Skip to main content
Log in

Large deviations conditioned on large deviations I: Markov chain and Langevin equation

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

Abstract

We present a systematic analysis of stochastic processes conditioned on an empirical observable \(Q_T\) defined in a time interval [0, T], for large T. We build our analysis starting with a discrete time Markov chain. Results for a continuous time Markov process and Langevin dynamics are derived as limiting cases. In the large T limit, we show how conditioning on a value of \(Q_T\) modifies the dynamics. For a Langevin dynamics with weak noise and conditioned on \(Q_T\), we introduce large deviation functions and calculate them using either a WKB method or a variational formulation. This allows us, in particular, to calculate the typical trajectory and the fluctuations around this trajectory when conditioned on a certain value of \(Q_T\), for large T.

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

References

  1. Mey, A.S.J.S., Geissler, P.L., Garrahan, J.P.: Rare-event trajectory ensemble analysis reveals metastable dynamical phases in lattice proteins. Phys. Rev. E 89, 032109 (2014)

    Article  ADS  Google Scholar 

  2. Delarue, M., Koehl, P., Orland, H.: Ab initio sampling of transition paths by conditioned Langevin dynamics. J. Chem. Phys. 147, 152703 (2017)

    Article  ADS  Google Scholar 

  3. Dykman, M.I., Mori, E., Ross, J., Hunt, P.M.: Large fluctuations and optimal paths in chemical kinetics. J. Chem. Phys. 100, 5735 (1994)

    Article  ADS  Google Scholar 

  4. Lauri, J., Bouchet, F.: Computation of rare transitions in the barotropic quasi-geostrophic equations. N. J. Phys. 17, 015009 (2015)

    Article  Google Scholar 

  5. 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  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. J. Phys. A 42, 075007 (2009)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  7. Dorlas, T.C., Wedagedera, J.R.: Large deviations and the random energy model. Int. J. Mod. Phys. B 15, 1 (2001)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  8. Varadhan, S.R.S.: Asymptotic probabilities and differential equations. Commun. Pure Appl. Math. 19, 261 (1966)

    Article  MathSciNet  MATH  Google Scholar 

  9. Varadhan, S.R.S.: The large deviation problem for empirical distributions of Markov processes. In: Large Deviations and Applications, p. 33. SIAM (1984). https://doi.org/10.1137/1.9781611970241.ch9

  10. Varadhan, S.R.S.: Large deviations for random walks in a random environment. Commun. Pure Appl. Math. 56, 1222 (2003)

    Article  MathSciNet  MATH  Google Scholar 

  11. Ellis, R.S.: Entropy, Large Deviations, and Statistical Mechanics. Die Grundlehren der mathematischen Wissenschaften in Einzeldarstellungen. Springer, Berlin (1985)

  12. Derrida, B.: Non-equilibrium steady states: fluctuations and large deviations of the density and of the current. J. Stat. Mech. P07023 (2007)

  13. den Hollander, F.: Large Deviations, Fields Institute Monographs. American Mathematical Society, Providence (2008)

    Google Scholar 

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

    Article  ADS  MathSciNet  Google Scholar 

  15. Dembo, A., Zeitouni, O.: Large Deviations Techniques and Applications Stochastic Modelling and Applied Probability. Springer, Berlin (2009)

    MATH  Google Scholar 

  16. Kurchan, J.: Fluctuation theorem for stochastic dynamics. J. Phys. A 31, 3719 (1998)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  17. Gallavotti, G., Cohen, E.G.D.: Dynamical ensembles in nonequilibrium statistical mechanics. Phys. Rev. Lett. 74, 2694 (1995)

    Article  ADS  Google Scholar 

  18. Lebowitz, J.L., Spohn, H.: A Gallavotti–Cohen-type symmetry in the large deviation functional for stochastic dynamics. J. Stat. Phys. 95, 333 (1999)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  19. Freidlin, M.I., Szücs, J., Wentzell, A.D.: Random Perturbations of Dynamical Systems. Grundlehren der mathematischen Wissenschaften. Springer, Berlin (2012)

    Book  Google Scholar 

  20. Graham, R., Tél, T.: Weak-noise limit of Fokker–Planck models and nondifferentiable potentials for dissipative dynamical systems. Phys. Rev. A 31, 1109 (1985)

    Article  ADS  MathSciNet  Google Scholar 

  21. Graham, R.: Statistical theory of instabilities in stationary nonequilibrium systems with applications to lasers and nonlinear optics. In: Springer Tracts in Modern Physics: Ergebnisse der exakten Naturwissenschaftenc, vol. 66, p.1. Springer, Berlin (1973). https://doi.org/10.1007/978-3-662-40468-3_1

  22. Bertini, L., De Sole, A., Gabrielli, D., Jona-Lasinio, G., Landim, C.: Fluctuations in stationary nonequilibrium states of irreversible processes. Phys. Rev. Lett. 87, 040601 (2001)

    Article  ADS  MathSciNet  Google Scholar 

  23. Derrida, B.: Microscopic versus macroscopic approaches to non-equilibrium systems. J. Stat. Mech. 2011, P01030 (2011)

