Skip to main content
Log in

Non-equivalence of Dynamical Ensembles and Emergent Non-ergodicity

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

Abstract

Dynamical ensembles have been introduced to study constrained stochastic processes. In the microcanonical ensemble, the value of a dynamical observable is constrained to a given value. In the canonical ensemble a bias is introduced in the process to move the mean value of this observable. The equivalence between the two ensembles means that calculations in one or the other ensemble lead to the same result. In this paper, we study the physical conditions associated with ensemble equivalence and the consequences of non-equivalence. For continuous time Markov jump processes, we show that ergodicity guarantees ensemble equivalence. For non-ergodic systems or systems with emergent ergodicity breaking, we adapt a method developed for equilibrium ensembles to compute asymptotic probabilities while caring about the initial condition. We illustrate our results on the infinite range Ising model by characterizing the fluctuations of magnetization and activity. We discuss the emergence of non-ergodicity by showing that the initial condition can only be forgotten after a time that scales exponentially with the number of spins.

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

Similar content being viewed by others

Notes

  1. The alternative viewpoint is to consider the non-differentiability of the CGF as the criterion for determining ensemble equivalence (merging together partial equivalence and non-equivalence).

  2. This property is stated in problem 5, page 174 of Ref. [30].

References

  1. Baek, Y., Kafri, Y., Lecomte, V.: Dynamical phase transitions in the current distribution of driven diffusive channels. J. Phys. A Math. Theor. 51(10), 105001 (2018). http://stacks.iop.org/1751-8121/51/i=10/a=105001

  2. Barato, A.C., Chetrite, R.: A formal view on level 2.5 large deviations and fluctuation relations. J. Stat. Phys. 160(5), 1154–1172 (2015). https://doi.org/10.1007/s10955-015-1283-0

    Article  ADS  MathSciNet  MATH  Google Scholar 

  3. Barré, J., Mukamel, D., Ruffo, S.: Inequivalence of ensembles in a system with long-range interactions. Phys. Rev. Lett. 87, 030,601 (2001). https://doi.org/10.1103/PhysRevLett.87.030601

    Article  Google Scholar 

  4. Bertini, L., Faggionato, A., Gabrielli, D.: From level 2.5 to level 2 large deviations for continuous time Markov chains. Markov Process. Relat. Fields. 20(3), 545–562 (2012)

    MathSciNet  MATH  Google Scholar 

  5. Biroli, G., Garrahan, J.P.: Perspective: the glass transition. J. Chem. Phys. 138(12), 12A301 (2013). https://doi.org/10.1063/1.4795539

    Article  Google Scholar 

  6. Bouchet, F., Dauxois, T., Mukamel, D., Ruffo, S.: Phase space gaps and ergodicity breaking in systems with long-range interactions. Phys. Rev. E 77(1), 011,125 (2008). https://doi.org/10.1103/PhysRevE.77.011125

    Article  Google Scholar 

  7. Bouchet, F., Gupta, S., Mukamel, D.: Thermodynamics and dynamics of systems with long-range interactions. Phys. A Stat. Mech. Appl. 389(20), 4389–4405 (2010). https://doi.org/10.1016/j.physa.2010.02.024. http://www.sciencedirect.com/science/article/pii/S0378437110001512. Proceedings of the 12th International Summer School on Fundamental Problems in Statistical Physics

  8. Boyd, S., Vandenberghe, L.: Convex Optimization. Cambridge University Press, Cambridge (2004)

    Book  MATH  Google Scholar 

  9. Budini, A.A., Turner, R.M., Garrahan, J.P.: Fluctuating observation time ensembles in the thermodynamics of trajectories. J. Stat. Mech. Theory Exp. 2014(3), P03012 (2014). http://stacks.iop.org/1742-5468/2014/i=3/a=P03012

  10. Callen, H.B.: Thermodynamics and an Introduction to Thermostatistics. Wiley (1985). https://www.wiley.com/en-us/Thermodynamics+and+an+Introduction+to+Thermostatistics%2C+2nd+Edition-p-9780471862567

  11. Campa, A., Dauxois, T., Ruffo, S.: Statistical mechanics and dynamics of solvable models with long-range interactions. Phys. Repo. 480(3–6), 57–159 (2009). https://doi.org/10.1016/j.physrep.2009.07.001. http://www.sciencedirect.com/science/article/pii/S0370157309001586

