Skip to main content
Log in

Effective Hamiltonians and Lagrangians for Conditioned Markov Processes at Large Volume

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

Abstract

When analysing statistical systems or stochastic processes, it is often interesting to ask how they behave given that some observable takes some prescribed value. This conditioning problem is well understood within the linear operator formalism based on rate matrices or Fokker–Planck operators, which describes the dynamics of many independent random walkers. Relying on certain spectral properties of the biased linear operators, guaranteed by the Perron–Frobenius theorem, an effective process can be found such that its path probability is equivalent to the conditional path probability. In this paper, we extend those results for nonlinear Markov processes that appear when the many random walkers are no longer independent, and which can be described naturally through a Lagrangian–Hamiltonian formalism within the theory of large deviations at large volume. We identify the appropriate spectral problem as being a Hamilton–Jacobi equation for a biased Hamiltonian, for which we conjecture that two special global solutions exist, replacing the Perron–Frobenius theorem concerning the positivity of the (left and right) dominant eigenvectors. We then devise a rectification procedure based on a canonical gauge transformation of the biased Hamiltonian, yielding an effective dynamics in agreement with the original conditioning. Along the way, we present simple examples in support of our conjecture, we examine its consequences on important physical objects such as the fluctuation symmetries of the biased and rectified processes as well as the dual dynamics obtained through time-reversal. We apply all those results to simple independent and interacting models, including a stochastic chemical reaction network and a population process called the Brownian Donkey.

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
Fig. 8
Fig. 9
Fig. 10
Fig. 11
Fig. 12

Similar content being viewed by others

Notes

  1. We use an over-tilde on the generator of the single process dynamics in order to distinguish it from the many-body process dynamics on which we focus in the remaining of this paper.

  2. A manifold is informally defined as a geometrical space generalizing the notion of curve or surface to arbitrary dimensions. For instance, a one-dimensional manifold is a curve and includes lines and circles. A two-dimensional manifold is a surface and includes planes, spheres and tori.

  3. We made here an abuse of language as attractor means that all trajectories converge toward it forward in time. Here, the strange attractor may be stable for some trajectories (attractor) and unstable for others (repeller).

References

  1. Huang, K.: Statistical Mechanics. Wiley, Chichester (1987)

    MATH  Google Scholar 

  2. Kubo, R., Toda, M., Hashitsume, N.: Statistical Physics II: Nonequilibrium Statistical Mechanics, 2nd edn. Springer, Berlin (1998)

    MATH  Google Scholar 

  3. Balian, R.: From Microphysics to Macrophysics: Methods and Applications of Statistical Physics, vol. II. Springer, Berlin (2007)

    MATH  Google Scholar 

  4. Matsoukas, T.: Generalized Statistical Thermodynamics: Thermodynamics of Probability Distributions and Stochastic Processes. Springer, Cham (2018)

    MATH  Google Scholar 

  5. Gibbs, J.W.: Elementary Principles in Statistical Mechanics, Developed With Especial Reference to the Rational Foundation of Thermodynamics. Charles Scribner’s Sons, New York (1902)

    MATH  Google Scholar 

  6. Callen, H.B.: Thermodynamics and an Introduction to Thermostatistics, 2nd edn. Wiley, New York (1985)

    MATH  Google Scholar 

  7. Van Kampen, N.G.: Stochastic Processes in Physics and Chemistry. Elsevier, Oxford (2007)

    MATH  Google Scholar 

  8. Sasa, S.-I.: Derivation of hydrodynamics from the Hamiltonian description of particle systems. Phys. Rev. Lett. 112, 100602 (2014)

    ADS  Google Scholar 

  9. Barrat, J.-L.: Elasticity and plasticity of disordered systems, a statistical physics perspective. Physica A 504, 20–30 (2018) (Lecture Notes of the 14th International Summer School on Fundamental Problems in Statistical Physics)

  10. Hartmann, S.: Effective field theories, reductionism and scientific explanation. Stud. Hist. Philos. Sci. Part B Stud. Hist. Philos. Mod. Phys. 32, 267–304 (2001)

    MathSciNet  MATH  ADS  Google Scholar 

  11. Peliti, L.: Statistical Mechanics in a Nutshell. Princeton University Press, Princeton (2011)

    MATH  Google Scholar 

  12. Touchette, H.: The large deviation approach to statistical mechanics. Phys. Rep. 478, 1–69 (2009)

    MathSciNet  ADS  Google Scholar 

  13. Sekimoto, K.: Stochastic Energetics, vol. 799. Springer, Berlin (2010)

    MATH  Google Scholar 

  14. Seifert, U.: Stochastic thermodynamics, fluctuation theorems and molecular machines. Rep. Prog. Phys. 75(12), 126001 (2012)

    ADS  Google Scholar 

  15. Esposito, M.: Stochastic thermodynamics under coarse graining. Phys. Rev. E 85, 041125 (2012)

    ADS  Google Scholar 

  16. Verley, G.: Nonequilibrium thermodynamic potentials for continuous-time Markov chains. Phys. Rev. E 93(1), 012111 (2016)

    ADS  Google Scholar 

  17. Barato, A.C., Chétrite, R.: A formal view on level 2.5 large deviations and fluctuation relations. J. Stat. Phys. 160, 1154–1172 (2015)

    MathSciNet  MATH  ADS  Google Scholar 

  18. Freidlin, M.I., Wentzell, A.D.: Random Perturbations of Dynamical Systems. Springer, Berlin (2012)

    MATH  Google Scholar 

  19. Touchette, H., Harris, R.J.: Deviation Approach to Nonequilibrium Systems, pp. 335–360. Wiley, Weinheim (2013)

  20. Bertini, L., De Sole, A., Gabrielli, D., Jona-Lasinio, G., Landim, C.: Macroscopic fluctuation theory. Rev. Mod. Phys. 87, 593–636 (2015)

    MathSciNet  MATH  ADS  Google Scholar 

  21. Weber, M.F., Frey, E.: Master equations and the theory of stochastic path integrals. Rep. Prog. Phys. 80, 046601 (2017)

    MathSciNet  ADS  Google Scholar 

  22. Assaf, M., Meerson, B.: WKB theory of large deviations in stochastic populations. J. Phys. A Math. Theor. 50(26), 263001 (2017)