    Article  MathSciNet  Google Scholar 

  24. Bertini, L., De Sole, A., Gabrielli, D., Jona-Lasinio, G., Landim, C.: Macroscopic fluctuation theory. Rev. Mod. Phys. 87, 593 (2015)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  25. Donsker, M.D., Varadhan, S.R.S.: Asymptotic evaluation of certain markov process expectations for large time, I. Commun. Pure Appl. Math. 28, 1 (1975)

    Article  MathSciNet  MATH  Google Scholar 

  26. Derrida, B., Lebowitz, J.L.: Exact large deviation function in the asymmetric exclusion process. Phys. Rev. Lett. 80, 209 (1998)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  27. Bodineau, T., Derrida, B.: Current fluctuations in nonequilibrium diffusive systems: an additivity principle. Phys. Rev. Lett. 92, 180601 (2004)

    Article  ADS  Google Scholar 

  28. Bertini, L., De Sole, A., Gabrielli, D., Jona-Lasinio, G., Landim, C.: Current fluctuations in stochastic lattice gases. Phys. Rev. Lett. 94, 030601 (2005)

    Article  ADS  MATH  Google Scholar 

  29. Hurtado, P.I., Garrido, P.L.: Large fluctuations of the macroscopic current in diffusive systems: a numerical test of the additivity principle. Phys. Rev. E 81, 041102 (2010)

    Article  ADS  Google Scholar 

  30. Hurtado, P.I., Espigares, C.P., del Pozo, J.J., Garrido, P.L.: Thermodynamics of currents in nonequilibrium diffusive systems: theory and simulation. J. Stat. Phys. 154, 214 (2014)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  31. Bertini, L., Faggionato, A., Gabrielli, D.: Large deviations of the empirical flow for continuous time Markov chains. Ann. Inst. H. Poincaré Prob. Stat. 51, 867 (2015)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  32. Touchette, H.: Introduction to dynamical large deviations of Markov processes. In: Lecture Notes of the 14th International Summer School on Fundamental Problems in Statistical Physics. Physica A 504, 5 (2018)

  33. Maes, C., Netocný, K.: Canonical structure of dynamical fluctuations in mesoscopic nonequilibrium steady states. Europhys. Lett. 82, 30003 (2008)

    Article  ADS  MathSciNet  Google Scholar 

  34. Maes, C., Netocnný, K., Wynants, B.: Steady state statistics of driven diffusions. Physica A 387, 2675 (2008)

    Article  ADS  MathSciNet  Google Scholar 

  35. Evans, R.M.L.: Rules for transition rates in nonequilibrium steady states. Phys. Rev. Lett. 92, 150601 (2004)

    Article  ADS  Google Scholar 

  36. Evans, R.M.L.: Detailed balance has a counterpart in non-equilibrium steady states. J. Phys. A 38, 293–313 (2004)

    Article  ADS  MATH  Google Scholar 

  37. Hartmann, C., Schütte, C.: Efficient rare event simulation by optimal nonequilibrium forcing. J. Stat. Mech P11004 (2012)

  38. Majumdar, S.N., Orland, H.: Effective Langevin equations for constrained stochastic processes. J. Stat. Mech P06039 (2015)

  39. Fleming, W.H.: Stochastic control and large deviations. In: Bensoussan, A., Verjus, J.P. (eds.) Future Tendencies in Computer Science, Control and Applied Mathematics, p. 291. Springer, Berlin (1992)

    Chapter  Google Scholar 

  40. Nemoto, T., Sasa, Si: Thermodynamic formula for the cumulant generating function of time-averaged current. Phys. Rev. E 84(6), 061113 (2011)

    Article  ADS  Google Scholar 

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

    Article  ADS  MathSciNet  MATH  Google Scholar 

  42. Strook, D.W.: An Introduction to Markov Processes. Graduate Texts in Mathematics, 2nd edn. Springer, Berlin (2014)

    Book  Google Scholar 

  43. Borkar, V.S., Juneja, S., Kherani, A.A.: Peformance analysis conditioned on rare events: an adaptive simulation scheme. Commun. Inf. Syst. 3, 259–278 (2003)

    MathSciNet  MATH  Google Scholar 

  44. Jack, R.L., Sollich, P.: Large deviations and ensembles of trajectories in stochastic models. Prog. Theor. Phys. Suppl. 184, 304 (2010)

    Article  ADS  MATH  Google Scholar 

  45. Jack, R.L., Sollich, P.: Effective interactions and large deviations in stochastic processes. Eur. Phys. J. Spec. Top. 224, 2351 (2015)

    Article  Google Scholar 

  46. Chetrite, R., Touchette, H.: Nonequilibrium microcanonical and canonical ensembles and their equivalence. Phys. Rev. Lett. 111, 120601 (2013)

    Article  ADS  Google Scholar 

  47. Chetrite, R., Touchette, H.: Nonequilibrium markov processes conditioned on large deviations. Ann. Henri Poincaré 16, 2005 (2015)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  48. Chetrite, R., Touchette, H.: Variational and optimal control representations of conditioned and driven processes. J. Stat. Mech P12001 (2015)

  49. Szavits-Nossan, J., Evans, M.R.: Inequivalence of nonequilibrium path ensembles: the example of stochastic bridges. J. Stat. Mech. P12008 (2015)

  50. Nyawo, P.T., Touchette, H.: Large deviations of the current for driven periodic diffusions. Phys. Rev. E 94(3), 032101 (2016)

