Skip to main content
Log in

Shocks, Rarefaction Waves, and Current Fluctuations for Anharmonic Chains

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

Abstract

The nonequilibrium dynamics of anharmonic chains is studied by imposing an initial domain-wall state, in which the two half lattices are prepared in equilibrium with distinct parameters. We analyse the Riemann problem for the corresponding Euler equations and, in specific cases, compare with molecular dynamics. Additionally, the fluctuations of time-integrated currents are investigated. In analogy with the KPZ equation, their typical fluctuations should be of size \(t^{1/3}\) and have a Tracy–Widom GUE distributed amplitude. The proper extension to anharmonic chains is explained and tested through molecular dynamics. Our results are calibrated against the stochastic LeRoux lattice gas.

This is a preview of subscription content, log in via an institution to check access.

Access this article

Price excludes VAT (USA)
Tax calculation will be finalised during checkout.

Instant access to the full article PDF.

Fig. 1
Fig. 2
Fig. 3
Fig. 4
Fig. 5
Fig. 6
Fig. 7
Fig. 8
Fig. 9
Fig. 10
Fig. 11
Fig. 12
Fig. 13
Fig. 14
Fig. 15
Fig. 16
Fig. 17

We’re sorry, something doesn't seem to be working properly.

Please try refreshing the page. If that doesn't work, please contact support so we can address the problem.

References

  1. Rigol, M., Dunjko, V., Olshanii, M.: Thermalization and its mechanism for generic isolated quantum systems. Nature 452, 854–858 (2008). doi:10.1038/nature06838

    Article  ADS  Google Scholar 

  2. Essler, F.H.L., Fagotti, M.: Quench dynamics and relaxation in isolated integrable quantum spin chains (2016). arXiv:1603.06452

  3. Bañuls, M.C., Cirac, J.I., Hastings, M.B.: Strong and weak thermalization of infinite nonintegrable quantum systems. Phys. Rev. Lett. 106, 050405 (2011). doi:10.1103/PhysRevLett.106.050405

    Article  ADS  Google Scholar 

  4. Sabetta, T., Misguich, G.: Nonequilibrium steady states in the quantum XXZ spin chain. Phys. Rev. B 88, 245114 (2013). doi:10.1103/PhysRevB.88.245114

    Article  ADS  Google Scholar 

  5. Langmann, E., Lebowitz, J.L., Mastropietro, V., Moosavi, P.: Steady states and universal conductance in a quenched Luttinger model. Commun. Math. Phys. (2016). doi:10.1007/s00220-016-2631-x

    MATH  Google Scholar 

  6. Karrasch, C., Ilan, R., Moore, J.E.: Nonequilibrium thermal transport and its relation to linear response. Phys. Rev. B 88, 195129 (2013). doi:10.1103/PhysRevB.88.195129

    Article  ADS  Google Scholar 

  7. Vasseur, R., Karrasch, C., Moore, J.E.: Expansion potentials for exact far-from-equilibrium spreading of particles and energy. Phys. Rev. Lett. 115, 267201 (2015). doi:10.1103/PhysRevLett.115.267201

    Article  ADS  Google Scholar 

  8. Castro-Alvaredo, O.A., Doyon, B., Yoshimura, T.: Emergent hydrodynamics in integrable quantum systems out of equilibrium (2016). arXiv:1605.07331

  9. Spohn, H.: Nonlinear fluctuating hydrodynamics for anharmonic chains. J. Stat. Phys. 154, 1191–1227 (2014). doi:10.1007/s10955-014-0933-y

    Article  ADS  MathSciNet  MATH  Google Scholar 

  10. Mendl, C.B., Spohn, H.: Dynamic correlators of Fermi-Pasta-Ulam chains and nonlinear fluctuating hydrodynamics. Phys. Rev. Lett. 111, 230601 (2013). doi:10.1103/PhysRevLett.111.230601

