Skip to main content
Log in

Nonlinear Fluctuating Hydrodynamics for Anharmonic Chains

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

Abstract

With focus on anharmonic chains, we develop a nonlinear version of fluctuating hydrodynamics, in which the Euler currents are kept to second order in the deviations from equilibrium and dissipation plus noise are added. The required model-dependent parameters are written in such a way that they can be computed numerically within seconds, once the interaction potential, pressure, and temperature are given. In principle the theory is applicable to any one-dimensional system with local conservation laws. The resulting nonlinear stochastic field theory is handled in the one-loop approximation. Some of the large scale predictions can still be worked out analytically. For more details one has to rely on numerical simulations of the corresponding mode-coupling equations. In this way we arrive at detailed predictions for the equilibrium time correlations of the locally conserved fields of an anharmonic chain.

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.

Similar content being viewed by others

References

  1. Campbell, D.K., Rosenau, P., Zaslavsky, G.: Introduction: the Fermi–Pasta–Ulam problem—the first fifty years. Chaos 1, 015101 (2005)

    Article  ADS  Google Scholar 

  2. Gallavotti, G.: The Fermi–Pasta–Ulam Problem: A Status Report. Lecture Notes in Physics 728. Springer, Berlin (2008)

  3. Lepri, S., Livi, R., Politi, A.: Heat conduction in chains of nonlinear oscillators. Phys. Rev. Lett. 78, 1896–1899 (1997)

    Article  ADS  Google Scholar 

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

    Article  ADS  MathSciNet  Google Scholar 

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

    Article  ADS  Google Scholar 

  6. van Beijeren, H.: Exact results for anomalous transport in one-dimensional Hamiltonian systems. Phys. Rev. Lett. 108, 180601 (2012)

    Article  Google Scholar 

  7. Mendl, C., Spohn, H.: Dynamic correlators of FPU chains and nonlinear fluctuating hydrodynamics. Phys. Rev. Lett. 111, 230601 (2013)

    Article  ADS  Google Scholar 

  8. Forster, D.: Hydrodynamic Fluctuations. Broken Symmetry and Correlation Functions. Benjamin, New York (1975)

  9. Landau, L.D., Lifshitz, E.M.: Fluid Dynamics. Pergamon Press, New York (1963)

  10. Résibois, P., De Leener, M.: Classical Kinetic Theory of Fluids. Wiley, New York (1977)

  11. Tremblay, A.-M.S., Arai, M., Siggia, E.D.: Fluctuations about simple nonequilibrium steady states. Phys. Rev. A 23, 1451–1480 (1981)

    Article  ADS  Google Scholar 

  12. Kardar, M., Parisi, G., Zhang, Y.-C.: Dynamic scaling of growing interfaces. Phys. Rev. Lett. 56, 889–892 (1986)

    Article  ADS  MATH  Google Scholar 

  13. Prähofer, M.: Exact scaling functions for one-dimensional stationary KPZ growth. http://www-m5.ma.tum.de/KPZ

  14. Prähofer, M., Spohn, H.: Exact scaling functions for one-dimensional stationary KPZ growth. J. Stat. Phys. 115, 255–279 (2004)

    Article  ADS  MATH  Google Scholar 

  15. van Beijeren, H., Kutner, R., Spohn, H.: Excess noise for driven diffusive systems. Phys. Rev. Lett. 54, 2026–2029 (1985)

    Article  ADS  MathSciNet  Google Scholar 

  16. Colaiori, F., Moore, M.A.: Numerical solution of the mode-coupling equations for the Kardar–Parisi–Zhang equation in one dimension. Phys. Rev. E 65, 017105 (2001)

    Article  ADS  Google Scholar 

  17. Alder, B.J., Wainwright, T.E.: Decay of the velocity autocorrelation function. Phys. Rev. A 1, 18–21 (1970)

    Article  ADS  Google Scholar 

  18. Ernst, M.H., Hauge, E.H., van Leeuwen, J.M.J.: Asymptotic time behavior of correlation functions. II. Kinetic and potential terms. J. Stat. Phys. 15, 7–22 (1976)

    Google Scholar 

  19. Forster, D., Nelson, D.R., Stephen, M.J.: Large-distance and long-time properties of a randomly stirred fluid. Phys. Rev. A 16, 732–749 (1977)

    Article  ADS  MathSciNet  Google Scholar 

  20. Andreev, A.F.: The hydrodynamics of two- and one-dimensional liquids. Sov. Phys. JETP 51, 1038 (1980)

    ADS  Google Scholar 

  21. Samokhin, K.V.: Lifetime of excitations in a clean Luttinger liquid. J. Phys. Condens. Matter 10, L533 (1998)

    Article  ADS  Google Scholar 

  22. Punk, M., Zwerger, W.: Collective mode damping and viscosity in a 1D unitary Fermi gas. New J. Phys. 8, 168 (2006)

    Article  ADS  Google Scholar 

  23. Kulkarni, M., Lamacraft, A.: From GPE to KPZ: finite temperature dynamical structure factor of the 1D Bose gas. Phys. Rev. A 88, 021603(R) (2013)

    Article  ADS  Google Scholar 

  24. Arzamasovs, M., Bovo, F., Gangardt, D.M.: Kinetics of mobile impurities and correlation functions in one-dimensional superfluids at finite temperature. arXiv:1309.2647

  25. Lepri, S., Livi, R., Politi, A.: Universality of anomalous one-dimensional heat conductivity. Phys. Rev. E 68, 067102 (2003)

    Article  ADS  Google Scholar 

  26. Lee-Dadswell, G.R., Nickel, B.G., Gray, C.G.: Thermal conductivity and bulk viscosity in quartic oscillator chains. Phys. Rev. E 72, 031202 (2005)

    Article  ADS  Google Scholar 

  27. Zhao, H.: Identifying diffusion processes in one-dimensional lattices in thermal equilibrium. Phys. Rev. Lett. 96, 140602 (2006)

    Article  ADS  Google Scholar 

  28. Chen, S., Zhang, Y., Wang, J., Zhao, H.: Diffusion of heat, energy, momentum and mass in one-dimensional systems. Phys. Rev. E 87, 032153 (2013)

    Article  ADS  Google Scholar 

  29. Wang, L., Wang, T.: Power-law divergent heat conductivity in one-dimensional momentum-conserving nonlinear lattices. EPL 93, 54002 (2011)

    Article  ADS  Google Scholar 

  30. Chen, S., Zhang, Y., Wang, J., Zhao, H.: Breakdown of the power-law decay prediction of the heat current correlation in one-dimensional momentum conserving lattices. arXiv:1204.5933

  31. Zaburdaev, V., Denisov, S., Hänggi, P.: Perturbation spreading in many-particle systems: a random walk approach. Phys. Rev. Lett. 106, 180601 (2011)

    Article  ADS  Google Scholar 

  32. Lee-Dadswell, G.R., Nickel, B.G., Gray, C.G.: Detailed examination of transport coefficients in cubic-plus-quartic oscillator chains. J. Stat. Phys. 132, 1–33 (2008)

    Article  ADS  MATH  MathSciNet  Google Scholar 

  33. Zhong, Y., Zhang, Y., Wang, J., Zhao, H.: Normal heat conduction in one-dimensional momentum conserving lattices with asymmetric interactions. Phys. Rev. E 85, 060102(R) (2012)

    Article  ADS  Google Scholar 

  34. Wang, J.-S., Li, B.: Intriguing heat conduction in chains with transverse motion. Phys. Rev. Lett. 92, 074302 (2004)

    Article  ADS  Google Scholar 

  35. Grassberger, P., Nadler, W., Yang, L.: Heat conduction and entropy production in a one-dimensional hard-particle gas. Phys. Rev. Lett. 89, 180601 (2002)

    Article  ADS  Google Scholar 

  36. Cipriani, P., Denisov, S., Politi, A.: From anomalous energy diffusion to Levy walks and heat conductivity in one-dimensional systems. Phys. Rev. Lett. 94, 244301 (2005)

    Article  ADS  Google Scholar 

  37. Delfini, L., Denisov, S., Lepri, S., Livi, R., Mohanty, P.K.: Energy diffusion in hard-point systems. Eur. Phys. J. Special Topics 146, 21–35 (2007)

    Article  ADS  Google Scholar 

  38. Politi, A.: Heat conduction of the hard point chain at zero pressure. JSTAT 2011, P03028 (2011)

  39. Deutsch, J.M., Narayan, O.: Correlations and scaling in one-dimensional heat conduction. Phys. Rev. E 68, 041203 (2003)

    Article  ADS  Google Scholar 

  40. Bernardin, C., Olla, S.: Thermodynamics and non-equilibrium macroscopic dynamics of chains of anharmonic oscillators (2014). https://www.ceremade.dauphine.fr/~olla/springs13.pdf

  41. van Beijeren, H.: Statistical physics of systems out of equilibrium. Course given at KIAS, Seoul, Fall (2011). www.staff.science.uu.nl/beije10/outofeqseoul.htm

  42. Scheipers, J., Schirmacher, W.: Mode-coupling theory for the lattice dynamics of anharmonic crystals: self-consistent damping and the 1d Lennard-Jones chain. Z. Phys. B Condens. Matter 103, 547–553 (1997)

    Article  ADS  Google Scholar 

  43. Lepri, S.: Relaxation of classical many-body Hamiltonians in one dimension. Phys. Rev. E 58, 7165–7171 (1998)

    Article  ADS  Google Scholar 

  44. Delfini, L., Lepri, S., Livi, R., Politi, A.: Self-consistent mode-coupling approach to one-dimensional heat transport. Phys. Rev. E 73, 060201(R) (2006)

  45. Delfini, L., Lepri, S., Livi, R.: Anomalous kinetics and transport from 1D self-consistent mode-coupling theory. J. Stat. Mech. P02007 (2007)

  46. Narayan, O., Ramaswamy, S.: Anomalous heat conduction in one-dimensional momentum conserving systems. Phys. Rev. Lett. 89, 200601 (2002)

    Article  ADS  Google Scholar 

  47. Bertini, L., Giacomin, G.: Stochastic Burgers and KPZ equations from particle systems. Commun. Math. Phys. 183, 571–607 (1997)

    Article  ADS  MATH  MathSciNet  Google Scholar 

  48. Amir, G., Corwin, I., Quastel, J.: Probability distribution of the free energy of the continuum directed random polymer in 1+1 dimensions. Commun. Pure Appl. Math. 64, 466–537 (2011)

