Skip to main content
Log in

Driven Interfaces: From Flow to Creep Through Model Reduction

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

Abstract

The response of spatially extended systems to a force leading their steady state out of equilibrium is strongly affected by the presence of disorder. We focus on the mean velocity induced by a constant force applied on one-dimensional interfaces. In the absence of disorder, the velocity is linear in the force. In the presence of disorder, it is widely admitted, as well as experimentally and numerically verified, that the velocity presents a stretched exponential dependence in the force (the so-called ‘creep law’), which is out of reach of linear response, or more generically of direct perturbative expansions at small force. In dimension one, there is no exact analytical derivation of such a law, even from a theoretical physical point of view. We propose an effective model with two degrees of freedom, constructed from the full spatially extended model, that captures many aspects of the creep phenomenology. It provides a justification of the creep law form of the velocity–force characteristics, in a quasistatic approximation. It allows, moreover, to capture the non-trivial effects of short-range correlations in the disorder, which govern the low-temperature asymptotics. It enables us to establish a phase diagram where the creep law manifests itself in the vicinity of the origin in the force–system-size–temperature coordinates. Conjointly, we characterise the crossover between the creep regime and a linear-response regime that arises due to finite system size.

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.

Institutional subscriptions

Fig. 1
Fig. 2
Fig. 3
Fig. 4
Fig. 5
Fig. 6

Similar content being viewed by others

References

  1. Brazovskii, S., Nattermann, T.: Pinning and sliding of driven elastic systems: from domain walls to charge density waves. Adv. Phys. 53, 177 (2004)

    Article  ADS  Google Scholar 

  2. Kleemann, W.: Universal domain wall dynamics in disordered ferroic materials. Annu. Rev. Mater. Res. 37, 415 (2007)

    Article  ADS  Google Scholar 

  3. Blatter, G., Feigel’man, M.V., Geshkenbein, V.B., Larkin, A.I., Vinokur, V.M.: Vortices in high-temperature superconductors. Rev. Mod. Phys. 66, 1125 (1994)

    Article  ADS  Google Scholar 

  4. Takeuchi, K.A., Sano, M.: Universal fluctuations of growing interfaces: evidence in turbulent liquid crystals. Phys. Rev. Lett. 104, 230601 (2010)

    Article  ADS  Google Scholar 

  5. Takeuchi, K., Sano, M.: Evidence for geometry-dependent universal fluctuations of the Kardar–Parisi–Zhang interfaces in liquid-crystal turbulence. J. Stat. Phys. 147, 853 (2012)

    Article  ADS  MATH  Google Scholar 

  6. Barabási, A.-L., Stanley, H.E.: Fractal Concepts in Surface Growth. Cambridge University Press, Cambridge (1995)

    Book  MATH  Google Scholar 

  7. Kardar, M., Parisi, G., Zhang, Y.-C.: Dynamic scaling of growing interfaces. Phys. Rev. Lett. 56, 889 (1986)

    Article  ADS  MATH  Google Scholar 

  8. Bouchaud, J.-P., Mézard, M., Parisi, G.: Scaling and intermittency in Burgers turbulence. Phys. Rev. E 52, 3656 (1995)

    Article  ADS  MathSciNet  Google Scholar 

  9. Halpin-Healy, T., Zhang, Y.-C.: Kinetic roughening phenomena, stochastic growth, directed polymers and all that. Aspects of multidisciplinary statistical mechanics. Phys. Rep. 254, 215 (1995)

    Article  ADS  Google Scholar 

  10. Corwin, I.: The Kardar–Parisi–Zhang equation and universality class. Random Matrices Theory Appl. 01, 1 (2012)

    Article  MathSciNet  MATH  Google Scholar 

  11. Agoritsas, E., Lecomte, V., Giamarchi, T.: Static fluctuations of a thick one-dimensional interface in the 1 + 1 directed polymer formulation. Phys. Rev. E 87, 042406 (2013)

    Article  ADS  Google Scholar 

  12. Halpin-Healy, T., Takeuchi, K.A.: A KPZ cocktail-shaken, not stirred. J. Stat. Phys. 160, 794 (2015)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  13. Ohm, G.S.: Die galvanische Kette: mathematisch bearbeitet. von Dr. G. S. Ohm, Berlin (1827)

