Skip to main content
Log in

Fluctuations When Driving Between Nonequilibrium Steady States

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

Abstract

Maintained by environmental fluxes, biological systems are thermodynamic processes that operate far from equilibrium without detailed-balanced dynamics. Yet, they often exhibit well defined nonequilibrium steady states (NESSs). More importantly, critical thermodynamic functionality arises directly from transitions among their NESSs, driven by environmental switching. Here, we identify the constraints on excess heat and dissipated work necessary to control a system that is kept far from equilibrium by background, uncontrolled “housekeeping” forces. We do this by extending the Crooks fluctuation theorem to transitions among NESSs, without invoking an unphysical dual dynamics. This and corresponding integral fluctuation theorems determine how much work must be expended when controlling systems maintained far from equilibrium. This generalizes thermodynamic feedback control theory, showing that Maxwellian Demons can leverage mesoscopic-state information to take advantage of the excess energetics in NESS transitions. We also generalize an approach recently used to determine the work dissipated when driving between functionally relevant configurations of an active energy-consuming complex system. Altogether, these results highlight universal thermodynamic laws that apply to the accessible degrees of freedom within the effective dynamic at any emergent level of hierarchical organization. By way of illustration, we analyze a voltage-gated sodium ion channel whose molecular conformational dynamics play a critical functional role in propagating action potentials in mammalian neuronal membranes.

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. We ignore nonergodicity to simplify the development. The approach, though, handles nonergodicity just as well. However, distracting nuances arise that we do not wish to dwell on. For example, if the Markov chain has more than one attracting component for a particular x, then \(\varvec{\pi _x}\) is not unique, but can be constructed as any one of infinitely many probability-normalized linear superpositions of left eigenvectors of \(\mathsf{{T}}^{( \varvec{\mathcal {S} } \rightarrow \varvec{\mathcal {S} } | x)}\) associated with the eigenvalue of unity.

  2. We start in a discrete-time setup, but later translate to continuous time.

  3. The sign conventions adopted for Q, \(Q_\text {hk}\), and \(Q_\text {ex}\) are slightly inharmonious. We take Q and \(Q_\text {ex}\) to be energy that spontaneously flows into a system at fixed x, whereas we have chosen for \(Q_\text {hk}\) to have the opposite sign convention, for easy comparison to the literature. As a result, our quantities technically satisfy \(Q_\text {ex}= Q + Q_\text {hk}\), rather than \(Q_\text {ex}= Q-Q_\text {hk}\).

  4. To be more precise, we write \(\Pr (\mathcal {S} _t = s | \mathcal {S} _0 \sim \varvec{\mu }_0 , x_{1:t+1})\) as \(\Pr _{\mathcal {S} _0 \sim \varvec{\mu }_0}(\mathcal {S} _t = s | x_{1:t+1})\), since the probability is not conditioned on \(\varvec{\mu }_0\)—a probability measure for subsequent state sequences. Here, we simply gloss over this nuance, later adopting the shorthand: \(\Pr (\mathcal {S} _t = s | \varvec{\mu }_0 , x_{1:t+1})\).

  5. Since \(e^{\langle {-Y}\rangle } \le \langle {e^{-Y}}\rangle \) and \(\ln (a)\) is monotonically increasing for positive-valued \(a \in \{ e^{- \langle {Y}\rangle }, \langle {e^{-Y}}\rangle \}\).

  6. The characteristic timescale is actually the net result of a combination of timescales from the inverse eigenvalues of G. Of necessity, these are the same timescales that determine the relaxation of the state distribution.

References

  1. Crooks, G.E.: On thermodynamic and microscopic reversibility. J. Stat. Mech. 2011(7), P07008 (2011)

    Article  Google Scholar 

  2. Crooks, G.E.: Nonequilibrium measurements of free energy differences for microscopically reversible Markovian systems. J. Stat. Phys. 90(5/6), 1481–1487 (1998)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  3. Sagawa, T., Ueda, M.: Nonequilibrium thermodynamics of feedback control. Phys. Rev. E 85, 021104 (2012)

    Article  ADS  Google Scholar 

  4. Wang, H., Oster, G.: Energy transduction in the F1 motor of ATP synthase. Nature 396(6708), 279–282 (1998)

    Article  ADS  Google Scholar 

  5. Polettini, M., Esposito, M.: Irreversible thermodynamics of open chemical networks. I. Emergent cycles and broken conservation laws. J. Chem. Phys. 141(2), 07B610 (2014)