    MathSciNet  MATH  ADS  Google Scholar 

  23. Monthus, C.: Microscopic fluctuation theory (mFT) for interacting Poisson processes. J. Phys. A Math. Theor. 52, 135003 (2019)

    MathSciNet  ADS  Google Scholar 

  24. Touchette, H.: Equivalence and nonequivalence of the microcanonical and canonical ensembles: a large deviations study. PhD thesis, Department of Physics, McGill University (2003)

  25. Tailleur, J., Kurchan, J., Lecomte, V.: Mapping nonequilibrium onto equilibrium: the macroscopic fluctuations of simple transport models. Phys. Rev. Lett. 99, 150602 (2007)

    ADS  Google Scholar 

  26. Lecomte, V., Imparato, A., van Wijland, F.: Current fluctuations in systems with diffusive dynamics, in and out of equilibrium. Prog. Theor. Phys. Suppl. 184, 276–289 (2010)

    MATH  ADS  Google Scholar 

  27. Mielke, A., Peletier, M., Renger, D.: On the relation between gradient flows and the large-deviation principle, with applications to Markov chains and diffusion. Potential Anal. 41(4), 1293–1327 (2014)

    MathSciNet  MATH  Google Scholar 

  28. Vroylandt, H., Verley, G.: Non-equivalence of dynamical ensembles and emergent non-ergodicity. J. Stat. Phys. 174(2), 404–432 (2019)

    MathSciNet  MATH  ADS  Google Scholar 

  29. Colangeli, M., Maes, C., Wynants, B.: A meaningful expansion around detailed balance. J. Phys. A Math. Theor. 44(9), 095001 (2011)

    MathSciNet  MATH  ADS  Google Scholar 

  30. Maes, C., Netočný, K.: Minimum entropy production principle from a dynamical fluctuation law. J. Math. Phys. 48(5),(2007)

  31. Maes, C., Netočný, K.: Canonical structure of dynamical fluctuations in mesoscopic nonequilibrium steady states. Europhys. Lett. 82(3), 30003 (2008)

  32. Derrida, B.: An exactly soluble non-equilibrium system: the asymmetric simple exclusion process. Phys. Rep. 301, 65–83 (1998)

    MathSciNet  ADS  Google Scholar 

  33. Lecomte, V., Rácz, Z., van Wijland, F.: Energy flux distribution in a two-temperature Ising model. J. Stat. Mech. Theory Exp. 2005(02), P02008 (2005)

  34. Wynants, B.: Structures of nonequilibrium fluctuations: dissipation and activity. PhD thesis, Universiteit K.U. Leuven (2010)

  35. Baule, A., Evans, R.M.L.: Nonequilibrium statistical mechanics of shear flow: invariant quantities and current relations. J. Stat. Mech. Theory Exp. 2010(03), P03030 (2010)

  36. Gorissen, M., Lazarescu, A., Mallick, K., Vanderzande, C.: Exact current statistics of the asymmetric simple exclusion process with open boundaries. Phys. Rev. Lett. 109, 170601 (2012)

    ADS  Google Scholar 

  37. Lazarescu, A., Mallick, K.: An exact formula for the statistics of the current in the Tasep with open boundaries. J. Phys. A Math. Theor. 44(31), 315001 (2011)

    MathSciNet  MATH  Google Scholar 

  38. Lazarescu, A.: The physicist’s companion to current fluctuations: one-dimensional bulk-driven lattice gases. J. Phys. A Math. Theor. 48(50), 503001 (2015)

    MathSciNet  MATH  Google Scholar 

  39. Garrahan, J.P.: Aspects of non-equilibrium in classical and quantum systems: Slow relaxation and glasses, dynamical large deviations, quantum non-ergodicity, and open quantum dynamics. Physica A 504, 130–154 (2018) (Lecture Notes of the 14th International Summer School on Fundamental Problems in Statistical Physics)

  40. Maes, C.: Frenesy: Time-symmetric dynamical activity in nonequilibria. Physics Reports 850, 1–33 (2020)

    MathSciNet  MATH  ADS  Google Scholar 

  41. Varadhan, S.R.S.: Large deviations. The Annals of Probability 36(2), 397–419 (2008)

    MathSciNet  MATH  Google Scholar 

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

    MathSciNet  MATH  ADS  Google Scholar 

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

    MathSciNet  MATH  ADS  Google Scholar 

  44. Lecomte, V.: Thermodynamique des histoires et fluctuations hors d’équilibre. PhD thesis, Université Paris-Diderot - Paris VII (2007)

  45. 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(19), 195702 (2007)

    MATH  ADS  Google Scholar 

  46. Lecomte, V., Appert-Rolland, C., van Wijland, F.: Thermodynamic formalism for systems with Markov dynamics. J. Stat. Phys. 127(1), 51–106 (2007)

    MathSciNet  MATH  ADS  Google Scholar 

  47. 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 Math. Theor. 42(7), 075007 (2009)

    MathSciNet  MATH  ADS  Google Scholar 

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

    ADS  Google Scholar 

  49. Giardina, C., Kurchan, J., Lecomte, V., Tailleur, J.: Simulating rare events in dynamical processes. J. Stat. Phys. 145(4), 787–811 (2011)

    MathSciNet  MATH  ADS  Google Scholar 

  50. Ferré, G., Touchette, H.: Adaptive sampling of large deviations. J. Stat. Phys. 172, 1525–1544 (2018)

    MathSciNet  MATH  ADS  Google Scholar 

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

    ADS  Google Scholar 

  52. Evans, R.M.L.: Detailed balance has a counterpart in non-equilibrium steady states. J. Phys. A Math. Gen. 38(2), 293 (2004)

    MATH  ADS  Google Scholar 

  53. Evans, R.M.L.: Statistical physics of shear flow: a non-equilibrium problem. Contemp. Phys. 51(5), 413–427 (2010)

    ADS  Google Scholar 

  54. Jack, R.L., Sollich, P.: Large deviations and ensembles of trajectories in stochastic models. Prog. Theor. Phys. Suppl. 184, 304–317 (2010)

