Skip to main content
Log in

Path Integral Derivation and Numerical Computation of Large Deviation Prefactors for Non-equilibrium Dynamics Through Matrix Riccati Equations

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

Abstract

For many non-equilibrium dynamics driven by small noise, in physics, chemistry, biology, or economy, rare events do matter. Large deviation theory then explains that the leading order term of the main statistical quantities have an exponential behavior. The exponential rate is often obtained as the infimum of an action, which is minimized along an instanton. In this paper, we consider the computation of the next order sub-exponential prefactors, which are crucial for a large number of applications. Following a path integral approach, we derive the dynamics of the Gaussian fluctuations around the instanton and compute from it the sub-exponential prefactors. As might be expected, the formalism leads to the computation of functional determinants and matrix Riccati equations. By contrast with the cases of equilibrium dynamics with detailed balance or generalized detailed balance, we stress the specific non locality of the solutions of the Riccati equation: the prefactors depend on fluctuations all along the instanton and not just at its starting and ending points. We explain how to numerically compute the prefactors. The case of statistically stationary quantities requires considerations of non trivial initial conditions for the matrix Riccati equation.

This is a preview of subscription content, log in via an institution to check access.

Access this article

Subscribe and save

Springer+ Basic
$34.99 /Month
  • Get 10 units per month
  • Download Article/Chapter or eBook
  • 1 Unit = 1 Article or 1 Chapter
  • Cancel anytime
Subscribe now

Buy Now

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

Similar content being viewed by others

Data Availability

This theoretical work uses no external dataset.

Notes

  1. Thermodynamic potentials are usually defined as static properties, independently of the dynamics, but they also appear as quasipotential in effective dynamical theory, a classical example being macroscopic fluctuation theory [3].

References

  1. Abbot, D.S., Webber, R.J., Hadden, S., Weare, J.: Rare event sampling improves mercury instability statistics. arXiv:2106.09091 (2021)

  2. Berglund, N.: Kramers’ law: validity, derivations and generalisations. Markov Process. Relat. Fields 19(3), 459–490 (2013)

    MathSciNet  MATH  Google Scholar 

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

    Article  ADS  MathSciNet  Google Scholar 

  4. Bouchet, F., Reygner, J.: Generalisation of the Eyring–Kramers transition rate formula to irreversible diffusion processes. Ann. Henri Poincaré 17(12), 3499–3532 (2016)

    Article  ADS  MathSciNet  Google Scholar 

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

    Article  ADS  MathSciNet  Google Scholar 

  6. Bouchet, F., Rolland, J., Simonnet, E.: Rare event algorithm links transitions in turbulent flows with activated nucleations. Phys. Rev. Lett. 122(7), 074,502 (2019)

    Article  Google Scholar 

  7. Bovier, A., Eckhoff, M., Gayrard, V., Klein, M.: Metastability in reversible diffusion processes. I. Sharp asymptotics for capacities and exit times. J. Euro. Math. Soc. 6(4), 399–424 (2004)

    Article  MathSciNet  Google Scholar 

  8. Callan, C.G., Jr., Coleman, S.R.: The fate of the false vacuum. 2. First quantum corrections. Phys. Rev. D 16, 1762–1768 (1977). https://doi.org/10.1103/PhysRevD.16.1762

    Article  ADS  Google Scholar 

  9. Cohen, J.K., Lewis, R.M.: A ray method for the asymptotic solution of the diffusion equation. IMA J. Appl. Math. 3(3), 266–290 (1967)

    Article  MathSciNet  Google Scholar 

  10. Coleman, S.R.: The uses of instantons. Subnucl. Ser. 15, 805 (1979)

    Google Scholar 

  11. Dematteis, G., Grafke, T., Vanden-Eijnden, E.: Rogue waves and large deviations in deep sea. Proc. Natl. Acad. Sci. USA 115(5), 855–860 (2018)