    Article  MATH  MathSciNet  Google Scholar 

  49. Hairer, M.: Solving the KPZ equation. Ann. Math. 178, 559–664 (2013)

    Article  MATH  MathSciNet  Google Scholar 

  50. Imamura, T., Sasamoto, T.: Stationary correlations for the 1D KPZ equation. J. Stat. Phys. 150, 908–939 (2013)

    Article  ADS  MATH  MathSciNet  Google Scholar 

  51. Prähofer, M., Spohn, H.: Universal distributions for growth processes in \(1+1\) dimensions and random matrices. Phys. Rev. Lett. 84, 4882–4885 (2000)

    Article  ADS  Google Scholar 

  52. Ferrari, P., Spohn, H.: Scaling limit for the space-time covariance of the stationary totally asymmetric simple exclusion process. Commun. Math. Phys. 265, 1–44 (2006)

    Article  ADS  MATH  MathSciNet  Google Scholar 

  53. Corwin, I.: The Kardar–Parisi–Zhang equation and universality class. Random Matrices Theory Appl 1, 113001 (2012)

    Article  MathSciNet  Google Scholar 

  54. Ertaş, D., Kardar, M.: Dynamic roughening of directed lines. Phys. Rev. Lett. 69, 929–932 (1992)

    Article  ADS  Google Scholar 

  55. Ertaş, D., Kardar, M.: Dynamic relaxation of drifting polymers: a phenomenological approach. Phys. Rev. E 48, 1228–1245 (1993)

    Article  ADS  Google Scholar 

  56. Kardar, M.: Nonequilibrium dynamics of interfaces and lines. Phys. Rep. 301, 85–112 (1998)

    Article  ADS  Google Scholar 

  57. Levine, A., Ramaswamy, S., Frey, E., Bruinsma, R.: Screened and unscreened phases in sedimenting suspensions. Phys. Rev. Lett. 81, 5944–5947 (1998)

    Article  ADS  Google Scholar 

  58. Lahiri, R., Ramaswamy, S.: Are steadily moving crystals unstable? Phys. Rev. Lett. 79, 1150–1153 (1997)

    Article  ADS  Google Scholar 

  59. Das, D., Basu, A., Barma, M., Ramaswamy, S.: Weak and strong dynamic scaling in a one-dimensional driven coupled-field model: effects of kinematic waves. Phys. Rev. E 64, 021402 (2001)

    Article  ADS  Google Scholar 

  60. Yanase, S.: New one-dimensional model equations of magnetohydrodynamic turbulence. Phys. Plasmas 4, 1010 (1997)

    Article  ADS  MathSciNet  Google Scholar 

  61. Fleischer, J., Diamond, P.H.: Compressible Alfven turbulence in one dimension. Phys. Rev. E 58, R2709–R2712 (1998)

    Article  ADS  MathSciNet  Google Scholar 

  62. Basu, A., Bhattacharjee, J.K., Ramaswamy, S.: Mean magnetic field and noise cross-correlation in magnetohydrodynamic turbulence: results from a one-dimensional model. Eur. Phys. J. B 9, 725–730 (1999)

    Article  ADS  Google Scholar 

  63. Arndt, P.F., Heinzel, T., Rittenberg, V.: Spontaneous breaking of translational invariance and spatial condensation in stationary states on a ring. I. The neutral system. J. Stat. Phys. 97, 1–65 (1999)

    Article  ADS  MATH  MathSciNet  Google Scholar 

  64. Ferrari, P., Sasamoto, S., Spohn, H.: Coupled Kardar–Parisi–Zhang equations in one dimension. J. Stat. Phys. 153, 377–399 (2013)

    Article  ADS  MathSciNet  Google Scholar 

  65. Basile, G., Bernardin, C., Olla, S.: Momentum conserving model with anomalous thermal conductivity in low dimensional systems. Phys. Rev. Lett. 96, 204303 (2006)

    Article  ADS  Google Scholar 

  66. Basile, G., Delfini, L., Lepri, S., Livi, R., Olla, S., Politi, A.: Anomalous transport and relaxation in classical one-dimensional models. Eur. J. Phys. Special Topics 151, 85–93 (2007)

    Article  ADS  Google Scholar 

  67. Gerschenfeld, A.: Fluctuations de courant hors d’quilibre. Thèse de Doctorat, Lab. Physique Statistique ENS, Paris (2012)

  68. Bernardin, C., Stoltz, G.: Anomalous diffusion for a class of systems with two conserved quantities. Nonlinearity 25, 1099–1133 (2012)

    Article  ADS  MATH  MathSciNet  Google Scholar 

  69. Bernardin, C., Gonçalves, P.: Anomalous fluctuations for a perturbed Hamiltonian system with exponential interactions. Commun. Math. Phys. 325, 291–332 (2014)

    Google Scholar 

  70. Spohn, H.: Large Scale Dynamics of Interacting Particles. Texts and Monographs in Physics. Springer, Heidelberg (1991)

  71. Künsch, H.: Non reversible stationary measures for infinite interacting particle systems. Z. Wahrscheinlichkeitstheorie verw. Geb. 66, 407–424 (1984)

    Article  MATH  Google Scholar 

  72. Grisi, R., Schütz, G.: Current symmetries for particle systems with several conservation laws. J. Stat. Phys. 145, 1499–1512 (2011)

    Article  ADS  MATH  MathSciNet  Google Scholar 

  73. Tóth, B., Valkó, B.: Onsager relations and Eulerian hydrodynamic limit for systems with several conservation laws. J. Stat. Phys. 112, 497–521 (2003)

    Article  MATH  Google Scholar 

  74. Uchaikin, V., Zolotarev, V.: Chance and Stability. Stable Distributions and Applications. W. de Gruyter, Berlin (1999)

Download references

Acknowledgments

The topic of my article reflects longstanding interests. The actual study was triggered, in fact, by two workshops in the fall 2012 on transport in one-dimensional systems, one at the ICTP Trieste, organized by A. Dhar, M.N. Kiselev, Y.A. Kosevich, R. Livi, and one at BIRS, Banff, organized by J.L. Lebowitz, S. Olla, G. Stoltz, for both of which I am most grateful. I thank H. van Beijeren for sharing his insights on mode-coupling theory, S. Olla for emphasizing the hydrodynamics of anharmonic chains, J. Krug for pointing at the early literature on coupled KPZ equations, P. Ferrari, C. Mendl, T. Sasamoto for constant help and encouragement, and C. Bernardin, S. Lepri, A. Politi, H. Posch, G. Schütz, H. Zhao for highly useful discussions. Support by Fund For Math is acknowledged.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Herbert Spohn.

Appendices

Appendix 1: Coupling Constants

1.1 (a) Equilibrium Susceptibilities, Currents

Averages with respect to \(Z^{-1}\exp [-\beta (V(y)+py)]dy\) are denoted in this appendix by \(\langle \cdot \rangle \) with the dependence on \(p,\beta \) being suppressed. \(\langle X_1;X_2\rangle = \langle X_1 X_2\rangle - \langle X_1\rangle \langle X_2\rangle \) is the second cumulant and

$$\begin{aligned} \langle X_1;X_2;X_3\rangle&= \langle X_1 X_2 X_3 \rangle \nonumber \\&\quad -\, \langle X_1 X_2\rangle \langle X_3\rangle - \langle X_1 X_3\rangle \langle X_2\rangle - \langle X_2 X_3\rangle \langle X_1\rangle + 2 \langle X_1\rangle \langle X_2\rangle \langle X_3\rangle \qquad \end{aligned}$$
(8.1)

the third cumulant. It holds

$$\begin{aligned}&\quad \partial _p \langle X_1\rangle =-\beta \langle X_1;y\rangle ,\quad \partial _\beta \langle X_1\rangle =- \langle X_1;V+py\rangle , \nonumber \\&\quad \partial _p \langle X_1;X_2\rangle =-\beta \langle X_1;X_2;y\rangle ,\quad \partial _\beta \langle X_1;X_2\rangle =- \langle X_1;X_2;V+py\rangle . \end{aligned}$$
(8.2)

In equilibrium, \(\{r_j,p_j, j\in \mathbb {Z}\}\) are independent random variables. \(p_j\) has a Gaussian density with mean zero and variance \(\beta ^{-1}\) and \(r_j\) has the density \(Z^{-1}\exp [-\beta (V(y)+py)]\). Hence \(S(j,0) = \delta _{0j} C\) with

$$\begin{aligned} C=\left( \begin{array}{l@{\quad }l@{\quad }l} \langle y;y\rangle &{} 0 &{} \langle y;V\rangle \\ 0 &{} \beta ^{-1} &{} 0 \\ \langle y;V\rangle &{} 0 &{} \tfrac{1}{2}\beta ^{-2}+\langle V;V\rangle \end{array}\,\right) . \end{aligned}$$
(8.3)

The macroscopic conserved fields are \((\ell ,\mathsf {u},\mathfrak {e})\). Their currents are

$$\begin{aligned} \big ({-}\mathsf {u},p(\ell ,\mathfrak {e}-\tfrac{1}{2} \mathsf {u}^2), \mathsf {u}p(\ell ,\mathfrak {e}-\tfrac{1}{2}\mathsf {u}^2)\big ). \end{aligned}$$
(8.4)

We linearize the Euler equations at \((\ell , \mathsf {u}=0,\mathsf {e})\), \(\mathsf {e}\) the internal energy, where

$$\begin{aligned} \ell =\langle y\rangle ,\quad \mathsf {e}= \tfrac{1}{2}\beta ^{-1} + \langle V\rangle , \end{aligned}$$
(8.5)

which defines \((p,\beta ) \mapsto (\ell (p,\beta ),\mathsf {e}(p,\beta ))\). Inverting this map yields

$$\begin{aligned} p=p(\ell ,\mathsf {e}),\quad \beta =\beta (\ell ,\mathsf {e}). \end{aligned}$$
(8.6)

1.2 (b) Linearization, Normal Modes

The linearized currents are

$$\begin{aligned} A=\left( \begin{array}{l@{\quad }l@{\quad }l} 0 &{} -1 &{} 0 \\ \partial _\ell p &{} 0 &{} \partial _\mathsf {e}p \\ 0 &{} p &{} 0 \end{array}\right) \end{aligned}$$
(8.7)

with the property that

$$\begin{aligned} AC=CA^{\mathrm {T}}. \end{aligned}$$
(8.8)

\(A\) has the eigenvalues \(c_\sigma =\sigma c\), \(c_0=0\), \(\sigma =\pm 1\), \(c\) the sound speed,

$$\begin{aligned} c^2= -\partial _{\ell } p+p \partial _\mathsf {e}p>0. \end{aligned}$$
(8.9)