    MATH  ADS  Google Scholar 

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

    ADS  Google Scholar 

  56. Chetrite, R., Touchette, H.: Nonequilibrium Markov processes conditioned on large deviations. Ann. Henri Poincaré 16, 2005–2057 (2015)

    MathSciNet  MATH  ADS  Google Scholar 

  57. Garrahan, J.P.: Classical stochastic dynamics and continuous matrix product states: gauge transformations, conditioned and driven processes, and equivalence of trajectory ensembles. J. Stat. Mech. Theory Exp. 2016, 073208 (2016)

    MathSciNet  MATH  Google Scholar 

  58. Ritort, F.: Work and heat fluctuations in two-state systems: a trajectory thermodynamics formalism. J. Stat. Mech. Theory Exp. (2004). https://doi.org/10.1088/1742-5468/2004/10/P10016

    Article  MATH  Google Scholar 

  59. Ross, J.: Thermodynamics and Fluctuations far from Equilibrium. Springer, Berlin (2008)

    Google Scholar 

  60. Wentzell, A.D.: Limit Theorems on Large Deviations for Markov Stochastic Processes, vol. 38. Mathematics and Its Applications. Springer, Dordrecht (1990)

  61. Vroylandt, H.: Thermodynamics and fluctuations of small machines. Université Paris-Saclay (COmUE). (2018). https://tel.archives-ouvertes.fr/tel-01968075

  62. Vroylandt, H., Esposito, M., Verley, G.: Efficiency fluctuations of stochastic machines undergoing a phase transition. Phys. Rev. Lett. 124(25), 250603 (2020)

    MathSciNet  ADS  Google Scholar 

  63. Mielke, A., Patterson, R.I.A., Peletier, M.A., Michiel Renger, D.R.: Non-equilibrium thermodynamical principles for chemical reactions with mass-action kinetics. SIAM J. Appl. Math. 77(4), 1562–1585 (2017)

    MathSciNet  MATH  Google Scholar 

  64. Lazarescu, A., Cossetto, T., Falasco, G., Esposito, M.: Large deviations and dynamical phase transitions in stochastic chemical networks. J. Chem. Phys. 151, 064117 (2019)

    ADS  Google Scholar 

  65. Rax, J.-M.: Mécanique analytique: Adiabaticité, résonances, chaos. Sciences Sup, Dunod (2020)

  66. Cleuren, B., Van den Broeck, C.: Ising model for a Brownian donkey. Europhys. Lett. (EPL) 54(1), 1 (2001)

    ADS  Google Scholar 

  67. Chetrite, R.: Pérégrinations sur les phénomènes aléatoires dans la nature. HDR Thesis, Université de Nice-Sophia-Antipolis, France (2018)

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

    MathSciNet  MATH  ADS  Google Scholar 

  69. Chetrite, R., Touchette, H.: Nonequilibrium microcanonical and canonical ensembles and their equivalence. Phys. Rev. Lett. 111(12), 120601 (2013)

    ADS  Google Scholar 

  70. Farina, L., Rinaldi, S.: Positive Linear Systems: Theory and Applications. Wiley, New York (2000)

    MATH  Google Scholar 

  71. Horn, R.A., Johnson, C.R.: Matrix Analysis, 2nd edn. Cambridge University Press, Cambridge (2012)

    Google Scholar 

  72. Bokharaie, V.S.: Stability analysis of positive systems with applications to epidemiology. PhD thesis, National University of Ireland Maynooth (2012)

  73. Chetrite, R., Touchette, H.: Variational and optimal control representations of conditioned and driven processes. J. Stat. Mech. Theory Exp. 2015(12), P12001 (2015)

    MathSciNet  MATH  Google Scholar 

  74. Poincaré, H.: Mémoire sur les courbes définies par une équation différentielle (I). J. Math. Pures Appl. 7, 375–422 (1881)

    MATH  Google Scholar 

  75. Poincaré, H.: Mémoire sur les courbes définies par une équation différentielle (II). J. Math. Pures Appl. 8, 251–296 (1882)

    MATH  Google Scholar 

  76. Hopf, H.: Vektorfelder in n-dimensionalen Mannigfaltigkeiten. Math. Ann. 96(1), 225–249 (1927)

    MathSciNet  MATH  Google Scholar 

  77. Richardson, T.J., Kwong, R.H.: On positive definite solutions to the algebraic Riccati equation. Syst. Control Lett. 7(2), 99–104 (1986)

    MathSciNet  MATH  Google Scholar 

  78. Lazarescu, A.: Generic dynamical phase transition in one-dimensional bulk-driven lattice gases with exclusion. J. Phys. A Math. Theor. 50(25), 254004 (2017)

    MathSciNet  MATH  ADS  Google Scholar 

  79. Bodineau, T., Derrida, B., Lecomte, V., Van Wijland, F.: Long range correlations and phase transitions in non-equilibrium diffusive systems. J. Stat. Phys. 133(6), 1013–1031 (2008)

    MathSciNet  MATH  ADS  Google Scholar 

  80. Baek, Y., Kafri, Y., Lecomte, V.: Dynamical symmetry breaking and phase transitions in driven diffusive systems. Phys. Rev. Lett. 118, 030604 (2017)

    ADS  Google Scholar 

  81. Bunin, G., Kafri, Y., Podolsky, D.: Non-differentiable large-deviation functionals in boundary-driven diffusive systems. J. Stat. Mech. Theory Exp. 2012(10), L10001 (2012)

    MathSciNet  MATH  Google Scholar 

  82. Donsker, M.D., Varadhan, S.S.: Asymptotic evaluation of certain Markov process expectations for large time, I. Commun. Pure Appl. Math. 28(1), 1–47 (1975)

    MathSciNet  MATH  Google Scholar 

  83. Donsker, M., Varadhan, S.: Asymptotic evaluation of certain Markov process expectations for large time, II. Commun. Pure Appl. Math. 28(2), 279–301 (1975)

    MathSciNet  MATH  Google Scholar 

  84. Donsker, M.D., Varadhan, S.S.: Asymptotic evaluation of certain Markov process expectations for large time—III. Commun. Pure Appl. Math. 29(4), 389–461 (1976)