    Article  Google Scholar 

  6. Landauer, R.: Statistical physics of machinery: forgotten middle-ground. Physica A 194(1–4), 551–562 (1993)

    Article  ADS  Google Scholar 

  7. Qian, H.: Nonequilibrium steady-state circulation and heat dissipation functional. Phys. Rev. E 64, 022101 (2001)

    Article  ADS  Google Scholar 

  8. Horsthemke, W.: Noise induced transitions. In: Vidal, C., Pacault, A. (eds.) Non-equilibrium Dynamics in Chemical Systems: Proceedings of the International Symposium. Bordeaux, France, pp. 150–160. Springer, Berlin, 3–7 Sept 1984

  9. Lindner, B., Garcia-Ojalvo, J., Neiman, A., Schimansky-Geier, L.: Effects of noise in excitable systems. Phys. Rep. 392(6), 321–424 (2004)

    Article  ADS  Google Scholar 

  10. Crutchfield, J.P., Aghamohammdi, C.: Not all fluctuations are created equal: Spontaneous variations in thermodynamic function. arXiv:1609.02519

  11. Crutchfield, J.P.: The calculi of emergence: Computation, dynamics, and induction. Physica D 75, 11–54 (1994)

    Article  ADS  MATH  Google Scholar 

  12. Riechers, P.M.: Exact results regarding the physics of complex systems via linear algebra, hidden Markov models, and information theory. PhD thesis, University of California, Davis (2016)

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

    Article  ADS  Google Scholar 

  14. Spinney, R., Ford, I.: Fluctuation relations: a pedagogical overview. Nonequilibrium Statistical Physics of Small Systems, pp. 3–56. Wiley, Weinheim (2013)

    Chapter  Google Scholar 

  15. Oono, Y., Paniconi, M.: Steady state thermodynamics. Prog. Theor. Phys. Suppl. 130, 29–44 (1998)

    Article  ADS  MathSciNet  MATH  Google Scholar 

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

    Article  ADS  Google Scholar 

  17. Trepagnier, E.H., Jarzynski, C., Ritort, F., Crooks, G.E., Bustamante, C.J., Liphardt, J.: Experimental test of Hatano and Sasa’s nonequilibrium steady-state equality. Proc. Natl. Acad. Sci. USA 101(42), 15038–15041 (2004)

    Article  ADS  Google Scholar 

  18. Mandal, D., Jarzynski, C.: Analysis of slow transitions between nonequilibrium steady states. J. Stat. Mech. 2016(6), 063204 (2016)

    Article  MathSciNet  Google Scholar 

  19. Evans, D.J., Searles, D.J., Williams, S.R.: The Evans–Searles fluctuation theorem. Fundamentals of Classical Statistical Thermodynamics, pp. 49–64. Wiley, Weinheim (2016)

    Chapter  Google Scholar 

  20. Evans, D.J., Cohen, E.G.D., Morriss, G.P.: Probability of second law violations in shearing steady states. Phys. Rev. Lett. 71(15), 2401 (1993)

    Article  ADS  MATH  Google Scholar 

  21. Gallavotti, G., Cohen, E.G.D.: Dynamical ensembles in stationary states. J. Stat. Phys. 80(5), 931–970 (1995)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  22. Evans, D.J., Searles, D.J., Rondoni, L.: Application of the Gallavotti-Cohen fluctuation relation to thermostated steady states near equilibrium. Phys. Rev. E 71(5), 056120 (2005)

    Article  ADS  MathSciNet  Google Scholar 

  23. Esposito, M., Van den Broeck, C.: Three detailed fluctuation theorems. Phys. Rev. Lett. 104, 090601 (2010)

    Article  ADS  MathSciNet  Google Scholar 

  24. Crooks, G.E.: Entropy production fluctuation theorem and the nonequilibrium work relation for free energy differences. Phys. Rev. E 60(3), 2721–2726 (1999)

    Article  ADS  Google Scholar 

  25. Roldán, E., Parrondo, J.M.R.: Estimating dissipation from single stationary trajectories. Phys. Rev. Lett. 105(15), 150607 (2010)

    Article  ADS  Google Scholar 

  26. Horowitz, J.M., Vaikuntanathan, S.: Nonequilibrium detailed fluctuation theorem for repeated discrete feedback. Phys. Rev. E 82, 061120 (2010)