  12. Chandler, D., Garrahan, J.P.: Dynamics on the way to forming glass: bubbles in space-time. Annu. Rev. Phys. Chem. 61(1), 191–217 (2010). https://doi.org/10.1146/annurev.physchem.040808.090405. https://doi.org/10.1146/annurev.physchem.040808.090405. PMID: 20055676

  13. Chetrite, R., Touchette, H.: Nonequilibrium microcanonical and canonical ensembles and their equivalence. Phys. Rev. Lett. 111, 120,601 (2013). https://doi.org/10.1103/PhysRevLett.111.120601

    Article  Google Scholar 

  14. Chetrite, R., Touchette, H.: Nonequilibrium Markov processes conditioned on large deviations. Ann. Henri Poincaré 16(9), 2005–2057 (2015). https://doi.org/10.1007/s00023-014-0375-8

    Article  ADS  MathSciNet  MATH  Google Scholar 

  15. Costeniuc, M., Ellis, R.S., Touchette, H.: Complete analysis of phase transitions and ensemble equivalence for the Curie–Weiss–Potts model. J. Math. Phys. 46(6), 063301 (2005). https://doi.org/10.1063/1.1904507

    Article  ADS  MathSciNet  MATH  Google Scholar 

  16. Csiszár, I., Shields, P.C.: Information Theory and Statistics: A Tutorial. Now Foundations and Trends (2004). https://doi.org/10.1561/0100000004

  17. Dinwoodie, I.H.: Identifying a large deviation rate function. Ann. Probab. 21(1), 216–231 (1993). https://doi.org/10.1214/aop/1176989402

    Article  MathSciNet  MATH  Google Scholar 

  18. Dinwoodie, I.H., Zabell, S.L.: Large deviations for exchangeable random vectors. Ann. Probab. 20(3), 1147–1166 (1992). https://doi.org/10.1214/aop/1176989683

    Article  MathSciNet  MATH  Google Scholar 

  19. Ellis, R.S., Touchette, H., Turkington, B.: Thermodynamic versus statistical nonequivalence of ensembles for the mean-field Blume–Emery–Griffiths model. Phys. A Stat. Mech. Appl. 335(3–4), 518–538 (2004). https://doi.org/10.1016/j.physa.2003.11.028. http://www.sciencedirect.com/science/article/pii/S0378437103011075

  20. Feller, W.: An Introduction to Probability Theory and Its Applications, vol. 2. Wiley, Hoboken (2008)

    MATH  Google Scholar 

  21. Ferré, G., Touchette, H.: Adaptive sampling of large deviations. J. Stat. Phys. (2018). https://doi.org/10.1007/s10955-018-2108-8

  22. Freidlin, M.I., Wentzell, A.D.: Random Perturbations of Dynamical Systems. Grundlehren der mathematischen Wissenschaften. Springer, New York (1984). https://doi.org/10.1007/978-3-642-25847-3

  23. 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(7), 073208 (2016). http://stacks.iop.org/1742-5468/2016/i=7/a=073208

  24. 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), 195,702 (2007). https://doi.org/10.1103/PhysRevLett.98.195702

    Article  MATH  Google Scholar 

  25. 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). http://stacks.iop.org/1751-8121/42/i=7/a=075007

  26. Glauber, R.J.: Time-dependant statistics of the Ising model. J. Math. Phys. 4, 294 (1963). https://doi.org/10.1063/1.1703954

    Article  ADS  MathSciNet  MATH  Google Scholar 

  27. Goldenfeld, N.: Lectures on Phase Transitions and the Renormalization Group. Perseus Books Publishing, L.L.C, New York (1992)

    MATH  Google Scholar 

  28. Grafke, T., Grauer, R., Schäfer, T.: The instanton method and its numerical implementation in fluid mechanics. J. Phys. A Math. Theor. 48(33), 333001 (2015). http://stacks.iop.org/1751-8121/48/i=33/a=333001

  29. Hedges, L.O., Jack, R.L., Garrahan, J.P., Chandler, D.: Dynamic order–disorder in atomistic models of structural glass formers. Science 323(5919), 1309–1313 (2009). https://doi.org/10.1126/science.1166665. http://science.sciencemag.org/content/323/5919/1309

