuppen, V. J., Wong, E.: Transformation of local martingales under a change of law. Ann. Probab. 2, 879–888 (1974)

    Article  MathSciNet  MATH  Google Scholar 

  31. Sion, M.: On general minimax theorems. Pac. J. Math. 8, 171–176 (1958)

    Article  MathSciNet  MATH  Google Scholar 

  32. Song, S., Yao, Q.: Computing the expected hitting time for the n-urn Ehrenfest model via two methods. Preprint, current version available at, arXiv:https://arxiv.org/pdf/1610.09745.pdf (2018)

  33. Whitt, W.: Proofs of the martingale FCLT. Probab. Surv. 4, 268–302 (2007)

    Article  MathSciNet  MATH  Google Scholar 

  34. Xue, X. F.: Moderate deviations of density-dependent Markov chains. arXiv:1908.03762 (2019)

  35. Xue, X. F., Zhao, L. J.: Non-equilibrium fluctuations of the weakly asymmetric normalized binary contact path process. arXiv:2008.10350 (2020)

  36. Xue, X. F., Zhao, L. J.: Moderate Deviations for the SSEP with a Slow Bond. arXiv:2005.14393 (2020)

Download references

Acknowledgements

The author is grateful to Dr. Zhengyao Sun, Dr. Linjie Zhao, Prof. Minzhi Zhao and Prof. Qiang Yao for useful suggestions and comments. The author is grateful to the reviewers. Their comments are great help for the improvement of this paper. The author is grateful to the financial support from the National Natural Science Foundation of China with grant number 11501542.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Xiaofeng Xue.

Additional information

Publisher’s Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Appendices

A Appendix

A.1 Proof of Lemma 2.2

Proof of Lemma 2.2

We first construct a weak solution μ of Eq. 2.1 to show the existence. We use \(\|f\|_{\infty }\) to denote the \(l_{\infty }\)-norm of f for any fC[0, 1] and \(\|\lambda \|_{\infty }\) to denote \(\sup _{0\leq x, y\leq 1}\lambda (x, y)\). For each fC[0, 1], we define

$$ \left( P_{3}f\right)(x)={{\int}_{0}^{1}}\lambda(y, x)f(y)dy $$

for any x ∈ [0, 1]. Since \(\|P_{2}f\|_{\infty }, \|P_{3}f\|_{\infty }\leq \|\lambda \|_{\infty }\|f\|_{\infty }\) for any fC[0, 1], the linear C[0, 1]-valued ODE

$$ \begin{cases} \frac{d}{dt}\rho_{t}=(P_{3}-P_{2})\rho_{t}, &\text{~}0\leq t\leq T,\\ \rho_{0}(x)=\phi(x), &\text{~}0\leq x\leq 1, \end{cases} $$

satisfies Lipschitz’ condition and hence has the unique solution

$$ \rho_{t}=e^{t(P_{3}-P_{2})}\phi=\sum\limits_{k=0}^{+\infty}\frac{t^{k}\left( P_{3}-P_{2}\right)^{k}}{k!}\phi. $$

Let μt(dx) = ρt(x)dx for any 0 ≤ tT, then μ = {μt}0≤tT is a weak solution of Eq. 2.1 since

$$ {{\int}_{0}^{1}}f(x)\left( P_{3}\rho_{t}\right)(x)dx={{\int}_{0}^{1}} \left( P_{1}f\right)(x)\rho_{t}(x)dx. $$

Now we only need to show the uniqueness. Let μ1,μ2 be two weak solutions of Eq. 2.1 and ν = μ1μ2, then

$$ \nu_{t}(f)={{\int}_{0}^{t}} \nu_{s}\left( (P_{1}-P_{2})f\right)ds $$

for any 0 ≤ tT. For any \({\mathscr{A}}\in \mathcal {S}\), we define \(\|{\mathscr{A}}\|_{\infty }=\sup \left \{{\mathscr{A}}(f):~\|f\|_{\infty }=1\right \}\). Then, for any f satisfying \(\|f\|_{\infty }=1\),

$$ |\nu_{s}\left( (P_{1}-P_{2})f\right)|\leq \|\nu_{s}\|_{\infty}\|(P_{1}-P_{2})f\|_{1}\leq 2\|\nu_{s}\|_{\infty}\|\lambda\|_{\infty}. $$