    Article  ADS  Google Scholar 

  27. England, J.L.: Statistical physics of self-replication. J. Chem. Phys. 139(12), 09B623 (2013)

    Article  Google Scholar 

  28. England, J.L.: Dissipative adaptation in driven self-assembly. Nat. Nanotech. 10(11), 919–923 (2015)

    Article  ADS  Google Scholar 

  29. Perunov, N., Marsland, R.A., England, J.L.: Statistical physics of adaptation. Phys. Rev. X 6(2), 021036 (2016)

    Google Scholar 

  30. Ruelle, D., Takens, F.: On the nature of turbulence. Commun. Math. Phys. 20(3), 167–192 (1971)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  31. Mackey, M.C.: Time’s Arrow: The Origins of Thermodynamic Behavior. Dover Publications, New York (2003)

    Google Scholar 

  32. Cover, T.M., Thomas, J.A.: Elements of Information Theory, 2nd edn. Wiley, New York (2006)

    MATH  Google Scholar 

  33. Still, S., Sivak, D.A., Bell, A.J., Crooks, G.E.: Thermodynamics of prediction. Phys. Rev. Lett. 109, 120604 (2012)

    Article  ADS  Google Scholar 

  34. Crutchfield, J.P., Young, K.: Inferring statistical complexity. Phys. Rev. Lett. 63, 105–108 (1989)

    Article  ADS  MathSciNet  Google Scholar 

  35. Lan, G., Sartori, P., Neumann, S., Sourjik, V., Tu, Y.: The energy-speed-accuracy trade-off in sensory adaptation. Nat. Phys. 8(5), 422–428 (2012)

    Article  Google Scholar 

  36. Sartori, P., Granger, L., Lee, C.F., Horowitz, J.M.: Thermodynamic costs of information processing in sensory adaptation. PLoS Comput. Biol. 10(12), e1003974 (2014)

    Article  ADS  Google Scholar 

  37. Hartich, D., Barato, A.C., Seifert, U.: Sensory capacity: an information theoretical measure of the performance of a sensor. Phys. Rev. E 93(2), 022116 (2016)

    Article  ADS  Google Scholar 

  38. Esposito, M., Harbola, U., Mukamel, S.: Entropy fluctuation theorems in driven open systems: application to electron counting statistics. Phys. Rev. E 76, 031132 (2007)

    Article  ADS  MathSciNet  Google Scholar 

  39. Bagci, G.B., Tirnakli, U., Kurths, J.: The second law for the transitions between the non-equilibrium steady states. Phys. Rev. E 87, 032161 (2013)

    Article  ADS  Google Scholar 

  40. Gaveau, B., Schulman, L.S.: A general framework for non-equilibrium phenomena: the master equation and its formal consequences. Phys. Lett. A 229(6), 347–353 (1997)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  41. Sivak, D.A., Crooks, G.E.: Near-equilibrium measurements of nonequilibrium free energy. Phys. Rev. Lett. 108(15), 150601 (2012)

    Article  ADS  Google Scholar 

  42. Deffner, S., Lutz, E.: Information free energy for nonequilibrium states (2012). arXiv:1201.3888

  43. Qian, H.: Cycle kinetics, steady state thermodynamics and motors: a paradigm for living matter physics. J. Phys. 17(47), S3783 (2005)

    Google Scholar 

  44. Liepelt, S., Lipowsky, R.: Steady-state balance conditions for molecular motor cycles and stochastic nonequilibrium processes. Euro Phys. Let. 77(5), 50002 (2007)

    Article  ADS  Google Scholar 

  45. Liepelt, S., Lipowsky, R.: Kinesin’s network of chemomechanical motor cycles. Phys. Rev. Lett. 98, 258102 (2007)

    Article  ADS  Google Scholar 

  46. Crooks, G.E.: Path-ensemble averages in systems driven far from equilibrium. Phys. Rev. E 61(3), 2361–2366 (2000)

    Article  ADS  Google Scholar 

  47. Chernyak, V.Y., Chertkov, M., Jarzynski, C.: Path-integral analysis of fluctuation theorems for general Langevin processes. J. Stat. Mech. 2006(8), P08001 (2006)

    Article  Google Scholar 

  48. Harris, R.J., Schutz, G.M.: Fluctuation theorems for stochastic dynamics. J. Stat. Mech. 2007, P07020 (2007)

    Article  MathSciNet  Google Scholar 

  49. Seifert, U.: Entropy production along a stochastic trajectory and an integral fluctuation theorem. Phys. Rev. Lett. 95, 040602 (2005)

    Article  ADS  Google Scholar 

  50. Lahiri, S., Jayannavar, A.M.: Fluctuation theorems for excess and housekeeping heat for underdamped Langevin systems. Euro Phys. J. B 87(9), 195 (2014)