    Article  ADS  Google Scholar 

  11. Lepri, S. (ed.): Thermal Transport in Low Dimensions From Statistical Physics to Nanoscale Heat Transfer. Lecture Notes in Physics 921. Springer, Berlin (2016)

  12. Pettini, M., Casetti, L., Cerruti-Sola, M., Franzosi, R., Cohen, E.G.D.: Weak and strong chaos in Fermi-Pasta-Ulam models and beyond. Chaos 15, 015106 (2005). doi:10.1063/1.1849131

    Article  ADS  MathSciNet  MATH  Google Scholar 

  13. Bressan, A.: Hyperbolic conservation laws: an illustrated tutorial, pp. 157–245. In: Modelling and Optimisation of Flows on Networks, Cetraro, Italy. Lecture Notes in Mathematics 2062 2013. Springer, Berlin (2009). doi:10.1007/978-3-642-32160-3_2

  14. Fritz, J., Tóth, B.: Derivation of the Leroux system as the hydrodynamic limit of a two-component lattice gas. Commun. Math. Phys. 249, 1–27 (2004). doi:10.1007/s00220-004-1103-x

    Article  ADS  MathSciNet  MATH  Google Scholar 

  15. Serre, D.: Systems of Conservation Laws 1: Hyperbolicity, Entropies. Shock Waves. Cambridge University Press, Cambridge (1999)

    Book  MATH  Google Scholar 

  16. Serre, D.: Systems of Conservation Laws 2: Geometric Structures, Oscillations, and Initial-Boundary Value Problems. Cambridge University Press, Cambridge (2000)

    MATH  Google Scholar 

  17. Arndt, P.F., Heinzel, T., Rittenberg, V.: Spontaneous breaking of translational invariance in one-dimensional stationary states on a ring. J. Phys. A 31, L45–L51 (1998). doi:10.1088/0305-4470/31/2/001

    Article  ADS  MATH  Google Scholar 

  18. Ferrari, P., Sasamoto, T., Spohn, H.: Coupled Kardar–Parisi–Zhang equations in one dimension. J. Stat. Phys. 153, 377–399 (2013). doi:10.1007/s10955-013-0842-5

    Article  ADS  MathSciNet  MATH  Google Scholar 

  19. Temple, B.: Global solution of the Cauchy problem for a class of \(2 \times 2\) nonstrictly hyperbolic conservation laws. Adv. Appl. Math. 3, 335–375 (1982). doi:10.1016/S0196-8858(82)80010-9

    Article  MathSciNet  MATH  Google Scholar 

  20. Mendl, C.B., Spohn, H.: Searching for the Tracy-Widom distribution in nonequilibrium processes. Phys. Rev. E 93, 060101(R) (2016). doi:10.1103/PhysRevE.93.060101

    Article  ADS  Google Scholar 

  21. Bethe, H.A.: On the theory of shock waves for an arbitrary equation of state, pp. 421–495. In: Classic Papers in Shock Compression Science. Springer, Berlin (1998). doi:10.1007/978-1-4612-2218-7_11

  22. Lax, P.D.: Hyperbolic systems of conservation laws II. Commun. Pure Appl. Math. 10, 537–566 (1957). doi:10.1002/cpa.3160100406

    Article  MathSciNet  MATH  Google Scholar 

  23. Glimm, J.: Solutions in the large for nonlinear hyperbolic systems of equations. Commun. Pure Appl. Math. 18, 697–715 (1965). doi:10.1002/cpa.3160180408

    Article  MathSciNet  MATH  Google Scholar 

  24. Menikoff, R., Plohr, B.J.: The Riemann problem for fluid flow of real materials. Rev. Mod. Phys. 61, 75–130 (1989). doi:10.1103/RevModPhys.61.75

    Article  ADS  MathSciNet  MATH  Google Scholar 

  25. Fermi, E., Pasta, J., Ulam, S.: Studies of non-linear problems (LA-1940). Technical report, Los Alamos Scientific Laboratory (1955)