  30. Horn, R.A., Johnson, C.R.: Matrix Analysis, 2nd edn. Cambridge University Press, Cambridge (2012). https://doi.org/10.1017/9781139020411

    Book  Google Scholar 

  31. Jack, R.L., Sollich, P.: Large deviations and ensembles of trajectories in stochastic models. Prog. Theor. Phys. Suppl. 184, 304–317 (2010). https://doi.org/10.1143/PTPS.184.304. http://ptps.oxfordjournals.org/content/184/304.abstract

  32. Jack, R.L., Garrahan, J.P., Chandler, D.: Space-time thermodynamics and subsystem observables in a kinetically constrained model of glassy materials. J. Chem. Phys. 125(18), 184509 (2006). https://doi.org/10.1063/1.2374885

    Article  ADS  Google Scholar 

  33. Keys, A.S., Chandler, D., Garrahan, J.P.: Using the \(s\) ensemble to probe glasses formed by cooling and aging. Phys. Rev. E 92, 022304 (2015). https://doi.org/10.1103/PhysRevE.92.022304

    Article  ADS  Google Scholar 

  34. Lawler, G.F., Sokal, A.D.: Bounds on the \(l^2\) spectrum for Markov chains and Markov processes: a generalization of Cheeger’s inequality. Trans. Am. Math. Soc. 309(2), 557–580 (1988). http://www.jstor.org/stable/2000925

  35. Lebowitz, J.L., Spohn, H.: A Gallavotti–Cohen-type symmetry in the large deviation functional for stochastic dynamics. J. Stat. Phys. 95(1), 333–365 (1999). https://doi.org/10.1023/A:1004589714161

    Article  ADS  MathSciNet  MATH  Google Scholar 

  36. Lecomte, V., Appert-Rolland, C., van Wijland, F.: Thermodynamic formalism for systems with Markov dynamics. J. Stat. Phys. 127(1), 51–106 (2007). https://doi.org/10.1007/s10955-006-9254-0

    Article  ADS  MathSciNet  MATH  Google Scholar 

  37. Levin, D.A., Yuval Peres, E.L.W.: Markov Chains and Mixing Times, vol. 58. American Mathematical Society, Providence (2008)

    Book  Google Scholar 

  38. Maes, C., Netočný, K.: Canonical structure of dynamical fluctuations in mesoscopic nonequilibrium steady states. EPL (Europhys. Lett.) 82(3), 30003 (2008). http://stacks.iop.org/0295-5075/82/i=3/a=30003

  39. Merolle, M., Garrahan, J.P., Chandler, D.: Space-time thermodynamics of the glass transition. Proc. Natl. Acad. Sci. 102(31), 10837–10840 (2005). https://doi.org/10.1073/pnas.0504820102. http://www.pnas.org/content/102/31/10837

  40. Monthus, C.: Non-equilibrium steady states: maximization of the Shannon entropy associated with the distribution of dynamical trajectories in the presence of constraints. J. Stat. Mech. 2011(03), P03008 (2011). http://stacks.iop.org/1742-5468/2011/i=03/a=P03008

  41. Mukamel, D., Ruffo, S., Schreiber, N.: Breaking of ergodicity and long relaxation times in systems with long-range interactions. Phys. Rev. Lett. 95, 240,604 (2005). https://doi.org/10.1103/PhysRevLett.95.240604

    Article  Google Scholar 

  42. Nemoto, T.: Zon–Cohen singularity and negative inverse temperature in a trapped-particle limit. Phys. Rev. E 85, 061,124 (2012). https://doi.org/10.1103/PhysRevE.85.061124

    Article  Google Scholar 

  43. Nemoto, T., Sasa, Si: Computation of large deviation statistics via iterative measurement-and-feedback procedure. Phys. Rev. Lett. 112(9), 090,602 (2014). https://doi.org/10.1103/PhysRevLett.112.090602

    Article  Google Scholar 

  44. Nyawo, P.T., Touchette, H.: A minimal model of dynamical phase transition. EPL (Europhys. Lett.) 116(5), 50009 (2016). http://stacks.iop.org/0295-5075/116/i=5/a=50009

  45. Onsager, L., Machlup, S.: Fluctuations and irreversible processes. Phys. Rev. 91(6), 1505–1512 (1953). https://doi.org/10.1103/PhysRev.91.1505