    Book  MATH  Google Scholar 

  14. Ioffe, L.B., Vinokur, V.M.: Dynamics of interfaces and dislocations in disordered media. J. Phys. C 20, 6149 (1987)

    Article  ADS  Google Scholar 

  15. Nattermann, T.: Interface roughening in systems with quenched random impurities. Europhys. Lett. 4, 1241 (1987)

    Article  ADS  Google Scholar 

  16. Feigel’man, M.V., Geshkenbein, V.B., Larkin, A.I., Vinokur, V.M.: Theory of collective flux creep. Phys. Rev. Lett. 63, 2303 (1989)

    Article  ADS  Google Scholar 

  17. Feigel’man, M.V., Vinokur, V.M.: Thermal fluctuations of vortex lines, pinning, and creep in high-\(T_c\) superconductors. Phys. Rev. B 41, 8986 (1990)

    Article  ADS  Google Scholar 

  18. Nattermann, T.: Scaling approach to pinning: charge density waves and giant flux creep in superconductors. Phys. Rev. Lett. 64, 2454 (1990)

    Article  ADS  Google Scholar 

  19. Huse, D.A., Henley, C.L., Fisher, D.S.: Huse, Henley, and Fisher respond. Phys. Rev. Lett. 55, 2924 (1985)

    Article  ADS  Google Scholar 

  20. Narayan, O., Fisher, D.S.: Threshold critical dynamics of driven interfaces in random media. Phys. Rev. B 48, 7030 (1993)

    Article  ADS  Google Scholar 

  21. Chauve, P., Giamarchi, T., Le Doussal, P.: Creep via dynamical functional renormalization group. Europhys. Lett. 44, 110 (1998)

    Article  ADS  Google Scholar 

  22. Chauve, P., Giamarchi, T., Le Doussal, P.: Creep and depinning in disordered media. Phys. Rev. B 62, 6241 (2000)

    Article  ADS  Google Scholar 

  23. Drossel, B., Kardar, M.: Scaling of energy barriers for flux lines and other random systems. Phys. Rev. E 52, 4841 (1995)

    Article  ADS  Google Scholar 

  24. Vinokur, V.M., Cristina Marchetti, M., Chen, L.-W.: Glassy motion of elastic manifolds. Phys. Rev. Lett. 77, 1845 (1996)

    Article  ADS  Google Scholar 

  25. Vinokur, V.M.: Glassy dynamics of driven elastic manifolds. Physica D 107, 411 (1997)

    Article  ADS  Google Scholar 

  26. Roters, L., Lübeck, S., Usadel, K.D.: Creep motion in a random-field Ising model. Phys. Rev. E 63, 026113 (2001)

    Article  ADS  MATH  Google Scholar 

  27. Kolton, A.B., Rosso, A., Giamarchi, T.: Creep motion of an elastic string in a random potential. Phys. Rev. Lett. 94, 047002 (2005)

    Article  ADS  Google Scholar 

  28. Giamarchi, T., Kolton, A.B., Rosso, A.: Dynamics of disordered elastic systems. Lect. Notes Phys. 688, 91 (2006)

    Article  ADS  Google Scholar 

  29. Kolton, A.B., Rosso, A., Giamarchi, T., Krauth, W.: Creep dynamics of elastic manifolds via exact transition pathways. Phys. Rev. B 79, 184207 (2009)

    Article  ADS  Google Scholar 

  30. Ferrero, E.E., Bustingorry, S., Kolton, A.B., Rosso, A.: Numerical approaches on driven elastic interfaces in random media. C. R. Phys. 14, 641 (2013) (disordered systems/Systèmes désordonnés)

  31. Lemerle, S., Ferré, J., Chappert, C., Mathet, V., Giamarchi, T., Le Doussal, P.: Domain wall creep in an Ising ultrathin magnetic film. Phys. Rev. Lett. 80, 849 (1998)

    Article  ADS  Google Scholar 

  32. Le Doussal, P., Vinokur, V.M.: Creep in one dimension and phenomenological theory of glass dynamics. Physica C 254, 63 (1995)