    Article  ADS  Google Scholar 

  51. Tizón-Escamilla, N., Lecomte, V., Bertin, E.: Effective driven dynamics for one-dimensional conditioned Langevin processes in the weak-noise limit. J. Stat. Mech. 2019, 013201 (2019)

    Article  MathSciNet  Google Scholar 

  52. Derrida, B., Sadhu, T.: Large deviations conditioned on large deviations II: fluctuating hydrodynamics (2019). arXiv:1905.07175

  53. Landau, L., Lifshitz, E.: Quantum Mechanics. MIR, Moskow (1967)

    MATH  Google Scholar 

  54. Derrida, B., Douçot, B., Roche, P.E.: Current fluctuations in the one-dimensional symmetric exclusion process with open boundaries. J. Stat. Phys. 115, 717 (2004)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  55. Hirschberg, O., Mukamel, D., Schütz, G.M.: Density profiles, dynamics, and condensation in the ZRP conditioned on an atypical current. J. Stat. Mech. P11023 (2015)

  56. Schütz, G.M.: Duality Relations for the Periodic ASEP Conditioned on a Low Current, p. 323. Springer, Cham (2016)

    MATH  Google Scholar 

  57. Popkov, V., Schütz, G.M.: Transition probabilities and dynamic structure function in the ASEP conditioned on strong flux. J. Stat. Phys. 142, 627 (2011)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  58. Carollo, F., Garrahan, J.P., Lesanovsky, I., Pérez-Espigares, C.: Making rare events typical in Markovian open quantum systems. Phys. Rev. A 98, 010103 (2018)

    Article  ADS  Google Scholar 

  59. Angeletti, F., Touchette, H.: Diffusions conditioned on occupation measures. J. Math. Phys. 57 (2016)

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

    MATH  Google Scholar 

  61. Popkov, V., Schütz, G.M., Simon, D.: ASEP on a ring conditioned on enhanced flux. P10007. J. Stat. Mech. (2010)

  62. Ellis, R.S.: Large deviations for a general class of random vectors. Ann. Probab. 12, 1–12 (1984)

    Article  MathSciNet  MATH  Google Scholar 

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

    Article  ADS  MathSciNet  Google Scholar 

  64. Harris, R.J., Rákos, A., Schütz, G.M.: Breakdown of Gallavotti–Cohen symmetry for stochastic dynamics. Eur. Phys. Lett. 75, 227–233 (2006)

    Article  ADS  MathSciNet  Google Scholar 

  65. Espigares, C.P., Garrido, P.L., Hurtado, P.I.: Dynamical phase transition for current statistics in a simple driven diffusive system. Phys. Rev. E 87, 032115 (2013)

    Article  ADS  Google Scholar 

  66. Touchette, H.: Equivalence and nonequivalence of ensembles: thermodynamic, macrostate, and measure levels. J. Stat. Phys. 159, 987–1016 (2015)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  67. McKean, H.P.: Stochastic Integrals. Probability and Mathematical Statistics: A Series of Monographs and Textbooks. Academic Press, Cambridge (1969). https://doi.org/10.1016/B978-1-4832-3054-2.50008-X

    Google Scholar 

  68. Mehl, J., Speck, T., Seifert, U.: Large deviation function for entropy production in driven one-dimensional systems. Phys. Rev. E 78, 011123 (2008)

    Article  ADS  Google Scholar 

  69. Speck, T., Engel, A., Seifert, U.: The large deviation function for entropy production: the optimal trajectory and the role of fluctuations. J. Stat. Mech. P12001 (2012)

  70. Risken, H.: The Fokker–Planck Equation: Methods of Solutions and Applications. Springer Series in Synergetics, 2nd edn. Springer, Berlin (1996)

    MATH  Google Scholar 

  71. Brownstein, K.R.: Criterion for existence of a bound state in one dimension. Am. J. Phys. 68, 160–161 (2000)

    Article  ADS  Google Scholar 

  72. Buell, W.F., Shadwick, B.A.: Potentials and bound states. Am. J. Phys. 63, 256–258 (1995)

    Article  ADS  Google Scholar 

  73. Ashbaugh, M.S., Benguria, R.D.: Optimal bounds for ratios of eigenvalues of one-dimensional Schrödinger operators with Dirichlet boundary conditions and positive potentials. Commun. Math. Phys. 124, 403–415 (1989)

    Article  ADS  MATH  Google Scholar 

  74. Andrews, B., Clutterbuck, J.: Proof fundamental gap conjecture. J. Am. Math. Soc. 24, 899–916 (2011)

    Article  MathSciNet  MATH  Google Scholar 

  75. Nickelsen, D., Engel, A.: Asymptotics of work distributions: the pre-exponential factor. Eur. Phys. J. B 82, 207–218 (2011)

    Article  ADS  Google Scholar 

  76. Engel, A.: Asymptotics of work distributions in nonequilibrium systems. Phys. Rev. E 80, 021120 (2009)

    Article  ADS  Google Scholar 

  77. Baule, A., Touchette, H., Cohen, E.G.D.: Stick-slip motion of solids with dry friction subject to random vibrations and an external field. Nonlinearity 24, 351 (2011)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  78. Sadhu, T., Derrida, B.: Correlations of the density and of the current in non-equilibrium diffusive systems. J. Stat. Mech. 113202 (2016)