Therefore, for any 0 ≤ tT,

$$ \|\nu_{t}\|_{\infty}\leq 2\|\lambda\|_{\infty}{{\int}_{0}^{t}} \|\nu_{s}\|_{\infty}ds. $$

Then, by Gronwall’s inequality, \(\|\nu _{t}\|_{\infty }\leq 0e^{2t\|\lambda \|_{\infty }}=0\) for any 0 ≤ tT and hence μ1 = μ2. □

A.2 The uniqueness of solution to Eq. 2.4

In this subsection we give an outline of the proof of the uniqueness of solution to Eq. 2.4. Our proof follows an analysis similar with that given in the proof of Theorem 1.4 of [15].

Proof of uniqueness to Eq. 2.4

For each t ≥ 0 and HC([0, 1]), we define

$$ A_{t}(H)=V_{t}(H)-V_{0}(H)-{{\int}_{0}^{t}} V_{s}\left( (P_{1}-P_{2})H\right)ds $$

and

$$ {\Xi}_{t}(H)={V_{t}^{2}}(H)-{V_{0}^{2}}(H)-{{\int}_{0}^{t}}2V_{s}(H)V_{s}\left( (P_{1}-P_{2})H\right)ds-{{\int}_{0}^{t}}\|b_{s}H\|_{2}^{2}ds. $$

According to the second statement in the definition of {Vt}t≥ 0, it is easy to check that {At(H)}t≥ 0 and {Ξt(H)}t≥ 0 are both local martingales. Since

$$ dV_{t}(H)=V_{t}(\left( P_{1}-P_{2}\right)H)dt+dA_{t}(H), $$

by Itô’s formula,

$$ \begin{array}{@{}rcl@{}} d{V_{t}^{2}}(H)&=&2V_{t}(H)dV_{t}(H)+d\langle A(H)\rangle_{t}\\ &=&2V_{t}(H)V_{t}(\left( P_{1}-P_{2}\right)H)dt+d\langle A(H)\rangle_{t}+2V_{t}(H)dA_{t}(H) \end{array} $$

and hence

$$ \left\{{V_{t}^{2}}(H)-{V_{0}^{2}}(H)-{{\int}_{0}^{t}}2V_{s}(H)V_{s}\left( (P_{1}-P_{2})H\right)ds-\langle A(H)\rangle_{t}\right\}_{t\geq 0} $$

is a local martingale. Then, since {Ξt}t≥ 0 is also a local martingale,

$$ \langle A(H)\rangle_{t}={{\int}_{0}^{t}} \|b_{s}H\|_{2}^{2}ds. $$

Let \(i=\sqrt {-1}\) and

$$ J_{t}(H)=\exp\left\{iA_{t}(H)+\frac{1}{2}\langle A(H)\rangle_{t}\right\}=\exp\left\{iA_{t}(H)+\frac{1}{2}{{\int}_{0}^{t}}\|b_{s}H\|_{2}^{2}ds\right\} $$

for any t ≥ 0, then {Jt(H)}0≤tT is a local martingale according to Itô’s formula and hence a martingale since {Jt(H)}0≤tT are uniformly bounded. As a result, let \({{\Upsilon }_{s}^{t}}(H)=\frac {J_{t}(H)}{J_{s}(H)}\) for any 0 ≤ s < t, then

$$ \mathbb{E}\left( {{\Upsilon}_{s}^{t}}(H)\big|V_{u}, u\leq s\right)=1 $$
(A.1)

for any 0 ≤ s < t and HC([0, 1]). For given 0 ≤ t1 < t2T and integers 0 ≤ kn, let \(\sigma _{n,k}=t_{1}+\frac {k(t_{2}-t_{1})}{n}\). For any t ≥ 0, let \(H_{t}=e^{t(P_{1}-P_{2})}H\), i.e., {Ht}t≥ 0 is the solution to

$$ \begin{cases} &\frac{d}{dt}H_{t}=(P_{1}-P_{2})H_{t}, \\ &H_{0}=H. \end{cases} $$