    Article  ADS  Google Scholar 

  33. Scheidl, S.: Mobility in a one-dimensional disorder potential. Z. Phys. B 97, 345 (1995)

    Article  ADS  Google Scholar 

  34. Ferrero, E.E., Foini, L., Giamarchi, T., Kolton, A.B., Rosso, A.: Spatio-temporal patterns in ultra-slow domain wall creep dynamics. arXiv:1604.03726 [cond-mat.dis-nn] (2016). Accessed 5 July 2016

  35. Kim, K.-J., Lee, J.-C., Ahn, S.-M., Lee, K.-S., Lee, C.-W., Cho, Y.J., Seo, S., Shin, K.-H., Choe, S.-B., Lee, H.-W.: Interdimensional universality of dynamic interfaces. Nature 458, 740 (2009)

    Article  ADS  Google Scholar 

  36. Nattermann, T., Renz, W.: Interface roughening due to random impurities at low temperatures. Phys. Rev. B 38, 5184 (1988)

    Article  ADS  Google Scholar 

  37. Agoritsas, E., Lecomte, V., Giamarchi, T.: Temperature-induced crossovers in the static roughness of a one-dimensional interface. Phys. Rev. B 82, 184207 (2010)

    Article  ADS  Google Scholar 

  38. Agoritsas, E., Lecomte, V., Giamarchi, T.: Disordered elastic systems and one-dimensional interfaces. Physica B 407, 1725 (2012)

    Article  ADS  Google Scholar 

  39. Agoritsas, E.: Temperature-dependence of a 1D interface fluctuations: role of a finite disorder correlation length. PhD Thesis, University of Geneva (2013)

  40. Canet, L., Chaté, H., Delamotte, B., Wschebor, N.: Nonperturbative renormalization group for the Kardar–Parisi–Zhang equation: general framework and first applications. Phys. Rev. E 84, 061128 (2011)

    Article  ADS  Google Scholar 

  41. Amir, G., Corwin, I., Quastel, J.: Probability distribution of the free energy of the continuum directed random polymer in 1 + 1 dimensions. Commun. Pure Appl. Math. 64, 466 (2011)

    Article  MathSciNet  MATH  Google Scholar 

  42. Quastel, J.: Introduction to KPZ. In: Current Developments in Mathematics, vol. 2011. International Press, Somerville, MA (2011)

  43. Gorokhov, D.A., Blatter, G.: Diffusion and creep of a particle in a random potential. Phys. Rev. B 58, 213 (1998)

    Article  ADS  Google Scholar 

  44. Gardiner, C.W.: Handbook of Stochastic Methods for Physics, Chemistry, and the Natural Sciences. Number 13 in Springer Series in Synergetics, 2nd edn. Springer, Berlin (1994)

  45. Risken, H.: The Fokker–Planck Equation: Methods of Solution and Applications. Number v. 18 in Springer Series in Synergetics, 2nd edn. Springer, New York (1996)

  46. Hänggi, P., Talkner, P., Borkovec, M.: Reaction-rate theory: fifty years after Kramers. Rev. Mod. Phys. 62, 251 (1990)

    Article  ADS  MathSciNet  Google Scholar 

  47. Pontryagin, L., Andronov, A., Vitt, A.: On the statistical investigation of dynamic systems. Zh. Eksp. Teor. Fiz. 3, 165 (1933)

    Google Scholar 

  48. Bouchet, F., Reygner, J.: Generalisation of the Eyring–Kramers transition rate formula to irreversible diffusion processes. Ann. Henri Poincaré (2016). doi:10.1007/s00023-016-0507-4

  49. Freidlin, M.I., Wentzell, A.D., Wentzell, A.D.: Random Perturbations of Dynamical Systems. Number 260 in Grundlehren der mathematischen Wissenschaften, 2nd edn. Springer, New York (1998)

  50. van Kampen, N.G.: Stochastic Processes in Physics and Chemistry. North-Holland Personal Library, 3rd edn. Elsevier, Amsterdam (2007)

  51. Agoritsas, E., Bustingorry, S., Lecomte, V., Schehr, G., Giamarchi, T.: Finite-temperature and finite-time scaling of the directed polymer free energy with respect to its geometrical fluctuations. Phys. Rev. E 86, 031144 (2012)