    MathSciNet  MATH  Google Scholar 

  85. Donsker, M.D., Varadhan, S.S.: Asymptotic evaluation of certain Markov process expectations for large time. IV. Commun. Pure Appl. Math. 36(2), 183–212 (1983)

    MathSciNet  MATH  Google Scholar 

  86. 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)

    MathSciNet  MATH  Google Scholar 

  87. Suda, T.: Application of Helmholtz–Hodge decomposition to the study of certain vector fields. J. Phys. A Math. Theor. 53(37), 375703 (2020)

    MathSciNet  Google Scholar 

  88. Suda, T.: Construction of Lyapunov functions using Helmholtz–Hodge decomposition. Discret. Contin. Dyn. Syst. 39(5), 2437–2454 (2019)

    MathSciNet  MATH  Google Scholar 

  89. Hatano, T., Sasa, S.-I.: Steady-state thermodynamics of Langevin systems. Phys. Rev. Lett. 86(16), 3463–3466 (2001)

    ADS  Google Scholar 

  90. Chernyak, V.Y., Chertkov, M., Jarzynski, C.: Path-integral analysis of fluctuation theorems for general Langevin processes. J. Stat. Mech. Theory Exp. 2006(08), P08001 (2006)

    MATH  Google Scholar 

  91. Blanchard, C.: Canonical transformations and time reversal. Am. J. Phys. 56(12), 1148–1149 (1988)

    ADS  Google Scholar 

  92. Nguyen, B., Seifert, U., Barato, A.C.: Phase transition in thermodynamically consistent biochemical oscillators. J. Chem. Phys. 149, 045101 (2018)

    ADS  Google Scholar 

  93. Assaf, M., Meerson, B.: Extinction of metastable stochastic populations. Phys. Rev. E 81, 021116 (2010)

    ADS  Google Scholar 

  94. Gaspard, P.: Fluctuation theorem for nonequilibrium reactions. J. Chem. Phys. 120, 8898–8905 (2004)

    ADS  Google Scholar 

  95. Chabane, L.: From rarity to typicality: the improbable journey of a large deviation. Université Paris-Saclay. (2021). https://tel.archives-ouvertes.fr/tel-03494870

  96. Chabane, L., Chétrite, R., Verley, G.: Periodically driven jump processes conditioned on large deviations. J. Stat. Mech. Theory Exp. 2020(3), 033208 (2020)

    MathSciNet  MATH  Google Scholar 

  97. Lau, A.W., Lubensky, T.C.: State-dependent diffusion: thermodynamic consistency and its path integral formulation. Phys. Rev. E 76(1), 011123 (2007)

    MathSciNet  ADS  Google Scholar 

  98. Gardiner, C.W.: Handbook of Stochastic Methods, vol. 3. Springer, Berlin (1985)

    Google Scholar 

  99. Dean, D.S.: Langevin equation for the density of a system of interacting Langevin processes. J. Phys. A Math. Gen. 29(24), L613 (1996)

    MathSciNet  ADS  Google Scholar 

  100. Bouchet, F., Gawȩdzki, K., Nardini, C.: Perturbative calculation of quasi-potential in non-equilibrium diffusions: a mean-field example. J. Stat. Phys. 163(5), 1157–1210 (2016)

    MathSciNet  MATH  ADS  Google Scholar 

Download references

Acknowledgements

A. Lazarescu was supported by the Belgian Excellence of Science (EOS) initiative through the Project 30889451 PRIMA Partners in Research on Integrable Systems and Applications.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Gatien Verley.

Additional information

Communicated by Julien Tailleur.

Publisher's Note

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

Appendices

Appendix

This appendix deals with the biasing and rectification of Markov diffusion processes. We first review the case of a single process in the linear operator framework, and then we study the case of \(N\) independent processes within the Lagrangian–Hamiltonian formalism.

Appendix A: Single diffusion process

We consider a diffusion process described by the following Langevin equation

$$\begin{aligned} \dot{x}_t = b(x_t) + \sigma (x_t) \circ \xi _t, \end{aligned}$$
(A.1)

where \(x_t\) is a random variable and \(\xi _t\) a Gaussian white noise of mean \(\langle \xi _t \rangle = 0\) and variance \(\langle \xi _t \xi _{t'} \rangle \,=\,\delta (t-t')\). The stochastic integrals are defined according to the mid-point Stratonovich convention, referred to by a circle \(\circ \). The drift b and the diffusion \(\sigma \) are functions of \(x_t\) and do not depend explicitly on time. The probability density \(\varrho (x,t)\) satisfies the Fokker–Planck equation

$$\begin{aligned} \frac{\partial \varrho (x,t)}{\partial t} = - \nabla J(x,t), \end{aligned}$$
(A.2)

with \(\nabla \) the derivative with respect to x and

$$\begin{aligned} J^{\varrho }(x,t) \equiv \hat{b}(x) \varrho (x,t) - \frac{1}{2} \sigma (x)^2 \nabla \varrho (x,t), \end{aligned}$$
(A.3)

where we introduced the modified drift \(\hat{b}(x) \equiv b(x) - \frac{1}{2} \sigma (x) \nabla \sigma (x)\). The Fokker–Planck equation appears to be a continuity equation that conserves the normalization of the probability density, i.e. \(\int \text {d}x \varrho (x,t) =1 , \forall t\). From now on, unspecified integrations are implicitly on x, e.g. \(\int \varrho =1\). One can rewrite the Fokker–Planck equation as

$$\begin{aligned} \frac{\partial \varrho }{\partial t} = \mathfrak {L}\varrho , \end{aligned}$$
(A.4)

where \(\mathfrak {L}\) is the Fokker–Planck operator in the Stratonovich convention, defined by its action on a function \(\varphi \):

$$\begin{aligned} \mathfrak {L}\varphi (x) \equiv - \nabla \left[ \hat{b}(x) \varphi (x) \right] + \frac{1}{2} \nabla \left[ \sigma (x)^2 \nabla \varphi (x) \right] \end{aligned}$$
(A.5)