    MATH  Google Scholar 

  26. Hurtado, P.I.: Breakdown of hydrodynamics in a simple one-dimensional fluid. Phys. Rev. Lett. 96, 010601 (2006). doi:10.1103/PhysRevLett.96.010601

    Article  ADS  Google Scholar 

  27. Balázs, M., Nagy, A.L., Tóth, B., Tóth, I.: Coexistence of shocks and rarefaction fans: complex phase diagram of a simple hyperbolic particle system (2016). arXiv:1601.02161

  28. Mendl, C.B., Spohn, H.: Equilibrium time-correlation functions for one-dimensional hard-point systems. Phys. Rev. E 90, 012147 (2014). doi:10.1103/PhysRevE.90.012147

    Article  ADS  Google Scholar 

  29. Wendroff, B.: The Riemann problem for materials with nonconvex equations of state I: isentropic flow. J. Math. Anal. Appl. 38, 454–466 (1972). doi:10.1016/0022-247X(72)90103-5

    Article  MathSciNet  MATH  Google Scholar 

  30. Johansson, K.: Shape fluctuations and random matrices. Commun. Math. Phys. 209, 437–476 (2000). doi:10.1007/s002200050027

    Article  ADS  MathSciNet  MATH  Google Scholar 

  31. Tracy, C.A., Widom, H.: Level-spacing distributions and the Airy kernel. Phys. Lett. B 305, 115–118 (1993). doi:10.1016/0370-2693(93)91114-3

    Article  ADS  MathSciNet  MATH  Google Scholar 

  32. Tracy, C.A., Widom, H.: Level-spacing distributions and the Airy kernel. Commun. Math. Phys. 159, 151–174 (1994). doi:10.1007/BF02100489

    Article  ADS  MathSciNet  MATH  Google Scholar 

  33. Prähofer, M., Spohn, H.: Current fluctuations for the totally asymmetric simple exclusion process, pp. 185–204. In: In and Out of Equilibrium: Probability with a Physics Flavor, Progress in Probability, vol. 51. Birkhäuser Boston (2002). doi:10.1007/978-1-4612-0063-5_7

  34. Ben Arous, G., Corwin, I.: Current fluctuations for TASEP: a proof of the Prähofer–Spohn conjecture. Ann. Probab. 39, 104–138 (2011). doi:10.1214/10-aop550

    Article  MathSciNet  MATH  Google Scholar 

  35. Tracy, C.A., Widom, H.: Asymptotics in ASEP with step initial condition. Commun. Math. Phys. 290, 129–154 (2009). doi:10.1007/s00220-009-0761-0

    Article  ADS  MathSciNet  MATH  Google Scholar 

  36. Mendl, C.B., Spohn, H.: Current fluctuations for anharmonic chains in thermal equilibrium. J. Stat. Mech. 2015, P03007 (2015). doi:10.1088/1742-5468/2015/03/P03007

    Article  MathSciNet  Google Scholar 

Download references

Acknowledgments

The work of HS has been supported as a Simons Distinguished Visiting Scholar, when visiting the KITP early 2016. CM acknowledges support from the Alexander von Humboldt Foundation and computing resources of the Leibniz-Rechenzentrum.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Christian B. Mendl.

Appendix: Coupling Matrices

Appendix: Coupling Matrices

We compute the G coupling matrices for the LeRoux model and anharmonic chains, following the derivation in [9].

