Skip to main content
Log in

Velocity-dependent forces and non-hydrodynamic initial conditions in quantum and classical fluids

  • Regular Article
  • Published:
The European Physical Journal Special Topics Aims and scope Submit manuscript

Abstract

We consider a fermionic fluid in a non-equilibrium steady state where the fluctuation–dissipation theorem is not valid and fields conjugate to the hydrodynamic variables are explicitly required to determine response functions. We identify velocity-dependent forces in the kinetic equation that are equivalent to such fields. They lead to driving terms in the hydrodynamic equations and to corrections to the hydrodynamic initial conditions.

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

Similar content being viewed by others

Data availability statement

No datasets were generated or analyzed during the current study.

Notes

  1. This approximation neglects effects of the temperature gradient that are less leading than the one caused by the \({\varvec{\nabla }}T\) term in Eq. (4b) [1].

  2. It is remarkable that this relation holds in complete generality, including systems that are not in equilibrium. We will elaborate on this elsewhere [7].

  3. The time scale on which Eq. (13) is valid is an example of a ‘slip time’, see Ref. [17] and references therein.

References

  1. T.R. Kirkpatrick, E.G.D. Cohen, J.R. Dorfman, Light scattering by a fluid in a nonequilibrium steady state. II. Large gradients. Phys. Rev. A 26, 995 (1982)

    Article  ADS  Google Scholar 

  2. J.R. Dorfman, T.R. Kirkpatrick, J.V. Sengers, Generic long-range correlations in molecular fluids. Ann. Rev. Phys. Chem. 45, 213 (1994)

    Article  CAS  ADS  Google Scholar 

  3. J.M. Ortiz de Zárate, J.V. Sengers, Hydrodynamic Fluctuations in Fluids and Fluid Mixtures (Elsevier, Amsterdam, 2007)

    Google Scholar 

  4. T.R. Kirkpatrick, D. Belitz, Fluctuating quantum kinetic theory. Phys. Rev. B 105, 245147 (2022)

    Article  CAS  ADS  Google Scholar 

  5. P.W. Anderson, Basic Notions of Condensed Matter Physics (Benjamin, Menlo Park, 1984)

    Google Scholar 

  6. T.R. Kirkpatrick, D. Belitz, J.R. Dorfman, Superfast signal propagation in fluids and solids in non-equilibrium steady states. J. Chem. Phys. B 125, 7499 (2021)

    Article  CAS  Google Scholar 

  7. T.R. Kirkpatrick, D. Belitz, unpublished results (2023)

  8. J.V. Sengers, J.M. Ortiz de Zárate, Chap. 3, Thermal fluctuations in non-equilibrium steady states, in Non-Equilibrium Thermodynamics with Applications. ed. by D. Bedeaux, S. Kjelstrup, J.V. Sengers (RSC Publishing, Cambridge, 2016), p.39

    Google Scholar 

  9. P.C. Hohenberg, B.I. Halperin, Theory of dynamic critical phenomena. Rev. Mod. Phys. 49, 435 (1977)

    Article  CAS  ADS  Google Scholar 

  10. P. Ehrenfest, T. Ehrenfest, The Conceptual Foundations of the Statistical Approach in Statistical Mechanics. Dover, New York (1990) . This is a translation of the article “Begriffliche Grundlagen der statistischen Auffassung in der Mechanik”, Enzyklopädie der mathematischen Wissenschaften, vol. IV/4, Art. 32, Teubner (Leipzig 1907-1914)

  11. E.A. Uehling, G.E. Uhlenbeck, Transport phenomena in Einstein-Bose and Fermi-Dirac gases I. Phys. Rev. 43, 552 (1933)

    Article  CAS  ADS  Google Scholar 

  12. J.R. Dorfman, H. van Beijeren, T.R. Kirkpatrick, Contemporary Kinetic Theory of Matter (Cambridge University Press, Cambridge, 2021)

    Book  Google Scholar 

  13. E.M. Lifshitz, L.P. Pitaevskii, Physical Kinetics (Butterworth-Heinemann, Oxford, 1981)

    Google Scholar 

  14. D. Belitz, T.R. Kirkpatrick, Soft modes in Fermi liquids at arbitrary temperatures. Phys. Rev. B 105, 245146 (2022)

    Article  CAS  ADS  Google Scholar 

  15. D. Forster, Hydrodynamic Fluctuations, Broken Symmetry, and Correlation Functions (Benjamin, Reading, 1975)

    Google Scholar 

  16. P. Chaikin, T.C. Lubensky, Principles of Condensed Matter Physics (Cambridge University Press, Cambridge, 1995)

    Book  Google Scholar 

  17. T.R. Kirkpatrick, D. Belitz, J.R. Dorfman, Non-hydrodynamic initial conditions are not soon forgotten. Phys. Rev. E 104, 024111 (2021)

    Article  CAS  PubMed  ADS  Google Scholar 

  18. J.M. Luttinger, Theory of thermal transport coefficients. Phys. Rev. 135, 1505 (1964)

    Article  MathSciNet  ADS  Google Scholar 

  19. L.D. Landau, E.M. Lifshitz, The Classical Theory of Fields (Pergamon, Oxford, 1971)

    Google Scholar 

  20. L.D. Landau, E.M. Lifshitz, Statistical Physics Part 1 (Pergamon, Oxford, 1980)

    Google Scholar 

  21. L.D. Landau, E.M. Lifshitz, Fluid Mechanics (Pergamon, Oxford, 1987)

    Google Scholar 

Download references

Acknowledgements

The authors would like to thank Thomas Schaefer for a discussion.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to D. Belitz.