    Article  ADS  MathSciNet  Google Scholar 

  12. Dembo, A., Zeitouni, O.: Large Deviations Techniques and Applications, Stochastic Modelling and Applied Probability, vol. 38. Springer, Berlin (2010). (Corrected reprint of the second edition)

  13. Dieci, L., Eirola, T.: Positive definiteness in the numerical solution of Riccati differential equations. Numer. Math. 67, 303–313 (1994)

    Article  MathSciNet  Google Scholar 

  14. Dieci, L., Eirola, T.: Preserving monotonicity in the numerical solution of Riccati differential equations. Numer. Math. 74, 35–47 (1996)

    Article  MathSciNet  Google Scholar 

  15. Dubois, F., Saïdi, A.: Unconditionnally stable scheme for Riccati equation. ESAIM Proc. 8, 39–52 (2000)

    Article  Google Scholar 

  16. Ferré, G., Grafke, T.: Approximate optimal controls via instanton expansion for low temperature free energy computation. arXiv:2011.10990

  17. Freidlin, M.I. and Wentzell, A.D.: Random perturbations of dynamical systems, Grundlehren der Mathematischen Wissenschaften, vol. 260. Springer, Heidelberg (2012). Translated from the 1979 Russian original by Joseph Szücs. Third edition

  18. Grafke, T., Vanden-Eijnden, E.: Numerical computation of rare events via large deviation theory. Chaos 29(6), 063118 (2019)

    Article  ADS  MathSciNet  Google Scholar 

  19. Grafke, T., Grauer, R., Schäfer, T.: Instanton filtering for the stochastic burgers equation. J. Phys. A 46(6), 062,002 (2013)

    Article  MathSciNet  Google Scholar 

  20. Grafke, T., Grauer, R., Schindel, S.: Efficient computation of instantons for multi-dimensional turbulent flows with large scale forcing. Commun. Comput. Phys. 18(3), 577–592 (2015)

    Article  MathSciNet  Google Scholar 

  21. Grafke, T., Schäfer, T., Vanden-Eijnden, E.: Sharp Asymptotic Estimates for Expectations, Probabilities, and Mean First Passage Times in Stochastic Systems with Small Noise. arXiv:2103.04837

  22. Graham, R.: Macroscopic potentials, bifurcations and noise in dissipative systems. Noise Nonlinear Dyn. Syst. 1, 225–278 (1988)

    Google Scholar 

  23. Heymann, M., Vanden-Eijnden, E.: The geometric minimum action method: a least action principle on the space of curves. Commun. Pure Appl. Math. 61(8), 1052–1117 (2008)

    Article  MathSciNet  Google Scholar 

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

    Google Scholar 

  25. Landim, C., Mariani, M., Seo, I.: Dirichlet’s and Thomson’s principles for non-selfadjoint elliptic operators with application to non-reversible metastable diffusion processes. Arch. Ration. Mech. Anal. 231(2), 887–938 (2019)

    Article  MathSciNet  Google Scholar 

  26. Langer, J.S.: Theory of the condensation point. Ann. Phys. 41, 108–157 (1967). https://doi.org/10.1016/0003-4916(67)90200-X

    Article  ADS  Google Scholar 

  27. Langer, J.: Excitation chains at the glass transition. Phys. Rev. Lett. 97(11), 115,704 (2006)

    Article  Google Scholar 

  28. Laurie, J., Bouchet, F.: Computation of rare transitions in the barotropic quasi-geostrophic equations. New J. Phys. (2015). https://doi.org/10.1088/1367-2630/17/1/015009

    Article  Google Scholar 

  29. Lee, J., Seo, I.: Non-reversible metastable diffusions with Gibbs invariant measure I: Eyring–Kramers formula. arXiv:2008.08291

  30. Lu, Y., Stuart, A.M., Weber, H.: Gaussian approximations for transition paths in molecular dynamics. SIAM J. Math. Anal. 49(4), 3005–3047 (2017)

