Skip to main content
Log in

A General Fluctuation–Response Relation for Noise Variations and its Application to Driven Hydrodynamic Experiments

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

Abstract

The effect of a change of noise amplitudes in overdamped diffusive systems is linked to their unperturbed behavior by means of a nonequilibrium fluctuation–response relation. This formula holds also for systems with state-independent nontrivial diffusivity matrices, as we show with an application to an experiment of two trapped and hydrodynamically coupled colloids, one of which is subject to an external random forcing that mimics an effective temperature. The nonequilibrium susceptibility of the energy to a variation of this driving is an example of our formulation, which improves an earlier version, as it does not depend on the time-discretization of the stochastic dynamics. This scheme holds for generic systems with additive noise and can be easily implemented numerically, thanks to matrix operations.

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

Similar content being viewed by others

Notes

  1. It can be modified by adding neutral density filters or by changing the time that the laser spend on each trap.

  2. For brevity of notation, we denote the elements of the inverse of a matrix as in \(D^{-1}_{ij}\) rather than the unambiguous but cumbersome \((D^{-1})_{ij}\).

  3. Note that \(\partial _i (h_{kl} x_l \partial _k a_i - h_{ik} a_k) = h_{kl} x_l \partial _k \partial _i a_i + h_{kl} \delta _{il} \partial _k a_i - h_{ik} \partial _i a_k = h_{kl} x_l \partial _k \partial _i a_i\).

  4. For this matter, we note here that we have intentionally left out integration limits in the action: The forcing \(h(s)\) is assumed to be temporally localized, and the integration domain is infinite in principle. Thus, \(\int \mathrm{d}s\, \dot{h} g = -\int \mathrm{d}s\, h\dot{g}\) for a generic function g(s), and likewise \(\int \mathrm{d}s\, \dot{h}_{kk} = 0\) in Eq. (22a).

References

  1. Agarwal, G.S.: Fluctuation–dissipation theorems for systems in non-thermal equilibrium and applications. Z. Phys. 252, 25–38 (1972)

    Article  ADS  Google Scholar 

  2. Altaner, B., Polettini, M., Esposito, M.: Fluctuation–dissipation relations far from equilibrium. Phys. Rev. Lett. 117, 180601 (2016)

    Article  ADS  Google Scholar 

  3. Baiesi, M., Maes, C., Wynants, B.: Fluctuations and response of nonequilibrium states. Phys. Rev. Lett. 103, 010602 (2009)

    Article  ADS  MATH  Google Scholar 

  4. Baiesi, M., Maes, C., Wynants, B.: Nonequilibrium linear response for Markov dynamics, I: jump processes and overdamped diffusions. J. Stat. Phys. 137, 1094–1116 (2009)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  5. Baiesi, M., Maes, C.: An update on the nonequilibrium linear response. New J. Phys. 15, 013004 (2013)

    Article  ADS  Google Scholar 

  6. Baiesi, M., Basu, U., Maes, C.: Thermal response in driven diffusive systems. Eur. Phys. J. B 87, 277 (2014)

    Article  ADS  MathSciNet  Google Scholar 

  7. Baiesi, M., Ciliberto, S., Falasco, G., Yolcu, C.: Thermal response of nonequilibrium \(RC\) circuits. Phys. Rev. E 94, 022144 (2016)

    Article  ADS  Google Scholar 

  8. Barrat, J.L., Kob, W.: Fluctuation–dissipation ratio in an aging Lennard-Jones glass. Europhys. Lett. 46, 637–642 (1999)

    Article  ADS  Google Scholar 

  9. Bertini, L., Gabrielli, D., Jona-Lasinio, G., Landim, C.: Clausius inequality and optimality of quasistatic transformations for nonequilibrium stationary states. Phys. Rev. Lett. 114, 020601 (2013)