Appendix 1: Derivation of Eq. (15)

Appendix 1: Derivation of Eq. (15)

Here, we show how to derive Eq. (15) from Eq. (14). We define a scalar product in the space of modes

$$\begin{aligned} \langle A_{\varvec{p}}\vert B_{\varvec{p}}\rangle = (1/V)\sum _{\varvec{p}} w({\varvec{p}}) A_{\varvec{p}} B_{\varvec{p}}, \end{aligned}$$
(16)

with \(w({\varvec{p}})\) as given after Eq. (6b) as the weight function. Let \({{\mathcal {L}}}_0\) be the hydrodynamic subspace spanned by the hydrodynamic modes \(a_{\rho }\), \(a_{\perp }\), and \(a_s\), and in addition the longitudinal component of the velocity and the second transverse component, and define a projector \({{\mathcal{P}}}\) that projects onto \({{\mathcal{L}}}_0\)

$$\begin{aligned} {{\mathcal {P}}} = \sum _{\alpha } \frac{\vert a_{\alpha }({\varvec{p}})\rangle \langle a_{\alpha }({\varvec{p}})\vert }{\langle a_{\alpha }({\varvec{p}})\vert a_{\alpha }({\varvec{p}})\rangle }. \end{aligned}$$
(17)

The projector onto the complementary space is \({{\mathcal {P}}}_{\perp } = \mathbb{1} - {{\mathcal {P}}}\), with \(\mathbb{1}\) the unit operator.

Now, consider Eq. (14), let \({\varvec{k}} = (k,0,0)\), and impose the initial conditions in the form of an initial shear velocity

$$\begin{aligned} u_{\perp }^{(0)}({\varvec{k}}) = \frac{1}{V} \sum _{\varvec{p}} v_{\varvec{p}}^y \delta f_{\varvec{p}}({\varvec{k}},t=0), \end{aligned}$$
(18a)

and an initial shear stress

$$\begin{aligned} \sigma _{\perp }^{(0)}({\varvec{k}}) = \frac{1}{V} \sum _{\varvec{p}} p_y v_{\varvec{p}}^x\, \delta f_{\varvec{p}}({\varvec{k}},t=0). \end{aligned}$$
(18b)

Operating on Eq. (14) from the left, we obtain

$$\begin{aligned} {{\mathcal {P}}} \delta f_{\varvec{p}}({\varvec{k}},t=0)= & {} \left( -iz + {{\mathcal {P}}} i{\varvec{k}}\cdot {\varvec{v}}_{\varvec{p}}\right) {{\mathcal {P}}} \delta f_{\varvec{p}}({\varvec{k}},z)\nonumber \\{} & {} + {{\mathcal {P}}} i{\varvec{k}}\cdot {\varvec{v}}_{\varvec{p}} {{\mathcal {P}}}_{\perp } \delta f_{\varvec{p}}({\varvec{k}},z). \end{aligned}$$
(19)

Here, we have used the projector properties \(\mathbb {1} = {{\mathcal {P}}} + {{\mathcal {P}}}_{\perp }\) and \({{\mathcal {P}}}\Lambda ({\varvec{p}}) = 0\). Performing the same operation with \({{\mathcal {P}}}_{\perp }\) instead of \({{\mathcal {P}}}\), we find

$$\begin{aligned} {{\mathcal {P}}}_{\perp } \delta f_{\varvec{p}}({\varvec{k}},t=0)= & {} -\left[ \Lambda _{\perp }({\varvec{p}}) + O(z,k)\right] {{\mathcal {P}}}_{\perp } \delta f_{\varvec{p}}({\varvec{k}},z)\nonumber \\{} & {} + {{\mathcal {P}}}_{\perp } i{\varvec{k}}\cdot {\varvec{v}}_{\varvec{p}} {{\mathcal {P}}} \delta f_{\varvec{p}}({\varvec{k}},z), \end{aligned}$$
(20)

where \(\Lambda _{\perp }({\varvec{p}}) = {{\mathcal {P}}}_{\perp } \Lambda ({\varvec{p}}) {{\mathcal {P}}}_{\perp }\). Note that \(\Lambda _{\perp }({\varvec{p}})\) has an inverse, since the zero eigenvalues of \(\Lambda ({\varvec{p}})\) have been projected out. Solving Eq. (20) for \({{\mathcal {P}}}_{\perp } \delta f_{\varvec{p}}({\varvec{k}},z)\), and inserting the result in Eq. (19), we find