    Article  ADS  Google Scholar 

  52. Agoritsas, E., Lecomte, V., Giamarchi, T.: Static fluctuations of a thick one-dimensional interface in the 1 + 1 directed polymer formulation: numerical study. Phys. Rev. E 87, 062405 (2013)

    Article  ADS  Google Scholar 

  53. Kolton, A.B., Bustingorry, S., Ferrero, E.E., Rosso, A.: Uniqueness of the thermodynamic limit for driven disordered elastic interfaces. J. Stat. Mech. 2013, P12004 (2013)

    Article  MathSciNet  Google Scholar 

  54. Malinin, S.V., Chernyak, V.Y.: Transition times in the low-noise limit of stochastic dynamics. J. Chem. Phys. 132, 014504 (2010)

    Article  ADS  Google Scholar 

  55. Anderson, P.W., Kim, Y.B.: Hard superconductivity: theory of the motion of Abrikosov flux lines. Rev. Mod. Phys. 36, 39 (1964)

    Article  ADS  Google Scholar 

  56. Leliaert, J., Van de Wiele, B., Vansteenkiste, A., Laurson, L., Durin, G., Dupré, L., Van Waeyenberge, B.: Creep turns linear in narrow ferromagnetic nanostrips. Sci. Rep. 6, 20472 (2016)

    Article  ADS  Google Scholar 

  57. Kes, P.H., Aarts, J., van den Berg, J., van der Beek, C.J., Mydosh, J.A.: Thermally assisted flux flow at small driving forces. Supercond. Sci. Technol. 1, 242 (1989)

    Article  ADS  Google Scholar 

  58. Larkin, A.I., Ovchinnikov, YuN: Pinning in type II superconductors. J. Low Temp. Phys. 34, 409 (1979)

    Article  ADS  Google Scholar 

  59. Bolech, C.J., Rosso, A.: Universal statistics of the critical depinning force of elastic systems in random media. Phys. Rev. Lett. 93, 125701 (2004)

    Article  ADS  Google Scholar 

  60. Jeudy, V., Mougin, A., Bustingorry, S., Torres, W.S., Gorchon, J., Kolton, A., Lemaître, A., Jamet, J.-P.: Universal pinning energy barrier for driven domain walls in thin ferromagnetic films. arXiv:1603.01674 [cond-mat] (2016). Accessed 5 March 2016

  61. Bustingorry, S., Kolton, A.B., Giamarchi, T.: Thermal rounding of the depinning transition in ultrathin Pt/Co/Pt films. Phys. Rev. B 85, 214416 (2012)

    Article  ADS  Google Scholar 

  62. Gorchon, J., Bustingorry, S., Ferré, J., Jeudy, V., Kolton, A., Giamarchi, T.: Pinning-dependent field-driven domain wall dynamics and thermal scaling in an ultrathin Pt/Co/Pt magnetic film. Phys. Rev. Lett. 113, 027205 (2014)

  63. Balents, L., Bouchaud, J.-P., Mézard, M.: The large scale energy landscape of randomly pinned objects. J. Phys. I 6, 1007 (1996)

    Google Scholar 

  64. Calabrese, P., Le Doussal, P., Rosso, A.: Free-energy distribution of the directed polymer at high temperature. EPL 90, 20002 (2010)

    Article  ADS  Google Scholar 

  65. Dotsenko, V.: Bethe Ansatz derivation of the Tracy–Widom distribution for one-dimensional directed polymers. EPL 90, 20003 (2010)

    Article  ADS  Google Scholar 

  66. Tanguy, A., Vettorel, T.: From weak to strong pinning I: a finite size study. Eur. Phys. J. B 38, 71 (2004)

    Article  ADS  Google Scholar 

  67. Patinet, S., Vandembroucq, D., Roux, S.: Quantitative prediction of effective toughness at random heterogeneous interfaces. Phys. Rev. Lett. 110, 165507 (2013)

    Article  ADS  Google Scholar 

  68. Giamarchi, T., Le Doussal, P.: Chapter 12: statics and dynamics of disordered elastic systems. In: Young, A.P. (ed.) Spin Glasses and Random Fields. World Scientific, Singapore (1997)