    Article  Google Scholar 

  10. Bérut, A., Petrosyan, A., Ciliberto, S.: Energy flow between two hydrodynamically coupled particles kept at different effective temperatures. Europhys. Lett. 107, 60004 (2014)

    Article  Google Scholar 

  11. Bérut, A., Imparato, A., Petrosyan, A., Ciliberto, S.: The role of coupling on the statistical properties of the energy fluxes between stochastic systems at different temperatures. J. Stat. Mech. 2016(5), 054002 (2016)

    Article  MathSciNet  Google Scholar 

  12. Bérut, A., Imparato, A., Petrosyan, A., Ciliberto, S.: Stationary and transient fluctuation theorems for effective heat fluxes between hydrodynamically coupled particles in optical traps. Phys. Rev. Lett. 116, 068301 (2016)

    Article  ADS  MATH  Google Scholar 

  13. Bohec, P., Gallet, F., Maes, C., Safaverdi, S., Visco, P., van Wijland, F.: Probing active forces via a fluctuation-dissipation relation: application to living cells. Europhys. Lett. 102, 50005 (2013)

    Article  ADS  Google Scholar 

  14. Boksenbojm, E., Maes, C., Netočný, K., Pešek, J.: Heat capacity in nonequilibrium steady states. Europhys. Lett. 96, 40001 (2011)

    Article  ADS  Google Scholar 

  15. Brandner, K., Saito, K., Seifert, U.: Thermodynamics of micro- and nano-systems driven by periodic temperature variations. Phys. Rev. X 5, 031019 (2016)

    Google Scholar 

  16. Cammarota, C., Cavagna, A., Gradenigo, G., Grigera, T.S., Verrocchio, P.: Numerical determination of the exponents controlling the relationship between time, length, and temperature in glass-forming liquids. J. Chem. Phys. 131, 194901 (2009)

    Article  ADS  Google Scholar 

  17. Chetrite, R., Falkovich, G., Gawȩdzki, K.: Fluctuation relations in simple examples of non-equilibrium steady states. J. Stat. Mech. 2008, 08005 (2008)

    Article  Google Scholar 

  18. Chetrite, R.: Fluctuation relations for diffusion that is thermally driven by a nonstationary bath. Phys. Rev. E 80, 051107 (2009)

    Article  ADS  Google Scholar 

  19. Ciliberto, S., Imparato, A., Naert, A., Tanase, M.: Heat flux and entropy produced by thermal fluctuations. Phys. Rev. Lett. 110, 180601 (2013)

    Article  ADS  Google Scholar 

  20. Ciliberto, S., Imparato, A., Naert, A., Tanase, M.: Statistical properties of the energy exchanged between two heat baths coupled by thermal fluctuations. J. Stat. Mech. 2013, 12014 (2013)

    Article  MathSciNet  Google Scholar 

  21. Crisanti, A., Puglisi, A., Villamaina, D.: Non-equilibrium and information: the role of cross-correlations. Phys. Rev. E 85, 061127 (2012)

    Article  ADS  Google Scholar 

  22. Crisanti, A., Picco, M., Ritort, F.: Fluctuation relation for weakly ergodic aging systems. Phys. Rev. Lett. 110, 080601 (2013)

    Article  ADS  Google Scholar 

  23. Cugliandolo, L., Kurchan, J., Parisi, G.: Off equilibrium dynamics and aging in unfrustrated systems. J. Phys. I 4, 1641 (1994)

    Google Scholar 

  24. Dhar, A.: Heat transport in low-dimensional systems. Adv. Phys. 57, 457–537 (2008)

    Article  ADS  Google Scholar 

  25. Doi, M., Edwards, S.: The Theory of Polymer Dynamics. International series of monographs on physics. Clarendon Press, Oxford (1988)

    Google Scholar 

  26. Falasco, G., Baiesi, M.: Nonequilibrium temperature response for stochastic overdamped systems. New J. Phys. 18, 043039 (2016)