It is easy to check that {Ht}0≤tT is continuous in C([0, 1]) and hence

$$ \begin{array}{@{}rcl@{}} \lim\limits_{n\rightarrow+\infty} \Bigg(\sum\limits_{k=0}^{n-1}{\int}_{\sigma_{n,k}}^{\sigma_{n, k+1}}V_{u}\left( (P_{1}-P_{2})H_{t_{2}-\sigma_{n, k+1}}\right)du\\ -\sum\limits_{k=0}^{n-1}{\int}_{\sigma_{n,k}}^{\sigma_{n, k+1}}V_{\sigma_{n,k}}\left( (P_{1}-P_{2})H_{t_{2}-u}\right)du\Bigg)=0 \end{array} $$

almost surely. Furthermore, since

$$ \begin{array}{@{}rcl@{}} \sum\limits_{k=0}^{n-1}{\int}_{\sigma_{n,k}}^{\sigma_{n, k+1}}V_{\sigma_{n,k}}\left( (P_{1}-P_{2})H_{t_{2}-u}\right)du &=&\sum\limits_{k=0}^{n-1}V_{\sigma_{n,k}}\left( {\int}_{\sigma_{n,k}}^{\sigma_{n, k+1}}(P_{1}-P_{2})H_{t_{2}-u}du\right)\\ &=&-\sum\limits_{k=0}^{n-1}V_{\sigma_{n,k}}\left( H_{t_{2}-\sigma_{n, k+1}}-H_{t_{2}-\sigma_{n,k}}\right), \end{array} $$

we have

$$ \lim\limits_{n\rightarrow+\infty}\prod\limits_{k=0}^{n-1}{\Upsilon}_{\sigma_{n,k}}^{\sigma_{n, k+1}}\left( H_{t_{2}-\sigma_{n, k+1}}\right) =\frac{Y_{t_{2}}}{Y_{t_{1}}} $$

almost surely, where

$$ Y_{t}=\exp\left\{iV_{t}(H_{t_{2}-t})+\frac{1}{2}{{\int}_{0}^{t}}\|b_{u}H_{t_{2}-u}\|_{2}^{2}du\right\} $$

for 0 ≤ tt2. Since \(\{\prod_{k=0}^{n-1}{\Upsilon }_{\sigma _{n,k}}^{\sigma _{n, k+1}}\left (H_{t_{2}-\sigma _{n, k+1}}\right )\}_{n\geq 1}\) are uniformly bounded, the convergence above is also in L1. As a result,

$$ \mathbb{E}\big(Y_{t_{2}}\big|V_{u}, u\leq t_{1}\big)=\lim_{n\rightarrow+\infty}\mathbb{E}\big(Y_{t_{1}}\prod\limits_{k=0}^{n-1}{\Upsilon}_{\sigma_{n,k}}^{\sigma_{n, k+1}}\left( H_{t_{2}-\sigma_{n, k+1}}\right)\big|V_{u}, u\leq t_{1}\big) $$

in L1. Repeatedly utilizing (A.1) by n times, we have

$$ \mathbb{E}\big(Y_{t_{1}}\prod\limits_{k=0}^{n-1}{\Upsilon}_{\sigma_{n,k}}^{\sigma_{n, k+1}}\left( H_{t_{2}-\sigma_{n, k+1}}\right)\big|V_{u}, u\leq t_{1}\big) =Y_{t_{1}} $$

and hence

$$ \mathbb{E}\left( \exp\{iV_{t_{2}}(H)\}\big|V_{u}, u\leq t_{1}\right)=e^{iV_{t_{1}}(H_{t_{2}-t_{1}})-\frac{1}{2}{\int}_{t_{1}}^{t_{2}}\|b_{u}H_{t_{2}-u}\|_{2}^{2}du} $$
(A.2)

for any 0 ≤ t1 < t2T and HC([0, 1]). The uniqueness of {Vt}t≥ 0 follows from Eq. A.2 and the given distribution of V0. □

A.3 A generalized N-urn Ehrenfest model with a metastable state

Throughout this subsection we assume that our model is under Assumptions (A) and (B) given in Section 2. We will show that in this case the model has a metastable state, i.e, \(\lim _{t\rightarrow +\infty }\mu _{t}=\mu _{\infty }\) for some \(\mu _{\infty }\in \mathcal {S}\), where μt≥ 0 satisfies that {μt}0≤tT is the solution to Eq. 2.1 for any T > 0.