$$\begin{aligned} & \left( -iz + {{\mathcal {P}}} i{\varvec{k}}\cdot {\varvec{v}}_{\varvec{p}} + {{\mathcal {P}}} i{\varvec{k}}\cdot {\varvec{v}}_{\varvec{p}} \Lambda _{\perp }^{-1}({\varvec{p}}) i{\varvec{k}}\cdot {\varvec{v}}_{\varvec{p}}\right) {{\mathcal {P}}} \delta f_{\varvec{p}}({\varvec{k}},z)\nonumber \\ & \quad = {{\mathcal {P}}} \delta f_{\varvec{p}}({\varvec{k}},t=0) \nonumber \\ & \qquad + {{\mathcal {P}}} i{\varvec{k}}\cdot {\varvec{v}}_{\varvec{p}} \Lambda _{\perp }^{-1}({\varvec{p}}) \delta f_{\varvec{p}}({\varvec{k}},t=0). \end{aligned}$$
(21)

Of the two initial-condition terms on the right-hand side of Eq. (21), the first one is in the hydrodynamic subspace, but the second one is not. Furthermore, the non-hydrodynamic initial condition involves a collision operator and hence is dissipative in nature, and Eq. (21) is a closed equation for \({{\mathcal {P}}} \delta f_{\varvec{p}}({\varvec{k}},z)\) only if one neglects the non-hydrodynamic initial condition.

By operating on Eq. (21) from the left with the hydrodynamic modes \(\langle a_{\alpha }({\varvec{p}})\vert\), we obtain the hydrodynamic equations as an initial-value problem. The only unusual aspect is the non-hydrodynamic initial condition. The latter does not contribute to the density equation, obtained by multiplying from the left with \(\langle 1\vert\), on account of the identity

$$\begin{aligned} \langle i{\varvec{k}}\cdot {\varvec{v}}_{\varvec{p}} \vert \Lambda _{\perp }^{-1}({\varvec{p}}) = 0, \end{aligned}$$
(22)

which holds, since \({\varvec{k}}\cdot {\varvec{v}}_{\varvec{p}} \in {{\mathcal {L}}}_0\). If the initial shear stress, Eq. (18b), is the only non-hydrodynamic initial condition, it does not contribute to the heat equation either, since the relevant matrix element vanishes by symmetry. It does, however, contribute to the equation for the shear velocity. The relevant matrix element is

$$\begin{aligned} M = \left\langle p_y\vert {{\mathcal {P}}} i{\varvec{k}}\cdot {\varvec{v}}_{\varvec{p}}^x \vert \Lambda _{\perp }^{-1}({\varvec{p}})\vert \delta f_{\varvec{p}}({\varvec{k}},t=0)\right\rangle . \end{aligned}$$
(23)

If we again take the initial shear stress to be the only non-hydrodynamic initial condition, we can project onto \(\sigma _{\perp }\) and have

$$\begin{aligned} M &= \left\langle p_y v_{\varvec{p}}^x \vert \Lambda _{\perp }^{-1}({\varvec{p}}) \vert p_y v_{\varvec{p}}^x \right\rangle \,\frac{ik}{\left\langle p_y v_{\varvec{p}}^x \vert p_y v_{\varvec{p}}^x \right\rangle }\,\\ &\quad\quad\left\langle p_y v_{\varvec{p}}^x \vert \delta f_{\varvec{p}}({\varvec{k}},t=0)\right\rangle . \end{aligned}$$
(24)

The first factor on the right-hand side of Eq. (24) we recognize as minus the shear viscosity \(\eta\), and if we remember that the mode \(a_{\perp }({\varvec{p}}) = p_y\) yields \(\rho\) times the transverse velocity \(u_{\perp }\), see Eq. (3), we obtain the transverse velocity equation in the form

$$\begin{aligned} \left( -iz + \nu {\varvec{k}}^2\right) \, u_{\perp }({\varvec{k}},z) = u_{\perp }^{(0)}({\varvec{k}}) - \frac{\nu i k}{\langle p_y v_{\varvec{p}}^x \vert p_y v_{\varvec{p}}^x \rangle }\,\sigma _{\perp }^{(0)}({\varvec{k}}), \end{aligned}$$
(25)

which is Eq. (15). Finally, the matrix element in the denominator is \(\langle p_y v_{\varvec{p}}^x \vert p_y v_{\varvec{p}}^x \rangle = 8(n\mu + Ts)/3\) by Eq. (A.23) in Ref. [14]. Ignoring the factor of 8/3 and the Ts correction to \(n\mu\), and using Eq. (12), we obtain the transverse velocity equation in the form of Eq. (15').

Rights and permissions

Springer Nature or its licensor (e.g. a society or other partner) 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

Kirkpatrick, T.R., Belitz, D. Velocity-dependent forces and non-hydrodynamic initial conditions in quantum and classical fluids. Eur. Phys. J. Spec. Top. 232, 3459–3466 (2023). https://doi.org/10.1140/epjs/s11734-023-01032-y

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1140/epjs/s11734-023-01032-y

Navigation