Skip to main content
Log in

A Path Integral Formalism for Non-equilibrium Hamiltonian Statistical Systems

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

Abstract

A path integral formalism for non-equilibrium systems is proposed based on a manifold of quasi-equilibrium densities. A generalized Boltzmann principle is used to weight manifold paths with the exponential of minus the information discrepancy of a particular manifold path with respect to full Liouvillean evolution. The likelihood of a manifold member at a particular time is termed a consistency distribution and is analogous to a quantum wavefunction. The Lagrangian here is of modified generalized Onsager-Machlup form. For large times and long slow timescales the thermodynamics is of Öttinger form. The proposed path integral has connections with those occuring in the quantum theory of a particle in an external electromagnetic field. It is however entirely of a Wiener form and so practical to compute. Finally it is shown that providing certain reasonable conditions are met then there exists a unique steady-state consistency distribution.

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

Similar content being viewed by others

Notes

  1. With the proviso that \(p\) is always Gaussian as was originally assumed by Onsager and Machlup.

  2. More general invariants than energy of the dynamical system may also be considered.

  3. Note that this Lagrangian is quite distinct from that applying to the original Hamiltonian dynamics. It can in some sense be regarded as a slow variable Lagrangian for the system since \(\lambda \) specifies the slow variable expectation values via a Legendre transform.

  4. The energy function for TBH is simply the sum of the squares of the spectral mode amplitudes meaning the Gibbs measure is a Gaussian with uncorrelated modes and equal variances proportional to the conserved energy of the system.

  5. Note that we use the terminology distribution here to avoid confusion with the approximating trial densities. The consistency distribution is a function (or distribution) of the coordinates \(\lambda \) which specify the position within the manifold of trial densities. The densities are defined on the original variables of the Hamiltonian system.

  6. If the trial distribution gives an invariant measure for the system this will not be the case.

  7. Strictly this identification as a projection is precise only in the limit as \(\Delta t\rightarrow 0\).

  8. We use \(M\) to denote the magnetic vector potential to avoid confusion with the resolved variable set \(A\).

  9. This can include both an electric potential and other potentials such as gravitation.

  10. Set by the static space-time Killing vector.

  11. Or Fourier/phase decompositions.

  12. This is strictly a Wick rotated Schrödinger equation i.e. a parabolic PDE of a diffusion-absorbtion type.

  13. Up to a scalar multiple and the addition of a function vanishing almost everywhere with respect to the Lebesgue measure.

References

  1. Amari, S., Nagaoka, H.: Methods of Information Geometry. Translations of Mathematical Monographs, AMS, Oxford University Press (2000)

  2. Battezzati, M.: Onsager principle for nonlinear mechanical systems modeled by stochastic dissipative equations. Arch. Mech. 64(2), 177–206 (2012)

    MATH  MathSciNet  Google Scholar 

  3. Ceperley, D.M.: Path integrals in the theory of condensed helium. Rev. Mod. Phys. 67(2), 279 (1995)

    Article  ADS  Google Scholar 

  4. Chaichian, M., Demichev, A.: Path integrals in physics. Vol. 1: Stochastic processes and quantum mechanics. IOP, London (2001)

  5. Conway, J.B.: A Course in Functional Analysis. Springer, Berlin (1990)

  6. Cover, T., Thomas, J.: Elements of Information Theory, 2nd edn. Wiley-Interscience, New York (2006)

    MATH  Google Scholar 

  7. Darve, E., Solomon, J., Kia, A.: Computing generalized Langevin equations and generalized Fokker-Planck equations. Proc. Natl Acad. Sci. 106(27), 10884–10889 (2009).

  8. Deininghaus, U., Graham, R.: Nonlinear point transformations and covariant interpretation of path integrals. Z. Phys. B Con. Mat. 34(2), 211–219 (1979)

    Article  ADS  MathSciNet  Google Scholar 

  9. Dekker, H.: On the path integral for diffusion in curved spaces. Phys. A 103(3), 586–596 (1980)

    Article  MathSciNet  Google Scholar 

  10. Du, Y.: Order Structure and Topological Methods in Nonlinear Partial Differential Equations: Vol. 1: Maximum Principles and Applications. World Scientific Publishing Company (2006)

  11. Einstein, A.: Theorie der Opaleszenz von homogenen Fluessigkeiten und Fluessigkeitsgemischen in der Naehe des kritischen Zustandes. Ann. Physik 33, 1275 (1910)