The proof of the existence of \(\mu _{\infty }\)

By Lemma 2.2, μt(dx) = ρ(t,x)dx, where

$$ \frac{d}{dt}\rho(t,x)=-m\lambda_{1}(x)\rho(t,x)+\lambda_{2}(x)r(t) $$

under Assumption (B), where \(m={{\int \limits }_{0}^{1}}\lambda _{2}(y)dy\) and \(r(t)={{\int \limits }_{0}^{1}}\lambda _{1}(y)\rho (t,y)dy\). Let \(K(t,x)=\frac {\lambda _{1}(x)}{\lambda _{2}(x)}\rho (t,x)\), \(\hat {\lambda }_{1}(x)=m\lambda _{1}(x)\), \(\hat {\lambda }_{2}(x)=\frac {\lambda _{2}(x)}{m}\), then

$$ \frac{d}{dt}K(t,x)=\hat{\lambda}_{1}(x){{\int}_{0}^{1}}\left( K(t,y)-K(t,x)\right)\hat{\lambda}_{2}(y)dy. $$
(A.3)

Let \(\{\mathcal {R}_{t}\}_{t\geq 0}\) be the Markov process with state space [0, 1] and generator \(\mathcal {A}\) given by

$$ \mathcal{A}f(x)=\hat{\lambda}_{1}(x){{\int}_{0}^{1}} \left( f(y)-f(x)\right)\hat{\lambda}_{2}(y)dy $$

for any fC[0, 1] and x ∈ [0, 1], i.e, conditioned on \(\mathcal {R}_{t}=x\), the process waits for an exponential time with rate \(\hat {\lambda }_{1}(x)\) to escape from x and choose the next state y according to the probability measure \(\hat {\lambda }_{2}(y)dy\). Then, by Eq. A.3,

$$ K(t,x)=\mathbb{E}_{x}K(0, \mathcal{R}_{t}). $$
(A.4)

It is easy to check that π(x)dx is a reversible distribution of \(\{\mathcal {R}_{t}\}_{t\geq 0}\), where

$$ \pi(x)=\frac{\frac{\hat{\lambda}_{2}(x)}{\hat{\lambda}_{1}(x)}}{{{\int}_{0}^{1}}\frac{\hat{\lambda}_{2}(y)}{\hat{\lambda}_{1}(y)}dy} $$

for any x ∈ [0, 1]. For any probability measure ν on [0, 1] absolutely continuous with respect to Lebesgue measure, we claim that \(\mathcal {R}_{t}\) converges weakly to π(x)dx as \(t\rightarrow +\infty \) conditioned on \(\mathcal {R}_{0}\) is with distribution ν. We prove this claim later. As a result, for any positive fC[0, 1] with \({{\int \limits }_{0}^{1}}f(x)dx=1\),

$$ \begin{array}{@{}rcl@{}} \lim\limits_{t\rightarrow+\infty}{{\int}_{0}^{1}}f(x)K(t,x)dx&=\lim_{t\rightarrow+\infty}{{\int}_{0}^{1}}f(x)\mathbb{E}_{x}K(0, \mathcal{R}_{t})dx\\ &={{\int}_{0}^{1}}\pi(x)K(0, x)dx \end{array} $$
(A.5)

by Eq. A.4. Since \(K(0, x)=\frac {\lambda _{1}(x)}{\lambda _{2}(x)}\rho (0, x)=\frac {\lambda _{1}(x)}{\lambda _{2}(x)}\phi (x)\) and \(\frac {h(x)}{{{\int \limits }_{0}^{1}}h(x)dx}\) is a probability density for any positive hC[0, 1], Eq. A.5 implies that \(\lim _{t\rightarrow +\infty }\mu _{t}=\mu _{\infty }\), where

$$ \mu_{\infty}(dx)=\frac{{{\int}_{0}^{1}}\phi(u)du}{{{\int}_{0}^{1}}\frac{\lambda_{2}(u)}{\lambda_{1}(u)}du}\frac{\lambda_{2}(x)}{\lambda_{1}(x)}dx. $$