    Google Scholar 

  69. Milnor, J.W.: Morse Theory. Number 51 in Annals of Mathematics Studies, 5th printing edition. Princeton Univ. Press, Princeton (1973)

  70. Tănase-Nicola, S., Kurchan, J.: Metastable states, transitions, basins and borders at finite temperatures. J. Stat. Phys. 116, 1201 (2004)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  71. Shapira, N., Lamhot, Y., Shpielberg, O., Kafri, Y., Ramshaw, B.J., Bonn, D.A., Liang, R., Hardy, W.N., Auslaender, O.M.: Disorder-induced power-law response of a superconducting vortex on a plane. Phys. Rev. B 92, 100501 (2015)

    Article  ADS  Google Scholar 

  72. Aragón, L.E., Kolton, A.B., Le Doussal, P., Wiese, K.J., Jagla, E.A.: Avalanches in tip-driven interfaces in random media. EPL 113, 10002 (2016)

  73. Brox, Th: A one-dimensional diffusion process in a Wiener medium. Ann. Probab. 14, 1206 (1986)

    Article  MathSciNet  MATH  Google Scholar 

  74. Halpin-Healy, T.: Extremal paths, the stochastic heat equation, and the three-dimensional Kardar-Parisi-Zhang universality class. Phys. Rev. E 88, 042118 (2013)

    Article  ADS  Google Scholar 

  75. Halpin-Healy, T.: (2 + 1)-Dimensional directed polymer in a random medium: scaling phenomena and universal distributions. Phys. Rev. Lett. 109, 170602 (2012)

    Article  ADS  Google Scholar 

  76. Le Doussal, P., Cristina Marchetti, M., Wiese, K.J.: Depinning in a two-layer model of plastic flow. Phys. Rev. B 78, 224201 (2008)

    Article  ADS  Google Scholar 

  77. Lecomte, V., Barnes, S.E., Eckmann, J.-P., Giamarchi, T.: Depinning of domain walls with an internal degree of freedom. Phys. Rev. B 80, 054413 (2009)

    Article  ADS  Google Scholar 

  78. Barnes, S.E., Eckmann, J.-P., Giamarchi, T., Lecomte, V.: Noise and topology in driven systems—an application to interface dynamics. Nonlinearity 25, 1427 (2012)

    Article  ADS  MathSciNet  MATH  Google Scholar 

Download references

Acknowledgments

We thank Thierry Giamarchi for fruitful discussions, especially in the early stages of this work. E. A. acknowledges financial support by a Fellowship for Prospective Researchers Grant No. P2GEP2-15586 from the Swiss National Science Foundation. V. L. thanks the Montbrun-les-Bains Centre for Theoretical Physics for its warm hospitality. E. A. and V. L. acknowledge support by the National Science Foundation under Grant No. NSF PHY11-25915 during a stay at KITP, UCSB where part of this research was performed. R. G. G. acknowledges financial support from Labex LaSIPS (No. ANR-10-LABX-0040-LaSIPS), managed by the French National Research Agency under the “Investissements d’avenir” Program (No. ANR-11-IDEX-0003-02).

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Vivien Lecomte.

Additional information

Elisabeth Agoritsas and Reinaldo García-García have equally contributed to this work.

Appendices

Appendix 1: Statistical Tilt Symmetry at \(f\ne 0\)

We derive in this appendix a form of the STS, which allows to decompose the point-to-point free energy into the sum of a deterministic contribution and of disorder-dependent one, which, in distribution, is invariant by translation along direction y. We first recall the known result at zero force before deriving a novel STS at \(f\ne 0.\)

1.1 A Reminder: The Zero Force STS (\(f=0\))

In the zero force situation, the linear change of coordinates \(y(t)= \tilde{y}(t) + y_{\text {i}}+ \frac{y_{\text {f}}-y_{\text {i}}}{t_{\text {f}}} t\) allows to relate the weight of trajectories starting in \((0,\,y_{\text {i}})\) and arriving in \((t_{\text {f}},\,y_{\text {f}})\) (see Fig.2) to those starting in (0, 0) and arriving in \((t_{\text {f}},\,0)\) in a translated disorder. One directly reads from (15) (with \(f=0\)) that