  79. Bertini, L., Sole, A.D., Gabrielli, D., Landim, C.: Macroscopic fluctuation theory for stationary non-equilibrium states. J. Stat. Phys. 107, 635 (2002)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  80. Meerson, B., Zilber, P.: Large deviations of a long-time average in the Ehrenfest urn model. J. Stat. Mech. 2018, 119901 (2018)

    Article  Google Scholar 

  81. Proesmans, K., Derrida, B.: Large-deviation theory for a Brownian particle on a ring: a WKB approach. J. Stat. Mech. 2019, 023201 (2019)

    Article  MathSciNet  Google Scholar 

  82. Fischer, L.P., Pietzonka, P., Seifert, U.: Large deviation function for a driven underdamped particle in a periodic potential. Phys. Rev. E 97, 1–10 (2018)

    Article  Google Scholar 

  83. Kubo, R., Matsuo, K., Kitahara, K.: Fluctuation and relaxation of macrovariables. J. Stat. Phys. 9, 51 (1973)

    Article  ADS  Google Scholar 

  84. Schrödinger, E.: Sur la théorie relativiste de l’électron et l’interprétation de la mécanique quantique. Ann. Henri Poincaré 2, 269–310 (1932)

    MathSciNet  MATH  Google Scholar 

  85. Zambrini, J.C.: Euclidean quantum mechanics. Phys. Rev. A 35(9), 3631–3649 (1987)

    Article  ADS  MathSciNet  Google Scholar 

  86. Cruzeiros, A.B., Zambrini J.C.: Euclidean quantum mechanics. An outline. In: Stochastic Analysis and Applications in Physics, pp. 59–97. Springer Netherlands, Dordrecht (1994). https://doi.org/10.1007/978-94-011-0219-3_4

Download references

Acknowledgements

We acknowledge the hospitality of ICTS-Bengaluru, India, where part of the work was completed during a workshop on Large deviation theory in August, 2017.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Tridib Sadhu.

Additional information

Publisher's Note

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

Appendices

Ensemble equivalence

In this appendix we show that, for large T, the equivalence of ensembles holds for an arbitrary time t. For an earlier derivation of the ensemble equivalence see [47].

As the reasoning is very similar in the five regions of Fig. 1, we will limit our discussion to the case of region II, i.e. for \(0\le t \ll T\). Let \(P_t(C_T,C,Q \vert C_0)\) be the joint probability of configuration \(C_T\) at time T, configuration C at time t, and of the observable \(Q_T\) to take value Q given its initial configuration \(C_0\) at time 0.

To establish the equivalence of ensembles in (17), we need to show that the conditioned probability in the microcanonical ensemble

$$\begin{aligned} {\mathcal {P}}_t(C \vert Q=qT)=\frac{\sum _{C_T}\sum _{C_0}P_t(C_T,C,Q=qT\vert C_0)R_{0}(C_0)}{\sum _{C'}\left[ \sum _{C_T}\sum _{C_0}P_t(C_T,C',Q=qT\vert C_0)R_{0}(C_0)\right] } \end{aligned}$$
(75)

and the canonical measure

$$\begin{aligned} P_t^{(\lambda )}(C)=\frac{\sum _{C_T}\sum _{C_0}\int dQ\, e^{\lambda Q}P_t(C_T,C,Q\vert C_0)R_{0}(C_0)}{\sum _{C'}\left[ \sum _{C_T}\sum _{C_0}\int dQ\, e^{\lambda Q}P_t(C_T,C',Q\vert C_0)R_{0}(C_0)\right] } \end{aligned}$$
(76)

converge to the same distribution for large T when \(\lambda \) and q are related by (12).

For this, we write, in terms of the probability (5),

$$\begin{aligned} P_t(C_T,C,Q=qT\vert C_0)=\int dQ_t\;P_{T-t}(C_T,qT-Q_t \vert C)\;P_t(C,Q_t \vert C_0) \end{aligned}$$
(77)

and use the large T asymptotics (13), which gives

$$\begin{aligned} P_{T-t}(C_T,qT-Q_t \vert C)\simeq e^{-(T-t)\phi (q)-(tq-Q_t)\phi '(q)}\sqrt{\frac{\phi ''(q)}{2\pi T}}R_{\phi '(q)}(C_T)L_{\phi '(q)}(C) \end{aligned}$$

Substituting in (75) and simplifying the expression for large T we get the microcanonical probability

$$\begin{aligned} P_t(C \vert Q=qT)\simeq \frac{L_{\phi '(q)}(C)\sum _{C_0}G_t^{(\phi '(q))}(C\vert C_0)\;R_{0}(C_0)}{\sum _{C'} L_{\phi '(q)}(C')\sum _{C_0}G_t^{(\phi '(q))}(C'\vert C_0)\;R_{0}(C_0)} \end{aligned}$$
(78)

where \(G_t^{(\lambda )}(C \vert C_0)\) is defined in (6). On the other hand, using (9) for large T we get the canonical probability

$$\begin{aligned} P_t^{(\lambda )}(C)\simeq \frac{L_{\lambda }(C)\sum _{C_0}G_t^{(\lambda )}(C\vert C_0)\;R_{0}(C_0)}{\sum _{C'} L_{\lambda }(C')\sum _{C_0}G_t^{(\lambda )}(C'\vert C_0)\;R_{0}(C_0)} \end{aligned}$$
(79)

Clearly the two probabilities in the two ensembles coincide for \(\lambda =\phi '(q)\). Replacing \(G_t^{(\lambda )}(C\vert C_0)\) by \(M_{\lambda }^t(C,C_0)\) in (79) leads to the canonical measure (18b).