The adjoint Fokker–Planck operator \(\mathfrak {L}^\dagger \) is given by

$$\begin{aligned} \mathfrak {L}^\dagger \varphi (x) = \hat{b}(x) \nabla \varphi (x) + \frac{1}{2} \nabla \left[ \sigma (x)^2 \nabla \varphi (x) \right] , \end{aligned}$$
(A.6)

both operators being related by

$$\begin{aligned} \int \left( \mathfrak {L}\varphi \right) \psi = \int \left( \mathfrak {L}^\dagger \psi \right) \varphi , \end{aligned}$$
(A.7)

for any functions \(\varphi \), \(\psi \). We denote a path by \([x_t]\), with \(x_t\) the state of the system at time t. The path probability of \([x_t]\) between the initial time 0 and the final time t within the Stratonovich convention reads [97]

$$\begin{aligned} \mathbb {P}_{b,\sigma ,\varrho (0)}[x_t] = \varrho (x_0,0) \exp \left\{ - \int _0^t \text {d}t' \left[ \left( \dot{x}_{t'} - \hat{b}(x_{t'}) \right) ^2 + \frac{1}{2} \nabla b(x_{t'}) \right] \right\} , \end{aligned}$$
(A.8)

Many dynamical and thermodynamic observables such as heat, matter currents, work, entropy production, etc. can be written as linear combinations of the empirical occupancy \(\tilde{\rho }_t\) and the empirical current \(\tilde{j}_t\). The function \(\tilde{\rho }_t(x)\) counts the rate of occupancy of the position x along the trajectory \([x_t]\):

$$\begin{aligned} \tilde{\rho }_t(x) = \frac{1}{t} \int _{0}^{t} \text {d}\tau \delta (x_\tau -x), \end{aligned}$$
(A.9)

while the function \(\tilde{j}_t(x)\) informs on the time-averaged local velocity at x [56]:

$$\begin{aligned} \tilde{j}_t(x) = \frac{1}{t} \int _0^t \text {d}\tau \delta (x_\tau -x) \circ \dot{x}_\tau , \end{aligned}$$
(A.10)

where the circle \(\circ \) refers to the Stratonovich convention. We would like to condition our original Markov process by filtering the ensemble of paths to select those leading to a chosen value of \({\tilde{\varvec{A}}}_t(x) = ( \tilde{j}_t(x), \tilde{\rho }_t(x))\), for each state x. This defines the conditioned process for which we aim to find an equivalent Markov process in the long-time limit, namely the driven process [67]. This process is described by the microcanonical path probability [56]

$$\begin{aligned} \mathbb {P}^{\text {micro}}_\mathbf{a ,\varrho (0)}[x_t]= \mathbb {P}_{\Lambda ,\varrho (0)}\left[ x_t ~\Big |~ {\tilde{\varvec{A}}}_t = \mathbf{a} \right] . \end{aligned}$$
(A.11)

To explicit the generator of the driven process, let us introduce the generating function

$$\begin{aligned} G_{\varvec{\gamma }}(x,t) \equiv \left\langle \text {e}^{t \varvec{\gamma }\cdot {\tilde{\varvec{A}}}_{t} } \delta (x_t - x) \right\rangle _{\varrho (0)}, \end{aligned}$$
(A.12)

where \(\langle \cdots \rangle _{\varrho (0)}\) is the path average based on (A.8), \(\varvec{\gamma }(x) \equiv \left( \gamma _1(x) , \gamma _2(x) \right) \) is the conjugate variable of \({\tilde{\varvec{A}}}_t(x)\) and where the dot stands for the scalar product \(\varvec{\gamma }\cdot {\tilde{\varvec{A}}}_{t} = \int \text {d}x \left[ \gamma _1(x) \tilde{j}_t(x) + \gamma _2(x) \tilde{\rho }_t(x) \right] \). From now on, we drop in the notation the x-dependency of the functions for clarity. The generating function (A.12) evolves according to

$$\begin{aligned} \dot{G}_{\varvec{\gamma }} = \Lambda _{\varvec{\gamma }} G_{\varvec{\gamma }}, \end{aligned}$$
(A.13)

where the biased Fokker–Planck operator \(\Lambda _{\varvec{\gamma }}\) is given by [56]:

$$\begin{aligned} \Lambda _{\varvec{\gamma }} \varphi \equiv (- \nabla + \gamma _1) ( \hat{b} \varphi ) + \frac{1}{2} (- \nabla + \gamma _1) \left[ \sigma ^2 (-\nabla + \gamma _1) \varphi \right] + \gamma _2 \varphi , \end{aligned}$$
(A.14)

One can compute its adjoint operator

$$\begin{aligned} \Lambda _{\varvec{\gamma }}^\dagger \varphi = \hat{b} (\nabla + \gamma _1) \varphi + \frac{1}{2} (\nabla + \gamma _1) \left[ \sigma ^2 (\nabla + \gamma _1) \varphi \right] + \gamma _2 \varphi . \end{aligned}$$
(A.15)

The biased Fokker–Planck operator \(\Lambda _{\varvec{\gamma }}\) generates a Markov process that is not norm-conserving since \(\int \Lambda _{\varvec{\gamma }} \varrho \ne 0\). As with Markov jump processes, we can build the operator of the driven process \(\mathcal {L}\) by taking the Doob transform of the biased operator \(\Lambda _{\varvec{\gamma }}\) associated with its dominant left eigenfunction:

$$\begin{aligned} \mathcal {L}\varphi \equiv l \Lambda _{\varvec{\gamma }} (l^{-1} \varphi ) - l (\Lambda _{\varvec{\gamma }}^\dagger l^{-1}) \varphi , \end{aligned}$$
(A.16)

with \(l \equiv l(x)\) being the left eigenfunction of \(\Lambda _{\varvec{\gamma }}\) for the highest eigenvalue \({\bar{\Gamma }}\)

$$\begin{aligned} \Lambda _{\varvec{\gamma }}^\dagger l = {\bar{\Gamma }}l. \end{aligned}$$
(A.17)

Since the Krein–Rutman theorem ensures the positivity of l, we introduce a new function \(u \equiv u(x)\) such that \(l \equiv \text {e}^u\). It follows from (A.17):

$$\begin{aligned} {\bar{\Gamma }}= \int \text {e}^{-u} (\Lambda _{\varvec{\gamma }}^\dagger \text {e}^u) \rho . \end{aligned}$$
(A.18)

Computing explicitly Eq. (A.16) using Eqs. (A.14, A.15) and \(l = \text {e}^u\), we finally find that the driven Fokker–Planck operator is given by

$$\begin{aligned} \mathcal {L}\varphi = - \nabla \left[ \hat{B}_\gamma \varphi - \frac{1}{2} \sigma ^2 \nabla \varphi \right] , \end{aligned}$$
(A.19)

where we introduced the rectified drift

$$\begin{aligned} \hat{B}_{\varvec{\gamma }} \equiv \hat{b} + \sigma ^2 (\nabla u + \gamma _1). \end{aligned}$$
(A.20)

The driven process is thus a diffusive process obeying the same stochastic equation (A.1) as the original process but with a new drift (A.20). Note that the dependence of u on \(\varvec{\gamma }\) is made implicit.