$$\begin{aligned} W_V\left( t_{\text {f}},\,y_{\text {f}}|0,\,y_{\text {i}}\right) = \text {e}^{-\frac{c}{T} \frac{(y_{\text {f}}-y_{\text {i}})^2}{2 t_{\text {f}}}} W_{\mathcal T_{y_{\text {f,i}}}^{t_{\text {f}}} V}\left( t_{\text {f}},\,0|0,\,0\right) , \end{aligned}$$
(A.77)

where the translated disorder \(\mathcal T_{y_{\text {f,i}}}^{t_{\text {f}}} V\) is defined as

$$\begin{aligned} \mathcal T_{y_{\text {f,i}}}^{t_{\text {f}}} V(t,\,\tilde{y})\equiv V\big (t,\,\tilde{y}+y_{\text {i}}+\tfrac{y_{\text {f}}-y_{\text {i}}}{t_{\text {f}}}t\big ). \end{aligned}$$
(A.78)

This remark enables to decompose the free energy \(F_V(t_{\text {f}},\,y_{\text {f}}|0,\,y_{\text {i}})={-}T\log W_V(t_{\text {f}},\,y_{\text {f}}|0,\,y_{\text {i}})\) as

$$\begin{aligned} F_V\left( t_{\text {f}},\,y_{\text {f}}|0,\,y_{\text {i}}\right) = F_{V\equiv 0}\left( t_{\text {f}},\,y_{\text {f}}-y_{\text {i}}\right) + \bar{F}_V\left( t_{\text {f}},\,y_{\text {f}}|0,\,y_{\text {i}}\right) , \end{aligned}$$
(A.79)

where

$$\begin{aligned} F_{V\equiv 0}\left( t_{\text {f}},\,y_{\text {f}}-y_{\text {i}}\right)&= F_{\text {th}}\left( t_{\text {f}},\,y_{\text {f}}-y_{\text {i}}\right) + \frac{T}{2} \log \frac{2\pi Tt}{c}, \end{aligned}$$
(A.80)
$$\begin{aligned} F_{\text {th}}\left( t_{\text {f}},\,y_{\text {f}}-y_{\text {i}}\right)&= c \frac{(y_{\text {f}}-y_{\text {i}})^2}{2 t_{\text {f}}}. \end{aligned}$$
(A.81)

Hence, \(\bar{F}_V(t_{\text {f}},\,y_{\text {f}}|0,\,y_{\text {i}})\) is invariant by translation in distribution, as we indeed observe that it is depending on \(y_{\text {i}}\) and \(y_{\text {f}}\) only through \(\mathcal T_{y_{\text {f,i}}}^{t_{\text {f}}}V.\)

$$\begin{aligned} \bar{F}_V\left( t_{\text {f}},\,y_{\text {f}}|0,\,y_{\text {i}}\right) = {-}T \log W_{\mathcal T_{y_{\text {f,i}}}^{t_{\text {f}}} V}\left( t_{\text {f}},\,0|0,\,0\right) . \end{aligned}$$
(A.82)

1.2 The Non-zero Force STS (\(f\ne 0\))

In the non-zero force situation, the key observation is that there exists a f-dependent transformation of the directed polymer trajectories

$$\begin{aligned} y(t)= \tilde{y}(t) + y_{\text {i}}+ \frac{y_{\text {f}}-y_{\text {i}}}{t_{\text {f}}} t + \frac{f}{2c}t\left( t_{\text {f}}-t\right) , \end{aligned}$$
(A.83)

which, remarkably, implies a f-STS generalising (A.79):

$$\begin{aligned} W^f_V\left( t_{\text {f}},\,y_{\text {f}}|0,\,y_{\text {i}}\right) = \exp \Big \{ -\frac{1}{T} \Big [ c \frac{(y_{\text {f}}-y_{\text {i}})^2}{2 t_{\text {f}}}-\frac{f}{2} t_{\text {f}}\left( y_{\text {f}}+y_{\text {i}}\right) -\frac{f^2}{24 c}t_{\text {f}}^3 \Big ] \Big \} W_{\mathcal T_{y_{\text {f,i}}}^{t_{\text {f}},f} V}\left( t_{\text {f}},\,0|0,\,0\right) . \end{aligned}$$
(A.84)