    Article  MathSciNet  Google Scholar 

  31. Ludwig, D.: Persistence of dynamical systems under random perturbations. SIAM Rev. 17(4), 605–640 (1975)

    Article  MathSciNet  Google Scholar 

  32. Maier, R.S., Stein, D.L.: Limiting exit location distributions in the stochastic exit problem. SIAM J. Appl. Math. 57(3), 752–790 (1997)

    Article  MathSciNet  Google Scholar 

  33. Paskal, N., Cameron, M.: An efficient jet marcher for computing the quasipotential for 2D SDEs. To appear in J. Sci. Comput

  34. Ragone, F., Wouters, J., Bouchet, F.: Computation of extreme heat waves in climate models using a large deviation algorithm. Proc. Natl. Acad. Sci. U.S.A. 115(1), 24–29 (2018). https://doi.org/10.1073/pnas.1712645115

    Article  ADS  MathSciNet  MATH  Google Scholar 

  35. Sanz-Alonso, D., Stuart, A.M.: Gaussian approximations of small noise diffusions in Kullback–Leibler divergence. Commun. Math. Sci. 15(7), 2087–2097 (2017)

    Article  MathSciNet  Google Scholar 

  36. Schuss, Z.: Theory and Applications of Stochastic Processes: An Analytical Approach, Applied Mathematical Sciences, vol. 170. Springer, New York (2010)

    Book  Google Scholar 

  37. Simonnet, E., Rolland, J., Bouchet, F.: Multistability and rare spontaneous transitions in barotropic \(\beta \)-plane turbulence. J. Atmos. Sci. 78(6), 1889–1911 (2021)

    ADS  Google Scholar 

  38. Vanden-Eijnden, E., Heymann, M.: The geometric minimum action method for computing minimum energy paths. J. Chem. Phys. 128(6), 61–103 (2008)

    Article  Google Scholar 

  39. Woillez, E., Bouchet, F.: Instantons for the destabilization of the inner solar system. Phys. Rev. Lett. 125(2), 021,101 (2020)

    Article  MathSciNet  Google Scholar 

  40. Woillez, E., Zhao, Y., Kafri, Y., Lecomte, V., Tailleur, J.: Activated escape of a self-propelled particle from a metastable state. Phys. Rev. Lett. 122, 258,001 (2019). https://doi.org/10.1103/PhysRevLett.122.258001

    Article  Google Scholar 

  41. Zinn-Justin, J.: Quantum Field Theory and Critical Phenomena. Clarendon Press, Oxford (1996)

    MATH  Google Scholar 

Download references

Acknowledgements

We would like to thank A. Alfonsi, A. Levitt, G. Stoltz for useful comments on the numerical integration of matrix Riccati equations. We also thank two anonymous referees for their careful reading of the paper which helped improving the presentation of our results. The research leading to these results has received funding from the European Research Council under the European Union’s seventh Framework Programme (FP7/2007-2013 Grant Agreement No. 616811). During the last stage of this work, F. Bouchet received support by a subagreement from the Johns Hopkins University with funds provided by Grant No. 663054 from Simons Foundation. Its contents are solely the responsibility of the authors and do not necessarily represent the official views of Simons Foundation or the Johns Hopkins University. J. Reygner is supported by the French National Research Agency (ANR) under the programs EFI (ANR-17-CE40-0030) and QuAMProcs (ANR-19-CE40-0010).

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Freddy Bouchet.

Ethics declarations

Conflict of interest

The authors have no relevant financial or non-financial interests to disclose, have no competing interests to declare that are relevant to the content of this article. The authors declare no conflict of interest relevant to the content of this article.

Additional information

Communicated by Jorge Kurchan.

Publisher's Note

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

Appendix

Appendix

The Appendix is organized as follows. In Sect. 1, the notion of time ordered exponential is introduced and a few properties are stated. Section 2 is dedicated to the resolution of the backward matrix Riccati equation appearing in (35). Finally, Sect. 3 presents the proof of (38).