At last we prove our claim. The proof follows the entropy method given in Section 2.4 of [22] and hence we only give an outline. For any ν absolutely continuous with respect to Lebesgue measure, let \(\mathcal {E}(\nu )\) be the entropy of ν relative to π, i.e,

$$ \mathcal{E}(\nu)={{\int}_{0}^{1}}q\left( \frac{f_{\nu}(x)}{\pi(x)}\right)\pi(x)dx, $$

where \(q(x)=x\log x\) and fν is the probability density of ν. For any t ≥ 0, let νt be the probability measure of \(\mathcal {R}_{t}\) conditioned on \(\mathcal {R}_{0}\) with distribution ν. By Chapman-Kolmogorov equation, νt is also absolutely continuous with respect to Lebesgue measure and its probability density {fν(t,x) : x ∈ [0, 1]} is the solution to the equation

$$ \begin{cases} \frac{d}{dt}f_{\nu}(t,x)=-\hat{\lambda}_{1}(x)f_{\nu}(t,x)+\hat{\lambda}_{2}(x){{\int}_{0}^{1}}f_{\nu}(t,y)\hat{\lambda}_{1}(y)dy,\\ f_{\nu}(0, x)=f_{\nu}(x). \end{cases} $$
(A.6)

For any y ∈ [0, 1], let pt(y,du) be the probability measure of \(\mathcal {R}_{t}\) conditioned on \(\mathcal {R}_{0}=y\), then for any Borel-measurable set B ⊂ [0, 1], we define

$$ m(t, B)={{\int}_{0}^{1}}p_{t}(y, B)dy, $$

i.e, m(t,du) is the probability measure of \(\mathcal {R}_{t}\) conditioned on \(\mathcal {R}_{0}\) is uniformly distributed on [0, 1]. Then, m(t,du) is absolutely continuous with respect to Lebesgue measure and its probability density fm(t,x) is the solution to Eq. A.6 with initial condition fm(0,x) = 1 for all x ∈ [0, 1]. By classic theory of measures on product spaces, for each u ∈ [0, 1], there exists probability measure m2(t,u,dy) such that

$$ {\int}_{A}p_{t}(y, B)dy={\int}_{B}m_{2}(t, u, A)m(t, du)={\int}_{B}m_{2}(t, u, A)f_{m}(t,u)du $$

for any Borel-measurable \(A, B\subseteq [0, 1]\). Since

$$ \mathbb{E}_{\nu} h(\mathcal{R}_{t})={{\int}_{0}^{1}} f_{\nu}(y)\left( {\int}_{[0, 1]}h(u)p_{t}(y, du)\right)dy $$

for any hC[0, 1], we have

$$ f_{\nu}(t,x)=f_{m}(t,x){\int}_{[0, 1]}f_{\nu}(y)m_{2}(t, x, dy) $$
(A.7)

and especially

$$ \pi(x)=f_{m}(t,x){\int}_{[0, 1]}\pi(y)m_{2}(t, x, dy) $$
(A.8)

since π(x)dx is an invariant measure of \(\{\mathcal {R}_{t}\}_{t\geq 0}\). By Eqs. A.7A.8 and Jensen’s inequality,

$$ \begin{array}{@{}rcl@{}} \mathcal{E}(\nu_{t})&=&{{\int}_{0}^{1}}q\left( \frac{f_{\nu}(t, x)}{\pi(x)}\right)\pi(x)dx\\ &=&{{\int}_{0}^{1}}q\left( {\int}_{[0, 1]}\frac{f_{m}(t,x)f_{\nu}(y)}{\pi(x)}m_{2}(t, x, dy)\right)\pi(x)dx\\ &=&{{\int}_{0}^{1}}q\left( {\int}_{[0, 1]}f_{m}(t,x)\frac{\pi(y)}{\pi(x)}\frac{f_{\nu}(y)}{\pi(y)}m_{2}(t,x,dy)\right)\pi(x)dx\\ &\leq& {{\int}_{0}^{1}}\left( {\int}_{[0, 1]}f_{m}(t, x)\frac{\pi(y)}{\pi(x)}q\left( \frac{f_{\nu}(y)}{\pi(y)}\right)m_{2}(t, x, dy)\right)\pi(x)dx\\ &=&{\int}_{[0, 1]}\left( {\int}_{[0, 1]}\pi(y)q\left( \frac{f_{\nu}(y)}{\pi(y)}\right)m_{2}(t, x, dy)\right)m(t, dx)\\ &=&{{\int}_{0}^{1}}\pi(y)q\left( \frac{f_{\nu}(y)}{\pi(y)}\right)\left( {\int}_{[0, 1]}1 p_{t}(y, dx)\right)dy\\ &=&{{\int}_{0}^{1}}\pi(y)q\left( \frac{f_{\nu}(y)}{\pi(y)}\right)dy=\mathcal{E}(\nu). \end{array} $$