The result is obtained by coming back to the path-integral definition of the f-dependent point-to-point partition function (15), and taking care of the boundary terms. The key observation in the computation is to recognise the following total derivative

$$\begin{aligned} \frac{ (2 c (y_{\text {f}}-y_{\text {i}})+f t_{\text {f}}(t_{\text {f}}-2 t))}{2 t_{\text {f}}}\partial _t \tilde{y}(t)-f \tilde{y}(t) = \partial _t\Big [ \frac{f t_{\text {f}}( t_{\text {f}}- 2 t) + 2 c (y_{\text {f}}- y_{\text {i}}) }{2 t_{\text {f}}} \tilde{y}(t) \Big ], \end{aligned}$$
(A.85)

for the terms linear in \(\tilde{y}(t)\) after the change of variable (A.84). The translated disorder \(\mathcal T_{y_{\text {f,i}}}^{t_{\text {f}},f} V^{\!f}\!\) is defined as

$$\begin{aligned} \mathcal T_{y_{\text {f,i}}}^{t_{\text {f}},f} V(t,\,\tilde{y})\equiv V\big (t,\,\tilde{y}+y_{\text {i}}+\tfrac{y_{\text {f}}-y_{\text {i}}}{t_{\text {f}}}t+ \tfrac{f}{2c}t\left( t_{\text {f}}-t\right) \big ). \end{aligned}$$
(A.86)

Note importantly that if the translated disorder (A.86) depends on the force through a f-dependent change of coordinate, the partition function on the right hand side of (A.84) is however the one at zero force. This enables to decompose the free energy \(F_V^f(t,\,y)\) as

$$\begin{aligned} F_V^f\left( t_{\text {f}},\,y_{\text {f}}|0,\,y_{\text {i}}\right)= & {} F_{\text {th}}\left( t_{\text {f}},\,y_{\text {f}}-y_{\text {i}}\right) -\frac{ft_{\text {f}}}{2} \left( y_{\text {f}}+y_{\text {i}}\right) \nonumber \\&-\,\frac{f^2}{24 c}t_{\text {f}}^3 + \bar{F}^f_V\left( t_{\text {f}},\,y_{\text {f}}|0,\,y_{\text {i}}\right) + \frac{T}{2} \log \frac{2\pi Tt}{c}, \end{aligned}$$
(A.87)

where

$$\begin{aligned} \bar{F}^f_V\left( t_{\text {f}},\,y_{\text {f}}|0,\,y_{\text {i}}\right) = {-}T \log W_{\mathcal T_{y_{\text {f,i}}}^{t_{\text {f}},f} V}\left( t_{\text {f}},\,0|0,\,0\right) . \end{aligned}$$
(A.88)

For uncorrelated disorder (\(\xi =0\)) this implies in particular that at large \(t_{\text {f}}\) the distribution of \(\bar{F}^f_V(t_{\text {f}},\,y_{\text {f}}|0,\,y_{\text {i}})\) adopts the Airy\(_2\) scaling [10] and goes to the distribution of a Brownian motion of coordinate \(y_{\text {f}}-y_{\text {i}}\) at infinite \(t_{\text {f}}\) (this corresponds to the steady state of the KPZ equation [19]). For correlated disorder (\(\xi >0\)), the picture is slightly changed at large \(t_{\text {f}}{\text {:}}\) the steady state is not distributed as a Brownian anymore but scales similarly (as long as \(y_{\text {f}}-y_{\text {i}}\gg \xi \)) with a different \(\xi \)-dependent amplitude \(\tilde{D}\) [11, 52] [see Eq. (42) in the main text].

Appendix 2: Effective Temperature \(\tilde{T}\) and Friction \(\tilde{\gamma }\) in the Absence of Disorder

In this Appendix, we determine the relation between the original friction \(\gamma \) and temperature T of the interface dynamics (1) and the ones \(\tilde{\gamma }\) and \(\tilde{T}\) of the effective model (3334), in the absence of disorder (\(V\equiv 0\)). Starting by the original dynamics, one averages (1) over thermal fluctuations and taking the long-time limit, one gets

$$\begin{aligned} \overline{v}(f)\big |_{V\equiv 0} = \frac{f}{\gamma }. \end{aligned}$$
(B.89)