Throughout the Appendix, we work under Assumptions (A1–4). Furthermore, as is argued in the beginning of Sect. 3, our overall purpose here is to emphasize the connections between fluctuation paths, large deviation prefactors, functional Gaussian determinants and matrix Riccati equations. Therefore, we chose not to obscure the exposition of our arguments with the exhaustive mathematical justification of technical details, which can however be checked on a case-by-case basis.

1.1 Time Ordered Exponentials

Throughout this section we let \(A=(A_t)_{t \le 0}\) be a bounded family of matrices of size \(d \times d\). The choice of \((-\infty ,0]\) as the set of times is convenient for our purpose but the contents of this section could easily be adapted to any interval.

For all \(t_0, t \le 0\), we denote by

$$\begin{aligned} M_t = \left\lfloor \exp \left( \int _{s=t_0}^t A_s \mathrm {d}s\right) \right\rfloor \in \mathbb {R}^{d \times d} \end{aligned}$$
(108)

the solution to the (two-sided) Cauchy problem

$$\begin{aligned} \left\{ \begin{aligned} \dot{M}_t&= A_t M_t, \quad t \le 0,\\ M_{t_0}&= I_d. \end{aligned}\right. \end{aligned}$$
(109)

It is called the time ordered exponential of A from \(t_0\) to t. It is related with the notion of path ordering which is in particular used in quantum field theory. However, in the context of the present article, we are only dealing with bounded, finite-dimensional matrices, therefore the following properties, which will be used in the sequel, are elementary consequences of the Cauchy–Lipschitz theorem for linear ordinary differential equations.

  1. (i)

    For all \(t_0, t_1, t_2 \le 0\), \(\lfloor \exp (\int _{s=t_0}^{t_2} A_s \mathrm {d}s)\rfloor = \lfloor \exp (\int _{s=t_1}^{t_2} A_s \mathrm {d}s)\rfloor \lfloor \exp (\int _{s=t_0}^{t_1} A_s \mathrm {d}s)\rfloor \).

  2. (ii)

    For all \(t_0, t_1 \le 0\), \(\lfloor \exp (\int _{s=t_0}^{t_1} A_s \mathrm {d}s)\rfloor \) is invertible and \(\lfloor \exp (\int _{s=t_0}^{t_1} A_s \mathrm {d}s)\rfloor ^{-1} = \lfloor \exp (\int _{s=t_1}^{t_0} A_s \mathrm {d}s)\rfloor \).

  3. (iii)

    For all \(t_0, t \le 0\), \(\det \lfloor \exp (\int _{s=t_0}^t A_s \mathrm {d}s)\rfloor = \det (\exp (\int _{s=t_0}^t A_s \mathrm {d}s)) = \exp (\int _{s=t_0}^t \text {tr}A_s \mathrm {d}s)\). As a consequence, if \((M_t)_{t \le 0}\) solves the matrix ordinary differential equation \(\dot{M}_t = A_t M_t\), then \(m_t = \det M_t\) satisfies \(\dot{m}_t = m_t \text {tr}A_t\).

1.2 Solving the Backward Matrix Riccati Equation

In this section, we solve the backward matrix Riccati equation which corresponds to the second line of (35). In order to alleviate the notations, we drop the superscript notation x on the quantities \(\varphi ^x\), \(Q^x_s\), \(R^x_s\), etc., therefore we are led to consider the backward Cauchy problem

$$\begin{aligned} \left\{ \begin{aligned}&\dot{K}_t = K_t^2 + Q_t^\top K_t + K_t Q_t - 2R_t, \quad t < 0,\\&\lim _{t \uparrow 0} K_t^{-1} = 0. \end{aligned}\right. \end{aligned}$$
(110)

We employ a quadrature method. Let us first define

$$\begin{aligned} K^0_t = -2\nabla ^2 V(\varphi _t). \end{aligned}$$
(111)

By Proposition 1 and Remark 6, we have

$$\begin{aligned} \dot{K}^0_t = \left( K^0_t\right) ^2 + Q_t^\top K^0_t + K^0_t Q_t - 2R_t; \end{aligned}$$
(112)

in other words, \(K^0_t\) is a particular solution to the backward Riccati equation (110), but with a different behavior when \(t \uparrow 0\).