As a result, \(\mathcal {E}(\nu _{t})\) is decreasing with t and hence

$$ \lim\limits_{t\rightarrow+\infty}\mathcal{E}(\nu_{t}) $$

exists. Furthermore, by Chapman-Kolmogorov equation,

$$ p_{t}(y, du)=l_{1}(t, y)\delta_{y}(du)+l_{2}(t, y, u)du, $$

where δy(du) is the Dirac measure concentrated on y and l1(t,y),l2(t,y,⋅) is the solution to the equation

$$ \begin{cases} &\frac{d}{dt}l_{1}(t, y)=-\hat{\lambda}_{1}(y)l_{1}(t,y),\\ &\frac{d}{dt}l_{2}(t, y, u)=\hat{\lambda}_{2}(u)l_{1}(t,y)\hat{\lambda}_{1}(y)+\hat{\lambda}_{2}(u){{\int}_{0}^{1}}l_{2}(t, y, x)\hat{\lambda}_{1}(x)dx-l_{2}(t, y, u)\hat{\lambda}_{1}(u),\\ &l_{1}(0, y)=1,\\ &l_{2}(0, y, u)=0. \end{cases} $$

Hence, \(l_{1}(t, y)=e^{-\hat {\lambda }_{1}(y)t}\) and l2(t,y,u) > 0 for any t > 0,u ∈ [0, 1]. According to the definition of m2(t,x,dy),

$$ m_{2}(t, x, dy)f_{m}(t,x)=l_{1}(x)\delta_{x}(dy)+l_{2}(t, y, x)dy. $$

As a result, m2(t,x,B) = 0 implies that the Lebesgue measure of B is 0 for any t > 0 and x ∈ [0, 1]. Consequently, \(\mathcal {E}(\nu _{t})=\mathcal {E}(\nu )\) when and only when ν(dx) = π(x)dx. Since {νt}t≥ 0 are probability measures on [0, 1], they are tight. Let \(\tilde {\nu }\) be the weak limit of a subsequence \(\{\nu _{t_{k}}\}_{k\geq 1}\) of {νt}t≥ 0. By Eq. A.6,

$$ f_{\nu}(t, x)=f_{\nu}(x)e^{-\hat{\lambda}_{1}(x)t}+{{\int}_{0}^{t}}e^{\hat{\lambda}_{1}(x)(s-t)}\hat{\lambda}_{2}(x)\mathbb{F}_{\nu}(s)ds $$
(A.9)

where \(\mathbb {F}_{\nu }(s)={{\int \limits }_{0}^{1}}\hat {\lambda }_{1}(y)f_{\nu }(s,y)dy\). By Eq. A.9, it is easy to check that {fν(t,⋅)}t≥ 0 are uniformly bounded and equicontinuous. As a result, \(\tilde {\nu }\) is absolutely continuous with respect to Lebesgue measure and its probability density \(f_{\tilde {\nu }}\) is a C[0, 1]-limit of a subsequence of {fν(t,⋅)}t≥ 0. Since q is continuous on [0, 1],

$$ \mathcal{E}(\tilde{\nu})=\lim\limits_{t\rightarrow+\infty}\mathcal{E}(\nu_{t}). $$

For any given s > 0, \(\tilde {\nu }_{s}\) is the weak-limit of the subsequence \(\{\nu _{t_{k}+s}\}_{k\geq 1}\) of {νt}t≥ 0. As a result,

$$ \mathcal{E}(\tilde{\nu}_{s})=\lim_{t\rightarrow+\infty}\mathcal{E}(\nu_{t})=\mathcal{E}(\tilde{\nu}) $$

and hence \(\tilde {\nu }(dx)=\pi (x)dx\). As a result, νt converges weakly to π(x)dx as \(t\rightarrow +\infty \) and the proof is complete. □

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Xue, X. Hydrodynamics of the Generalized N-urn Ehrenfest Model. Potential Anal 59, 613–649 (2023). https://doi.org/10.1007/s11118-021-09980-7

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s11118-021-09980-7

Keywords

Mathematics Subject Classification (2010)

Navigation