Skip to main content
Log in

Finite Size Scaling of the Dynamical Free-Energy in a Kinetically Constrained Model

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

Abstract

We determine the finite size corrections to the large deviation function of the activity in a kinetically constrained model (the Fredrickson-Andersen model in one dimension), in the regime of dynamical phase coexistence. Numerical results agree with an effective model where the boundary between active and inactive regions is described by a Brownian interface.

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

Similar content being viewed by others

References

  1. Berthier, L., Biroli, G.: Theoretical perspective on the glass transition and amorphous materials. Rev. Mod. Phys. 83, 587 (2011)

    Article  ADS  Google Scholar 

  2. Ritort, F., Sollich, P.: Glassy dynamics of kinetically constrained models. Adv. Phys. 52, 219 (2003)

    Article  ADS  Google Scholar 

  3. Garrahan, J.P., Sollich, P., Toninelli, C.: Kinetically constrained models. In: Berthier, L., Biroli, G., Bouchaud, J.-P., Cipelletti, L., van Saarloos, W. (eds.) Dynamical Heterogeneities in Glasses, Colloids, and Granular Media. Oxford University Press, London (2011). arXiv:1009.6113 (2010)

    Google Scholar 

  4. Merolle, M., Garrahan, J.P., Chandler, D.: Space-time thermodynamics of the glass transition. Proc. Natl. Acad. Sci. USA 102, 10837 (2005)

    Article  ADS  Google Scholar 

  5. Lecomte, V., Appert-Rolland, C., Wijland, F.: Thermodynamic formalism for systems with Markov dynamics. J. Stat. Phys. 127, 51 (2007)

    Article  MathSciNet  ADS  MATH  Google Scholar 

  6. Garrahan, J.P., Jack, R.L., Lecomte, V., Pitard, E., van Duijvendijk, K., van Wijland, F.: Dynamical first-order phase transition in kinetically constrained models of glasses. Phys. Rev. Lett. 98, 195702 (2007)

    Article  ADS  Google Scholar 

  7. Garrahan, J.P., Jack, R.L., Lecomte, V., Pitard, E., van Duijvendijk, K., van Wijland, F.: First-order dynamical phase transition in models of glasses: an approach based on ensembles of histories. J. Phys. A 42, 075007 (2009)

    Article  MathSciNet  ADS  Google Scholar 

  8. Bodineau, T., Toninelli, C.: Activity phase transition for constrained dynamics. Commun. Math. Phys. (2011). doi:10.1007/s00220-012-1449-4. arXiv:1101.1760

    MATH  Google Scholar 

  9. Maes, C., Netočný, K., Wynants, B.: On and beyond entropy production: the case of Markov jump processes. Markov Process. Relat. Fields 14, 445 (2008)

    MATH  Google Scholar 

  10. Hedges, L.O., Jack, R.L., Garrahan, J.P., Chandler, D.: Dynamic order-disorder in atomistic models of structural glass formers. Science 323, 1309 (2009)

    Article  ADS  Google Scholar 

  11. Bodineau, T., Lefevere, R.: Large deviations of lattice hamiltonian dynamics coupled to stochastic thermostats. J. Stat. Phys. 133, 1 (2008)

    Article  MathSciNet  ADS  MATH  Google Scholar 

  12. Pitard, E., Lecomte, V., van Wijland, F.: Dynamic transition in an atomic glass former: a molecular dynamics evidence. Europhys. Lett. 96, 184207 (2011)

    Article  Google Scholar 

  13. Chandler, D., Garrahan, J.P.: Dynamics on the way to forming glass: bubbles in space-time. Annu. Rev. Phys. Chem. 61, 191 (2010)

    Article  Google Scholar 

  14. Jack, R.L., Garrahan, J.P., Chandler, D.: Space-time thermodynamics and subsystem observables in a kinetically constrained model of glassy materials. J. Chem. Phys. 125, 184509 (2006)

    Article  ADS  Google Scholar 

  15. Appert-Rolland, C., Derrida, B., Lecomte, V., van Wijland, F.: Universal cumulants of the current in diffusive systems on a ring. Phys. Rev. E 78, 021122 (2008)

    Article  ADS  Google Scholar 

  16. Lecomte, V., Imparato, A., van Wijland, F.: Current fluctuations in systems with diffusive dynamics, in and out of equilibrium. Prog. Theor. Phys. Suppl. 184, 276 (2010)