    Article  ADS  MathSciNet  MATH  Google Scholar 

  46. Pérez-Espigares, C., Carollo, F., Garrahan, J.P., Hurtado, P.I.: Dynamical criticality in driven systems: non-perturbative results, microscopic origin and direct observation. ArXiv e-prints (2018)

  47. Speck, T.: Thermodynamic formalism and linear response theory for nonequilibrium steady states. Phys. Rev. E 94, 022,131 (2016). https://doi.org/10.1103/PhysRevE.94.022131

    Article  Google Scholar 

  48. Speck, T., Garrahan, J.P.: Space-time phase transitions in driven kinetically constrained latticemodels. Eur. Phys. J. B 79(1), 1–6 (2011). https://doi.org/10.1140/epjb/e2010-10800-x

    Article  ADS  Google Scholar 

  49. Szavits-Nossan, J., Evans, M.R.: Inequivalence of nonequilibrium path ensembles: the example of stochastic bridges. J. Stat. Mech. Theory Exp. 2015(12), P12008 (2015). http://stacks.iop.org/1742-5468/2015/i=12/a=P12008

  50. Touchette, H.: Equivalence and nonequivalence of the microcanonical and canonical ensembles : a large deviations study. Ph.D. thesis, PhD Thesis Department of Physics, McGill University (2003). http://digitool.Library.McGill.CA:80/R/-?func=dbin-jump-full&object_id=84851&silo_library=GEN01

  51. Touchette, H.: Simple spin models with non-concave entropies. Am. J. Phys. 76, 26–30 (2008). https://doi.org/10.1119/1.2794350

    Article  ADS  Google Scholar 

  52. Touchette, H.: The large deviation approach to statistical mechanics (2009). https://doi.org/10.1016/j.physrep.2009.05.002. http://www.sciencedirect.com/science/article/pii/S0370157309001410

  53. Touchette, H.: Methods for calculating nonconcave entropies. J. Stat. Mech. 05008 (2010). http://stacks.iop.org/1742-5468/2010/i=05/a=P05008

  54. Touchette, H., Ellis, R.S., Turkington, B.: An introduction to the thermodynamic and macrostate levels of nonequivalent ensembles. Phys. A Stat. Mech. Appl. 340(1–3), 138–146 (2004). https://doi.org/10.1016/j.physa.2004.03.088. http://www.sciencedirect.com/science/article/pii/S037843710400408X. News and Expectations in Thermostatistics

  55. Tsobgni Nyawo, P., Touchette, H.: Large deviations of the current for driven periodic diffusions. Phys. Rev. E 94, 032,101 (2016). https://doi.org/10.1103/PhysRevE.94.032101

    Article  Google Scholar 

  56. Van Kampen, N.: Stochastic Processes in Physics and Chemistry (Third Edition). Elsevier, Amsterdam (2007). https://www.elsevier.com/books/stochastic-processes-in-physics-and-chemistry/van-kampen/978-0-444-52965-7

Download references

Acknowledgements

We acknowledge Massimiliano Esposito for his advice on this project that started when GV was a post-doctoral fellow in his research team. We thank V. Lecomte and A. Lazarescu for the insightful discussions in connection with this work.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Hadrien Vroylandt.

Appendices

Appendix A: Derivation of the Propagator for Generating Function

We look for an expression of the propagator of the generating function \(G(x_{\mathrm {f}},x_{\mathrm {i}},\kappa ,\gamma )= \big \langle e^{NT(\kappa {m}+\gamma a)}\big \rangle _{x_{\mathrm {f}},x_{\mathrm {i}}}\) as a path integral over all trajectories for the system state n. We use the discrete time \(t_k=k \mathrm {d}t\) corresponding to the k’s time step of duration \(\mathrm {d}t\), with the final time \(T=t_L = L \mathrm {d}t \). We write the system state \(n_{t_k}=n_k\) for short. This state at time \(t_k\) changes by \(n_{k}-n_{k-1}=2\epsilon _k = -2, 0\) or 2 when a spin jumps from up to down, doesn’t jump or jumps from down to up. The sum over all paths for a given initial condition is

$$\begin{aligned} \sum _{[\epsilon _k]} = \sum _{\epsilon _0}\sum _{\epsilon _1} \cdots \sum _{\epsilon _{N-1}}. \end{aligned}$$
(85)