    Article  ADS  MATH  Google Scholar 

  12. Eyink, G.L.: Action principle in nonequilibrium statistical dynamics. Phys. Rev. E 54(4), 3419–3435 (1996)

    Article  ADS  MathSciNet  Google Scholar 

  13. Feynman, R.P., Hibbs, A.R., Styer, D.F.: Quantum Mechanics and Path Integrals. McGraw-Hill, New York (1965)

  14. Gardiner, C.W.: Handbook of Stochastic Methods for Physics. Springer, Chemistry and the Natural Sciences (2004)

    Book  MATH  Google Scholar 

  15. Graham, R.: Path integral formulation of general diffusion processes. Z. Phys. B Con. Mat. 26(3), 281–290 (1977)

    Article  ADS  Google Scholar 

  16. Haken, H.: Generalized Onsager-Machlup function and classes of path integral solutions of the Fokker-Planck equation and the Master equation. Z. Phys. B Con. Mat. 24(3), 321–326 (1976)

    Article  ADS  MathSciNet  Google Scholar 

  17. Hanche-Olsen, H., Holden, H.: The Kolmogorov-Riesz compactness theorem. Expositiones Mathematicae 28, 385–394 (2010)

    Article  MATH  MathSciNet  Google Scholar 

  18. Kleeman, R., Turkington, B.E.: A nonequilibrium statistical model of spectrally truncated Burgers-Hopf dynamics. Commun. Pure Appl. Math. 67(12), 1905–1946 (2014)

  19. Kraichnan, R.H.: Variational method in turbulence theory. Phys. Rev. Lett. 42(19), 1263–1266 (1979)

    Article  ADS  Google Scholar 

  20. Landau, L.D., Lifshitz, E.M.: Quantum Mechanics Non-relativistic Theory. Pergamon, London (1965)

    MATH  Google Scholar 

  21. Lavenda, B.H.: On the validity of the Onsager-Machlup postulate for nonlinear stochastic processes. Found. Phys. 9(5–6), 405–420 (1979)

    Article  ADS  MathSciNet  Google Scholar 

  22. Lieb, E.H., Loss, M.: Analysis. American Mathematical Society, Providence (2001)

  23. Onsager, L., Machlup, S.: Fluctuations and irreversible processes. Phys. Rev. 91(6), 1505–1512 (1953)

    Article  ADS  MATH  MathSciNet  Google Scholar 

  24. Öttinger, H.C.: Beyond Equilibrium Thermodynamics. Wiley-Interscience (2005)

  25. Roncadelli, M.: New path integral representation of the quantum mechanical propagator. J. Phys. A-Math. Gen. 25(16), L997 (1992)

    Article  ADS  MATH  MathSciNet  Google Scholar 

  26. Taniguchi, T., Cohen, E.G.D.: Onsager-Machlup theory for nonequilibrium steady states and fluctuation theorems. J. Stat. Phys. 126(1), 1–41 (2007)

    Article  ADS  MATH  MathSciNet  Google Scholar 

  27. Turkington, B.: An optimization principle for deriving nonequilibrium statistical models of hamiltonian dynamics. J. Stat. Phys. 152, 569–597 (2013)

    Article  ADS  MATH  MathSciNet  Google Scholar 

  28. Ventsel, A.D., Freidlin, M.I.: On small random perturbations of dynamical systems. Russ. Math. Surv. 25(1), 1–55 (1970)

  29. Wald, R.M.: General Relativity. University of Chicago Press, Chicago (1984)

    Book  MATH  Google Scholar 

  30. Zubarev, D.N.: Nonequilibrium Statistical Thermodynamics. Plenum Press, New York (1974)

    Google Scholar 

  31. Zwanzig, R.: Nonequilibrium Statistical Mechanics. Oxford University Press, New York (2001)

    MATH  Google Scholar 

Download references

Acknowledgments

Comprehensive discussions with Bruce Turkington on matters related to the present contribution are very gratefully acknowledged. Useful discussions on related matters over many years with Andy Majda are also acknowledged. This paper is dedicated to my mother Annette.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Richard Kleeman.

Appendices

Appendix A: Some Useful Relations

Define the expectation bracket

$$\begin{aligned} \left\langle F\right\rangle \equiv \int F\hat{p} \end{aligned}$$

for a general function of the state variables and time \(F\). We have now

$$\begin{aligned} \frac{\partial \left\langle F\right\rangle }{\partial t}-\left\langle L(F)\right\rangle&= \left\langle \frac{\partial F}{\partial t}\right\rangle +\int (F\hat{p}_{t}-L(F)\hat{p})\nonumber \\&= \left\langle \frac{\partial F}{\partial t}\right\rangle +\int (F(\partial _{t}+L)\hat{p}\nonumber \\&= \left\langle \frac{\partial F}{\partial t}\right\rangle +\int FR\hat{p}\nonumber \\&= \left\langle \frac{\partial F}{\partial t}+FR\right\rangle \end{aligned}$$
(50)

where we are using the anti-Hermitian nature of \(L\) on the second line. Setting \(F=1\) we obtain immediately that

$$\begin{aligned} \left\langle R\right\rangle =0 \end{aligned}$$
(51)