\(A\) has right eigenvectors defined by \(A|\psi _\alpha \rangle = c_\alpha |\psi _\alpha \rangle \) and given by

$$\begin{aligned} \psi _0=Z^{-1}_0\left( \begin{array}{l} \partial _\mathsf {e} p\\ 0\\ -\partial _\ell p \\ \end{array}\right) , \quad \psi _\sigma =Z^{-1}_\sigma \left( \begin{array}{l} -1\\ \sigma c\\ p \\ \end{array}\right) \end{aligned}$$
(8.10)

and left eigenvectors defined by \(\langle \tilde{\psi }_\alpha |A = c_\alpha \langle \tilde{\psi }_\alpha |\) and given by

$$\begin{aligned} \tilde{\psi }_0=\tilde{Z}^{-1}_0 \left( \begin{array}{l} p\\ 0\\ 1 \\ \end{array}\right) , \quad \tilde{\psi }_\sigma =\tilde{Z}^{-1}_\sigma \left( \begin{array}{l} \partial _\ell p \\ \sigma c\\ \partial _\mathsf {e} p \\ \end{array}\right) , \end{aligned}$$
(8.11)

which satisfy \(\langle \tilde{\psi }_\alpha |\psi _\beta \rangle =0\) for \(\alpha \ne \beta \). The linear transformation to normal modes, \(\vec {\phi }=R\vec {u}\), is defined through

$$\begin{aligned} RAR^{-1}= \mathrm {diag}(-c,0,c),\quad RCR^\mathrm {T}=1. \end{aligned}$$
(8.12)

The first identity can be achieved by setting

$$\begin{aligned}&\quad R= \left( \begin{array}{l} \langle \tilde{\psi }_-|\\ \langle \tilde{\psi }_0|\\ \langle \tilde{\psi }_+| \\ \end{array}\right) = (\tilde{Z}_{1})^{-1} \left( \begin{array}{l@{\quad }l@{\quad }l} \partial _\ell p&{}- c&{}\partial _ \mathsf {e}p\\ \tilde{\kappa }p&{}0&{} \tilde{\kappa }\\ \partial _\ell p&{} c &{}\partial _ \mathsf {e}p\\ \end{array}\right) ,\end{aligned}$$
(8.13)
$$\begin{aligned}&\quad R^{-1}=\big (|\psi _-\rangle |\psi _0\rangle |\psi _+\rangle \big ) = (Z_{1})^{-1} \left( \begin{array}{l@{\quad }l@{\quad }l} -1&{}\kappa \partial _ \mathsf {e}p&{}- 1\\ - c&{}0&{}c\\ p&{}-\kappa \partial _\ell p&{}p \end{array}\right) . \end{aligned}$$
(8.14)

The normalization factors are still free, but up to an overall factor of \(-1\) determined by the second identity of (8.12),

$$\begin{aligned}&\quad Z_0\tilde{Z}_0=c^2,\quad \tilde{Z}^2_0=\tfrac{1}{2}\beta ^{-2}+\langle V+py;V+py\rangle = \Gamma c^2, \nonumber \\&\quad Z_\sigma \tilde{Z}_\sigma =2c^2,\quad \tilde{Z}^2_\sigma =2 \beta ^{-1}c^2,\quad \kappa = Z_1/Z_0, \quad \tilde{\kappa } = \tilde{Z}_1/ \tilde{Z}_0. \end{aligned}$$
(8.15)

The Euler part of the equations of motion reads

$$\begin{aligned} \partial _t u_\alpha + \partial _x\big ((A\vec {u})_\alpha + \tfrac{1}{2} \langle \vec {u}, H^{\alpha }\vec {u}\rangle \big )= 0. \end{aligned}$$
(8.16)

Using \(\vec {\phi } = R \vec {u}\), one arrives at

$$\begin{aligned} \partial _t \phi _\alpha + \partial _x\big (c_\alpha \phi _\alpha + \tfrac{1}{2}\sum _{\alpha ' =1}^3 R_{\alpha \alpha '} \langle R^{-1}\vec {\phi }, H^{\alpha '} R^{-1}\vec {\phi }\rangle \big )= 0, \end{aligned}$$
(8.17)

which implies

$$\begin{aligned} \partial _t \phi _\alpha + \partial _x\big ( c_\alpha \phi _\alpha + \langle \vec {\phi }, G^{\alpha }\vec {\phi }\rangle \big )= 0 \end{aligned}$$
(8.18)

with the normal mode coupling constants

$$\begin{aligned} G^\alpha _{\beta \gamma }=\tfrac{1}{2}\sum _{\alpha '=1}^3R_{\alpha \alpha '} \langle \psi _\beta , H^{\alpha '} \psi _\gamma \rangle . \end{aligned}$$
(8.19)

1.3 (c) Hessians and \(G\) Couplings

The three Hessians are

$$\begin{aligned} H^\ell =0,\quad H^\mathsf {u}= \left( \begin{array}{l@{\quad }l@{\quad }l} \partial ^2_\ell p &{} 0 &{} \partial _\ell \partial _\mathsf {e}p \\ 0 &{} -\partial _\mathsf {e}p &{} 0 \\ \partial _\ell \partial _\mathsf {e}p &{} 0 &{} \partial ^2_\mathsf {e}p \end{array}\right) , \quad H^\mathsf {e}= \left( \begin{array}{l@{\quad }l@{\quad }l} 0 &{} \partial _\ell p &{} 0 \\ \partial _\ell p &{} 0 &{} \partial _\mathsf {e}p \\ 0 &{} \partial _\mathsf {e}p &{} 0 \end{array}\,\right) . \end{aligned}$$
(8.20)

The \(H^\mathsf {u}\) matrix elements are given by

$$\begin{aligned}&\qquad \langle \psi _0,H^\mathsf {u} \psi _0\rangle =\frac{1}{Z^2_0} \big (\partial ^2_\ell p\partial _\mathsf {e}p \partial _\mathsf {e}p-2\partial _\ell \partial _\mathsf {e}p \partial _\ell p \partial _\mathsf {e}p +\partial ^2_\mathsf {e}p \partial _\ell p\partial _\ell p\big ),\end{aligned}$$
(8.21)
$$\begin{aligned}&\qquad \langle \psi _0,H^\mathsf {u} \psi _\sigma \rangle =\frac{1}{Z_0 Z_\sigma } \big (-\partial ^2_\ell p\partial _\mathsf {e}p-p \partial ^2_\mathsf {e}p\partial _\ell p +p \partial _\ell \partial _\mathsf {e}p \partial _\mathsf {e}p +\partial _\ell \partial _\mathsf {e} p \partial _\ell p\big ),\qquad \quad \end{aligned}$$
(8.22)
$$\begin{aligned}&\qquad \langle \psi _{\sigma },H^\mathsf {u} \psi _{\sigma '}\rangle =\frac{1}{Z_\sigma Z_{\sigma '}} \big (\partial ^2_\ell p- 2p\partial _\ell \partial _\mathsf {e}p +p^2 \partial ^2_\mathsf {e}p-\sigma \sigma ' c^2 \partial _\mathsf {e}p \big ) \end{aligned}$$
(8.23)

and the \(H^\mathsf {e}\) matrix elements are

$$\begin{aligned} \langle \psi _0,H^\mathsf {e} \psi _0\rangle = 0,\quad \langle \psi _0,H^\mathsf {e}\psi _\sigma \rangle = 0,\quad \langle \psi _{\sigma },H^\mathsf {e} \psi _{\sigma '}\rangle = c(2\beta )^{-1}(\sigma + \sigma '). \end{aligned}$$
(8.24)

Denoting the standard basis vectors by \(\mathrm {e}_1,\mathrm {e}_2,\mathrm {e}_3\), one has

$$\begin{aligned} R\mathrm {e}_2 = (\tfrac{1}{2}\beta )^{1/2}(- \mathrm {e}_1 + \mathrm {e}_3),\quad R\mathrm {e}_3 = c^{-1}(\tfrac{1}{2}\beta )^{1/2}\partial _\mathsf {e}p (\mathrm {e}_1 + \mathrm {e}_3) + (\tilde{Z}_0)^{-1} \mathrm {e}_2 \end{aligned}$$
(8.25)

and arrives at the coupling constants

$$\begin{aligned}&\qquad G^\sigma _{\beta \gamma }=\tfrac{1}{2}\sigma (\tfrac{1}{2}\beta )^{1/2} \langle \psi _\beta ,H^\mathsf {u}\psi _\gamma \rangle + \tfrac{1}{2}\partial _\mathsf {e}p (\tfrac{1}{2}\beta )^{1/2}{c} \langle \psi _\beta ,H^\mathsf {e}\psi _\gamma \rangle , \end{aligned}$$
(8.26)
$$\begin{aligned}&\qquad G^0_{\beta \gamma }=\frac{1}{2\tilde{Z}_0} \langle \psi _\beta ,H^\mathsf {e}\psi _\gamma \rangle . \end{aligned}$$
(8.27)

They have the symmetries

$$\begin{aligned}&\qquad G^\alpha _{\beta \gamma } = G^\alpha _{\gamma \beta }, \quad G^\sigma _{\alpha \beta } = - G^{-\sigma }_{-\alpha -\beta },\quad G^\sigma _{-10} = G^\sigma _{01}, \nonumber \\&\qquad G^0_{\sigma \sigma } = G^0_{-\sigma -\sigma },\quad G^0_{\alpha \beta } = 0\,\, \mathrm {otherwise}. \end{aligned}$$
(8.28)

1.4 (d) Transformation to Canonical Variables

We transform the couplings \(G\) from microcanonical to canonical variables. In canonical variables, all \(G\) coefficients are given by up to third order cumulants in \(y,V,V+py\), which can be completely expressed in terms of one-dimensional integrals. A Mathematica program computes all \(G\) coefficients for specified \(V,p,\beta \) and by the same procedure also the \(R,R^{-1}\) matrices.