    Article  ADS  Google Scholar 

  27. Falasco, G., Baiesi, M.: Temperature response in nonequilibrium stochastic systems. Europhys. Lett. 113, 20005 (2016)

    Article  ADS  Google Scholar 

  28. Falasco, G., Baldovin, F., Kroy, K., Baiesi, M.: Mesoscopic virial equation for nonequilibrium statistical mechanics. New J. Phys. 18(9), 093043 (2016)

    Article  ADS  Google Scholar 

  29. Falcioni, M., Isola, S., Vulpiani, A.: Correlation functions and relaxation properties in chaotic dynamics and statistical mechanics. Phys. Lett. A 144, 341 (1990)

    Article  ADS  MathSciNet  Google Scholar 

  30. Gautrais, J., Ginelli, F., Fournier, R., Blanco, S., Soria, M., Chate, H., Theraulaz, G.: Deciphering interactions in moving animal groups. PLoS Comput. Biol. 8, e1002678 (2012)

    Article  ADS  MathSciNet  Google Scholar 

  31. Girsanov, I.V.: On transforming a certain class of stochastic processes by absolutely continuous substitution of measures. Theor. Probab. Appl. 5(3), 285–301 (1960)

    Article  MathSciNet  MATH  Google Scholar 

  32. Gomez-Solano, J.R., Petrosyan, A., Ciliberto, S., Chetrite, R., Gawȩdzki, K.: Experimental verification of a modified fluctuation-dissipation relation for a micron-sized particle in a non-equilibrium steady state. Phys. Rev. Lett. 103, 040601 (2009)

    Article  ADS  Google Scholar 

  33. Gomez-Solano, J.R., Petrosyan, A., Ciliberto, S., Maes, C.: Fluctuations and response in a non-equilibrium micron-sized system. J. Stat. Mech. 2011(1), P01008 (2011)

    Article  Google Scholar 

  34. Gomez-Solano, J.R., Petrosyan, A., Ciliberto, S.: Fluctuations, linear response and heat flux of an aging system. Europhys. Lett. 98, 10007 (2012)

    Article  ADS  Google Scholar 

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

    Article  ADS  Google Scholar 

  36. Kailath, T.: The structure of Radon–Nikodym derivatives with respect to Wiener and related measures. Ann. Math. Stat. 42(3), 1054–1067 (1971)

    Article  MathSciNet  MATH  Google Scholar 

  37. Katz, S., Lebowitz, J.L., Spohn, H.: Phase transitions in stationary nonequilibrium states of model lattice systems. Phys. Rev. B 28, 1655–1658 (1983)

    Article  ADS  Google Scholar 

  38. Komatsu, T.S., Nakagawa, N., Sasa, S., Tasaki, H.: Exact equalities and thermodynamic relations for nonequilibrium steady states. J. Stat. Phys. 159, 1237–1285 (2015)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  39. Lepri, S., Livi, R., Politi, A.: Thermal conduction in classical low-dimensional lattices. Phys. Rep. 377, 1–80 (2003)

    Article  ADS  MathSciNet  Google Scholar 

  40. Lippiello, E., Corberi, F., Zannetti, M.: Fluctuation dissipation relations far from equilibrium. J. Stat. Mech., P07002 (2007)

  41. Loi, D., Mossa, S., Cugliandolo, L.F.: Non-conservative forces and effective temperatures in active polymers. Soft Matter 7, 10193–10209 (2011)

    Article  ADS  Google Scholar 

  42. Maes, C., Netočný, K.: A nonequilibrium extension of the Clausius heat theorem. J. Stat. Phys. 154, 188–203 (2014)

    Article  MathSciNet  MATH  Google Scholar 

  43. Maggi, C., Di Leonardo, R., Dyre, J.C., Ruocco, G.: Generalized fluctuation-dissipation relation and effective temperature in off-equilibrium colloids. Phys. Rev. B 81, 104201 (2010)