    Article  ADS  Google Scholar 

  51. Jarzynski, C.: Nonequilibrium equality for free energy differences. Phys. Rev. Lett. 78(14), 2690–2693 (1997)

    Article  ADS  Google Scholar 

  52. Speck, T., Seifert, U.: Integral fluctuation theorem for the housekeeping heat. J. Phys. A 38(34), L581 (2005)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  53. Vaikuntanathan, S., Jarzynski, C.: Dissipation and lag in irreversible processes. Europhys. Lett. 87(6), 60005 (2009)

    Article  ADS  Google Scholar 

  54. O’Leary, T., Williams, A.H., Franci, A., Marder, E.: Cell types, network homeostasis, and pathological compensation from a biologically plausible ion channel expression model. Neuron 82(4), 809–821 (2014)

    Article  Google Scholar 

  55. Turrigiano, G.G., Nelson, S.B.: Homeostatic plasticity in the developing nervous system. Nat. Rev. Neurosci. 5(2), 97–107 (2004)

    Article  Google Scholar 

  56. Sengupta, B., Stemmler, M.B.: Power consumption during neuronal computation. Proc. IEEE 102(5), 738–750 (2014)

    Article  Google Scholar 

  57. Howarth, C., Peppiatt-Wildman, C.M., Attwell, D.: The energy use associated with neural computation in the cerebellum. J. Cereb. Blood Flow Metab. 30(2), 403–414 (2010)

    Article  Google Scholar 

  58. Dayan, P., Abbott, L.F.: Theoretical Neuroscience: Computational and Mathematical Modeling of Neural Systems. Computational Neuroscience Series, revised edn. MIT Press, Boston (2005)

    MATH  Google Scholar 

  59. Attwell, D., Laughlin, S.B.: An energy budget for signaling in the grey matter of the brain. J. Cereb. Blood Flow Metab. 21(10), 1133–1145 (2001)

    Article  Google Scholar 

  60. Izhikevich, E.M.: Dynamical Systems in Neuroscience. Computational Neuroscience Series. MIT Press, Boston (2010)

    Google Scholar 

  61. Rieke, F., Warland, D., de Ruyter van Steveninck, R., Bialek, W.: Spikes: Exploring the Neural Code. Bradford Books, New York (1999)

    MATH  Google Scholar 

  62. Hodgkin, A.L., Huxley, A.F.: A quantitative description of membrane current and its application to conduction and excitation in nerve. J. Physio 117(4), 500 (1952)

    Article  Google Scholar 

  63. Patlak, J.: Molecular kinetics of voltage-dependent Na\(^+\) channels. Physiol. Rev. 71(4), 1047–1080 (1991)

    Google Scholar 

  64. Crutchfield, J.P., Ellison, C.J., Riechers, P.M.: Exact complexity: Spectral decomposition of intrinsic computation. Phys. Lett. A 380(9–10), 998–1002 (2016)

    Article  ADS  MATH  Google Scholar 

  65. Riechers, P.M., Crutchfield, J.P.: Beyond the spectral theorem: Decomposing arbitrary functions of nondiagonalizable operators (2016). arXiv:1607.06526 [math-ph]

  66. Lacoste, D., Lau, A.W.C., Mallick, K.: Fluctuation theorem and large deviation function for a solvable model of a molecular motor. Phys. Rev. E 78, 011915 (2008)

    Article  ADS  Google Scholar 

  67. Murashita, Y., Funo, K., Ueda, M.: Nonequilibrium equalities in absolutely irreversible processes. Phys. Rev. E 90, 042110 (2014)

    Article  ADS  Google Scholar 

  68. Altaner, B., Wachtel, A., Vollmer, J.: Fluctuating currents in stochastic thermodynamics II: energy conversion and nonequilibrium response in kinesin models (2015). arXiv:1504.03648

  69. Colquhoun, D., Hawkes, A.G.: Relaxation and fluctuations of membrane currents that flow through drug-operated channels. Proc. R. Soc. Lond. B 199(1135), 231–262 (1977)

    Article  ADS  Google Scholar 

  70. Lahiri, S., Ganguli, S.: A memory frontier for complex synapses. In: Burges C.J.C., Bottou L., Welling M., Ghahramani Z., Weinberger K.Q. (eds.) Advances in Neural Information Processing System 26, pp. 1034–1042. Curran Associates, Inc. (2013)

  71. Shen, Q., Hao, Q., Gruner, S.M.: Macromolecular phasing. Phys. Today 59(3), 46–52 (2006)

    Article  Google Scholar 

  72. Anderson, P.W.: More is different. Science 177(4047), 393–396 (1972)

    Article  ADS  Google Scholar 

  73. Boyd, A.B., Mandal, D., Crutchfield, J.P.: Correlation-powered information engines and the thermodynamics of self-correction. Phys. Rev. E 95(1), 012152 (2017)