Now it is easily derived from the definition of \(R\) and the form of the exponential family of distributions that

$$\begin{aligned} R=\dot{\lambda }^{t}(A-a)+\lambda {}^{t}LA \end{aligned}$$
(52)

which when combined with (51) yields

$$\begin{aligned} \lambda ^{t}\left\langle LA\right\rangle =0 \end{aligned}$$
(53)

The anti-Hermitian nature of \(L\) also allows us to deduce the following two useful relations (using the summation convention and vector/matrix indices for clarity):

(54)

where we are using the fact that \(L\) annihilates \(E\) and (53) for the second last step. Combining (53) and (54) we obtain

$$\begin{aligned} \lambda ^{t}h\lambda =0 \end{aligned}$$

In a completely analogous way to (54) we deduce that

$$\begin{aligned} \left\langle L^{2}(A_{i})\right\rangle =-\lambda _{j}\left\langle LA_{j}LA_{i}\right\rangle \equiv -k_{ij}\lambda _{j}. \end{aligned}$$

and more generally

$$\begin{aligned} \left\langle L^{n}A_{j}\right\rangle =-\lambda _{i}\left\langle LA_{i}L^{n-1}A_{j}\right\rangle \end{aligned}$$
(55)

It is easily shown also that

$$\begin{aligned} \frac{\partial M_{i}}{\partial \lambda _{j}}=\left\langle L(A_{i})\left( A_{j}-a_{j}\right) \right\rangle =h_{ji} \end{aligned}$$

Appendix B: Section 8 Theorem Proof

We first establish that the operator \(K\) is bounded with respect to the \(L_{1}\) norm. Consider the effect of \(K\) on a distribution \(\psi \) with \(L_{1}\) norm unity:

$$\begin{aligned} \left\| K\psi \right\| _{1}&= \int \left| \int R\psi d\kappa \right| d\lambda \\&\le \int \left[ \int \left| R\psi \right| d\kappa \right] d\lambda \\&= \iint N\exp \left[ -IL_{irr}-IL_{rev}\right] \left| \psi \right| d\kappa d\lambda \\&\le \iint N\exp \left[ -IL_{irr}\right] \left| \psi \right| d\kappa d\lambda \\&= \int \left| \psi \right| d\kappa =1 \end{aligned}$$

where on line 4 we have used the fundamental fact derived in section 2 that \(IL_{rev}\ge 0\) while the last line follows after switching variables of integration and using the normalization condition which also holds for \(\exp \left[ -IL_{irr}\right] \).

We further establish that the image of \(K\) is totally bounded which means establishing the additional two properties (48) and (49).

From condition 1. of the Theroem we deduce that there exists a \(\left| \kappa _{0}\right| \) such that

$$\begin{aligned} \exp \left( -IL_{rev}\right) <\frac{\epsilon }{2}\qquad if\;\left| \kappa \right| >\left| \kappa _{0}\right| \end{aligned}$$

Consider now the bounded region \(\left| \kappa \right| \le \left| \kappa _{0}\right| \). From condition 2. of the Theorem; the region boundedness and the fact that \(\exp \left( -IL_{irr}\right) \) is Gaussian in \(\lambda \), if follows that there exists an \(r(\epsilon )>0\) such that for all \(\left| \kappa \right| \le \left| \kappa _{0}\right| \)

$$\begin{aligned} \int \limits _{\left| \lambda \right| >R(\epsilon )}N\exp \left( -IL_{irr}\right) d\lambda <\frac{\epsilon }{2} \end{aligned}$$

Thus

$$\begin{aligned} \int \limits _{\left| \lambda \right| >r(\epsilon )}\left| K\psi \right| d\lambda =&\int \limits _{\left| \lambda \right| >r(\epsilon )}d\lambda \left| \int d\kappa N\exp \left( -IL_{irr}-IL_{rev}\right) \psi \right| \\ \le&\int \limits _{\left| \lambda \right| >r(\epsilon )}d\lambda \int d\kappa N\exp \left( -IL_{irr}-IL_{rev}\right) \left| \psi \right| \\ \le&\int \limits _{\left| \lambda \right| >r(\epsilon )}d\lambda \int \limits _{\left| \kappa \right| \le \left| \kappa _{0}\right| }d\kappa N\exp \left( -IL_{irr}\right) \left| \psi \right| \\&+\int \limits _{\left| \kappa \right| >\left| \kappa _{0}\right| }d\kappa \exp \left( -IL_{rev}\right) \left| \psi \right| \\ <&\frac{\epsilon }{2}+\frac{\epsilon }{2} \end{aligned}$$

which establishes (48).