Since jump probabilities depend on the magnetization, the path integration should include a sum on the system state \(n_k\) for all k

$$\begin{aligned} \sum _{[n_k]}= \sum _{n_1}\sum _{n_2} \cdots \sum _{n_{L-1}}, \end{aligned}$$
(86)

with the condition that one spin at most jumps at each time step. In the above sums, the index n goes from \(-N\) to N. All in all, the propagator of the generating function writes

$$\begin{aligned} G(n_{\mathrm {i}},n_{\mathrm {f}},\kappa ,\gamma )= & {} \sum _{[n_k,\epsilon _k]} \left[ \prod _{k=1}^{L} p(n_{k}|n_{k-1}) \delta (n_{k}-n_{k-1}-2\epsilon _{k-1}) \right] \nonumber \\&\times \,e^{\kappa \left( \sum _{k=1}^L n_{k-1} \mathrm {d}t \right) +\gamma \left( \sum _{k=1\, ,\epsilon _k \ne 0}^L 1 \right) }, \end{aligned}$$
(87)

with \(p(n_{k}|n_{k-1}) \) the probability of the transition \( n_{k-1} \rightarrow n_{k}\). Let’s use the Laplace representation of the Dirac distribution

$$\begin{aligned} \delta (n_{k}-n_{k-1}-2\epsilon _{k-1}) = \int _{i{b}_k-\infty }^{i{b}_k+\infty } \frac{\mathrm {d}\mathfrak {p}_{k}}{2 \pi }e^{ i \mathfrak {p}_{k}(n_{k}-n_{k-1}-2\epsilon _{k-1})}, \quad {b}_k \in \mathbb {R}, \end{aligned}$$
(88)

for every integer k between 1 and L to get

$$\begin{aligned} G(n_{\mathrm {i}},n_{\mathrm {f}},\kappa ,\gamma )= & {} \sum _{[n_k,\epsilon _k]} \int \left( \prod _{k=1}^{L} \frac{\mathrm {d}\mathfrak {p}_k}{2 \pi } \right) \nonumber \\&\times \,\left[ \prod _{k=1}^{L} p(n_{k}|n_{k-1}) e^{i \mathfrak {p}_k(n_{k}-n_{k-1}-2\epsilon _{k-1}) +\kappa n_{k-1} \mathrm {d}t + \gamma } \right] . \end{aligned}$$
(89)

In the next step, we sum over \([\epsilon _k]\) to obtain

$$\begin{aligned} G(n_{\mathrm {i}},n_{\mathrm {f}},\kappa ,\gamma )= & {} \sum _{[n_k]} \int \left( \prod _{k=1}^{L} \frac{\mathrm {d}\mathfrak {p}_k}{2 \pi } e^{i \mathfrak {p}_k(n_{k}-n_{k-1})} \right) \nonumber \\&\times \,\prod _{k=1}^{L} \left[ p(n_{k-1}|n_{k-1}) + \sum _{\epsilon =\pm 1} p(n_{k-1}+\epsilon |n_{k-1}) e^{-2i\mathfrak {p}_k \epsilon +\gamma } \right] e^{\kappa n_{k-1}\mathrm {d}t}.\nonumber \\ \end{aligned}$$
(90)

The propagator for staying in the same state during a time step is \(p(n_{k}|n_{k}) = 1- \left( \sum \limits _{\epsilon =\pm 1}K(n_k+2\epsilon ,n_k)\right) \mathrm {d}t\), and when changing of state writes \(p(n_{k}+\epsilon |n_{k}) = K(n_k+2\epsilon ,n_k) \mathrm {d}t\). Exponentiating the product yields

$$\begin{aligned} G(n_{\mathrm {i}},n_{\mathrm {f}},\kappa ,\gamma )= & {} \sum _{[n_k]} \int \left( \prod _{k=1}^{L} \frac{\mathrm {d}\mathfrak {p}_k}{2 \pi } \right) \exp { \left( i \mathrm {d}t \sum _{k=1}^{L} \mathfrak {p}_k \mathchoice{\dfrac{(n_{k}-n_{k-1})}{\mathrm {d}t}}{\dfrac{(n_{k}-n_{k-1})}{\mathrm {d}t}}{\frac{(n_{k}-n_{k-1})}{\mathrm {d}t}}{\frac{(n_{k}-n_{k-1})}{\mathrm {d}t}} \right) } \nonumber \\&\times \,\exp \left[ \sum _{k=1}^{L} \mathrm {d}t \left( \sum _{\epsilon } K(n_{k-1}+2\epsilon ,n_{k-1}) \left( e^{-2i \mathfrak {p}_k\epsilon + \gamma } -1 \right) +\kappa n_{k-1} \right) \right] .\nonumber \\ \end{aligned}$$
(91)