    Article  ADS  Google Scholar 

  74. Boyd, A.B., Mandal, D., Crutchfield, J.P.: Leveraging environmental correlations: The thermodynamics of requisite variety. J. Stat. Phys. 167(6), 1555–1585 (2017)

  75. Speck, T., Seifert, U.: The Jarzynski relation, fluctuation theorems, and stochastic thermodynamics for non-Markovian processes. J. Stat. Mech. 09, L09002 (2007)

    MathSciNet  Google Scholar 

Download references

Acknowledgements

We thank Tony Bell, Alec Boyd, Gavin Crooks, Sebastian Deffner, Chris Jarzynski, John Mahoney, Dibyendu Mandal, and Adam Rupe for useful feedback. We thank the Santa Fe Institute for its hospitality during visits. JPC is an SFI External Faculty member. This material is based upon work supported by, or in part by, the U. S. Army Research Laboratory and the U. S. Army Research Office under contracts W911NF-12-1-0234, W911NF-13-1-0390, and W911NF-13-1-0340.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to James P. Crutchfield.

Appendices

Appendix A: Extension to Non-Markovian Instantaneous Dynamics

Commonly, theoretical developments assume state-to-state transitions are instantaneously Markovian given the input. This assumption works well for many cases, but fails in others with strong coupling between system and environment. Fortunately, we can straightforwardly generalize the results of stochastic thermodynamics by considering a system’s observable states to be functions of latent variables \({\varvec{{\mathcal {R}}}}\). The goal in the following is to highlight the necessary changes, so that it should be relatively direct to adapt our derivations to the non-Markovian dynamics. See Ref. [75] for an alternative approach to addressing non-Markovian dynamics.

1.1 Latent States, System States, and Their Many Distributions

Even with constant environmental input, the dynamic over a system’s states need not obey detailed balance nor exhibit any finite Markov order. We assume that the classical observed states \( \varvec{\mathcal {S} } \) are functions \(f: {\varvec{{\mathcal {R}}}}\rightarrow \varvec{\mathcal {S} } \) of a latent Markov chain. We also assume that the stochastic transitions among latent states are determined by the current environmental input \(x \in \mathcal {X}\), which can depend arbitrarily on all previous input and system-state history. The Perron–Frobenius theorem guarantees that there is a stationary distribution over latent states associated with each fixed input x; the function of the Markov chain maps this stationary distribution over latent states into the stationary distribution over system states. These are the stationary distributions associated with system NESSs.

We assume too that the \({\varvec{{\mathcal {R}}}}\)-to-\({\varvec{{\mathcal {R}}}}\) transitions are Markovian given the input. However, different inputs induce different Markov chains over the latent states. This can be described by a (possibly infinite) set of input-conditioned transition matrices over the latent state set \({\varvec{{\mathcal {R}}}}\): \(\{ \mathsf{{T}}^{({\varvec{{\mathcal {R}}}}\rightarrow {\varvec{{\mathcal {R}}}}| x)} \}_{x \in \mathcal {X}}\), where \(\mathsf{{T}}^{({\varvec{{\mathcal {R}}}}\rightarrow {\varvec{{\mathcal {R}}}}| x)}_{i,j} = \Pr ({\mathcal {R}}_{t} = r^j | {\mathcal {R}}_{t-1} = r^i , X_t = x)\). Probabilities regarding actual state paths can be obtained from the latent-state-to-state transition dynamic together with the observable-state projectors, which we now define.

We denote distributions over the latent states as bold Greek symbols, such as \(\varvec{\mu }\). As in the main text, it is convenient to cast \(\varvec{\mu }\) as a row-vector, in which case it appears as the bra \(\langle {\varvec{\mu }}|\). The distribution over latent states \({\varvec{{\mathcal {R}}}}\) implies a distinct distribution over observable states \( \varvec{\mathcal {S} } \). A sequence of driving inputs updates the distribution: \(\varvec{\mu }_{t+n}(\varvec{\mu }_{t}, x_{t:t+n})\). In particular:

$$\begin{aligned} \langle {\varvec{\mu }_{t+n}}|&= \langle {\varvec{\mu }_{t}}| \mathsf{{T}}^{({\varvec{{\mathcal {R}}}}\rightarrow {\varvec{{\mathcal {R}}}}| x_{t:t+n} )} \\&= \langle {\varvec{\mu }_{t}}| \mathsf{{T}}^{({\varvec{{\mathcal {R}}}}\rightarrow {\varvec{{\mathcal {R}}}}| x_{t} )} \mathsf{{T}}^{({\varvec{{\mathcal {R}}}}\rightarrow {\varvec{{\mathcal {R}}}}| x_{t+1} )} \cdots \mathsf{{T}}^{({\varvec{{\mathcal {R}}}}\rightarrow {\varvec{{\mathcal {R}}}}| x_{t+n-1} )} . \end{aligned}$$