    Article  ADS  Google Scholar 

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

    Article  MathSciNet  Google Scholar 

  45. Marchetti, M.C., Joanny, J.F., Ramaswamy, S., Liverpool, T.B., Rao, M., Simha, R.A.: Hydrodynamics of soft active matter. Rev. Mod. Phys. 85, 1143 (2013)

    Article  ADS  Google Scholar 

  46. Nakamura, T., Sasa, S.: A fluctuation–response relation of many Brownian particles under non-equilibrium conditions. Phys. Rev. E 77, 021108 (2008)

    Article  ADS  Google Scholar 

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

    Article  ADS  MathSciNet  MATH  Google Scholar 

  48. Proesmans, K., Cleuren, B., Van den Broeck, C.: Linear stochastic thermodynamics for periodically driven systems. J. Stat. Mech. 2016(2), 023202 (2016)

    Article  MathSciNet  Google Scholar 

  49. Prost, J., Joanny, J.F., Parrondo, J.M.: Generalized fluctuation–dissipation theorem for steady-state systems. Phys. Rev. Lett. 103, 090601 (2009)

    Article  ADS  Google Scholar 

  50. Sagawa, T., Hayakawa, H.: Geometrical expression of excess entropy production. Phys. Rev. E 84, 051110 (2011)

    Article  ADS  Google Scholar 

  51. Seifert, U., Speck, T.: Fluctuation–dissipation theorem in nonequilibrium steady states. Europhys. Lett. 89, 10007 (2010)

    Article  ADS  Google Scholar 

  52. Sekimoto, K.: Microscopic heat from the energetics of stochastic phenomena. Phys. Rev. E 76, 060103(R) (2007)

    Article  ADS  Google Scholar 

  53. Speck, T., Seifert, U.: Restoring a fluctuation–dissipation theorem in a nonequilibrium steady state. Europhys. Lett. 74, 391–396 (2006)

    Article  ADS  Google Scholar 

  54. Speck, T., Seifert, U.: Extended fluctuation–dissipation theorem for soft matter in stationary flow. Phys. Rev. E 79, 040102 (2009)

    Article  ADS  Google Scholar 

  55. Verley, G., Mallick, K., Lacoste, D.: Modified fluctuation–dissipation theorem for non-equilibrium steady states and applications to molecular motors. Europhys. Lett. 93, 10002 (2011)

    Article  ADS  Google Scholar 

  56. Villamaina, D., Baldassarri, A., Puglisi, A., Vulpiani, A.: Fluctuation dissipation relation: how does one compare correlation functions and responses? J. Stat. Mech. 2009(7), P07024 (2009)

    Article  Google Scholar 

  57. Yolcu, C., Baiesi, M.: Linear response of hydrodynamically-coupled particles under a nonequilibrium reservoir. J. Stat. Mech. 2016, 033209 (2016)

    Article  MathSciNet  Google Scholar 

  58. Zinn-Justin, J.: Quantum Field Theory and Critical Phenomena, 4th edn. Clarendon Press, Oxford (2002)

    Book  MATH  Google Scholar 

Download references

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Marco Baiesi.

Appendices

Appendix 1: Stationary Distribution

Here we briefly sketch the calculation of the stationary distribution of the system, for the sake of completeness.

With a drift \(a(x)=-\mu \kappa x\) linear in \(x\), the equation of motion (1) describes an Ornstein-Uhlenbeck process, which we rewrite here as

$$\begin{aligned} \dot{x} = - \Omega x+ \sqrt{2D} \xi \ , \end{aligned}$$
(38)

with \(\Omega = \mu \kappa \). According to the corresponding Fokker–Planck equation, \(\dot{\rho } = -\partial _i J_i\), stationarity requires the ensemble current \(J = -\rho \Omega x- D\nabla \rho \) to be divergenceless, yielding

$$\begin{aligned} 0 = \rho \Omega _{ii} + \Omega _{ij} x_j \partial _i \rho +D_{ij} \partial _i \partial _j \rho \ . \end{aligned}$$
(39)