    Article  ADS  MATH  Google Scholar 

  17. Gorissen, M., Vanderzande, C.: Finite size scaling of current fluctuations in the totally asymmetric exclusion process. J. Phys. A 44, 115005 (2011)

    Article  MathSciNet  ADS  Google Scholar 

  18. Bodineau, T., Derrida, B.: Distribution of current in nonequilibrium diffusive systems and phase transitions. Phys. Rev. E 72, 066110 (2005)

    Article  MathSciNet  ADS  Google Scholar 

  19. Bertini, L., De Sole, A., Gabrielli, D., Jona-Lasinio, G., Landim, C.: Non equilibrium current fluctuations in stochastic lattice gases. J. Stat. Phys. 123, 237 (2006)

    Article  MathSciNet  ADS  MATH  Google Scholar 

  20. Fredrickson, G.H., Andersen, H.C.: Kinetic Ising model of the glass transition. Phys. Rev. Lett. 53, 1244 (1984)

    Article  ADS  Google Scholar 

  21. Hryniv, O., Velenik, Y.: Universality of critical behaviour in a class of recurrent random walks. Probab. Theory Relat. Fields 130, 222 (2004)

    Article  MathSciNet  MATH  Google Scholar 

  22. Velenik, Y.: Entropic repulsion of an interface in an external field. Probab. Theory Relat. Fields 129, 83 (2004)

    Article  MathSciNet  MATH  Google Scholar 

  23. Majumdar, S.N., Comtet, A.: Airy distribution function: from the area under a brownian excursion to the maximal height of fluctuating interfaces. J. Stat. Phys. 119, 777 (2005)

    Article  MathSciNet  ADS  MATH  Google Scholar 

  24. Lecomte, V., Tailleur, J.: A numerical approach to large deviations in continuous time. J. Stat. Mech. P03004 (2007)

  25. Tailleur, J., Lecomte, V.: Simulation of large deviation functions using population dynamics. AIP Conf. Proc. 1091, 212 (2009)

    Article  ADS  Google Scholar 

  26. Giardinà, C., Kurchan, J., Peliti, L.: Direct evaluation of large-deviation functions. Phys. Rev. Lett. 96, 120603 (2006)

    Article  ADS  Google Scholar 

  27. Giardinà, C., Kurchan, J., Lecomte, V., Tailleur, J.: Simulating rare events in dynamical processes. J. Stat. Phys. 145, 787 (2011)

    Article  ADS  MATH  Google Scholar 

  28. Hurtado, P.I., Garrido, P.L.: Current fluctuations and statistics during a large deviation event in an exactly solvable transport model. J. Stat. Mech. P02032 (2009)

  29. Turci, F., Pitard, E.: Large deviations and heterogeneities in a driven kinetically constrained model. Europhys. Lett. 94, 10003 (2011)

    Article  ADS  Google Scholar 

  30. Touchette, H.: The large deviation approach to statistical mechanics. Phys. Rep. 478, 1 (2009)

    Article  MathSciNet  ADS  Google Scholar 

  31. Garrahan, J.P., Chandler, D.: Geometrical explanation and scaling of dynamical heterogeneities in glass forming systems. Phys. Rev. Lett. 89, 035704 (2002)

    Article  ADS  Google Scholar 

  32. Berthier, L., Biroli, G., Bouchaud, J.-P., Kob, W., Miyazaki, K., Reichman, D.R.: Spontaneous and induced dynamic correlations in glass formers. II. Model calculations and comparison to numerical simulations. J. Chem. Phys. 126, 184504 (2007)

    Article  ADS  Google Scholar 

  33. Candelier, R., Widmer-Cooper, A., Kummerfeld, J.K., Dauchot, O., Biroli, G., Harrowell, P., Reichman, D.R.: Spatiotemporal hierarchy of relaxation events, dynamical heterogeneities, and structural reorganization in a supercooled liquid. Phys. Rev. Lett. 105, 135702 (2010)

    Article  ADS  Google Scholar 

  34. Keys, A., Hedges, L., Garrahan, J., Glotzer, S., Chandler, D.: Excitations are localized and relaxation is hierarchical in glass-forming liquids. Phys. Rev. X 1, 021013 (2011)

    Article  Google Scholar 

Download references

Acknowledgements

We would like to thank Frédéric van Wijland for useful discussions, and Christophe Berthod and Thierry Giamarchi for the Mafalda cluster at DPMC, University of Geneva, where part of the simulations were run. T.B., V.L. and C.T acknowledge funding from ANR SHEPI and C.T. from ERC Advanced Grant PTRELSS 228032.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Vivien Lecomte.