Changing to the reference frame of the centre of mass, one recognises that \(y(t,\,\tau )-\frac{f}{\gamma }\tau \) obeys the Edwards–Wilkinson equation, whose steady state is Gaussian, implying that

$$\begin{aligned} \big \langle \left[ y\left( t_{\text {f}}\right) -y(0)\right] ^2\big \rangle \big |_{V\equiv 0} = \frac{T}{c}t_{\text {f}}. \end{aligned}$$
(B.90)

On the other hand, the effective equations (3334) are written as follows in the absence of disorder, as seen from the expression (A.87) of the tilted free energy:

$$\begin{aligned} \tilde{\gamma } \partial _\tau y_{\text {i}}(\tau )&= {+}\,c\frac{y_{\text {f}}-y_{\text {i}}}{t_{\text {f}}}+\frac{1}{2} ft_{\text {f}}+ \sqrt{2\tilde{\gamma }\tilde{T}}\tilde{\eta }_\text {i}(\tau ), \end{aligned}$$
(B.91)
$$\begin{aligned} \tilde{\gamma } \partial _\tau y_{\text {f}}(\tau )&= {-}c\frac{y_{\text {f}}-y_{\text {i}}}{t_{\text {f}}}+\frac{1}{2} ft_{\text {f}}+ \sqrt{2\tilde{\gamma }\tilde{T}}\tilde{\eta }_\text {f}(\tau ). \end{aligned}$$
(B.92)

Summing these equations, averaging over thermal noise and imposing that both \(y_{\text {i}}(\tau )\) and \(y_{\text {i}}(\tau )\) move at an average velocity \(\overline{v}(f)|_{V\equiv 0}\) at large times, one gets \(2\tilde{\gamma }\overline{v}(f)|_{V\equiv 0} = ft_{\text {f}};\) hence, comparing to (B.89), one obtains

$$\begin{aligned} \tilde{\gamma } = \frac{1}{2} t_{\text {f}}\gamma . \end{aligned}$$
(B.93)

Note that to get this result, we relaxed the conditions (walls) that would enforce the model to reach a zero-velocity equilibrium steady state at large times. However, a MFPT analysis in such conditions would also yield the result (B.93) for the velocity in a finite window of the system far from the walls. Subtracting now (B.91) and (B.92) one obtains a closed equation for \(\delta \!y(\tau )=y_{\text {f}}(\tau )-y_{\text {i}}(\tau )\) in the absence of disorder:

$$\begin{aligned} \tilde{\gamma }\partial _\tau \delta \!y(\tau ) = {-}2\frac{c}{t_{\text {f}}}\delta \!y(\tau ) + \sqrt{4\tilde{\gamma }\tilde{T}}\tilde{\eta }_1, \end{aligned}$$
(B.94)

where \(\tilde{\eta }_1\) is a white noise of unit variance. The force term of this Langevin equation derives from the energy \(\frac{c}{t_{\text {f}}}(\delta \!y)^2\) and the noise term has temperature \(2\tilde{T}.\) This shows that the steady-state distribution of \(\delta \!y\) is Gaussian \(\propto \exp \left[ -\frac{c}{2\tilde{T} t_{\text {f}}}(\delta \!y)^2\right] .\) This implies in turn that \( \langle (\delta \!y)^2\rangle |_{V\equiv 0}=\langle [y_{\text {f}}-y_{\text {i}}]^2\rangle |_{V\equiv 0}=\frac{\tilde{T}}{c}t_{\text {f}}.\) Finally, comparing to (B.90) one obtains that the temperatures of the original and of the effective models are equal:

$$\begin{aligned} \tilde{T} = T. \end{aligned}$$
(B.95)

To summarise, in the absence of disorder, the effective model with friction (B.93) and temperature (B.95) presents the same velocity and the same Gaussian end-point distribution as the original dynamics. We assume that this correspondence between original and effective parameters also holds in the presence of disorder.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Agoritsas, E., García-García, R., Lecomte, V. et al. Driven Interfaces: From Flow to Creep Through Model Reduction. J Stat Phys 164, 1394–1428 (2016). https://doi.org/10.1007/s10955-016-1588-7

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s10955-016-1588-7

Keywords

Navigation