Appendix B: \(N\) Independent Diffusion Processes

We consider \(N\) independent and identical systems, each one modeled by a time-homogeneous Markov diffusion process of time-independent drift b and diffusion coefficient \(\sigma \). We denote by \(\nu \in \{1, 2, \dots N\}\) the \(\nu ^\text {th}\) system and by \(x^\nu _t\) the stochastic process of the system \(\nu \) which evolves according to the Langevin equation

$$\begin{aligned} \dot{x}^\nu _t = b(x_t^\nu ) + \sigma (x_t^\nu ) \xi ^\nu _t, \end{aligned}$$
(B.1)

We are interested in the empirical occupation density:

$$\begin{aligned} \rho (x,t) = \frac{1}{N} \sum _{\nu =1}^N \delta (x^\nu _t-x) , \end{aligned}$$
(B.2)

and the empirical current:

$$\begin{aligned} j(x,t) = \frac{1}{N} \sum _{\nu =1}^N \delta (x^\nu _t-x) \circ \dot{x}^\nu _t, \end{aligned}$$
(B.3)

playing respectively the role of the variables \({\varvec{z}}\) and \(\varvec{\lambda }\) in our general framework of section 3. The empirical occupation density gives the density of systems being at a state in \([x,x+\text {d}x[\) at time t, and the empirical current measures the density of systems performing a displacement between x and \(x+\text {d}x\) within the time interval \([t,t+\text {d}t[\). Both variables are related by

$$\begin{aligned} \dot{\rho }(x,t) = - \nabla j(x,t), \end{aligned}$$
(B.4)

with \(- \nabla \) playing the role of \(\mathcal {D}\). Notice that these observables are related to the empirical occupancy \(\tilde{\rho }_t^\nu \) (A.9) and the empirical transition current \(\tilde{j}_t^\nu \) (A.10) for a single process through

$$\begin{aligned} \begin{aligned} \frac{1}{t} \int _0^t \text {d}\tau \, \rho (\tau )&= \frac{1}{N} \sum _{\nu = 1}^N \tilde{\rho }_t^\nu , \\ \frac{1}{t} \int _0^t \text {d}\tau \, j(\tau )&= \frac{1}{N} \sum _{\nu = 1}^N \tilde{j}_t^\nu , \end{aligned} \end{aligned}$$
(B.5)

where the superscript \(\nu \) indicates that the empirical occupancy or current are those for the trajectory of the \(\nu ^{\text {th}}\) system.

1.1 Stochastic equation for the empirical occupation density

We aim to give a coarse-grained description of the global system by deriving the stochastic equation for \(\rho \). To do so, we compute the quantity: \(\Delta x^\nu ~\equiv ~x^\nu _{t+\Delta t}~-~x^\nu _t\). Using the Langevin equation (A.1) for \(x^\nu _t\), we get:

$$\begin{aligned} \Delta x^\nu = \int _t^{t+\Delta t} \text {d}\tau \left[ b(x^\nu _\tau ) + \sigma (x_\tau ^\nu ) \circ \xi ^\nu _\tau \right] . \end{aligned}$$
(B.6)

For an infinitesimal \(\Delta t\), we have in the Stratonovich convention [97]:

$$\begin{aligned} \int _t^{t+\Delta t} \text {d}\tau \, b(x^\nu _\tau )&\simeq b (x^\nu _t + \frac{\Delta x^\nu }{2} ) \Delta t, \end{aligned}$$
(B.7)
$$\begin{aligned} \int _t^{t+\Delta t} \text {d}\tau \, \left[ \sigma (x_\tau ^\nu ) \xi ^\nu _\tau \right]&\simeq \sigma (x^\nu _t + \frac{\Delta x^\nu }{2} ) \int _t^{t+\Delta t} \text {d}\tau \, \xi ^\nu _\tau . \end{aligned}$$
(B.8)

Expanding up to order \(\Delta t\) and using the identity \((\int _t^{t+\Delta t} \xi _t \text {d}t)^2 = \Delta t\) when \(\Delta t \rightarrow 0\) [13, 98], it follows

$$\begin{aligned} \Delta x^\nu&\simeq b(x^\nu _t) \Delta t + \sigma (x^\nu _t) \int _t^{t+\Delta t} \xi ^\nu _{\tau } \text {d}\tau + \frac{1}{2} \sigma (x^\nu _t) \nabla \sigma (x^\nu _t) \Delta t, \end{aligned}$$
(B.9)
$$\begin{aligned} (\Delta x^\nu )^2&\simeq \sigma (x^\nu _t)^2 \Delta t. \end{aligned}$$
(B.10)