The propagator of the generating function becomes in the continuous limit

$$\begin{aligned} G(x_{\mathrm {f}},x_{\mathrm {i}},\kappa ,\gamma )= & {} \int \mathcal {D} [x] \int \mathcal {D}[\mathfrak {p}] \exp \left[ N \int _{0}^{T} \mathrm {d}t i \mathfrak {p}_t \dot{x}_{t} \right. \nonumber \\&\left. - \int _{0}^{T} \mathrm {d}t \left( \sum _{\epsilon } J_\epsilon (x_{t}) \left( 1- e^{-2 i \mathfrak {p}_t\epsilon + \gamma } \right) +\kappa x_{t} \right) \right] \end{aligned}$$
(92)

where \(\int \mathcal {D} [x]\) (resp. \(\int \mathcal {D} [\mathfrak {p}]\)) is a short notation for the integral on the paths of density (resp. momentum) with given initial and final values. Finally, the path integral expression of the generating function is given by

$$\begin{aligned} G(x_{\mathrm {f}},x_{\mathrm {i}},\kappa ,\gamma )= \int \mathcal {D} [x] \int \mathcal {D}[\mathfrak {p}] \exp \left( N \mathcal {A}_{\kappa ,\gamma } [x,\mathfrak {p},T] \right) \end{aligned}$$
(93)

with the action \(\mathcal {A}_{\kappa ,\gamma }\) being a functional of the paths \([x,\mathfrak {p}]\)

$$\begin{aligned} \mathcal {A}_{\kappa ,\gamma } [x,\mathfrak {p},T] = \int _{0}^{T} \mathrm {d}t i \mathfrak {p}_t \dot{x}_{t} - \int _{0}^{T} \mathrm {d}t \mathcal {H}(x_{t},\mathfrak {p}_t,\kappa ,\gamma ). \end{aligned}$$
(94)

We have introduced the Hamiltonian function

$$\begin{aligned} \mathcal {H}(x,\mathfrak {p},\kappa ,\gamma ) = \sum _{\epsilon =\pm 1} \left[ 1-\exp { \left( -2 i \mathfrak {p} \epsilon +\gamma \right) } \right] J_\epsilon (x) -\kappa x. \end{aligned}$$
(95)

This function is the representation in the continuous limit of a tridiagonal Metzler matrix, whose spectrum is real.Footnote 2 As a consequence, \(i \mathfrak {p}\) is a real number.

Appendix B: Wentzel–Kramers–Brillouin Approximation

We now use a saddle point integration to obtain the leading order in N of the propagator of the generating function \(G(x_{\mathrm {f}},x_{\mathrm {i}},\kappa ,\gamma )\) which is asymptotically exact in the large size limit. For each integration, we deform the contour so that it goes through the saddle point of \(\mathcal {A}_{\kappa ,\gamma } [x,\mathfrak {p},t]\). Therefore, the saddle point is the minimum of \(\mathcal {A}_{\kappa ,\gamma } [x,\mathfrak {p},t]\) solving

$$\begin{aligned} \mathchoice{\dfrac{\delta }{\delta \mathfrak {p}_t}}{\dfrac{\delta }{\delta \mathfrak {p}_t}}{\frac{\delta }{\delta \mathfrak {p}_t}}{\frac{\delta }{\delta \mathfrak {p}_t}} \mathcal {A}_{\kappa ,\gamma } [x,\mathfrak {p},T] = 0, \end{aligned}$$
(96)
$$\begin{aligned} \mathchoice{\dfrac{\delta }{\delta x_t}}{\dfrac{\delta }{\delta x_t}}{\frac{\delta }{\delta x_t}}{\frac{\delta }{\delta x_t}} \mathcal {A}_{\kappa ,\gamma } [x,\mathfrak {p},T] = 0, \end{aligned}$$
(97)

where \( \mathchoice{\dfrac{\delta }{\delta x_t}}{\dfrac{\delta }{\delta x_t}}{\frac{\delta }{\delta x_t}}{\frac{\delta }{\delta x_t}} \) denotes a functional derivative.