1.1 LeRoux Model

The linearization matrix A and its left and right eigenvectors are stated in Eqs. (2.7) and (2.9). The transformation to normal modes is accomplished through the matrix R defined by

$$\begin{aligned} R = \begin{pmatrix} \langle \tilde{\psi }_{-1} |\\ \langle \tilde{\psi }_{1} |\end{pmatrix}, \quad R^{-1} = \Big ( |\psi _{-1} \rangle \ |\psi _{1} \rangle \Big ). \end{aligned}$$
(8.1)

By construction one has

$$\begin{aligned} R A R^{-1} = \mathrm {diag}(c_{-1}, c_1). \end{aligned}$$
(8.2)

As usual, the static susceptibility matrix, C, is given by

$$\begin{aligned} C = \begin{pmatrix} \langle 1 - |\eta _j|; 1 - |\eta _j |\rangle &{} \langle 1 - |\eta _j |; \eta _j\rangle \\ \langle 1 - |\eta _j |; \eta _j \rangle &{} \langle \eta _j ; \eta _j\rangle \end{pmatrix} = \begin{pmatrix} \rho (1 - \rho ) &{} -\rho v \\ -\rho v &{} 1 - \rho - v^2 \\ \end{pmatrix}. \end{aligned}$$
(8.3)

In addition we then require

$$\begin{aligned} R C R^{\mathrm {T}} = \mathbbm {1}, \end{aligned}$$
(8.4)

thereby fixing the normalizations \(\tilde{Z}_{\sigma }\), \(Z_{\sigma }\) in (2.9) to

$$\begin{aligned} \tilde{Z}_{\sigma } = \sqrt{2} \left( 4 \rho (1 - \rho ) + v^2 \big (1 - 5 \rho -v^2\big ) + \sigma \,v \big (3 \rho + v^2 - 1\big )\sqrt{4 \rho + v^2}\right) ^{1/2}, \end{aligned}$$
(8.5)
$$\begin{aligned} Z_{\sigma } = 2\,\tilde{Z}_{\sigma }^{-1} \left( 4 \rho - v \big (\sigma \sqrt{4 \rho + v^2} - v\big )\right) . \end{aligned}$$
(8.6)

To obtain the nonlinear couplings G, in particular \(G^1_{11}\), we first compute the Hessians of the current as second derivatives of \(\vec {\mathsf {j}}(\vec {u})\),

$$\begin{aligned} H^{\rho } = - \begin{pmatrix} 0 &{}\quad 1 \\ 1 &{}\quad 0 \end{pmatrix}, \qquad H^{v} = - \begin{pmatrix} 0 &{}\quad 0 \\ 0 &{}\quad 2 \end{pmatrix}. \end{aligned}$$
(8.7)

In normal coordinates

$$\begin{aligned}&\langle \psi _{\sigma } \vert H^{\rho } \vert \psi _{\tau } \rangle = \rho \,v \begin{pmatrix} 1 &{}\quad 0 \\ 0 &{}\quad 1 \end{pmatrix} + \frac{2 \rho }{\sqrt{4 \rho + v^2}} \big (1 - \rho - \tfrac{1}{2} v^2\big ) \begin{pmatrix} -1 &{}\quad 0 \\ 0 &{}\quad 1 \end{pmatrix} \nonumber \\&\qquad \qquad + \frac{v}{\sqrt{4 \rho + v^2}} \sqrt{\rho \big ((1 - \rho )^2 - v^2\big )} \begin{pmatrix} 0 &{}\quad 1 \\ 1 &{}\quad 0 \end{pmatrix} \end{aligned}$$
(8.8)

and

$$\begin{aligned}&\langle \psi _{\sigma } \vert H^{v} \vert \psi _{\tau } \rangle = -\big (1 - \rho - v^2\big ) \begin{pmatrix} 1 &{}\quad 0 \\ 0 &{}\quad 1 \end{pmatrix} + \frac{v}{\sqrt{4 \rho + v^2}} \big (1 - 3 \rho - v^2\big ) \begin{pmatrix} -1 &{}\quad 0 \\ 0 &{}\quad 1 \end{pmatrix} \nonumber \\&\qquad \qquad - \frac{2}{\sqrt{4 \rho + v^2}} \sqrt{\rho \big ((1 - \rho )^2 - v^2\big )} \begin{pmatrix} 0 &{}\quad 1 \\ 1 &{}\quad 0 \end{pmatrix}. \end{aligned}$$
(8.9)