Differentiating the identities \(p(\ell (p,\beta ),\mathsf {e}(p,\beta ))= p\) and \(\beta (\ell (p,\beta ),\mathsf {e}(p,\beta ))= \beta \) with respect to \(p\) and \(\beta \) yields

$$\begin{aligned} \left( \begin{array}{l} 1 \\ 0 \\ \end{array}\right) = \left( \begin{array}{l@{\quad }l} \partial _p\ell &{} \partial _p\mathsf {e} \\ \partial _\beta \ell &{} \partial _\beta \mathsf {e} \\ \end{array}\right) \left( \begin{array}{l} \partial _\ell p \\ \partial _\mathsf {e} p \\ \end{array}\,\right) , \end{aligned}$$
(8.29)
$$\begin{aligned} \left( \begin{array}{l} 0 \\ 1\\ \end{array}\right) = \left( \begin{array}{l@{\quad }l} \partial _p\ell &{} \partial _p\mathsf {e} \\ \partial _\beta \ell &{} \partial _\beta \mathsf {e} \\ \end{array}\right) \left( \begin{array}{l} \partial _\ell \beta \\ \partial _\mathsf {e} \beta \\ \end{array}\,\right) . \end{aligned}$$
(8.30)

Hence

$$\begin{aligned}&\partial _\ell p = \Gamma ^{-1}\partial _\beta \mathsf {e} =-\Gamma ^{-1} \big ( \tfrac{1}{2}\beta ^{-2} +\langle V;V+py\rangle \big ), \nonumber \\&\partial _\mathsf {e} p = -\Gamma ^{-1}\partial _\beta \ell =\Gamma ^{-1}\langle y;V+py\rangle ,\nonumber \\&\Gamma = \beta \big (\langle y;y\rangle \langle V;V\rangle - \langle y;V\rangle ^2\big ) +\tfrac{1}{2}\beta ^{-1}\langle y;y\rangle , \end{aligned}$$
(8.31)

and for the velocity of sound

$$\begin{aligned} c^2= \frac{1}{\Gamma }\big ( \tfrac{1}{2}\beta ^{-2}+\langle V+py;V+py\rangle \big ). \end{aligned}$$
(8.32)

We collect first and second derivatives of \(\ell ,\mathsf {e}\) and first derivatives of \(\Gamma \),

$$\begin{aligned}&\partial _p \ell = - \beta \langle y;y\rangle ,\quad \partial _\beta \ell = - \langle y;V + py\rangle ,\end{aligned}$$
(8.33)
$$\begin{aligned}&\partial _p \mathsf {e} = -\beta \langle y;V\rangle ,\quad \partial _\beta \mathsf {e}= - \tfrac{1}{2} \beta ^{-2}- \langle V;V +py \rangle ,\end{aligned}$$
(8.34)
$$\begin{aligned}&\partial _p\partial _\beta \ell = - \langle y;y\rangle +\beta \langle y;y;V+ py\rangle ,\quad \partial _\beta ^2 \ell = \langle y;V+py;V +py\rangle ,\end{aligned}$$
(8.35)
$$\begin{aligned}&\partial _p\partial _\beta \mathsf {e} = - \langle y;V\rangle + \beta \langle y;V; V +py\rangle , \quad \partial _\beta ^2\mathsf {e}= \beta ^{-3} + \langle V;V+py;V+py\rangle ,\end{aligned}$$
(8.36)
$$\begin{aligned}&\partial _p \Gamma \!=\! \beta ^2\big ( -\langle y;y;y\rangle \langle V;V\rangle \!-\! \langle y;y\rangle \langle y;V;V\rangle \!+\! 2 \langle y;V\rangle \langle y;y;V\rangle \big ) -\tfrac{1}{2}\langle y;y;y\rangle ,\qquad \end{aligned}$$
(8.37)
$$\begin{aligned}&\partial _\beta \Gamma = \langle y;y\rangle \langle V;V\rangle - \langle y;V\rangle ^2 - \tfrac{1}{2}\beta ^{-2}\langle y;y\rangle +\beta \big ( - \langle y;y;V+py\rangle \langle V;V\rangle \nonumber \\&\qquad \quad \,\,\, - \langle y;y\rangle \langle V;V;V\!+\!py\rangle +2 \langle y;V\rangle \langle y;V;V\!+\!py\rangle ) \!-\! \tfrac{1}{2} \beta ^{-1} \langle y;y;V+py\rangle \big ).\qquad \quad \end{aligned}$$
(8.38)

To complete our task we still need the second derivatives \(\partial _\ell \partial _\ell p\), \(\partial _\ell \partial _\mathsf {e} p\), \(\partial _\mathrm {e}\partial _\mathsf {e} p\), where we start from

$$\begin{aligned} \left( \begin{array}{l} \partial _p(\Gamma ^{-1} \partial _\beta \mathsf {e}) \\ \partial _\beta (\Gamma ^{-1} \partial _\beta \mathsf {e} )\\ \end{array}\right) = \left( \begin{array}{l@{\quad }l} \partial _p\ell &{} \partial _p\mathsf {e} \\ \partial _\beta \ell &{} \partial _\beta \mathsf {e} \\ \end{array}\right) \left( \begin{array}{l} \partial _\ell \partial _\ell p \\ \partial _\mathsf {e} \partial _\ell p \\ \end{array}\right) , \end{aligned}$$
(8.39)
$$\begin{aligned} -\left( \begin{array}{l} \partial _p(\Gamma ^{-1} \partial _\beta \ell )\\ \partial _\beta ( \Gamma ^{-1} \partial _\beta \ell )\\ \end{array}\right) = \left( \begin{array}{l@{\quad }l} \partial _p\ell &{} \partial _p\mathsf {e} \\ \partial _\beta \ell &{} \partial _\beta \mathsf {e} \\ \end{array}\right) \left( \begin{array}{l} \partial _\ell \partial _\mathsf {e} p \\ \partial _\mathsf {e} \partial _\mathsf {e} p \\ \end{array}\right) . \end{aligned}$$
(8.40)

Inverting (8.39), (8.40) we arrive at

$$\begin{aligned}&\partial ^2_\ell p= - \Gamma ^{-2}(\partial _p\mathsf {e}\partial _\beta ^2\mathsf {e} -\partial _\beta \mathsf {e} \partial _p\partial _\beta \mathsf {e} )+ \Gamma ^{-3}(\partial _p\mathsf {e}\partial _\beta \mathsf {e}\partial _\beta \Gamma -\partial _\beta \mathsf {e}\partial _\beta \mathsf {e}\partial _p\Gamma ),\end{aligned}$$
(8.41)
$$\begin{aligned}&\partial _\ell \partial _\mathsf {e} p = -\Gamma ^{-2}(- \partial _p\mathsf {e} \partial _\beta ^2 \mathsf {e}+ \partial _\beta \mathsf {e} \partial _p\partial _\beta \ell ) + \Gamma ^{-3}(-\partial _p\mathsf {e} \partial _\beta \ell \partial _\beta \Gamma + \partial _\beta \mathsf {e} \partial _\beta \ell \partial _p\Gamma ),\qquad \quad \end{aligned}$$
(8.42)
$$\begin{aligned}&\partial _\mathsf {e}^2p = \Gamma ^{-2}(- \partial _p\ell \partial _\beta ^2\ell + \partial _\beta \ell \partial _p\partial _\beta \ell ) - \Gamma ^{-3}(-\partial _\beta \ell \partial _p \ell \partial _\beta \Gamma + \partial _\beta \ell \partial _\beta \ell \partial _p\Gamma ). \end{aligned}$$
(8.43)

By successive substitutions, the coupling constants in (8.26), (8.26) are expressed in terms of cumulants in \(y, V, V+py\) at most of order three.

1.5 (e) Even Potential, Zero Pressure

For an even potential and \(p=0\) by symmetry \(\langle y;V\rangle =0\), \(\langle y;V;V\rangle =0\), \(\langle y;y;y\rangle =0\). This simplifies the expressions for the couplings and \(c^2\). It holds

$$\begin{aligned} c^2 = (\beta \langle y;y\rangle )^{-1}. \end{aligned}$$
(8.44)

The only non-zero couplings are

$$\begin{aligned}&G^\sigma _{0 \sigma '} = G^\sigma _{\sigma '0}= \tfrac{1}{2}\sigma ' c^3 \big ((2\beta ^2)^{-1}+ \langle V;V\rangle \big )^{-1/2},\quad \sigma ,\sigma '=\pm 1,\nonumber \\&G^0_{\sigma \sigma } = \tfrac{1}{2}\sigma c \beta ^{-1} \big ((2\beta ^2)^{-1}+ \langle V;V\rangle \big )^{-1/2}. \end{aligned}$$
(8.45)

1.6 (f) Second Sum Rule

As claimed in Eq. (3.9), the linearized Euler currents, \(A\), and the susceptibility matrix, \(C\), satisfy

$$\begin{aligned} AC = CA^\mathrm {T}. \end{aligned}$$
(8.46)

This relation is well known for classical fluids, see e.g. [70]. In fact, (8.46) is very general and relies only on space–time stationarity. Let us denote, independently of any particular model, the conserved fields by \(\eta _\alpha (j,t)\), \(j\in \mathbb {Z}\), \(t \in \mathbb {R}\), \(\alpha = 1,\ldots ,n\), which are assumed to be space–time stationary with zero mean. The corresponding currents are denoted by \(\mathcal {J}_\alpha (j,t)\). By stationarity

$$\begin{aligned} S_{\alpha \beta }(j,t) = \langle \eta _\alpha (j,t) \eta _\beta (0,0)\rangle = S_{\beta \alpha }(-j,-t). \end{aligned}$$
(8.47)

Using the conservation law,

$$\begin{aligned} \begin{aligned} \frac{d}{dt}\sum _{j \in \mathbb {Z}} jS_{\alpha \beta }(j,t)&= \sum _{j \in \mathbb {Z}} j \langle (\mathcal {J}_\alpha (j-1,t) - \mathcal {J}_\alpha (j,t))\eta _\beta (0,0)\rangle \\&= \sum _{j \in \mathbb {Z}} \langle \mathcal {J}_\alpha (j,t)\eta _\beta (0,0)\rangle = \sum _{j \in \mathbb {Z}} \langle \mathcal {J}_\alpha (0,0)\eta _\beta (-j,-t)\rangle \\&= \sum _{j \in \mathbb {Z}} \langle \mathcal {J}_\alpha (0,0)\eta _\beta (j,0)\rangle . \end{aligned} \end{aligned}$$
(8.48)

As standard for mechanical systems in thermal equilibrium, but also valid for stochastic lattice gases with conservation laws [71, 72], the infinite volume average can be obtained from a system on a ring by introducing a chemical potential, \(\mu _\alpha \), for the density \(\rho _\alpha \). Hence

$$\begin{aligned} \sum _{j \in \mathbb {Z}} \langle \mathcal {J}_\alpha (0,0)\eta _\beta (j,0)\rangle&= \frac{\partial }{\partial \mu _{\beta }} \langle \mathcal {J}_\alpha (0,0)\rangle _{\vec {\mu }}\end{aligned}$$
(8.49)
$$\begin{aligned}&= \sum _{\gamma =1}^{n} \frac{\partial }{\partial \rho _{\gamma }} \langle \mathcal {J}_\alpha (0,0)\rangle _{\vec {\rho }} \,\frac{\partial \rho _{\gamma }}{\partial \mu _{\beta }} = (AC)_{\alpha \beta } \end{aligned}$$
(8.50)

and

$$\begin{aligned} \sum _{j \in \mathbb {Z}} jS_{\alpha \beta }(j,t) = (AC)_{\alpha \beta }t +\sum _{j \in \mathbb {Z}} jS_{\alpha \beta }(j,0). \end{aligned}$$
(8.51)

Summing in Eq. (8.47) over \(j\) yields

$$\begin{aligned} (AC)_{\alpha \beta }t = (AC)_{\beta \alpha }t, \end{aligned}$$
(8.52)

which is the desired identity.