The same reasoning can be easily adapted in the other regions of Fig. 1.

Continuous time Markov process

In this Appendix, we describe a continuous time limit of the Markov process, illustrated in Fig. 7. In this, the empirical observable analogous to (4) is the \(dt\rightarrow 0\) limit of

$$\begin{aligned} Q_T=dt\sum _{i=0}^{\frac{T}{dt}-1}f(C_i)+\sum _{n}g(C_n^+,C_n^-) \end{aligned}$$
(80)

where \(t=i\, dt\), and \((C_n^-, C_n^+)\) are the configurations before and after the nth jump during the time interval [0, T].

Fig. 7
figure 7

A schematic of a time evolution in a Markov process with discrete time steps dt. The continuous time limit is obtained by taking \(dt\rightarrow 0\) limit

From (7) we get

$$\begin{aligned} G_{T}^{(\lambda )}(C' \vert C_0)=\sum _{C}M_{\lambda }(C',C)\,G_{T-dt}^{(\lambda )}(C\vert C_0) \end{aligned}$$

where

$$\begin{aligned} M_\lambda (C',C)={\left\{ \begin{array}{ll} M_0(C',C)e^{\lambda [dt\, f(C)+g(C',C)]} &{} \text {for }C'\ne C,\\ M_0(C,C)e^{\lambda \, dt \, f(C)} &{} \text {for }C'= C.\end{array}\right. } \end{aligned}$$

Using the construction (27) for \(M_0(C',C)\) we take the continuous time limit \(dt\rightarrow 0\) and get

$$\begin{aligned} \frac{d}{dT} G_T^{(\lambda )}(C'\vert C_0)=\sum _{C}\mathcal {M}_{\lambda }(C',C)G_T^{(\lambda )}(C\vert C_0) \end{aligned}$$
(81)

where we recover an earlier result [6, 47] for the tilted matrix \(\mathcal {M}_{\lambda }\) for the continuous time process, given by

$$\begin{aligned} \mathcal {M}_{\lambda }(C',C)={\left\{ \begin{array}{ll} e^{\lambda g(C',C)}\mathcal {M}_0(C',C) &{} \text {for } C'\ne C,\\ \lambda f(C)-\sum _{C^{''}\ne C}\mathcal {M}_0(C^{''},C) &{} \text {for } C'= C. \end{array}\right. } \end{aligned}$$
(82)

This shows that the generating function is the \((C,C_0)\)th element of \(e^{T\mathcal {M}_{\lambda }}\), i.e.

$$\begin{aligned} G_T^{(\lambda )}(C\vert C_0)=e^{T\mathcal {M}_{\lambda }}(C,C_0) \end{aligned}$$
(83)

Although (81) resembles a Master equation, the tilted matrix \(\mathcal {M}_{\lambda }\) is not a Markov matrix as \(\sum _{C'} \mathcal {M}_{\lambda }(C',C)\) does not necessarily vanish.

For large T, one would get \(G_{T}^{(\lambda )}(C \vert C_0)\simeq e^{T \mu (\lambda )}R_{\lambda }(C)L_{\lambda }(C_0)\) where the cumulant generating function \(\mu (\lambda )\) is the largest eigenvalue of \(\mathcal {M}_\lambda \) with \(L_{\lambda }(C)\) and \(R_\lambda (C)\) being the left and right eigenvectors, respectively. (Note the difference with the discrete time case (9), where \(\mu (\lambda )\) is the logarithm of the largest eigenvalue of the tilted matrix \(M_\lambda \) in (8).)

In a similar construction, one could get the continuous time limit of the canonical measure (18a18d) and its time evolution (20a20d). The analysis is straightforward and we present only the final result.

The time evolution of the canonical measure \(P_t^{(\lambda )}(C)\) for a continuous time Markov process is also a Markov process

$$\begin{aligned} \frac{d}{dt} P_t^{(\lambda )}(C')=\sum _C\mathcal {W}_t^{(\lambda )}(C',C)P_t^{(\lambda )}(C) \end{aligned}$$
(84)

where \(\mathcal {W}_t^{(\lambda )}(C',C)\) is the transition rate from C to \(C'\) at time t in the canonical ensemble. The canonical measure and transition rate have different expressions in the five regions indicated in Fig. 1. Their expression is given below, where we use a matrix product notation, e.g. \([L_\lambda \mathcal {M}_\lambda ](C)\equiv \sum _{C'}L_\lambda (C') \mathcal {M}_\lambda (C',C)\).

  1. 1.

    Region I.

    $$\begin{aligned} P_t^{(\lambda )}(C)=&\frac{[L_{\lambda }e^{-t\mathcal {M}_0}](C)R_{0}(C)}{\sum _{C'}L_{\lambda }(C')R_{0}(C')} \end{aligned}$$
    (85a)
    $$\begin{aligned} \mathcal {W}_t^{(\lambda )} (C',C)=&{[L_\lambda e^{-t\mathcal {M}_0}](C') \over [L_\lambda e^{-t\mathcal {M}_0}](C) }\mathcal {M}_0(C',C) -{\left[ L_\lambda e^{-t\mathcal {M}_0}\mathcal {M}_0\right] (C) \over [L_\lambda e^{-t\mathcal {M}_0}](C) }\;\delta _{C',C} \end{aligned}$$
    (85b)
  2. 2.