We now compute \(\rho (x,t+\Delta t) = \frac{1}{N} \sum _{\nu =1}^N \delta (x^\nu _{t+\Delta t}-x)\). Let \(\varphi \) be a test function, then

$$\begin{aligned} \int \text {d}x \varphi (x) \rho (x,t+\Delta t)&= \frac{1}{N} \sum _{\nu = 1}^N \varphi (x_{t+\Delta t}^\nu ) \end{aligned}$$
(B.11)
$$\begin{aligned}&= \frac{1}{N} \sum _{\nu = 1}^N \varphi (x_{t}^\nu + \Delta x^\nu ) \end{aligned}$$
(B.12)
$$\begin{aligned}&\simeq \frac{1}{N} \sum _{\nu = 1}^N \varphi (x_t^\nu ) + \frac{1}{N} \sum _{\nu = 1}^N \Delta x^\nu \varphi '(x_t^\nu ) + \frac{1}{N} \sum _{\nu = 1}^N \frac{1}{2} (\Delta {x^\nu })^2 \varphi ''(x_t^\nu ), \end{aligned}$$
(B.13)

where we used Taylor’s formula around \(x^\nu _t\) up to second order in \(\Delta x^\nu \) in the last equation. Using Eqs. (B.9, B.10) and the fact that \(\frac{1}{N} \sum _{\nu =1}^N \varphi (x_t^\nu ) = \int \text {d}x \varphi (x) \rho (x,t)\), Eq. (B.13) gives

$$\begin{aligned}&\int \text {d}x \varphi (x) \dot{\rho }(x,t) \nonumber \\&\quad = \int \text {d}x \varphi (x)\left\{ - \nabla \left[ \hat{b}(x) \rho (x,t) - \frac{1}{2} \sigma (x)^2 \nabla \rho (x,t) + \sigma (x) \sqrt{\frac{\rho (x,t)}{N}} \eta (x,t) \right] \right\} , \nonumber \\ \end{aligned}$$
(B.14)

with \(\dot{\rho }(x,t) = \lim _{\Delta t \rightarrow 0} \frac{\rho (x,t+\Delta t) - \rho (x,t)}{\Delta t}\) and where we introduced

$$\begin{aligned} \eta (x,t) \equiv \frac{1}{\sqrt{N\, \rho (x,t)} } \sum _{\nu =1}^N\delta (x-x_t^\nu ) \bar{\xi }^\nu _t, \end{aligned}$$
(B.15)

with \(\bar{\xi }^\nu _t \equiv \lim _{\Delta t \rightarrow 0} \frac{1}{\Delta t} \int _t^{t+\Delta t} \text {d}\tau \xi _\tau ^\nu \). The stochastic process \(\eta \) is a Gaussian white noise in time and space [99, 100] with mean and variance

$$\begin{aligned} \langle \eta (x,t) \rangle= & {} 0, \end{aligned}$$
(B.16)
$$\begin{aligned} \langle \eta (x,t) \eta (x',t') \rangle= & {} \delta (x-x') \delta (t-t'). \end{aligned}$$
(B.17)

Since (B.14) is valid for any function \(\varphi \), we obtain the stochastic equation for the density \(\rho \):

$$\begin{aligned} \dot{\rho }(x,t) = - \nabla \left[ \hat{b}(x) \rho (x,t) - \frac{1}{2} \sigma (x)^2 \nabla \rho (x,t) + \sigma (x) \sqrt{\frac{\rho (x,t)}{N}} \eta (x,t) \right] . \end{aligned}$$
(B.18)

Eq. (B.18) is known as the Dean equation.

1.2 Derivation of the Lagrangians and Hamiltonians

In order to obtain the Lagrangian \(\mathscr {L}(j,\rho )\), we compute the conditional probability \(P_{\delta t}(j \,|\, \rho )\) using Eqs. (B.4, B.18):

$$\begin{aligned} P_{\delta t}(j \,|\, \rho ) = \prod _x \left\langle \delta \left[ j - \hat{b} \rho + \frac{1}{2} \sigma ^2 \nabla \rho - \sigma \sqrt{\frac{\rho }{N}} \eta \right] \right\rangle _{\eta }. \end{aligned}$$
(B.19)

The continuous product \(\prod _x\) runs over the states \(x_\ell \equiv \ell \delta x\) with \(\ell \) an integer and \(\delta x\) an infinitesimal space step, and \(\langle \cdots \rangle _\eta \) is the average over the noise \(\eta \):

$$\begin{aligned} \left\langle \mathcal {O}(\eta ) \right\rangle _\eta \equiv \frac{1}{\mathsf {N}} \int \mathcal {O}(\eta ) \text {e}^{-\frac{1}{2} \delta t \delta x \eta ^2} \text {d}\eta , \end{aligned}$$
(B.20)

with \(\mathsf {N}\) the normalization factor and \(O(\eta )\) an arbitrary function of \(\eta \). We have dropped the (xt)-dependency in all functions for clarity. It follows

$$\begin{aligned} P_{\delta t}(j \,|\, \rho )&= \frac{1}{\mathsf {N}} \int \text {d}\eta \prod _{x} \text {e}^{-\frac{1}{2} \delta t \delta x \eta ^2} \delta \left[ j - \hat{b} \rho + \frac{1}{2} \sigma ^2 \nabla \rho - \sigma \sqrt{\frac{\rho }{N}} \eta \right] \nonumber \\&= \frac{1}{\mathsf {N}} \int \text {d}\eta \prod _{x} e^{-\frac{1}{2} \delta t \delta x \eta ^2} \frac{\sqrt{N}}{\sigma \sqrt{\rho }} \delta \left[ \eta - \frac{ j - \hat{b} \rho + \frac{1}{2} \sigma ^2 \nabla \rho }{\frac{1}{\sqrt{N}} \sigma \sqrt{\rho }} \right] \nonumber \\&= \frac{1}{\mathsf {N}} \exp { \left[ - \delta t \int \frac{N}{2\sigma ^2 \rho } \left( j - \hat{b} \rho + \frac{1}{2} \sigma ^2 \nabla \rho \right) ^2 + \frac{1}{2} \int \ln \left( \frac{N}{\rho \sigma ^2}\right) \right] }, \end{aligned}$$
(B.21)

where we used in the second equality the relation \(\delta (\varphi (y)) = |\varphi '(y_0) |^{-1} \delta (y-y_0)\), for any smooth test function \(\varphi \) and any root \(y_0\) of \(\varphi (y)=0\). In the limit of large \(N\), the last term in the exponential is asymptotically dominated by \(N\) and we obtain

$$\begin{aligned} P_{\delta t}(j \,|\, \rho ) \underset{N \rightarrow \infty }{\asymp }\text {e}^{- N\delta t \mathscr {L}(j,\rho )}, \end{aligned}$$
(B.22)

where the Lagrangian is given by

$$\begin{aligned} \mathscr {L}(j,\rho ) = \int \frac{1}{2\sigma ^2 \rho } \left( j - J^\rho \right) ^2, \end{aligned}$$
(B.23)

with \(J^\rho = {\hat{b}} \rho - \frac{1}{2} \sigma ^2 \nabla \rho \). Computing the Legendre transform of \(\mathscr {L}\) with respect to j yields the detailed Hamiltonian

$$\begin{aligned} \mathscr {H}(f,\rho ) = \int f \left[ \frac{1}{2} \sigma ^2 f \rho + J^\rho \right] . \end{aligned}$$
(B.24)