For many-component lattice gases Eq. (3.9) was noted by Tóth and Valkó [73] in a special case and proved in generality by Grisi and Schütz [72].

Appendix 2: The Hard-Point Gas with Alternating Masses

As the name suggests, the hard-point gas consists of point particles which collide elastically. This amounts to merely exchanging the labels and thus ideal gas dynamics. To introduce more chaotic elements one modifies the model to have alternating masses, say \(m_0\) and \(m_1\). Numerically the choice \(m_1/m_0 = 3\) seems to have good time mixing. The simulation runs from collision to collision which is much faster than solving differential equations. Regarded as a chain, the hard-point gas has maximal simplicity and appears to be a favorable candidate the check our predictions. At the same time it serves as a nice illustration of the method.

As a novel feature, the unit cell consists of two particles. Let us first consider the case of hard-points with a general potential \(V\) and let us reintroduce the mass \(m\) of a particle. Then the hydrodynamic currents from (2.22) are modified as

$$\begin{aligned} -\tfrac{1}{m}\mathsf {u}, p, \tfrac{1}{m}\mathsf {u}p, \quad p = p(\ell ,\mathfrak {e} - \tfrac{1}{2m}\mathsf {u}^2), \end{aligned}$$
(9.1)

\(\mathsf {u}\) the momentum density. Alternating masses modify the currents. We claim that in (9.1) one merely has to substitute for \(m\) the average mass

$$\begin{aligned} \bar{m} = \tfrac{1}{2}(m_0 +m _1). \end{aligned}$$
(9.2)

To verify the claim we use the relation \(AC = CA\), which holds in generality. The static correlator \(C\) is modified to

$$\begin{aligned} C= \left( \begin{array}{l@{\quad }l@{\quad }l} \langle y;y\rangle &{} 0 &{} \langle y;V\rangle \\ 0 &{} \beta ^{-1}\bar{m} &{} 0 \\ \langle y;V\rangle &{} 0 &{} \tfrac{1}{2}\beta ^{-2}+\langle V;V\rangle \end{array}\right) , \end{aligned}$$
(9.3)

since

$$\begin{aligned} \tfrac{1}{2}\langle (p_0 + p_1)^2\rangle = \beta ^{-1}\bar{m}. \end{aligned}$$
(9.4)

Hence the linearization \(A\) is modified to

$$\begin{aligned} A= \left( \begin{array}{l@{\quad }l@{\quad }l} 0 &{} -\tfrac{1}{\bar{m}} &{} 0 \\ \partial _\ell p &{} 0 &{} \partial _\mathsf {e}p \\ 0 &{} \tfrac{1}{\bar{m}}p &{} 0 \end{array}\right) \end{aligned}$$
(9.5)

as claimed.

The hard-point gas has ideal gas thermodynamics, which means

$$\begin{aligned} p = \frac{\mathsf {2e}}{\ell },\quad \beta = \frac{1}{2\mathsf {e}}. \end{aligned}$$
(9.6)

Thus \(\partial _\ell p= -\beta p^2\), \(\partial _\mathsf {e}p= 2\beta p\) and the sound speed reads

$$\begin{aligned} c_{\bar{m}} = (3\beta /\bar{m})^{1/2}p. \end{aligned}$$
(9.7)

The transformation matrix is obtained to

$$\begin{aligned} R = \frac{1}{\sqrt{6}} \left( \begin{array}{l@{\quad }l@{\quad }l} -\beta p&{}- \sqrt{3\beta /\bar{m}}&{}2\beta \\ 2\beta p&{}0&{} 2\beta \\ -\beta p&{} \sqrt{3\beta /\bar{m}}&{}2\beta \\ \end{array}\right) , \end{aligned}$$
(9.8)
$$\begin{aligned} R^{-1}= \frac{1}{\sqrt{6}\beta p} \left( \begin{array}{l@{\quad }l@{\quad }l} -1&{}2&{}-1\\ -\sqrt{3\beta \bar{m}} \,p&{}0&{} \sqrt{3\beta \bar{m}} \,p\\ p&{}p&{}p\\ \end{array}\right) . \end{aligned}$$
(9.9)

The correlations of the physical fields are then given through \(S= R^{-1}S^\sharp R^{-1\mathrm {T}}\), where \(S^\sharp \) is assumed to be approximately diagonal,

$$\begin{aligned} S^\sharp _{\alpha \beta } = \delta _{\alpha \beta }f_\alpha . \end{aligned}$$
(9.10)

Using (9.9) one obtains

$$\begin{aligned}&\ell {-}\ell \;\;\mathrm {correlations}{:}\quad \frac{1}{6\beta ^2 p^2}(f_{-1} + 4 f_0 +f_1),\nonumber \\&\mathsf {u}{-}\mathsf {u} \;\;\mathrm {correlations}{:}\quad \frac{\bar{m}}{2\beta }(f_{-1} + f_1),\nonumber \\&\mathsf {e}{-}\mathsf {e} \;\;\mathrm {correlations}{:}\quad \frac{1}{6\beta ^2}(f_{-1} + f_0 +f_1). \end{aligned}$$
(9.11)

Next we compute the \(G\) matrices. Firstly, by direct differentiation of \(p\),

$$\begin{aligned} H^\ell =0,\quad H^\mathsf {u}= \frac{2}{\ell ^3} \left( \begin{array}{l@{\quad }l@{\quad }l} 2 \mathsf {e} &{} 0 &{} -\ell \\ 0 &{} -\bar{m}^{-1}\ell ^2 &{} 0 \\ -\ell &{} 0 &{} 0 \end{array}\right) ,\quad H^\mathsf {e}= \frac{2}{\bar{m} \ell ^2} \left( \begin{array}{l@{\quad }l@{\quad }l} 0 &{} - \mathsf {e}&{} 0 \\ - \mathsf {e} &{} 0 &{} \ell \\ 0 &{} \ell &{} 0 \end{array}\right) \end{aligned}$$
(9.12)

and transformed as

$$\begin{aligned} (R^{-1})^{\mathrm {T}}H^\mathsf {u} R^{-1}= p D_0, \quad (R^{-1})^{\mathrm {T}}H^\mathsf {e} R^{-1} = \frac{c_{\bar{m}}}{\beta } D_1 \end{aligned}$$
(9.13)

with the matrices

$$\begin{aligned} D_0= \left( \begin{array}{l@{\quad }l@{\quad }l} 0 &{} -1 &{} 2 \\ -1 &{} 0 &{} -1 \\ 2 &{} -1 &{} 0 \end{array}\right) ,\quad D_1= \left( \begin{array}{l@{\quad }l@{\quad }l} -1 &{} 0&{} 0 \\ 0 &{} 0 &{} 0 \\ 0 &{} 0 &{} 1 \end{array}\right) . \end{aligned}$$
(9.14)

We conclude that

$$\begin{aligned} G^{\pm 1} = \frac{c_{\bar{m}}}{2 \sqrt{6}}(\pm D_0 + 2D_1), \quad G^0 = \frac{c_{\bar{m}}}{ \sqrt{6}} D_1 \end{aligned}$$
(9.15)

and more explicitly

$$\begin{aligned} G^{\pm 1} = \pm \frac{c_{\bar{m}}}{ 2\sqrt{6}} \left( \begin{array}{lll} \mp 2 &{} -1 &{} 2 \\ -1 &{} 0 &{} -1 \\ 2 &{} -1 &{} \pm 2 \end{array}\right) , \quad G^0= \frac{c_{\bar{m}}}{ \sqrt{6}} \left( \begin{array}{lll} -1 &{} 0&{} 0 \\ 0 &{} 0 &{} 0 \\ 0 &{} 0 &{} 1 \end{array}\right) . \end{aligned}$$
(9.16)

As a special feature of the hard-point gas, all non-universal parameters are accounted for by the sound speed \(c_{\bar{m}}\). Most importantly, it should be noted that \(G^1_{00} = 0\), which implies that to leading order there is no coupling of the heat mode to the sound mode. This is an indication that for the sound mode peaks the finite time corrections are less severe than in asymmetric FPU chains.

Appendix 3: Discretized Multi-Component KPZ, Mode-Coupling

1.1 (a) Discretized KPZ

The Langevin equation (3.21) is somewhat formal. To have a well-defined evolution, we lattice discretize space by a lattice of \(N\) sites. The field \(\phi (x,t)\) then becomes \(\phi _j(t)\) with components \(\phi _{j,\alpha }(t)\), \(j=1,\ldots ,N\), \(\alpha =1,\ldots ,n\). The spatial finite difference operator is denoted by \(\tau \), \(\tau f_j=f_j-f_{j-1}\), with transpose \(\tau ^\mathrm {T} f_j=f_j-f_{j+1}\). Then the discretized KPZ equation reads

$$\begin{aligned} \partial _t \phi _{j,\alpha }+\tau \big (c_\alpha \phi _{j,\alpha } +\mathcal {N}_{j,\alpha } + \tau ^\mathrm {T} D\phi _{j,\alpha } + \sqrt{2D}\xi _{j,\alpha }\big )=0 \end{aligned}$$
(10.1)

with \(\phi _{0}=\phi _N\), \(\phi _{N+1}=\phi _1\), \(\xi _0=\xi _N\), where \(\xi _{j,\alpha }\) are independent Gaussian white noises with covariance

$$\begin{aligned} \langle \xi _{j,\alpha }(t) \xi _{j',\alpha '} (t')\rangle =\delta _{jj'} \delta _{\alpha \alpha '} \delta (t-t'). \end{aligned}$$
(10.2)

The diffusion matrix \(D\) acts on components, while the difference \(\tau \) acts on the lattice site variable \(j\).