(Recall that time indexing is denoted by subscript ranges n : m that are left-inclusive and right-exclusive.) An infinite driving history induces a distribution over the state space, and \(\varvec{\pi _x}\) is the specific steady-state distribution over \({\varvec{{\mathcal {R}}}}\) induced by tireless repetition of the single environmental drive x. Explicitly:

$$\begin{aligned} \langle {\varvec{\pi _x}}| = \lim _{n \rightarrow \infty } \langle {\varvec{\mu }_0}| \left( \mathsf{{T}}^{({\varvec{{\mathcal {R}}}}\rightarrow {\varvec{{\mathcal {R}}}}| x)} \right) ^n . \end{aligned}$$

Usefully, \(\varvec{\pi _x}\) can also be found as the left eigenvector of \(\mathsf{{T}}^{({\varvec{{\mathcal {R}}}}\rightarrow {\varvec{{\mathcal {R}}}}| x)}\) associated with the eigenvalue of unity:

$$\begin{aligned} \langle {\varvec{\pi _x}}| = \langle {\varvec{\pi _x}}| \mathsf{{T}}^{({\varvec{{\mathcal {R}}}}\rightarrow {\varvec{{\mathcal {R}}}}| x)} . \end{aligned}$$
(A1)

The physically relevant steady-state probabilities are this vector’s projection onto observable states: \(\pi _x(s) = \langle {\varvec{\pi _{x}} | s}\rangle \), where \(|{s}\rangle = |{\delta _{s,f(r)}}\rangle \) has a vector-representation in the latent-state basis with elements of all 0s except 1s where the latent state maps to the observable state s.

Assuming latent-state-to-state transitions are Markovian allows the distribution \(\varvec{\mu }\) over these latent states to summarize the causal relevance of the entire driving history.

1.2 Implications

A semi-infinite history induces a particular distribution over system latent states and implies another particular distribution over its observable states. This can be usefully recast in terms of the “start” (or initial) distribution \(\varvec{\mu }_0\) induced by the path \(x_{-\infty :1}\) and the driving history \(x_{1:t+1}\) since then, giving the entropy of the induced state distribution:

$$\begin{aligned} h^{(s | \varvec{\mu }_0 , x_{1:t+1} ) }&= - \ln \Pr (\mathcal {S} _t = s | \varvec{\mu }_0 , x_{1:t+1}) \\&= - \ln \langle {\varvec{\mu }_0}| \mathsf{{T}}^{({\varvec{{\mathcal {R}}}}\rightarrow {\varvec{{\mathcal {R}}}}| x_{1:t+1})} |{s}\rangle . \end{aligned}$$

Or, employing the new distribution and the driving history since then, the path entropy (functional of state and driving history) can be expressed simply in terms of the current distribution over latent states and the candidate observable state s:

$$\begin{aligned} h^{(s | \varvec{\mu } ) }&= - \ln \Pr (\mathcal {S} _t = s | {\mathcal {R}}_{t} \sim \varvec{\mu } ) \\&= - \ln \langle {\varvec{\mu } | s}\rangle . \end{aligned}$$

Averaging the path-conditional state entropy over observable states again gives a genuine input-conditioned Shannon state entropy:

$$\begin{aligned} \langle {h^{(s_t | {\overleftarrow{x}}_{t})} }\rangle _{\Pr (s_t | {\overleftarrow{x}}_{t})}&= {\text {H}}[\mathcal {S} _t | {\overleftarrow{X}}_{t}= {\overleftarrow{x}}_{t}] . \end{aligned}$$

It is again easy to show that the state-averaged path entropy \(k_\text {B}{\text {H}}[\mathcal {S} _t | {\overleftarrow{x}}_{t}]\) is an extension of the system’s steady-state nonequilibrium entropy. In steady-state, the state-averaged path entropy reduces to:

$$\begin{aligned} k_\text {B}{\text {H}}[\mathcal {S} _t | {\overleftarrow{X}}_{t}= \dots xxx ]&= - k_\text {B}{\text {H}}[\mathcal {S} _t | {\mathcal {R}}_t \sim \varvec{\pi _x} ] \\&= - k_\text {B}\sum _{s \in \varvec{\mathcal {S} } } \pi _x(s) \ln \pi _x(s) \\&= S_\text {ss}(x) . \end{aligned}$$