Appendices

Appendix A: Scaling of ψ L (s) in the Interface Model

We determine in this appendix the finite size corrections to the function \(\hat{\varphi}_{L}(\lambda)\), defined in (9), in the effective interface description discussed in part 2.2. Let us first focus on a system with one boundary x(t) between the active and inactive regions. It performs a random walk of jump rate p (resp. q) to the left (resp. right), starting from x=0 at time 0, and constrained to come back to 0 at final time t. The walk x(t) is reflected at 0 so that x(t)≥0. Appropriate values of p and q are discussed in Sect. 4.1. The computation is done for generic values of p and q.

In the original model, the cost of maintaining the interface (that is, for x(τ) to come back in 0 at time τ=t) is given by the surface tension Σ. In this effective description

$$Z_{\mathrm{eff}}(s,t)\equiv\frac{ \langle e^{-s{\mathbb{K}}\int_0^td\tau\,x(\tau)}\delta (x(t)=0) \rangle_{p,q}}{ \langle\delta(x(t)=0) \rangle_{p,q}}$$
(43)

where \({\mathbb{K}}= 4 c^{2}(1-c)\) is the mean density of activity and 〈⋅〉 p,q denotes the average on trajectories x(τ)0≤τt without constraint at final time. The normalisation 〈δ(x(t)=0)〉 p,q is e.g. fixed from Z eff(0,t)=1.

Let us denote by P(x,X,t) the probability of being in x at time t, having observed a value X of the area \(\int_{0}^{t} d\tau\,x(\tau)\). The initial condition is P(x,X,0)=δ x,0 δ(X). Defining \(\hat{P}(x,s,t)=\int dX\,e^{-s{\mathbb{K}}X} P(x,X,t)\), one has (using the constraint at final time x(t)=0):

$$Z_{\mathrm{eff}}(s,t) = \frac{\hat{P}(0,s,t)}{\hat{P}(0,0,t)}$$
(44)

Moreover, from Feynman-Kac formula, the time evolution of \(\hat{P}(x,s,t)\) is given by

(45)

and a reflection at x=0. To symmetrise the walk, we now set \(\hat{Q}(x,s,t)=(q/p)^{\frac{x}{2}}\hat{P}(x,s,t)\). One has again \(Z_{\mathrm{eff}}(s,t) = \frac{\hat{Q}(0,s,t)}{\hat{Q}(0,0,t)}\) and the evolution of \(\hat{Q}(x,s,t)\) writes

(46)

The normalisation has changed but we see that apart from the constant loss rate \((2\sqrt{pq}-p-q )\), \(\hat{Q}\) describes a symmetric walk with the same term \(s{\mathbb{K}}x\hat{Q}(x,s,t)\) corresponding to weighting trajectories by the area \(\int_{0}^{t} d\tau\,x(\tau)\). In other words, defining at last \(\tilde{Q}(x,s,t) = \hat{Q}(x,s,t) e^{-t(2\sqrt{pq}-p-q)}\), one has

$$Z_{\mathrm{eff}}(s,t) = \frac{\tilde{Q}(0,s,t)}{\tilde{Q}(0,0,t)}$$
(47)

and from the equation of evolution

(48)

we see that \(\tilde{Q}(0,0,t)\) does not increase exponentially in time since for s=0 the equation describes a simple symmetric random walk. Moreover, in the large size limit (s→0), the evolution of \(\tilde{Q}(x,s,t)\) is governed by the continuous in space operator

$$\sqrt{pq}\partial_x^2 - s{\mathbb{K}}x = 2 \sqrt{pq}\biggl[ \frac{1}{2} \partial_x^2 - \frac{s{\mathbb{K}}}{2\sqrt{pq}} x\biggr]$$
(49)

with reflecting boundary condition in 0. For bridges, the spectrum is known and its largest eigenvalue is given by [23]

$$\psi^{\mbox{\scriptsize{non-per}}}_{\mathrm{eff}}(s)=2\sqrt{pq} \biggl(\frac {s{\mathbb{K}}}{2\sqrt{pq}} \biggr)^{\frac{2}{3}} 2^{-\frac{1}{3}}\alpha_1$$
(50)

where α 1≈2.3381… is the first zero of the Airy function on the negative real axis. In periodic boundary conditions, one has two interfaces and the equivalent jump rates are multiplied by 2. Finally, this gives in the s→0 limit:

$$\psi^{\mathrm{per}}_{\mathrm{eff}}(s)=4\sqrt{pq} \biggl(\frac{s{\mathbb{K}}}{4\sqrt{pq}}\biggr)^{\frac{2}{3}} 2^{-\frac{1}{3}}\alpha_1$$
(51)