    Region II.

    $$\begin{aligned} P_t^{(\lambda )}(C)&=\frac{L_{\lambda }(C)[e^{t\mathcal {M}_\lambda }R_0](C)}{e^{t\mu (\lambda )} \sum _{C'}L_{\lambda }(C')R_{0}(C')} \end{aligned}$$
    (86a)
    $$\begin{aligned} \mathcal {W}_t^{(\lambda )}(C',C)&= { L_\lambda (C') \over L_\lambda (C) }\mathcal {M}_\lambda (C',C)-\mu (\lambda )\delta _{C',C} \end{aligned}$$
    (86b)
  3. 3.

    Region III.

    $$\begin{aligned} P_t^{(\lambda )}(C)&=L_{\lambda }(C)R_\lambda (C) \end{aligned}$$
    (87a)
    $$\begin{aligned} \mathcal {W}_t^{(\lambda )}(C',C)&= { L_\lambda (C') \over L_\lambda (C) }\mathcal {M}_\lambda (C',C) -\mu (\lambda )\delta _{C',C} \end{aligned}$$
    (87b)
  4. 4.

    Region IV.

    $$\begin{aligned} P_t^{(\lambda )}(C)=&\frac{[L_0\, e^{(T-t)\mathcal {M}_\lambda }](C)R_{\lambda }(C)}{e^{(T-t)\mu (\lambda )} \sum _{C'}R_{\lambda }(C')} \end{aligned}$$
    (88a)
    $$\begin{aligned} \mathcal {W}_t^{(\lambda )}(C',C)=&{[L_0 \, e^{(T-t)\mathcal {M}_\lambda }](C') \over [L_0\, e^{(T-t)\mathcal {M}_\lambda }](C) }\mathcal {M}_\lambda (C',C) - { \left[ L_0\, e^{(T-t)\mathcal {M}_\lambda } \mathcal {M}_\lambda \right] (C)\over \left[ L_0\, e^{(T-t)\mathcal {M}_\lambda }\right] (C) } \delta _{C',C} \end{aligned}$$
    (88b)

    where the left eigenvector \(L_0\) for the original (unconditioned) evolution is a unit vector such that \([L_0\, \mathcal {M}_\lambda ](C)\equiv \sum _{C'}\mathcal {M}_\lambda (C',C)\).

  5. 5.

    Region V.

    $$\begin{aligned} P_t^{(\lambda )}(C)=&\,\frac{[e^{(t-T)\mathcal {M}_0}R_{\lambda }](C)}{\sum _{C'}R_{\lambda }(C')} \end{aligned}$$
    (89a)
    $$\begin{aligned} \mathcal {W}_t^{(\lambda )}(C',C)=&\,\mathcal {M}_0(C',C) \end{aligned}$$
    (89b)

These expressions of \(P_{t}^{(\lambda )}\) and \(\mathcal {W}_t^{(\lambda )}\), particularly (86a) and (87), have been derived earlier in [47]. The results for \(\mathcal {W}_t^{(\lambda )}\) can be viewed as a generalization of the Doob’s h-transformation [42, 47].

One can verify the property \(\sum _{C'}\mathcal {W}_t^{(\lambda )}(C',C)=0\) in all five regions. Moreover, setting \(\lambda =0\), and \(L_0(C)=1\), gives \(\mathcal {W}_t^{(0)}(C',C)=\mathcal {M}_0(C',C)\), as one would expect.

Langevin dynamics as a limit of a Markov process

In this appendix, we show how the case of Langevin dynamics in Sect. 3 can be obtained as a continuous limit of the discrete time Markov process in Sect. 2.

Let us consider a jump process on a one-dimensional lattice where a configuration C is given by the site index i as indicated in Fig. 8. Only nearest neighbor jumps are allowed with transition rates that we take of the form

$$\begin{aligned} M_0(i\pm 1,i)=\frac{\epsilon }{2}\pm \frac{a}{2}F(a\,i) \end{aligned}$$
(90)

with \(M_0(i,i)=1-\epsilon \), where a is the unit lattice spacing, \({\epsilon <1}\) is a fixed parameter, and F(x) is an arbitrary function defined on the lattice.

Fig. 8
figure 8

A jump process on a one-dimensional chain where a particle jumps to its nearest neighbour site with rates indicated in the figure

The probability \(P_{t,i}\) of the jump process to be in site i at time t satisfies the Master equation

$$\begin{aligned} P_{t+1,i}=M_0(i,i+1)P_{t,i+1}+M_0(i,i-1)P_{t,i-1}+M_0(i,i)P_{t,i} \end{aligned}$$
(91)

Taking the limit \(a\rightarrow 0\), keeping \(\epsilon \) arbitrary, one can easily see that \(P_{a^2 t}(a\, i)\equiv P_{t,i}\) follows the Fokker–Planck equation (29). This shows that the continuous limit of the jump process is indeed identical to the Langevin dynamics (28).