To establish the other required property consider an arbitrary \(\rho _{t}>0\) and all \(\gamma \) with \(\left| \gamma \right| <\rho _{t}\). The triangle inequality plus (48) implies that

$$\begin{aligned} \int \limits _{\left| \lambda \right| >r(\frac{\epsilon }{4})+\rho _{t}}\left| K\psi \left( \lambda +\gamma \right) -K\psi (\lambda )\right| d\lambda&\le \int \limits _{\left| \lambda \right| >r(\frac{\epsilon }{4})+\rho _{t}}\left| K\psi \left( \lambda +\gamma \right) \right| d\lambda \nonumber \\ +\int \limits _{\left| \lambda \right| >r(\frac{\epsilon }{4})+\rho _{t}}\left| K\psi (\lambda )\right| d\lambda&\le \frac{\epsilon }{4}+\frac{\epsilon }{4} \end{aligned}$$
(56)

Set \(S(\epsilon ,\rho )=R(\frac{\epsilon }{4})+\rho \) and \(V_{\epsilon \rho _{t}}\) the volume of the region \(Z:\left| \lambda \right| \le S(\epsilon ,\rho _{t})\). Let \(\kappa _{0}\) be such that \(\left| \kappa \right| >\left| \kappa _{0}\right| \) implies that

$$\begin{aligned} \exp \left( -IL_{rev}\right) <\frac{\epsilon }{8V_{\epsilon \rho _{t}}} \end{aligned}$$
(57)

We have

$$\begin{aligned}&\int \limits _{\left| \lambda \right| \le S(\epsilon .\rho _{t})}\left| K\psi \left( \lambda +\gamma \right) -K\psi (\lambda )\right| d\lambda \nonumber \\&\quad =\int \limits _{\left| \lambda \right| \le S(\epsilon .\rho _{t})}\int \limits _{\left| \kappa \right| >\left| \kappa _{0}\right| }\exp \left( -IL_{red}\right) N\left| \exp \left( -IL_{irr}(\lambda +\gamma \right) -\exp \left( -IL_{irr}(\lambda \right) \right| \left| \psi \right| d\kappa d\lambda \nonumber \\&\qquad +\int \limits _{\left| \lambda \right| \le S(\epsilon .\rho _{t})}\left| f_{\kappa _{0}}(\lambda +\gamma )-f_{\kappa _{0}}(\lambda )\right| d\lambda \end{aligned}$$
(58)

with

$$\begin{aligned} f_{\kappa _{0}}\left( \lambda \right) \equiv \int \limits _{\left| \kappa \right| \le \left| \kappa _{0}\right| }R(\lambda ,\kappa )\psi (\lambda )d\kappa \end{aligned}$$

an integral transform defined on a bounded domain. The first integral on the RHS of (58) is easily shown using the triangle inequality; the inequality (57) and the non-negativity of the \(IL\) terms to be less than \(\frac{\epsilon }{4}\).

The function \(f_{\kappa _{0}}\) can be shown by standard arguments to be continuous since the integral transform is defined on a bounded domain and the function \(R\) is continuous with respect to the first argument. By the Heine-Cantor theorem it is therefore uniformly continuous on the bounded region \(Z\). It follows that there exists a \(\rho _{U}\) such that for all \(\gamma :\left| \gamma \right| \le \rho _{U}\) and all \(\lambda \in Z\)

$$\begin{aligned} \left| f_{\kappa _{0}}(\lambda +\gamma )-f_{\kappa _{0}}(\lambda )\right| <\frac{\epsilon }{4V_{\epsilon \rho _{t}}} \end{aligned}$$
(59)

and so for such \(\gamma \) the second integral from (58) is also less than \(\frac{\epsilon }{4}.\) Compare now \(\rho _{U}\) and \(\rho _{t}\). If \(\rho _{U}\ge \rho _{t}\) then we can replace \(\rho _{U}\) with \(\rho _{t}\) in the last integral inequality discussed and obtain the required inequality (49) by combining the three inequalities derived from (56) and (58). Conversely if \(\rho _{U}<\rho _{t}\) then \(V_{\epsilon \rho }<V_{\epsilon \rho _{t}}\) Thus inequalities (57) and (59) still hold if we use \(\rho _{U}\) in place of \(\rho _{t}\). Furthermore the newly defined bounded \(Z\) is a subset of the old \(Z\) whence the uniform continuity just discussed holds with the same \(\rho _{U}\) and hence we are done.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Kleeman, R. A Path Integral Formalism for Non-equilibrium Hamiltonian Statistical Systems. J Stat Phys 158, 1271–1297 (2015). https://doi.org/10.1007/s10955-014-1149-x

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s10955-014-1149-x

Keywords

Navigation