\(\mathcal {N}_{j,\alpha }\) is quadratic in \(\phi \). But let us first consider the case \(\mathcal {N}_{j,\alpha } =0\). Then, since according to  (10.1) the drift is linear in \(\phi \), \(\phi _{j,\alpha }(t)\) a Gaussian process. The noise strength has been chosen such that one invariant measure is the Gaussian

$$\begin{aligned} \prod ^N_{j=1} \prod ^n_{\alpha =1} \big (\exp [-\tfrac{1}{2}\phi ^2_{j,\alpha }] (2\pi )^{-1/2} d \phi _{j,\alpha }\big )= \rho _\mathrm {G} (\phi ) \prod ^N_{j=1} \prod ^n_{\alpha =1} d \phi _{j,\alpha }. \end{aligned}$$
(10.3)

The set of all extremal invariant measures are obtained by conditioning (10.3) on the hyperplanes

$$\begin{aligned} \sum ^N_{j=1} \phi _{j,\alpha }=N\rho _\alpha , \end{aligned}$$
(10.4)

which for large \(N\) would become independent Gaussians with mean \(\rho _\alpha \). In the following we fix \(\rho _\alpha =0\) and denote the average with respect to \(\rho _\mathrm {G}\) by \(\langle \cdot \rangle _{\mathrm {eq}}\).

The generator corresponding to (10.1) with \(\mathcal {N}_{j,\alpha }=0\) is given by

$$\begin{aligned} L_0=\sum ^N_{j=1} \Big (-\sum ^n_{\alpha =1} \tau \big (c_\alpha \phi _{j,\alpha }+\tau ^\mathrm {T} D\phi _{j,\alpha }\big ) \partial _{\phi _{j,\alpha }}+\sum ^n_{\alpha ,\beta =1} 2D_{\alpha \beta } (\tau \partial _{\phi _{j,\alpha }} )(\tau \partial _{\phi _{j,\beta }})\Big ). \end{aligned}$$
(10.5)

The invariance of \(\rho _\mathrm {G}(\phi )\) can be checked through

$$\begin{aligned} L^*_0 \rho _\mathrm {G}(\phi )=0, \end{aligned}$$
(10.6)

where \(^*\) is the adjoint with respect to the flat volume measure. Furthermore linear functions evolve to linear functions according to

$$\begin{aligned} \mathrm {e}^{L_0 t}\phi _{j,\alpha }= \sum ^N_{j'=1} \sum ^n_{\alpha '=1} (\mathrm {e}^{tB})_{j\alpha ,j'\alpha '} \phi _{j',\alpha '}, \end{aligned}$$
(10.7)

where the matrix \(B=-\tau \otimes \mathrm {diag} (c_1,\ldots , c_n) -\tau \tau ^\mathrm {T} \otimes D\).

We now add the nonlinearity \(\mathcal {N}_{j,\alpha }\). In general, this will modify the invariant measure and we have little control how. Therefore we propose to choose \(\mathcal {N}_{j,\alpha }\) such that \(\rho _\mathrm {G}\) is left invariant under the deterministic flow generated the evolution equation \(\tfrac{d}{dt}\phi = -\tau \mathcal {N}\), i.e. under the generator

$$\begin{aligned} L_1= - \sum ^N_{j=1} \sum ^n_{\alpha =1} \tau \mathcal {N}_{j,\alpha } \partial _{\phi _{j,\alpha }}. \end{aligned}$$
(10.8)

The invariance of \(\rho _\mathrm {G}\) under \(L_1\) is equivalent to the condition

$$\begin{aligned} \sum ^N_{j=1} \sum ^n_{\alpha =1} \phi _{j,\alpha } \tau \mathcal {N}_{j,\alpha }=0. \end{aligned}$$
(10.9)

If \(\mathcal {N}_{j,\alpha }\) depends only on the field at sites \(j\) and \(j+1\), then the most general solution to (10.9) reads

$$\begin{aligned} \mathcal {N}_{j,\alpha }=\tfrac{1}{3}\sum ^n_{\beta ,\gamma =1} G^\alpha _{\beta \gamma }\big (\phi _{j,\beta }\phi _{j,\gamma }+\phi _{j,\beta }\phi _{j+1,\gamma }+\phi _{j+1,\beta }\phi _{j+1,\gamma }\big ) \end{aligned}$$
(10.10)

under the constraint that

$$\begin{aligned} G^\alpha _{\beta \gamma }=G^\alpha _{\gamma \beta }=G^\beta _{\alpha \gamma } \end{aligned}$$
(10.11)

for all \(\alpha ,\beta ,\gamma =1,\ldots ,n\). Denoting the generator of the Langevin equation (10.1) by

$$\begin{aligned} L=L_0+L_1, \end{aligned}$$
(10.12)

one concludes \(L^*\rho _\mathrm {G}=0\), i.e. the invariance of \(\rho _\mathrm {G}\).

In the continuum limit the condition (10.9) reads

$$\begin{aligned} \sum ^n_{\alpha ,\beta ,\gamma =1} G^\alpha _{\beta \gamma } \int dx\phi _\alpha (x) \partial _x \big (\phi _\beta (x) \phi _\gamma (x)\big )=0, \end{aligned}$$
(10.13)

where \(G^\alpha _{\beta \gamma }=G^\alpha _{\gamma \beta }\). By partial integration

$$\begin{aligned} 2 \sum ^n_{\alpha ,\beta ,\gamma =1} G^\alpha _{\beta \gamma } \int dx \phi _\alpha (x) \phi _\beta (x) \partial _x \phi _\gamma (x) =- \sum ^n_{\alpha ,\beta ,\gamma } G^\alpha _{\beta \gamma } \int dx \phi _\beta (x) \phi _\gamma (x) \partial _x \phi _\alpha (x)\nonumber \\ \end{aligned}$$
(10.14)

and (10.13) is satisfied only if \(G^\gamma _{\beta \alpha }=G^\alpha _{\beta \gamma }\), which is the condition (10.11) obtained already in the discrete setting.

We henceforth assume (10.11) although it will not hold for anharmonic chains, in general. The leading coefficients \(G^\alpha _{\alpha \alpha }\)’s are not constrained and one can still freely choose the sub-leading \(G^\alpha _{\beta \beta }\). The symmetry thus restricts the coefficients corresponding to sub-sub-leading terms. Appealing to universality we expect that the true invariant measure for general \(G\) will have short range correlations and nonlinear fluctuating hydrodynamics remains a valid approximation to the microscopic dynamics.

1.2 (b) Mode-Coupling

We consider the stationary \(\phi _{j,\alpha }(t)\) process, average denoted by \(\langle \cdot \rangle \), governed by (10.1) with \(\rho _\mathrm {G}\) as \(t=0\) measure. For \(t\ge 0\) the stationary covariance reads

$$\begin{aligned} S_{\alpha \beta }(j,t)=\langle \phi _{j,\alpha }(t)\phi _{0,\beta }(0)\rangle =\langle \phi _{0,\beta }\mathrm {e}^{Lt}\phi _{j,\alpha }\rangle _{\mathrm {eq}}, \end{aligned}$$
(10.15)

where for easier reading we leave out the superscript \(^{\sharp \phi }\). By construction

$$\begin{aligned} S_{\alpha \beta }(j,0)=\delta _{\alpha \beta } \delta _{j0}. \end{aligned}$$
(10.16)

The time derivative reads

$$\begin{aligned} \frac{d}{dt} S_{\alpha \beta }(j,t)=\langle \phi _{0,\beta }(\mathrm {e}^{Lt}L_0 \phi _{j,\alpha })\rangle _{\mathrm {eq}}+ \langle \phi _{0,\beta }(\mathrm {e}^{Lt}L_1 \phi _{j,\alpha })\rangle _{\mathrm {eq}}. \end{aligned}$$
(10.17)

We insert

$$\begin{aligned} \mathrm {e}^{Lt}=\mathrm {e}^{L_0 t}+\int ^t_0 ds\, \mathrm {e}^{L_0(t-s)} L_1 \mathrm {e}^{Ls} \end{aligned}$$
(10.18)

in the second summand of (10.17). The term containing \(\mathrm {e}^{L_0 t}\) does not show, since it is cubic in the time zero fields and the average \(\langle \cdot \rangle _\mathrm {eq}\) vanishes. Therefore one arrives at

$$\begin{aligned} \frac{d}{dt} S_{\alpha \beta }(j,t)= \sum _{j'\in \mathbb {Z}} \sum ^n_{\alpha '=1} \big (B_{\alpha j,\alpha ' j'} S_{\alpha '\beta }(j',t) + \int ^t_0 ds \langle \phi _{0,\beta } \mathrm {e}^{L_0(t-s)}L_1(\mathrm {e}^{Ls} L_1 \phi _{j,\alpha }) \rangle _\mathrm {eq}\big ). \end{aligned}$$
(10.19)

For the adjoint of \(\mathrm {e}^{L_0(t-s)}\) we use (10.7) and for the adjoint of \(L_1\) we use

$$\begin{aligned} \langle \phi _{j,\alpha } L_1 F(\phi )\rangle _\mathrm {eq} = - \langle (L_1 \phi _{j,\alpha }) F(\phi )\rangle _\mathrm {eq}. \end{aligned}$$
(10.20)

Furthermore

$$\begin{aligned} L_1 \phi _{j,\alpha } = - \tau \mathcal {N}_{j,\alpha }. \end{aligned}$$
(10.21)

Inserting in (10.19) one arrives at the identity

$$\begin{aligned} \frac{d}{dt} S_{\alpha \beta }(j,t)&= \sum _{j'\in \mathbb {Z}} \sum ^n_{\alpha '=1} \big (B_{\alpha j,\alpha ' j'} S_{\alpha '\beta }(j',t)\nonumber \\&\quad - \int ^t_0 ds (\mathrm {e}^{B^\mathrm {T}(t-s)})_{0\beta ,j'\alpha '} \langle \tau \mathcal {N}_{j',\alpha '}(\mathrm {e}^{Ls} \tau \mathcal {N}_{j,\alpha }) \rangle _\mathrm {eq}\big ). \end{aligned}$$
(10.22)

To obtain a closed equation for \(S\) we note that the average \(\langle \tau \mathcal {N}_{j,\alpha }(s)\tau \mathcal {N}_{j',\alpha '}(0)\rangle \) appearing in (10.22) is a four-point correlation, for which we invoke a Gaussian factorization as

$$\begin{aligned} \langle \phi (s)\phi (s)\phi (0)\phi (0)\rangle \cong \langle \phi (s)\phi (s)\rangle \langle \phi (0)\phi (0)\rangle +2\langle \phi (s)\phi (0)\rangle \langle \phi (s)\phi (0)\rangle . \end{aligned}$$
(10.23)