The nonsteady-state addition to free energy is:

$$\begin{aligned} \beta ^{-1} \gamma (s | \varvec{\mu } , x) \equiv \beta ^{-1} \ln \frac{\Pr (\mathcal {S} _t = s | {\mathcal {R}}_{t-1} \sim \varvec{\mu }, X_{t} = x )}{ \pi _{x}(s ) } . \end{aligned}$$

Averaging over observable states this becomes the relative entropy:

$$\begin{aligned} \langle {\gamma (s | \varvec{\mu } , x) }\rangle = D_\text {KL} \left[ \Pr (\mathcal {S} _t | {\mathcal {R}}_{t-1} \sim \varvec{\mu } , X_{t} = x ) || \varvec{\pi _x} \right] , \end{aligned}$$

which is always nonnegative.

Using this setup and decomposing:

$$\begin{aligned} \frac{\Pr (\mathcal {S} _{0:N} = s^0 \mathbf {s}| {\mathcal {R}}_{-1} \sim {\varvec{\mu }}_\text {F}, X_{0:N} = x^0 \mathbf {x})}{\Pr (\mathcal {S} _{0:N} = s^{N-1} \mathbf {s}_{\leftarrow }^\text {R}| {\mathcal {R}}_{-1} \sim {\varvec{\mu }}_\text {R}, X_{0:N} = x^N \mathbf {x}^\text {R})} \end{aligned}$$

in analogy with Eq. (25), it is straightforward to extend the remaining results of the main body to the setting in which observed states are functions of a Markov chain. Notably, the path dependencies pick up new contributions from non-Markovity. Also, knowledge of distributions over latent states provides a thermodynamic advantage to Maxwellian Demons.

Appendix B: Integral Fluctuation Theorems with Auxiliary Variables

Recall that we quantify how much the auxiliary variable independently informs the state sequence via the nonaveraged conditional mutual information:

$$\begin{aligned} i[{\overrightarrow{s}}; {\overrightarrow{y}}| {\overrightarrow{x}}, {\varvec{\mu }}_\text {F}]&\equiv \ln \frac{\Pr ({\overrightarrow{s}}, {\overrightarrow{y}}| {\overrightarrow{x}}, {\varvec{\mu }}_\text {F})}{ \Pr ( {\overrightarrow{y}}| {\overrightarrow{x}}, {\varvec{\mu }}_\text {F}) \Pr ( {\overrightarrow{s}}| {\overrightarrow{x}}, {\varvec{\mu }}_\text {F}) } \\&= \ln \frac{\Pr ({\overrightarrow{s}}, {\overrightarrow{y}}, {\overrightarrow{x}}| {\varvec{\mu }}_\text {F})}{ \Pr ( {\overrightarrow{y}}, {\overrightarrow{x}}, {\varvec{\mu }}_\text {F}) \Pr ( {\overrightarrow{s}}| {\overrightarrow{x}}, {\varvec{\mu }}_\text {F}) } . \end{aligned}$$

Note that averaging over the input, state, and auxiliary sequences gives the familiar conditional mutual information:

$$\begin{aligned} \text {I}[\mathcal {S} _{0:N}; Y_{0:N}&| X_{0:N} , {\varvec{\mu }}_\text {F}] = \left\langle i[{\overrightarrow{s}}; {\overrightarrow{y}}| {\overrightarrow{x}}, {\varvec{\mu }}_\text {F}] \right\rangle _{\Pr (x_{0:N} , s_{0:N} , y_{0:N} | {\varvec{\mu }}_\text {F})} . \end{aligned}$$

(Averaging over distributions is the same as being given the distribution, since the distribution over distributions is assumed to be peaked at \({\varvec{\mu }}_\text {F}\).)