One can similarly obtain the tilted Langevin dynamics from the continuous limit of the jump process when weighted by \(e^{\lambda Q}\) with the observable Q in (24). For this we define

$$\begin{aligned} f_t(i)=a^2 f(a\, i, a^2 t)\qquad \text {and}\qquad {g_t(j,i)=(j-i)\, a\, \left\{ \alpha \, h(a\, j, a^ 2 t)+(1-\alpha ) \,h(a\, i, a^ 2 t)\right\} } \end{aligned}$$

where \(\alpha \) is the parameter, which specifies the prescription (Îto or Stratonovich) as in (31). Then, the continuous limit of (24) corresponds to an observable Q of the Langevin dynamics

$$\begin{aligned} Q=\int dt\, f(X_t,t)+{\int dX_t\, h(X_t,t)} \end{aligned}$$
(92)

In the expression (25a) for the canonical measure if we define \(H^{(\lambda )}_{a^2 t}(a\, i)\equiv Z_t^{(\lambda )}(i)\) and \({\mathbb {H}}^{(\lambda )}_{a^2 t}(a\, i)\equiv {\mathbb {Z}}_t^{(\lambda )}(i)\), then in the continuous limit \(a\rightarrow 0\) we get the canonical measure for the Langevin dynamics weighted by \(e^{\lambda Q}\) with Q in (92):

$$\begin{aligned} P_t^{(\lambda )} (x)=\frac{ H^{(\lambda )}_t(x)\, {\mathbb {H}}^{(\lambda )}_t(x)}{\int dy\, H^{(\lambda )}_t(y) \, {\mathbb {H}}^{(\lambda )}_t(y)} \end{aligned}$$
(93)

The time evolution of \(H^{(\lambda )}_t(x)\) and \({\mathbb {H}}^{(\lambda )}_t(x)\) are obtained from (25b25c) for the jump process by taking the \(a\rightarrow 0\) limit, keeping \(\epsilon \) fixed. We get

$$\begin{aligned} \frac{d}{dt}H^{(\lambda )}_t (x)&= \lambda f(x,t)H^{(\lambda )}_t(x)-\left( \frac{d}{dx}-\lambda \,h(x,t) \right) F(x)H^{(\lambda )}_t(x) +\frac{\epsilon }{2} \bigg (\frac{d^2}{dx^2}H^{(\lambda )}_t(x)\\&\quad {-}2\lambda \,h(x,t)\frac{d}{dx}H^{(\lambda )}_t(x){-}2(1{-}\alpha )\lambda \, \partial _xh(x,t)H^{(\lambda )}_t(x)+\lambda ^2 h(x,t)^2H^{(\lambda )}_t(x)\bigg ) \end{aligned}$$
(94a)
$$\begin{aligned} -\frac{d}{dt} {\mathbb {H}}^{(\lambda )}_t (x)&= \lambda f(x,t) {\mathbb {H}}^{(\lambda )}_t(x)+F(x)\left( \frac{d}{dx}+\lambda \,h(x,t)\right) {\mathbb {H}}^{(\lambda )}_t(x) + \frac{\epsilon }{2}\bigg (\frac{d^2}{dx^2}{\mathbb {H}}^{(\lambda )}_t(x)\\&\quad +2\lambda \,h(x,t)\frac{d}{dx}{\mathbb {H}}^{(\lambda )}_t(x)+2\alpha \lambda \, \partial _xh(x,t){\mathbb {H}}^{(\lambda )}_t(x) +\lambda ^2 h(x,t)^2{\mathbb {H}}_t(x)\bigg ) \end{aligned}$$
(94b)

Similarly, the continuous limit of (2126) gives the Fokker–Planck equation

$$\begin{aligned} \frac{d}{dt}P_t^{(\lambda )}(x)=-\frac{d}{dx}\left[ F_t^{(\lambda )}(x)P_t^{(\lambda )}(x) \right] +\frac{\epsilon }{2}\frac{d^2}{dx^2}P_t^{(\lambda )}(x) \end{aligned}$$
(95a)

where the modified force

$$\begin{aligned} F_t^{(\lambda )}(x)=F(x)+\epsilon \left( \lambda \,h(x,t)+\frac{d}{dx}\log {\mathbb {H}}_t^{(\lambda )}(x)\right) \end{aligned}$$
(95b)

This gives the time evolution of the Langevin dynamics when it is weighted by the observable (92).

Remarks

  1. 1.

    In the derivation of (95a) we have used that the denominator in (93) is time independent, which can be checked using (94a94b).

  2. 2.

    The Fokker–Planck equation (95a) shows that the effect of biasing a Langevin dynamics by \(e^{\lambda Q}\) with an arbitrary time dependent observable (92) is described by another Langevin dynamics with a modified force (95b), but the noise strength \(\epsilon \) remains unchanged. This works even without a large parameter T (see [38, 49] for earlier examples of conditioned dynamics).