The first summand vanishes because of the difference operator \(\tau \). Secondly we replace the bare propagator \(\mathrm {e}^{B(t-s)}\) by the interacting propagator \(S(t-s)\). Finally we take a limit of zero lattice spacing. This step could be avoided, and is done so in numerical schemes. We could also maintain the ring geometry. As one example of interest, thereby one could investigate collisions between the moving peaks. Universality is only expected for large \(j,t\), hence in the limit of zero lattice spacing. The continuum limit of \(S(j,t)\) is denoted by \(S(x,t)\), \(x\in \mathbb {R}\). With these steps we arrive at the mode-coupling equation

$$\begin{aligned} \partial _t S_{\alpha \beta }(x,t)&= \sum ^n_{\alpha '=1} \big ({-}c_\alpha \delta _{\alpha \alpha '}\partial _x +D_{\alpha \alpha '}\partial ^2_x\big ) S_{\alpha '\beta }(x,t)\nonumber \\&+ \int ^t_0 ds \int _{\mathbb {R}} dy \partial ^2_y M_{\alpha \alpha '}(y,s) S_{\alpha '\beta }(x-y,t-s) \end{aligned}$$
(10.24)

with the memory kernel

$$\begin{aligned} M_{\alpha \alpha '}(x,t)= 2\sum ^n_{\beta ',\beta '',\gamma ',\gamma ''=1} G^\alpha _{\beta '\gamma '} G^{\alpha '}_{\beta ''\gamma ''} S_{\beta '\beta ''}(x,t) S_{\gamma '\gamma ''}(x,t). \end{aligned}$$
(10.25)

1.3 (c) Fourier Transform Conventions

Our equations of motion are written mostly in terms of the index \(j\) and its continuum approximation \(x\). For numerical simulations, and also for the asymptotic analysis, it is convenient to use Fourier space. We list here the conventions used throughout.

We naively discretize (10.24) in replacing \(x\in \mathbb {R}\) by \(j\in \mathbb {Z}\), \(\partial _x f(x)\) by \(\frac{1}{2}( f(j+1)-f(j-1))\), and \(\partial ^2_x f(x)\) by \(f(j+1)-2f(j)+f(j-1)\). Then (10.24) becomes

$$\begin{aligned} \partial _t S_{\alpha \beta } (j,t)&= -c_\alpha \tfrac{1}{2}\big (S_{\alpha \beta }(j+1,t)-S_{\alpha \beta }(j-1,t)\big )\nonumber \\&\quad +\sum ^n_{\alpha '=1}\Big (D_{\alpha \alpha '} \big (S_{\alpha '\beta }(j+1,t) -2S_{\alpha '\beta }(j,t)+S_{\alpha '\beta }(j-1,t)\big ) \nonumber \\&\quad +\int ^t_0 ds \sum _{j'\in \mathbb {Z}} \big (M_{\alpha \alpha '}(j' +1,s) {-}2 M_{\alpha \alpha '}(j',s)\nonumber \\&\quad +\, M_{\alpha \alpha '}(j'-1,s)\big ) S_{\alpha '\beta }(j-j',t-s)\Big ). \end{aligned}$$
(10.26)

For finite \(N\), \(j=1,\ldots ,N\), periodic boundary conditions are understood.

We adopt the standard discrete Fourier transform,

$$\begin{aligned} \hat{f}(k)= \sum _{j\in \mathbb {Z}} f(j) \mathrm {e}^{-\mathrm {i} 2\pi jk},\quad f(j)=\int ^{1/2}_{-1/2} dk \hat{f}(k) \mathrm {e}^{\mathrm {i}2\pi jk}. \end{aligned}$$
(10.27)

\(\hat{f}\) is one-periodic and the standard Brillouin zone is \(k\in [-\frac{1}{2},\frac{1}{2}]\). For finite \(N\),

$$\begin{aligned} \hat{f}(k)= \sum ^N_{j=1} f(j) \mathrm {e}^{-\mathrm {i}2\pi jk},\quad f(j)=\frac{1}{N}\sum _{k\in [0,1]_N} \hat{f}(k) \mathrm {e}^{\mathrm {i}2\pi jk} \end{aligned}$$
(10.28)

with \([0,1]_N = \{k|k=0,1,\ldots ,N-1\}\). In Fourier space (10.26) becomes

$$\begin{aligned} \partial _t \hat{S}_{\alpha \beta }(k,t)&= -\mathrm {i} c_\alpha \sin (2\pi k) \hat{S}_{\alpha \beta }(k,t) -2(1-\cos (2\pi k)) \nonumber \\&\sum ^n_{\alpha '=1} \Big (D_{\alpha \alpha '}\hat{S}_{\alpha '\beta }(k,t)+\int ^t_0 ds \hat{M}_{\alpha \alpha '}(k,s) \hat{S}_{\alpha '\beta }(k,t-s)\Big ), \end{aligned}$$
(10.29)

where

$$\begin{aligned} \hat{M}_{\alpha \alpha '} (k,s)= 2 \sum ^n_{\beta ,\beta ',\gamma ,\gamma '=1} G^\alpha _{\beta \gamma } G^{\alpha '}_{\beta '\gamma '} \int ^{1/2}_{-1/2} dq \hat{S}_{\beta \beta '} (k-q,s) \hat{S}_{\gamma \gamma '} (q,s). \end{aligned}$$
(10.30)

Since \(S\) refers here to the normal mode covariance, the initial conditions are

$$\begin{aligned} S_{\alpha \beta } (j,0)=\delta _{\alpha \beta } \delta _{j0},\quad \hat{S}_{\alpha \beta } (k,0)=\delta _{\alpha \beta }. \end{aligned}$$
(10.31)

Correspondingly we use for the continuum Fourier transform

$$\begin{aligned} \hat{f}(k)= \int dx f(x) \mathrm {e}^{-\mathrm {i}2\pi xk},\quad f(x)=\int dk \hat{f}(k) \mathrm {e}^{\mathrm {i}2\pi xk}. \end{aligned}$$
(10.32)

Convolutions come with no extra factor of \(\pi \), while \(\partial _x f(x)\) is transformed to \(\mathrm {i}2\pi k \hat{f}(k)\). Also \(\hat{f}(0)=\int dx f(x)\).

Appendix 4: Levy Asymptotics for the Heat Mode

The starting equation reads

$$\begin{aligned} \partial _t \hat{f}(k,t)=- (2\pi k\lambda _0)^2 \int ^t_0 ds \hat{f}(k,t-s) \int _\mathbb {R} dq \hat{g}(k-q,s)\hat{g}(q,s), \end{aligned}$$
(11.1)

\(\hat{f}(k,0)=1\). The function \(\hat{g}\) is assumed to have the scaling form

$$\begin{aligned} \hat{g}(k,t)= \mathrm {e}^{\mathrm {i}2\pi kct} \hat{g}_0\big ((\lambda _\mathrm {s} t)^\beta k\big ),\quad \hat{g}_0(0)=1, \end{aligned}$$
(11.2)

with \(0<\beta <1\), \(\lambda _\mathrm {s} >0\). We claim that with \(\gamma =1+\beta \)

$$\begin{aligned} \lim _{k\rightarrow 0} \hat{f}(k,|k|^{-\gamma }t)= \mathrm {e}^{-\lambda _{\beta } t}, \end{aligned}$$
(11.3)

where the non-universal coefficient, \(\lambda _\beta \), reads

$$\begin{aligned} \lambda _\beta = (2\pi \lambda _0)^2(\lambda _\mathrm {s})^{-\beta } \int ^\infty _0 dt t^{-\beta } \mathrm {e}^{\mathrm {i}2\pi \mathrm {sgn}(k)ct} \int _\mathbb {R} dk |\hat{g}_0(k)|^2. \end{aligned}$$
(11.4)

To prove (11.3), (11.4) we look for a scaling function \(\hat{h}\) such that, for \(t\) of order \(|k|^\gamma \),

$$\begin{aligned} \hat{f}(k,t )= \hat{h}(|k|^\gamma t). \end{aligned}$$
(11.5)

We first set \(k> 0\). Inserting in (11.1) one obtains

$$\begin{aligned} \hat{h}'(|k|^\gamma t) |k|^\gamma = - (2\pi k \lambda _0)^2 \int ^t_0 ds \hat{h}(|k|^\gamma (t -s)) (\lambda _\mathrm {s}s)^{-\beta } \mathrm {e}^{\mathrm {i}2\pi kcs} \int _\mathbb {R} dq |\hat{g}_0(q)|^2. \end{aligned}$$
(11.6)

We substitute \(w = |k|^\gamma t\) and \(s\) by \(s/|k|\). Then

$$\begin{aligned}&\hat{h} '(w) |k|^\gamma \nonumber \\&\!=\! - (2\pi \lambda _0)^2|k|^{(1+\beta )} \int ^{|k|^{(-\gamma +1)}w}_0 ds \hat{h}(w\! -\! |k|^{\gamma - 1}s) s^{-\beta } \mathrm {e}^{\mathrm {i}2\pi cs} (\lambda _\mathrm {s})^{-\beta }\int _\mathbb {R} dq |\hat{g}_0(q)|^2.\qquad \quad \end{aligned}$$
(11.7)

Taking the limit \(k \rightarrow 0\), \(\hat{h}\) can be taken out of the \(s\)-integral with the result \(\hat{h}' = -\lambda \hat{h}\), which yields (11.3). In repeating the same computation for \(k< 0\), only the sign in the oscillating factor changes.

Working out the integration in (11.4), the scaling function at \(t=1\) becomes

$$\begin{aligned} \hat{h}(|k|^\gamma )= \exp \big [-\lambda _\beta |k|^\gamma \big (1 - \mathrm {i}\,\mathrm {sgn}(k)\tan (\tfrac{1}{2}\pi \gamma ) \big )\big ], \end{aligned}$$
(11.8)

where

$$\begin{aligned} \lambda _\beta = (2 \pi \lambda _0)^2(\lambda _\mathrm {s})^{-\beta } (2\pi c)^{-1+\beta } \tfrac{1}{2}\pi \frac{1}{\Gamma (\beta )} \frac{1}{\cos (\tfrac{1}{2}\pi \beta )} \int _\mathbb {R} dk |\hat{g}_0(k)|^2. \end{aligned}$$
(11.9)

One recognizes (11.9) as the Fourier transform of asymmetric \(\alpha \)-stable law with \(\alpha = \gamma \). The asymmetry parameter is at its limiting allowed value. This implies a decay in position space as \(|x|^{-\gamma -1}\) to the left and as \(\exp [-|x|^{\gamma /(\gamma -1)}]\) to the right, see [74] for the asymptotics. In the context of fluctuating hydrodynamics this feature implies that outside the sound cone the correlations are suppressed faster than exponential.