Appendix B: A Generic Identity Between Large Deviation Functions

In this appendix we prove an identity used in part 3.2 between large deviation functions associated to the activity K t and to the escape rate R t defined below. We consider a Markov process on a finite number of configurations \(\{\mathcal{C}\}\), with transition rates \(W(\mathcal{C}\to\mathcal{C}')\) between configurations. The activity K t is a history-dependent observable increasing by 1 upon jumping from \(\mathcal{C}\) to \(\mathcal{C}'\). The probability \(P(\mathcal{C},K,t)\) of being in \(\mathcal{C}\) at time t having observed a value K of the observable K t thus evolves in time through

$$\partial_t P(\mathcal{C},K,t) = \sum_{\mathcal{C}'}W\bigl(\mathcal{C}'\to\mathcal{C}\bigr) P\bigl(\mathcal{C}',K-1,t\bigr) - r(\mathcal{C}) P(\mathcal{C},K,t)$$
(52)

with \(r(\mathcal{C})= \sum_{\mathcal{C}'} W(\mathcal{C}\to\mathcal{C}')\) the escape rate from configuration \(\mathcal{C}\). The Laplace transform \(P(\mathcal{C},s,t) =\sum_{K} e^{-sK}P(\mathcal{C},K,t)\) verifies

$$\partial_t P(\mathcal{C},s,t) = \sum_{\mathcal{C}'}e^{-s}W\bigl(\mathcal{C}'\to\mathcal{C}\bigr) P\bigl(\mathcal{C}',s,t\bigr) - r(\mathcal{C}) P(\mathcal{C},s,t)$$
(53)

The cumulant generating function ψ(s) defined in the infinite time limit as \(\langle e^{-sK_{t}}\rangle\sim e^{t\psi(s)}\) is the largest eigenvalue of the operator \(\mathbb{W}_{s}\) of elements

$$(\mathbb{W}_s )_{\mathcal{C}\mathcal{C}'}= e^{-s} W\bigl(\mathcal{C}'\to\mathcal{C}\bigr) - r(\mathcal{C})\delta_{\mathcal{C}\mathcal{C}'} $$
(54)

since \(\langle e^{-sK_{t}}\rangle=\sum_{\mathcal{C}}P(\mathcal{C},s,t) \). Both K t and \(R_{t}=\int_{0}^{t}d\tau\,r(\mathcal{C}(\tau))\) quantify the activity of the histories. Their large deviation functions (ldf) are closely related. Indeed, let’s consider the joint ldf

$$\varPsi(s,\sigma)= \lim_{t\to\infty}\frac{1}{t} \log\bigl\langle e^{-sK_t-\sigma R_t}\bigr\rangle $$
(55)

As previously, one checks that Ψ(s,σ) is given by the maximum eigenvalue of the operator \(\mathbb{W}_{s,\sigma}\) of elements [7]

$$(\mathbb{W}_{s,\sigma} )_{\mathcal{C}\mathcal{C}'}= e^{-s} W\bigl (\mathcal{C}'\to\mathcal{C}\bigr) - (1+\sigma)r(\mathcal{C})\delta_{\mathcal{C}\mathcal{C}'}$$
(56)

It verifies the symmetry \(\mathbb{W}_{s,\sigma} = (1+\sigma)\mathbb{W}_{s+\log(1+\sigma),0}\) and so does the ldf:

$$\varPsi(s,\sigma)= (1+\sigma)\varPsi\bigl(s+\log(1+\sigma),0\bigr) $$
(57)

Besides, the mean values of K t and R t in the s-state are given by

$$\frac{1}{t} \langle K_t\rangle_s = -\partial_s\varPsi(s,\sigma) |_{\sigma=0},\qquad \frac{1}{t} \langle R_t\rangle_s = -\partial_\sigma\varPsi(s,\sigma) |_{\sigma=0}$$
(58)

Differentiating the symmetry (57) with respect to σ and sending σ to 0, one gets

$$\partial_\sigma\varPsi(s,\sigma) |_{\sigma=0} = \varPsi(s,0)+\partial_s\varPsi(s,\sigma) |_{\sigma=0}$$
(59)

which implies

$$\psi(s) = \frac{1}{t} \langle K_t\rangle_s - \frac{1}{t}\langle R_t\rangle_s $$
(60)

This relation is generic. It leads to the relation (14) between scaling exponents for the FA model in the inactive regime s=λ/L.