Noting that:

$$\begin{aligned} e^{\beta W_\text {diss} + i({\overrightarrow{s}}; {\overrightarrow{y}}| {\overrightarrow{x}}, {\varvec{\mu }}_\text {F}) + \Psi }&= e^{\Omega + i({\overrightarrow{s}}; {\overrightarrow{y}}| {\overrightarrow{x}}, {\varvec{\mu }}_\text {F}) + \Psi + (\gamma _\text {F} - \gamma _\text {R} )} \\&= \frac{\Pr ({\overrightarrow{s}}, {\overrightarrow{y}}, {\overrightarrow{x}}| {\varvec{\mu }}_\text {F})}{ \Pr ( {\overrightarrow{y}}, {\overrightarrow{x}}| {\varvec{\mu }}_\text {F}) \Pr ( s^{N-1} \mathbf {s}_{\leftarrow }^\text {R}| \mathbf {x}^\text {R}x^0 , {\varvec{\mu }}_\text {R}) } \\&= \frac{\Pr ({\overrightarrow{s}}, {\overrightarrow{y}}, {\overrightarrow{x}}| {\varvec{\mu }}_\text {F})}{ \Pr ( {\overrightarrow{y}}, {\overrightarrow{x}}| {\varvec{\mu }}_\text {F}) \Pr ( {\overleftarrow{s}}| {\overleftarrow{x}}, {\varvec{\mu }}_\text {R}) } , \end{aligned}$$

where , we have the integral fluctuation theorem (IFT):

$$\begin{aligned} \left\langle e^{-\beta W_\text {diss} - i({\overrightarrow{s}}; {\overrightarrow{y}}| {\overrightarrow{x}}, {\varvec{\mu }}_\text {F}) - \Psi } \right\rangle _{\Pr ({\overrightarrow{s}}, {\overrightarrow{y}}, {\overrightarrow{x}}| {\varvec{\mu }}_\text {F})}&= \sum _{{\overrightarrow{x}}, {\overrightarrow{s}}, {\overrightarrow{y}}} \Pr ({\overrightarrow{s}}, {\overrightarrow{y}}, {\overrightarrow{x}}| {\varvec{\mu }}_\text {F}) \frac{\Pr ({\overrightarrow{y}}, {\overrightarrow{x}}| {\varvec{\mu }}_\text {F}) \Pr ( {\overleftarrow{s}}| {\overleftarrow{x}}, {\varvec{\mu }}_\text {R})}{\Pr ({\overrightarrow{s}}, {\overrightarrow{y}}, {\overrightarrow{x}}| {\varvec{\mu }}_\text {F})} \\&= \sum _{{\overrightarrow{x}}, {\overrightarrow{s}}, {\overrightarrow{y}}} \Pr ( {\overrightarrow{y}}, {\overrightarrow{x}}| {\varvec{\mu }}_\text {F}) \Pr ( {\overleftarrow{s}}| {\overleftarrow{x}}, {\varvec{\mu }}_\text {F}) \\&= \sum _{{\overrightarrow{x}}, {\overrightarrow{y}}} \Pr ( {\overrightarrow{y}}, {\overrightarrow{x}}| {\varvec{\mu }}_\text {F}) \sum _{{\overleftarrow{s}}} \Pr ( {\overleftarrow{s}}| {\overleftarrow{x}}, {\varvec{\mu }}_\text {R}) \\&= \sum _{{\overrightarrow{x}}, {\overrightarrow{y}}} \Pr ( {\overrightarrow{y}}, {\overrightarrow{x}}| {\varvec{\mu }}_\text {F}) \\&= 1 . \end{aligned}$$

Notably, this relation holds arbitrarily far from equilibrium and allows for the starting and ending distributions to both be nonsteady-state.

It is tempting to conclude that the revised Second Law of Thermodynamics should read:

(B1)

which includes the effects of both irreversibility and conditional mutual information between state-sequence and auxiliary sequence, given input-sequence. However, we expect that \(\left\langle Q_\text {hk}\right\rangle > 0\), so Eq. (B1) is not the strongest bound derivable. Dropping \(\Psi \) from the IFT still yields a valid equality. However, the derivation runs differently since it depends on the normalization of the dual dynamic—that is, using quantities of the form:

$$\begin{aligned} \frac{\pi _{x^n}(s^{n-1})}{\pi _{x^n}(s^n)} \Pr (s^n | s^{n-1}, x^n) . \end{aligned}$$

These are mathematically-sound transition probabilities \(\widetilde{\mathsf{{T}}}_{s^n, s^{n-1}}^{( \varvec{\mathcal {S} } \rightarrow \varvec{\mathcal {S} } | x^n)}\), but only of a nonphysical artificial dynamic. Although IFTs with \(\Psi \) may be useful for other reasons, it is the non-\(\Psi \) IFTs that seem to yield the tighter bound for the revised Second Laws of information thermodynamics without detailed balance.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Riechers, P.M., Crutchfield, J.P. Fluctuations When Driving Between Nonequilibrium Steady States. J Stat Phys 168, 873–918 (2017). https://doi.org/10.1007/s10955-017-1822-y

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s10955-017-1822-y

Keywords

Navigation