The coupling matrices are thus obtained as

$$\begin{aligned} \begin{array}{ll} G^{\sigma } &{}= \tfrac{1}{2} \sum \limits _{i=\{\rho ,v\}} R_{\sigma i} \, R^{-\mathrm {T}} H^{i} R^{-1} \\ &{}= Z_{\sigma }^{-1} \big (\sqrt{4 \rho +v^2} - \sigma v\big ) \begin{pmatrix} \tfrac{1}{2}(1 - \sigma ) &{}\quad 0 \\ 0 &{}\quad \tfrac{1}{2}(1 + \sigma ) \end{pmatrix}\\ &{}\quad + \tilde{Z}_{\sigma }^{-1} \sqrt{\rho \big ((1 - \rho )^2 - v^2\big )} \begin{pmatrix} 0 &{}\quad 1 \\ 1 &{}\quad 0\end{pmatrix}. \end{array} \end{aligned}$$
(8.10)

In particular, comparison with (2.10) shows that

$$\begin{aligned} G^{\sigma }_{\sigma \sigma } = \tfrac{1}{2} \psi _\sigma \cdot D c_{\sigma }. \end{aligned}$$
(8.11)

1.2 General Anharmonic Chain

To set the scale for the Tracy–Widom distribution, one has to compute \(G_{11}^1\). For a general anharmonic chain, in the special case \(v = 0\), the coupling matrices are derived in [9], a result which should be extended to \(v \ne 0\). In fact, it turns out that the coupling matrices do not depend on v.

Following the notation of [9], the static susceptibility matrix is given by

$$\begin{aligned} C = \begin{pmatrix} \langle r_j; r_j \rangle &{} \langle r_j; v_j \rangle &{} \langle r_j; e_j \rangle \\ \langle r_j; v_j \rangle &{} \langle v_j; v_j \rangle &{} \langle v_j; e_j \rangle \\ \langle r_j; e_j \rangle &{} \langle v_j; e_j \rangle &{} \langle e_j; e_j \rangle \end{pmatrix} = \begin{pmatrix} \langle y; y \rangle &{} 0 &{} \langle y; V \rangle \\ 0 &{} 1/(m \beta ) &{} v/\beta \\ \langle y; V \rangle &{} v/\beta &{} \frac{1}{2}\beta ^{-2} + m v^2 \beta ^{-1} + \langle V; V\rangle \end{pmatrix}.\nonumber \\ \end{aligned}$$
(8.12)

The linearization matrix A in (3.14) and its right and left eigenvectors in (3.16) and (3.17), respectively, define the transformation to normal modes via

$$\begin{aligned} R = \begin{pmatrix} \langle \tilde{\psi }_{-1} |\\ \langle \tilde{\psi }_{0} |\\ \langle \tilde{\psi }_{1} |\end{pmatrix}, \quad R^{-1} = \Big ( |\psi _{-1} \rangle \ |\psi _{0} \rangle \ |\psi _{1} \rangle \Big ) \end{aligned}$$
(8.13)

such that

$$\begin{aligned} R A R^{-1} = \mathrm {diag}(-c, 0, c), \quad R C R^{\mathrm {T}} = \mathbbm {1}. \end{aligned}$$
(8.14)

To have \(R R^{-1} = \mathbbm {1}\), the normalization constants of the eigenvectors must satisfy

$$\begin{aligned} Z_{0} \tilde{Z}_{0} = m c^2, \qquad Z_{\sigma } \tilde{Z}_{\sigma } = 2 m c^2 \ \ \text {for}\ \ \sigma = \pm 1. \end{aligned}$$
(8.15)

An explicit computation of the diagonal entries of \(R C R^{\mathrm {T}}\) shows that the velocity terms cancel. Hence the relations

$$\begin{aligned} \tilde{Z}_0 = \sqrt{m \Upsilon }\,c, \qquad \tilde{Z}_{\sigma } = \sqrt{2 m/\beta }\,c \ \ \text {for}\ \ \sigma = \pm 1 \end{aligned}$$
(8.16)

from [9] remain valid in general, where \(\Upsilon = \beta \left( \langle y; y \rangle \langle V; V \rangle - \langle y; V \rangle ^2\right) + \frac{1}{2} \beta ^{-1} \langle y; y \rangle \).