For all \(t \le 0\), we now let

$$\begin{aligned} A_t = \nabla ^2 V(\varphi _t) + \nabla \ell (\varphi _t) = 2\nabla ^2 V(\varphi _t) - Q_t, \end{aligned}$$
(113)

and introduce

$$\begin{aligned} Z_t = \int _{s=t}^0 \left\lfloor \exp \left( \int _{r=s}^t A_r \mathrm {d}r\right) \right\rfloor \left\lfloor \exp \left( \int _{r=s}^t A_r \mathrm {d}r\right) \right\rfloor ^{\top }\mathrm {d}s. \end{aligned}$$
(114)

Notice that since \(\nabla V\) and \(\ell \) are assumed to be smooth, and by definition, the fluctuation path \(\varphi ^x\) is bounded in \(\mathbb {R}^d\), the family of matrices \((A_t)_{t \le 0}\) is bounded and therefore matches the setting of Sect. 1. Then \(Z_t\) solves the time-dependent Lyapunov equation

$$\begin{aligned} \dot{Z}_t = -I_d + A_t Z_t + Z_t A_t^\top , \end{aligned}$$
(115)

so that

$$\begin{aligned} \frac{\mathrm {d}}{\mathrm {d}t} Z^{-1}_t = -Z_t^{-1} \dot{Z}_t Z^{-1}_t = (Z^{-1}_t)^2 - A_t^{\top } Z_t^{-1} - Z_t^{-1} A_t. \end{aligned}$$
(116)

We may now define

$$\begin{aligned} K_t = K^0_t + Z_t^{-1}, \end{aligned}$$
(117)

for all \(t<0\), and check that \(\dot{K}_t = K_t^2 + Q_t^\top K_t + K_t Q_t - 2R_t\) using (111), (112), (113) and (116). Furthermore, \(K^0_t\) remains bounded when \(t \uparrow 0\) whereas \(|t|Z_t^{-1}\) converges to \(I_d\), which implies that \(K_t^{-1}\) converges to 0.

As a conclusion, \(K_t\) is the solution to (110), which yields the formula (36) in Sect. 3.2.

Remark 8

It follows from the expression of \(K_t\) that \(|t|K_t\) converges to \(I_d\) when \(t \uparrow 0\). This implies that \(|t|^d\det K_t\) converges to 1, so that the third condition in (34) takes the more explicit form

$$\begin{aligned} \lim _{t \uparrow 0} |t|^{-d} \eta _t = 1. \end{aligned}$$
(118)

1.3 Proof of (38)

In this section we prove the identity (38). The solution \((K_t)_{t<0}\) to the backward matrix Riccati equation was constructed in the previous section, and it is easily observed that \((\eta _t)_{t<0}\) is defined up to a multiplicative constant by

$$\begin{aligned} \forall t_1, t_2 < 0, \quad \frac{\eta _{t_2}}{\eta _{t_1}} = \exp \left( -\int \limits _{s=t_1}^{t_2} \text {tr}K_s \mathrm {d}s\right) . \end{aligned}$$
(119)

The appropriate multiplicative constant shall be chosen in accordance with the third condition of (34), which shall then provide the correct \(t \rightarrow -\infty \) limit for \(\eta _t\).

We first introduce the notation

$$\begin{aligned} U_t = \int \limits _{s=t}^0 \left\lfloor \exp \left( \int _{r=s}^0 A_r \mathrm {d}r\right) \right\rfloor \left\lfloor \exp \left( \int \limits _{r=s}^0 A_r \mathrm {d}r\right) \right\rfloor ^{\top }\mathrm {d}s, \end{aligned}$$
(120)

for all \(t \le 0\).