We are interested in the observable

$$\begin{aligned} \varvec{A}_{t}(x) \equiv \frac{N}{t} \left( \begin{array}{c} \int _0^t \text {d}\tau j(x,\tau ) \\ \int _0^t \text {d}\tau \rho (x,\tau ) \end{array} \right) . \end{aligned}$$
(B.25)

Using the results of Sect. 3.2, the dynamical fluctuations of \(\varvec{A}\) are encoded in the biased Lagrangian and Hamiltonian

$$\begin{aligned} \mathscr {L}_{\varvec{\gamma }}(j,\rho )&= \mathscr {L}(j,\rho ) - \gamma _1 \cdot j - \gamma _2 \cdot \rho , \end{aligned}$$
(B.26)
$$\begin{aligned} \mathscr {H}_{\varvec{\gamma }}(f,\rho )&= \int (f+\gamma _1) \left[ \frac{1}{2} \sigma ^2 (f+\gamma _1) \rho + J^\rho \right] + \gamma _2 \cdot \rho = \mathscr {H}(f+\gamma _1,\rho ) + \gamma _2 \rho , \end{aligned}$$
(B.27)

where \(\gamma _1(x)\) (resp. \(\gamma _2(x)\)) is conjugated to the first (resp. second) component of \(\varvec{A}(x)\).

1.3 SCGF and HJ equation

In order to derive the rectified Hamiltonian, we first need to translate the spectral properties of the biased generator from the linear operator formalism to the Hamiltonian formalism. Because of the independence of the \(N\) processes, it suffices to look at a single process. Indeed, we can relate the SCGF \(\Gamma \) of the global system to the SCGF \({\bar{\Gamma }}\) of the single process by:

$$\begin{aligned} \Gamma&= \lim _{t \rightarrow \infty } \frac{1}{t} \ln \left\langle \text {e}^{t \varvec{\gamma }\cdot \varvec{A}_t} \right\rangle _{x_0} \end{aligned}$$
(B.28)
$$\begin{aligned}&= \lim _{t \rightarrow \infty } \frac{1}{t} \ln \left\langle \text {e}^{N\int _0^t \text {d}\tau \gamma _1 \cdot j(\tau ) + N\int _0^t \text {d}\tau \gamma _2 \cdot \rho (\tau )} \right\rangle _{x_0} \end{aligned}$$
(B.29)
$$\begin{aligned}&= \lim _{t \rightarrow \infty } \frac{1}{t} \ln \left\langle \text {e}^{t \sum _{\nu = 1}^N \left( \gamma _1 \cdot \tilde{j}_t^\nu + \gamma _2 \cdot \tilde{\rho }_t^\nu \right) } \right\rangle _{x_0} \end{aligned}$$
(B.30)
$$\begin{aligned}&= N\lim _{t \rightarrow \infty } \frac{1}{t} \ln \left\langle \text {e}^{t \left( \gamma _1 \cdot \tilde{j}_t^\nu + \gamma _2 \cdot \tilde{\rho }_t^\nu \right) } \right\rangle _{x_0} \end{aligned}$$
(B.31)
$$\begin{aligned}&= N{\bar{\Gamma }}, \end{aligned}$$
(B.32)

where we used Eq. (B.25) in Eqs. (B.29), (B.5) in Eqs. (B.30) and the fact that the \(N\) processes are independent and identically distributed in Eq. (B.31). Computing explicitly the right-hand-side of Eq. (A.18), we find

$$\begin{aligned} {\bar{\Gamma }}= \mathscr {H}_{\varvec{\gamma }}(f=\nabla u, \rho ), \end{aligned}$$
(B.33)

with \(\nabla = (-\nabla )^\dag \). As expected, the function u appearing in the left eigenfunction of the single-process biased Fokker–Planck operator is the solution of the HJ equation and we write \(u = \partial _{\rho } W_{\text {s}}\).

1.4 Rectified Hamiltonian

The rectified Hamiltonian follows from Eq. (5.6):

$$\begin{aligned} \mathscr {H}^r(f,\rho ;\varvec{\gamma }) = \mathscr {H}_{\varvec{\gamma }}(f+\nabla u, \rho ) - \mathscr {H}_{\varvec{\gamma }}(\nabla u, \rho ), \end{aligned}$$
(B.34)

leading after explicit computation to

$$\begin{aligned} \mathscr {H}^{r}(f,\rho ;\varvec{\gamma }) = \int f \left[ \frac{1}{2} \sigma ^2 f \rho + J^{{\text {r}},\rho }_{\varvec{\gamma }} \right] , \end{aligned}$$
(B.35)

with \(J^{{\text {r}},\rho }_{\varvec{\gamma }} \equiv \hat{B}_{\varvec{\gamma }} - \frac{1}{2} \sigma ^2 \nabla \rho \), where the rectified drift \(\hat{B}_{\varvec{\gamma }}\) is defined in Eq. (A.20). Unsurprisingly, the rectified Hamiltonian corresponds to an unbiased Hamiltonian associated with the drift \(\hat{B}_{\varvec{\gamma }}\) of the driven process obtained from a Doob transform as seen in Eq. (A.20). This illustrates the fact that the rectification of biased Hamiltonians is equivalent to the rectification of biased generators in the linear operator formalism using the Doob transform.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Chabane, L., Lazarescu, A. & Verley, G. Effective Hamiltonians and Lagrangians for Conditioned Markov Processes at Large Volume. J Stat Phys 187, 6 (2022). https://doi.org/10.1007/s10955-022-02898-3

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1007/s10955-022-02898-3

Keywords

Navigation