In this case, the saddle point calculation is equivalent to the WKB approximation of quantum mechanics. The propagator of the generating function will be given by the paths maximizing the action, i.e. the classical paths starting at \(x_{\mathrm {i}}\) and ending at \(x_{\mathrm {f}}\). These paths solve the Hamilton’s equations

$$\begin{aligned} i \dot{x}_t =&\,\partial _\mathfrak {p} \mathcal {H}(x_t,\mathfrak {p}_t,\kappa ,\gamma ), \end{aligned}$$
(98)
$$\begin{aligned} i \dot{\mathfrak {p}}_t =&- \partial _x \mathcal {H}(x_t,\mathfrak {p}_t,\kappa ,\gamma ). \end{aligned}$$
(99)

The propagator of the generating function can be written as

$$\begin{aligned} G(x_{\mathrm {f}},x_{\mathrm {i}},\kappa ,\gamma ) \simeq _{ N \rightarrow +\infty } \exp \left[ N\int _{0}^{T} \mathrm {d}t \left( i \mathfrak {p}_t \dot{x}_{t} - \mathcal {H}(x_{t},\mathfrak {p}_t,\kappa ,\gamma ) \right) \right] , \end{aligned}$$
(100)

where \((x,\mathfrak {p})\) correspond to the solutions of Eqs. (98)–(99) with initial and final conditions \(x_{\mathrm {i}}\) and \(x_{\mathrm {f}}\). Moreover the Hamiltonian (95) is time independent, and is hence a conserved quantity along the trajectory. Maupertuis’ action \(\int _{0}^{T} \mathrm {d}t i \mathfrak {p}_t \dot{x}_{t}\) is a function of the Hamiltonian and writes

$$\begin{aligned} \int _{0}^{T}\mathrm {d}t i \mathfrak {p}_t \dot{x}_{t} = \int _{x_{\mathrm {i}}}^{x_{\mathrm {f}}}\mathrm {d}x i \mathfrak {p}(x). \end{aligned}$$
(101)

where \(\mathfrak {p}(x)\) can be obtained by inverting Eq. (95). We observe that Maupertuis’ action is always negative in view of the clockwise motion along the orbits that can be justified by inspection of Eq. (98).

In the large size limit, we can also use a saddle point integration on \(\mathfrak {p}\) to switch to the Lagrangian framework. To do so, we look for the solution of Eq. (98). We derive a quadratic equation either for \(e^{i\mathfrak {p}_{s}}\) or for \(e^{-i\mathfrak {p}_{s}}\) whose solutions are

$$\begin{aligned} e^{\pm 2i\mathfrak {p}_{s}}= \mathchoice{\dfrac{\mp \dot{x}_{s} + \sqrt{\dot{x_{s}}^2+\varphi (x_{s},\gamma )}}{2 J_\mp (x_{s}) e^{\gamma }}}{\dfrac{\mp \dot{x}_{s} + \sqrt{\dot{x_{s}}^2+\varphi (x_{s},\gamma )}}{2 J_\mp (x_{s}) e^{\gamma }}}{\frac{\mp \dot{x}_{s} + \sqrt{\dot{x_{s}}^2+\varphi (x_{s},\gamma )}}{2 J_\mp (x_{s}) e^{\gamma }}}{\frac{\mp \dot{x}_{s} + \sqrt{\dot{x_{s}}^2+\varphi (x_{s},\gamma )}}{2 J_\mp (x_{s}) e^{\gamma }}} , \end{aligned}$$
(102)

with \(\varphi (x,\gamma )\) defined in Eq. (59). Inserting this into Eq. (93) leads to the propagator of the generating function of Eq. (57) with Lagrangian (58).

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Vroylandt, H., Verley, G. Non-equivalence of Dynamical Ensembles and Emergent Non-ergodicity. J Stat Phys 174, 404–432 (2019). https://doi.org/10.1007/s10955-018-2186-7

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s10955-018-2186-7

Keywords

Navigation