Clearly, a Gaussian distribution of the form

$$\begin{aligned} \rho (x) = Z^{-1} e^{-\frac{1}{2} x^\dagger Gx} \end{aligned}$$
(40)

satisfies the zero divergence condition above. To find the inverse covariance matrix \(G\), one substitutes the ansatz, upon which a few lines of algebra implies the condition

$$\begin{aligned} \Omega G^{-1} + G^{-1} \Omega ^\dagger = 2 D\ . \end{aligned}$$
(41)

This is a system of linear equations in the elements of the covariance matrix \(G^{-1}\), which can be solved, for instance, by rewriting it in terms of a vector composed of the columns of \(G^{-1}\). Here, we simply quote the resulting matrix: Recalling \(\Omega = \mu \kappa \) with \(\kappa \) diagonal and \(\mu \) given in Eq. (3), and \(D\) given in Eq. (6), one finds

$$\begin{aligned} G^{-1} = \begin{bmatrix} \frac{T+(1-\varepsilon ^2)\Delta T}{\kappa _{11}} + \frac{\varepsilon ^2\Delta T}{\kappa _{11}+\kappa _{22}}&\frac{\varepsilon \Delta T}{\kappa _{11} + \kappa _{22}} \\ \frac{\varepsilon \Delta T}{\kappa _{11} + \kappa _{22}}&\frac{T}{\kappa _{22}} + \frac{\varepsilon ^2\Delta T}{\kappa _{11}+\kappa _{22}} \end{bmatrix} \ . \end{aligned}$$
(42)

Inversion of this matrix thus determines the stationary distribution (40).

Appendix 2: Form of the Perturbation Parameter

Even though it may sometimes be more convenient mathematically to work in terms of the noise amplitude \(B\), it is the diffusivity matrix \(D\) which is physically relevant, since the noise amplitude is fixed only up to an orthogonal transformation. Hence the perturbation parameter, \(h= \delta B\tilde{B}^{-1}\), defined in Sec. 3.1 in terms of the noise amplitude should eventually be expressed in terms of a perturbation of the diffusivity. With \(D= \tilde{D}+ \delta D\) and \(B= \tilde{B}+ \delta B\), in line with Sec. 3.1, the definitions \(2 D= BB^\dagger \) and \(2 \tilde{D}= \tilde{B}\tilde{B}^\dagger \) imply that

$$\begin{aligned} 2 \delta D= \tilde{B}\delta B^\dagger + \delta B\tilde{B}^\dagger \ . \end{aligned}$$
(43)

Note that one should not expect to solve this equation for \(\delta B\) uniquely given a specific \(\delta D\); many noise coefficient matrices \(B\) map into the same diffusion matrix \(D\).

The relation above can now be used to express the perturbation parameter \(h= \delta B\tilde{B}^{-1}\) in terms of \(\tilde{D}\) and \(\delta D\), but not uniquely, which is not a problem. Multiplying Eq. (43) by \(\tilde{D}^{-1}\) from the left and right, one easily finds

$$\begin{aligned} \tilde{D}^{-1} \delta D\tilde{D}^{-1} = \tilde{D}^{-1} h+ h^\dagger \tilde{D}^{-1}\ . \end{aligned}$$
(44)

The left hand side of this equation is equal to \(-\delta D^{-1}\) (verified easily by varying the identity \(1 = DD^{-1}\)) which is determined by the physical description of the perturbation. Meanwhile, the right hand side of the equation is twice the symmetric part of \(\tilde{D}^{-1} h\). In other words, it is only the symmetric part of \(\tilde{D}^{-1} h\) that is fixed by the physical form of the perturbation, and the antisymmetric part is left undetermined. It therefore behooves one to choose \(\tilde{D}^{-1} h\) to be purely symmetric, whence one obtains \(-\delta D^{-1} = 2 \tilde{D}^{-1} h\), or

$$\begin{aligned} h= -\tfrac{1}{2} \tilde{D}\delta D^{-1} = -\tfrac{1}{2} D\delta D^{-1} \ , \end{aligned}$$
(45)

where the second equality is valid due to the overarching first order approximation of linear response.