Appendix C: Bernoulli Approximation to Determine φ L (λ)

In this appendix, we obtain the expression of the Courant-Fischer optimisation principle of part 3.3 for Bernoulli states. Before this, one needs to symmetrise the evolution operator \(\mathbb{W}_{s}\) introduced in (54). We take the notation of Appendix B. Assuming that the jump rates verify the detailed balance symmetry \(W(\mathcal{C}\to\mathcal{C}')P_{\mathrm{eq}}(\mathcal{C})=W(\mathcal{C}'\to\mathcal{C})P_{\mathrm{eq}}(\mathcal{C}')\), it is generically possible to symmetrise the operator of evolution \(\mathbb{W}_{s}\) through the similarity transformation \(\mathbb{W}^{\mathrm{sym}}_{s}\equiv \hat{P}^{-\frac{1}{2}}_{\mathrm{eq}} \mathbb{W}_{s} \hat{P}^{\frac{1}{2}}_{\mathrm{eq}} \), where \(\hat{P}_{\mathrm{eq}}\) is the diagonal operator of elements \(P_{\mathrm{eq}}(\mathcal{C})\). Upon symmetrisation, we have that ψ(s) is also the largest eigenvalue of the symmetric operator \(\mathbb{W}^{\mathrm{sym}}_{s}\) of elements

$$\bigl(\mathbb{W}^{\mathrm{sym}}_s \bigr)_{\mathcal{C}\mathcal{C}'}=e^{-s} \bigl[W\bigl(\mathcal{C}'\to\mathcal{C}\bigr) W\bigl(\mathcal{C}\to\mathcal{C}'\bigr) \bigr]^{\frac{1}{2}} - r(\mathcal{C})\delta_{\mathcal{C}\mathcal{C}'}$$
(61)

It is convenient to represent the operator of evolution in terms of spin \(\frac{1}{2}\) operators σ ± and \(\hat{n}\). On each site i, σ ± is the creation/annihilation operator and \(\hat{n}\) is the counting operator. They are defined by

(62)
(63)

Where |0〉 and |1〉 are the vectors for empty and occupied states. One has

$$\mathbb{W}_s = \sum_{1\leq i\leq L}\bigl[e^{-s} \bigl(c\sigma^+_i+(1-c)\sigma_i^-\bigr)-c(1-\hat{n}_i)-(1-c)\hat{n}_i \bigr](\hat{n}_{i+1}+\hat{n}_{i-1})$$
(64)

Transition rates obey detailed balance with respect to the product Bernoulli measure of uniform density c (conditioned to exclude the fully inactive configuration). The symmetrised operator of evolution writes

$$\mathbb{W}^{\mathrm{sym}}_s = \sum_{1\leq i\leq L}\bigl[e^{-s}\sqrt{c(1-c)}\bigl(\sigma^+_i+\sigma_i^-\bigr)-c(1-\hat{n}_i)-(1-c)\hat{n}_i \bigr](\hat{n}_{i+1}+\hat{n}_{i-1})$$
(65)

Defining the vector |ρ〉 corresponding to the Bernoulli distribution of density \(\sqrt{\rho}\):

$$|\rho\rangle= \sqrt{\rho} |1\rangle+ (1-\sqrt{\rho}\,) |1\rangle $$
(66)

and taking |X〉 in (17) to be the state |ρ 1ρ L cond conditioned to exclude the fully empty configuration:

$$|\rho_1\ldots\rho_L\rangle^{\mathrm{cond}} = \sum _{n_i:\sum_in_i\neq0}\langle n_1\ldots n_L|\rho_1 \ldots\rho_L\rangle |n_1\ldots n_L\rangle $$
(67)

one obtains (18) by direct computation.

Rights and permissions

Reprints and permissions

About this article

Cite this article

Bodineau, T., Lecomte, V. & Toninelli, C. Finite Size Scaling of the Dynamical Free-Energy in a Kinetically Constrained Model. J Stat Phys 147, 1–17 (2012). https://doi.org/10.1007/s10955-012-0458-1

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s10955-012-0458-1

Keywords

Navigation