Appendix 5: Anharmonic Chains and One-Dimensional Fluids

One-dimensional fluids are governed by the hamiltonian

$$\begin{aligned} H_\mathrm {f}=\sum ^N_{j=1} \tfrac{1}{2}p^2_j +\tfrac{1}{2}\sum ^N_{i\ne j=1} V(q_i-q_j) \end{aligned}$$
(12.1)

with an even interaction potential, \(V(x)=V(-x)\). \(V\) must be chosen such that the system is thermodynamically stable, to say \(V\) may have a hard core and \(V\) decays sufficiently fast at infinity, minimal requirements being \(V(x)\rightarrow 0\) as \(|x|\rightarrow \infty \), faster than \(- |x|^{-2}\) to avoid phase transitions, \(V\) is bounded from below, and not “too negative”. Particles move on the interval \([0,L]\) with periodic boundary conditions. Particle number, momentum, and energy are locally conserved. Their corresponding microscopic fields are written as

$$\begin{aligned} \sum ^N_{j=1}\delta (q_j-x),\quad \sum ^N_{j=1}\delta (q_j-x)p_j,\quad \sum ^N_{j=1}\delta (q_j-x) \Big (\tfrac{1}{2}p^2_j +\tfrac{1}{2}\sum ^N_{i=1,i\ne j} V(q_i-q_j)\Big ). \end{aligned}$$
(12.2)

From the evolution equations one finds the corresponding microscopic currents as

$$\begin{aligned} J_1(x)&= \sum ^N_{j=1}\delta (q_j-x)p_j,\nonumber \\ J_2(x)&= \sum ^N_{j=1}\delta (q_j-x)p^2_j -\tfrac{1}{2}\sum ^N_{i=1,i\ne j} V'(q_i-q_j)(q_i-q_j) \int ^1_0 dx \delta (\lambda q_i +(1-\lambda ) q_j-x),\nonumber \\ J_3(x)&= \sum ^N_{j=1}\delta (q_j-x)p_j\Big (\tfrac{1}{2}p^2_j +\tfrac{1}{2}\sum ^N_{i=1,i\ne j} V(q_i-q_j)\Big ),\nonumber \\&+\,\tfrac{1}{2}\sum ^N_{i,j=1}\tfrac{1}{2}(p_i+p_j) V'(q_i-q_j)(q_i-q_j) \int ^1_0 d\lambda \delta (\lambda q_i +(1-\lambda ) q_j-x). \end{aligned}$$
(12.3)

Even without averaging, the densities from (12.2) and the currents from (12.3) satisfy a system of conservation laws.

If the fluid starts and approximately remains in local equilibrium, one can equilibrium average the microscopic conservation laws and arrives at the Euler equations of a one-dimensional fluid. The hydrodynamic fields are \(\rho \), \(\rho v\), \(\rho \mathfrak {e}_\mathrm {f}\), which depend on \(x,t\). \(\rho \) is the local particle density, \(v\) the local velocity per particle and \(\mathfrak {e}_\mathrm {f}\) the local energy per particle. Then the Euler equations read

$$\begin{aligned}&\partial _t \rho +\partial _x (\rho v) =0,\nonumber \\&\partial _t (\rho v)+\partial _x (\rho v^2 +p_\mathrm {f}) =0,\nonumber \\&\partial _t (\rho \mathfrak {e}_\mathrm {f})+ \partial _x (\rho \mathfrak {e}_\mathrm {f} v +p_\mathrm {f}v)=0. \end{aligned}$$
(12.4)

Here \(p_\mathrm {f}\) is the local thermodynamic pressure, which depends on \(\rho \) and the internal energy. We use the subscript “f” to distinguish from the corresponding quantity for the anharmonic chain. To construct nonlinear fluctuating hydrodynamics we may proceed as in the main text. In particular, \(G^0_{00}=0\) always and generically \(G^1_{11}= -G^{-1}_{-1-1}\ne 0\). Thus up to model-dependent coefficients the overall structure remains unaltered. There is one technical difference. To determine the pressure \(p_\mathrm {f}\) requires a full many-body computation, in general. One has then to resort to series expansions and Monte Carlo techniques. In particular, to obtain with sufficient accuracy the coupling constants \(G\) will be more costly than for the chain.

If one introduces a hard core, the ordering of particles is preserved and by a suitable choice of parameters one can achieve that only the nearest neighbor contribution to the potential term remains. An explicit example is

$$\begin{aligned} V_{\mathrm {hc}}(x)= \left\{ {\begin{array}{ll} \infty , &{} |x|\le a/2, \\ \text {``arbitrary''}, &{} a/2\le |x|\le a, \\ 0, &{} a\le |x|. \\ \end{array} } \right. \end{aligned}$$
(12.5)

Then, on the restricted configuration space,

$$\begin{aligned} \tfrac{1}{2}\sum ^N_{i=1,i\ne j} V_{\mathrm {hc}}(q_i-q_j) = \sum ^N_{j=1} V_{\mathrm {hc}}(q_{j+1}-q_j) \end{aligned}$$
(12.6)

Thus for \(V_{\mathrm {hc}}\) one may adopt the field theory point of view with coupling only between nearest neighbors.

At second thoughts one is puzzled, since the hydrodynamic equations for a fluid look different from the ones of the field theory. One reason is that \(\rho =\ell ^{-1}\). More importantly \(\partial _x\) above refers to per unit length while the \(\partial _x\) in (2.21), (2.22) refers to the particle label. Still, because it is the same physical system, there must be a transformation which converts (12.4) to (2.21), (2.22). I did not find a discussion in the literature and hence explain how the transformation is done.

First one has to properly distinguish between “\(x\)” in (2.21) and in (12.5). To do so, only in this section, we replace \(x\) in (2.21) by \(y\). Let us define

$$\begin{aligned} \int ^x_{-\infty } dx' \rho (x',t)= y_t(x),\quad \int ^y_{-\infty } dy' \ell (y',t)= x_t(y). \end{aligned}$$
(12.7)

\(y_t\) is the inverse function of \(x_t\). It follows that

$$\begin{aligned} \rho (x,t) \ell (y,t)=1, \end{aligned}$$
(12.8)

where we used

$$\begin{aligned} \partial _x y_t(x)=\rho (x,t),\quad \partial _y x_t(y)=\ell (y,t). \end{aligned}$$
(12.9)

If in a single equation, as (12.8), both \(x\) and \(y\) appear, they are understood to be related as in (12.9). Using the conservation of mass one obtains

$$\begin{aligned} \partial _t y_t(x)=-\rho (x,t) v(x,t). \end{aligned}$$
(12.10)

Other quantities flow along, i.e.

$$\begin{aligned} v(x,t)= u(y,t),\quad \mathfrak {e}_\mathrm {f}(x,t)=\mathfrak {e}(y,t),\quad p_\mathrm {f}(x,t)=p(y,t). \end{aligned}$$
(12.11)

We first discuss mass conservation and write

$$\begin{aligned} 0&= \partial _t \rho + \partial _x(\rho v)=\partial _t\big (\ell (y_t(x),t)^{-1}\big )+\rho \partial _x u(y_t(x),t)+v \partial _x\big (\ell (y_t(x),t)^{-1}\big )\nonumber \\&= -\ell ^{-2} (\partial _y\ell )(\partial _t y_t(x))-\ell ^{-2} \partial _t \ell +\rho \big (\partial _y u(y,t)\big )\big (\partial _x y_t(x)\big ) -v\ell ^{-2}\big (\partial _y \ell (y,t)\big )\big (\partial _x y_t(x)\big )\nonumber \\&= \ell ^{-3} v \partial _y\ell -\ell ^{-2} \partial _t\ell +\rho ^2 \partial _y u -v\ell ^{-3}\partial _y \ell \nonumber \\&= -\ell ^{-2} (\partial _t \ell - \partial _y u), \end{aligned}$$
(12.12)

as to be shown. We turn to momentum conservation

$$\begin{aligned} 0&= \partial _t (\rho v) + \partial _x(\rho v^2) + \partial _x p_\mathrm {f}= \rho \partial _t v + \rho v \partial _x v +\partial _x p_\mathrm {f} \nonumber \\&= \rho \partial _t u (y_t(x),t) + \rho v\partial _x u(y_t(x),t) +\partial _x p_\mathrm {f}\nonumber \\&= -\rho (\partial _y u)\rho v +\rho \partial _t u+\rho v (\partial _y u) \rho +\partial _x p_\mathrm {f}\nonumber \\&= \rho \partial _t u + \partial _x p_\mathrm {f}. \end{aligned}$$
(12.13)

Now for any pair of functions \(g\), \(g_\mathrm {f}\) such that \(g_\mathrm {f}(x,t)=g(y,t)\), one has

$$\begin{aligned} \partial _x g_\mathrm {f}(x,t)= \partial _x g\big (y_t(x),t\big )=\big (\partial _y g(y,t)\big )\big (\partial _x y_t(x)\big )=\big (\partial _y g(y,t)\big ) \rho (x,t). \end{aligned}$$
(12.14)

From (12.13) one concludes that

$$\begin{aligned} \partial _t u +\partial _y p=0. \end{aligned}$$
(12.15)

Finally we turn to energy conservation, with similar steps as in (12.15)

$$\begin{aligned} 0&= \partial _t (\rho \mathfrak {e}_\mathrm {f}) + \partial _x(\rho \mathfrak {e}_\mathrm {f} v) + \partial _x (vp_\mathrm {f}) \nonumber \\&= \rho \partial _t \mathfrak {e}_\mathrm {f} + \rho v\partial _x \mathfrak {e}_\mathrm {f} +\partial _x (v p_\mathrm {f})\nonumber \\&= \rho \partial _t \mathfrak {e} +\partial _x (v p_\mathrm {f})= \rho \big (\partial _t \mathfrak {e}+\partial _y (p u)\big ), \end{aligned}$$
(12.16)

where we used (12.14) with \(g_\mathrm {f}=v p_\mathrm {f}\), \(g= up\). The pressure \(p_\mathrm {f}\) depends on \(x,t\) through \(\rho \) and \(\mathsf {e}_\mathrm {f}\) and the pressure \(p\) depends on \(y,t\) through \(\ell \) and \(\mathsf {e}\).

Rights and permissions

Reprints and permissions

About this article

Cite this article

Spohn, H. Nonlinear Fluctuating Hydrodynamics for Anharmonic Chains. J Stat Phys 154, 1191–1227 (2014). https://doi.org/10.1007/s10955-014-0933-y

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s10955-014-0933-y

Keywords

Navigation