Appendix 3: Asymptotic Values of the Susceptibility

In this appendix, we sketch how the asymptotic values for the susceptibilities in Figs. 2 and 3 were obtained.

One can derive a host of relations valid in the stationary regime by requiring that time derivatives of state observables vanish [28]. For the present discussion, the relevant observable is the tensor \(x_i x_j\), i.e.,

$$\begin{aligned} 0 = \frac{\mathrm{d}}{\mathrm{d}t} \langle x_i x_j \rangle = \langle \mathbb {L}x_i x_j \rangle \ , \end{aligned}$$
(46)

with the backward generator \(\mathbb {L}= a_i \partial _i + D_{ij} \partial _i \partial _j\). (One should keep in mind that the average is in the stationary regime, although we leave it unlabeled). When evaluated explicitly, with \(a= \mu F\), one finds the relation

$$\begin{aligned} 0 = \mu \langle Fx^\dagger \rangle + \langle xF^\dagger \rangle \mu + 2 D\ . \end{aligned}$$
(47)

This matrix relation entails a set of equations for the independent components of the tensor \(\langle x_i F_j \rangle \), of which there are 3 in our case with 2 degrees of freedom.

We note that for our system, \(U(x) = -(1/2) \mathrm{Tr}(Fx)\). Thus, multiplying Eq. (47) from the left by \(\mu ^{-1}\) and taking the trace, we find the stationary average of the potential energy to be

$$\begin{aligned} \langle U(x) \rangle = -\frac{1}{2} \mathrm{Tr}\langle Fx \rangle = \frac{1}{2} \mathrm{Tr}(\mu ^{-1} D) \ . \end{aligned}$$
(48)

Hence, the stationary (asymptotic) value for the susceptibility is found as

$$\begin{aligned} \frac{\partial \langle U \rangle }{\partial \Delta T} = \frac{1}{2} \mathrm{Tr}\! \left( \mu ^{-1} \frac{\partial D}{\partial \Delta T}\right) = \frac{1}{2} \ , \end{aligned}$$
(49)

which was evaluated using Eq. (6) for the diffusion matrix. This asymptotic value for \(\chi \) was indicated in Fig. 2.

The susceptibility of the energy of the second particle \(U_2 (x_2) = -(1/2) F_2 x_2\) can be extracted similarly from Eq. (47), with the exception that one has to go through the tedium of actually solving for the component \(\langle F_2 x_2 \rangle \). We quote only the result of this straightforward exercise:

$$\begin{aligned} \frac{\partial \langle U_2 \rangle }{\partial \Delta T} =&-\frac{1}{2} \frac{\partial }{\partial \Delta T} \langle F_2 x_2 \rangle = \frac{\varepsilon ^2}{2\left( 1+\frac{\kappa _{11}}{\kappa _{22}}\right) } \ . \end{aligned}$$
(50)

This was evaluated as 0.01899 for the actual experimental values of \(\kappa _{11} = {3.3745} \,\mathrm{pN}/\upmu \mathrm{m} \), \(\kappa _{22} = {3.3285} \,\mathrm{pN}/\upmu \mathrm{m} \), and \(\varepsilon = 0.2766\), and indicated in Fig. 3 as the asymptotic value of the susceptibility \(\chi \).

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Yolcu, C., Bérut, A., Falasco, G. et al. A General Fluctuation–Response Relation for Noise Variations and its Application to Driven Hydrodynamic Experiments. J Stat Phys 167, 29–45 (2017). https://doi.org/10.1007/s10955-017-1732-z

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s10955-017-1732-z

Keywords

Navigation