Lemma 1

For all \(t_1, t_2 < 0\),

$$\begin{aligned} \frac{\eta _{t_2}}{\eta _{t_1}} = \frac{\det U_{t_2}}{\det U_{t_1}}\exp \left( 2 \int _{s=t_1}^{t_2} \Delta V(\varphi _s) \mathrm {d}s\right) . \end{aligned}$$
(121)

Proof

We first inject the formula (117) for \(K_s\) into (119) and obtain

$$\begin{aligned} \frac{\eta _{t_2}}{\eta _{t_1}} = \exp \left( 2 \int \limits _{s=t_1}^{t_2} \Delta V(\varphi _s) \mathrm {d}s - \int \limits _{s=t_1}^{t_2} \text {tr}Z_s^{-1} \mathrm {d}s\right) . \end{aligned}$$
(122)

By Property (i) of time ordered exponentials, for all \(t<0\),

$$\begin{aligned} Z_t = \left\lfloor \exp \left( \int \limits _{r=0}^t A_r \mathrm {d}r\right) \right\rfloor U_t\left\lfloor \exp \left( \int \limits _{r=0}^t A_r \mathrm {d}r\right) \right\rfloor ^\top , \end{aligned}$$
(123)

so that

$$\begin{aligned} \left\lfloor \exp \left( \int \limits _{r=t}^0 A_r \mathrm {d}r\right) \right\rfloor ^\top = Z_t^{-1}\left\lfloor \exp \left( \int \limits _{r=0}^t A_r \mathrm {d}r\right) \right\rfloor U_t, \end{aligned}$$
(124)

where we have used Property (ii) of time ordered exponentials. On the other hand,

$$\begin{aligned} \dot{U}_t = -\left\lfloor \exp \left( \int \limits _{r=t}^0 A_r \mathrm {d}r\right) \right\rfloor \left\lfloor \exp \left( \int \limits _{r=t}^0 A_r \mathrm {d}r\right) \right\rfloor ^{\top }, \end{aligned}$$
(125)

therefore

$$\begin{aligned} \dot{U}_t = -\left\lfloor \exp \left( \int \limits _{r=t}^0 A_r \mathrm {d}r\right) \right\rfloor Z_t^{-1}\left\lfloor \exp \left( \int \limits _{r=0}^t A_r \mathrm {d}r\right) \right\rfloor U_t \end{aligned}$$
(126)

and by Property (iii) of time ordered exponentials, \(m_t = \det U_t\) satisfies

$$\begin{aligned} \dot{m}_t = - m_t\text {tr}\left( \left\lfloor \exp \left( \int \limits _{r=t}^0 A_r \mathrm {d}r\right) \right\rfloor Z_t^{-1}\left\lfloor \exp \left( \int \limits _{r=0}^t A_r \mathrm {d}r\right) \right\rfloor \right) \end{aligned}$$
(127)

which reduces to

$$\begin{aligned} \dot{m}_t = - m_t\text {tr}Z_t^{-1} \end{aligned}$$
(128)

thanks to Property (ii) of time ordered exponentials again. As a consequence,

$$\begin{aligned} \exp \left( -\int \limits _{s=t_1}^{t_2} \text {tr}Z_s^{-1} \mathrm {d}s\right) = \frac{m_{t_2}}{m_{t_1}} = \frac{\det U_{t_2}}{\det U_{t_1}}, \end{aligned}$$
(129)

which completes the proof.

In the next lemma we describe the \(t_2 \uparrow 0\) limit of the ratio \(\eta _{t_2}/\eta _{t_1}\).

Lemma 2

For all \(t_1 < 0\),

$$\begin{aligned} \eta _{t_1} = \det U_{t_1} \exp \left( -2 \int \limits _{s=t_1}^0 \Delta V(\varphi _s) \mathrm {d}s\right) . \end{aligned}$$
(130)

Proof

We fix \(t_1<0\) and let \(t_2\) grow to 0 in (121). By the definition of \(U_t\), \(|t_2|^{-1}U_{t_2}\) converges to \(I_d\), so that by Remark 8,

$$\begin{aligned} \lim _{t_2 \uparrow 0} \frac{\eta _{t_2}}{\det U_{t_2}} = 1, \end{aligned}$$
(131)

which completes the proof.