Our results in Sect. 3 belongs to a particular case, where the observable (92) is defined in a large time interval [0, T]. This corresponds to [see (30)]

$$\begin{aligned} f(x,t)={\left\{ \begin{array}{ll} f(x) \text { for } t\in [0,T],\\ 0 \text { otherwise,}\end{array}\right. } \quad \text {and}\qquad h(x,t)={\left\{ \begin{array}{ll} h(x) \text { for }t\in [0,T],\\ 0 \text { otherwise.}\end{array}\right. } \end{aligned}$$

In this case, (94a94b) gives

$$\begin{aligned} \frac{d}{dt}H^{(\lambda )}_t (x)={\mathcal {L}}(t)\cdot H^{(\lambda )}_t (x),\qquad \frac{d}{dt}{\mathbb {H}}^{(\lambda )}_t (x)=-{\mathcal {L}}^\dagger (t)\cdot {\mathbb {H}}^{(\lambda )}_t (x) \end{aligned}$$
(96)

where \({\mathcal {L}}(t)={\mathcal {L}}_\lambda \) for \(t\in [0,T]\) and \({\mathcal {L}}(t)={\mathcal {L}}_0\) outside this time window, with the operators defined in (29) and (33); similar for the conjugate operator \({\mathcal {L}}^\dagger (t)\).

This gives, for example, for \(t\le 0\), \(H^{(\lambda )}_t (x)=r_0(x)\) (defined in (35)), whereas \({\mathbb {H}}^{(\lambda )}_t (x)\sim e^{-t{\mathcal {L}}_0^\dagger }\cdot e^{T{\mathcal {L}}_\lambda ^\dagger }\cdot \ell _0(x)\), (upto a constant pre-factor) which in the large T limit, gives \({\mathbb {H}}^{(\lambda )}_t (x)\sim e^{T\mu (\lambda )}\left[ e^{-t{\mathcal {L}}_0^\dagger }\cdot \ell _\lambda \right] (x)\). Substituting these in (93) and (95b) we get the expression for the canonical measure (38a) and effective force (39a), respectively, in region I of Fig. 1. Results for rest of the regions in Sect. 3.2 can be obtained similarly.

Lastly, from (25b) one could see that for the observable (4), the generating function \(G_T^{(\lambda )}(C\vert C_0)\) in (6) is identical to \(Z_T^{(\lambda )}(C)\) if one sets \(Z_0^{(\lambda )}(C)=\delta _{C,C_0}\). Then from the above calculation it is straightforward to show that in the continuous limit one would get (32).

Path integral formulation

The path integral formulation of a Fokker–Planck equation is standard [83]. The Fokker–Planck equation (29) can be written as

$$\begin{aligned} \frac{dP_t(x)}{dt}=-\frac{d}{dx}\left[ F(x)P_t(x)\right] +\frac{\epsilon }{2}\frac{d^2}{dx^2}P_t(x)\equiv -\mathcal {H}\left( x,-i \frac{d}{dx} \right) P_t(x) \end{aligned}$$

such that \(H(x,p)=F'(x)+iF(x)p+\frac{\epsilon }{2}p^2\). Considering a small increment dt in time, we get

$$\begin{aligned} P_{t+dt}(x)\simeq&\int dx'\left[ 1-dt\,\mathcal {H}\left( x,-i \frac{d}{dx} \right) \right] \delta (x-x') P_t(x') \\\simeq&\int \frac{dp\, dx'}{2\pi }\left[ 1-dt\,\mathcal {H}\left( x,p \right) \right] e^{i\,p(x-x')} P_t(x') \end{aligned}$$

where we used the Fourier transform of the Dirac delta function \(\delta (x-x')\). Iterating the evolution and taking \(dt\rightarrow 0\) limit we get a path integral representation

$$\begin{aligned} P_T(x)=\int _{z(0)=y}^{z(T)=x}\mathcal {D}[z,p]e^{\int _0^Tdt\, [ip{\dot{z}}-H(z,p)]} \end{aligned}$$

with an initial condition \(P_0(z)=\delta (z-y)\). The H(zp) is quadratic in p, and the corresponding path integral can be evaluated exactly, giving

$$\begin{aligned} P_T(x)=\int _{z(0)=y}^{z(T)=x}\mathcal {D}[z]e^{-\frac{1}{2\epsilon }\int _0^Tdt\, ({\dot{z}}-F(z))^2-\int _0^Tdt\,F'(z)} \end{aligned}$$

This is the path integral representation of the Fokker–Planck equation (29).

It is straightforward to generalize the above analysis for the generating function (32) and we get

$$\begin{aligned} G_T^{(\lambda )}(x\vert y)=\int _{z(0)=y}^{z(T)=x}\mathcal {D}[z]e^{{\mathbb {S}}_T^{(\lambda )}[z(t)]} \end{aligned}$$
(97a)

where the Action

$$\begin{aligned} {\mathbb {S}}_T^{(\lambda )}[z]=\int _0^T dt\left[ \lambda f(z)+\lambda {\dot{z}} h(z)-\frac{({\dot{z}}-F(z))^2}{2\epsilon }-F'(z)-\epsilon \lambda \left( \alpha -\frac{1}{2}\right) h'(z)\right] \nonumber \\ \end{aligned}$$
(97b)

Taking small \(\epsilon \) limit, we get \({\mathbb {S}}_T^{(\frac{\kappa }{\epsilon })}[z]\simeq \frac{1}{\epsilon }S_T^{(\kappa )}[z]\) with the latter given in (56) where we used \(h(x)=0\).

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Derrida, B., Sadhu, T. Large deviations conditioned on large deviations I: Markov chain and Langevin equation. J Stat Phys 176, 773–805 (2019). https://doi.org/10.1007/s10955-019-02321-4

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s10955-019-02321-4

Keywords

Navigation