We finally address the \(t_1 \rightarrow -\infty \) limit of the expression obtained for \(\eta _{t_1}\).

Lemma 3

We have

$$\begin{aligned} \lim _{t_1 \rightarrow -\infty } \eta _{t_1} = \frac{1}{2^d \det (\nabla ^2 V(\bar{x}))}\exp \left( 2\int \limits _{s=-\infty }^0 \text {div}\ell (\varphi _s)\mathrm {d}s\right) . \end{aligned}$$
(132)

Proof

Let \(t_1<0\). Using Properties (i) and (iii) of time ordered exponentials and (123), we rewrite

$$\begin{aligned} \begin{aligned} \eta _{t_1}&= \det \left( \exp \left( -\int \limits _{s=t_1}^0 \nabla ^2 V(\varphi _s)\mathrm {d}s\right) U_{t_1} \exp \left( -\int \limits _{s=t_1}^0 \nabla ^2 V(\varphi _s)\mathrm {d}s\right) \right) \\&= \exp \left( 2 \int \limits _{s=t_1}^0 \text {div}\ell (\varphi _s)\mathrm {d}s\right) \det \left( \exp \left( \int \limits _{s=t_1}^0 A_s \mathrm {d}s\right) U_{t_1} \exp \left( \int _{s=t_1}^0 A_s \mathrm {d}s\right) ^\top \right) \\&= \exp \left( 2 \int \limits _{s=t_1}^0 \text {div}\ell (\varphi _s)\mathrm {d}s\right) \det Z_{t_1}. \end{aligned} \end{aligned}$$
(133)

We are therefore led to compute the \(t_1 \rightarrow -\infty \) limit of \(Z_{t_1}\). To this aim we recall that \((Z_t)_{t \le 0}\) solves the time-dependent Lyapunov equation (115). In addition to the notation \(\bar{H} = \nabla ^2 V(\bar{x})\) introduced in Sect. 5.1, let us denote \(\bar{D} = \nabla \ell (\bar{x})\), so that taking the \(t \rightarrow -\infty \) limit of (115) shows that

$$\begin{aligned} \bar{Z} = \lim _{t \rightarrow -\infty } Z_t \end{aligned}$$
(134)

satisfies the stationary Lyapunov equation

$$\begin{aligned} (\bar{H}+\bar{D})\bar{Z} + \bar{Z}(\bar{H}+\bar{D})^\top = I_d. \end{aligned}$$
(135)

In addition,

  • evaluating the identity (12) at \(\bar{x}\) yields \(\bar{D}^\top \bar{H} + \bar{H}\bar{D} = 0\),

  • \(\bar{H}\) is assumed to be positive-definite.

As a consequence,

$$\begin{aligned} \bar{Z} = \frac{1}{2} \bar{H}^{-1}, \end{aligned}$$
(136)

from which we deduce that

$$\begin{aligned} \lim _{t_1 \rightarrow -\infty } \eta _{t_1} = \exp \left( 2 \int \limits _{s=-\infty }^0 \text {div}\ell (\varphi _s)\mathrm {d}s\right) \det \left( \frac{1}{2} \bar{H}^{-1}\right) , \end{aligned}$$
(137)

and the proof is completed.

Rights and permissions

Springer Nature or its licensor holds exclusive rights to this article under a publishing agreement with the author(s) or other rightsholder(s); author self-archiving of the accepted manuscript version of this article is solely governed by the terms of such publishing agreement and applicable law.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Bouchet, F., Reygner, J. Path Integral Derivation and Numerical Computation of Large Deviation Prefactors for Non-equilibrium Dynamics Through Matrix Riccati Equations. J Stat Phys 189, 21 (2022). https://doi.org/10.1007/s10955-022-02983-7

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1007/s10955-022-02983-7

Keywords

Navigation