As in [9], we denote the Hessian matrices of the average current by

$$\begin{aligned} H^{i}_{\alpha \beta } = \partial _{u_\alpha } \partial _{u_\beta }\,\mathsf {j}_{i} \end{aligned}$$
(8.17)

with the conserved fields \(\vec {u} = (r, v, \mathfrak {e})\) and the current vector defined in (3.10). The coupling matrices are then given by

$$\begin{aligned} G^{\sigma } = \tfrac{1}{2} \sum _{i=1}^3 R_{\sigma i} \, R^{-\mathrm {T}} H^{i} R^{-1} \end{aligned}$$
(8.18)

for \(\sigma = -1, 0, 1\). While the Hessian matrices \(H^{i}\) depend on v, the coupling matrices are actually independent of v. Thus using the formulas in [9] one arrives at

$$\begin{aligned} G^0 = \frac{1}{2 \beta \sqrt{m \Upsilon }} \begin{pmatrix} -1 &{}\quad 0 &{}\quad 0 \\ 0 &{}\quad 0 &{}\quad 0 \\ 0 &{}\quad 0 &{}\quad 1 \end{pmatrix} \end{aligned}$$
(8.19)

and for \(\sigma = \pm 1\)

$$\begin{aligned} \begin{array}{lll} G^{\sigma } &{}= \frac{P\,\partial _e c - \partial _r c }{2 \sqrt{2 m \beta }\,c} \begin{pmatrix} 1 &{}\quad 0 &{}\quad -1 \\ 0 &{}\quad 0 &{}\quad 0 \\ -1 &{}\quad 0 &{}\quad 1 \end{pmatrix} - \frac{\partial _e P}{\sqrt{2 m \beta }} \begin{pmatrix} \frac{1}{2}(1+\sigma ) &{}\quad 0 &{}\quad 0 \\ 0 &{}\quad 0 &{}\quad 0 \\ 0 &{}\quad 0 &{}\quad \frac{1}{2}(1-\sigma ) \end{pmatrix} \\ &{}\quad + \frac{\Upsilon }{2 \sqrt{2m/\beta }\,m c^2} \left[ (\partial _r P)^2 (\partial _e^2 P) - 2 (\partial _r P) (\partial _r \partial _e P) (\partial _e P) + (\partial _r^2 P) (\partial _e P)^2 \right] \begin{pmatrix} 0 &{}\quad 0 &{}\quad 0 \\ 0 &{}\quad 1 &{}\quad 0 \\ 0 &{}\quad 0 &{}\quad 0 \end{pmatrix} \\ &{}\quad + \frac{\sqrt{\Upsilon }}{2 \sqrt{m}\,c} \left[ (\partial _r P) (\partial _e c) - (\partial _e P) (\partial _r c) \right] \begin{pmatrix} 0 &{}\quad 1 &{}\quad 0 \\ 1 &{}\quad 0 &{}\quad -1 \\ 0 &{}\quad -1 &{}\quad 0 \end{pmatrix}. \end{array}\nonumber \\ \end{aligned}$$
(8.20)

The relation (6.12) follows by using on both sides the expressions provided above.

Note that the signs of some entries in \(G^{\sigma }\) are flipped compared to [9], which is due to different sign conventions for the eigenvectors of A.

1.3 Hard-Point and Square-Well Potential

The coupling constants for these models have been discussed already in Appendix A of [28]. For completeness, here we adapt to the current sign convention for the eigenvectors and using the velocity (instead of momentum) as field variable. The linearization matrix A and its right eigenvectors are stated in Eqs. (5.7) and (5.9). The corresponding left eigenvectors of A are

$$\begin{aligned} \tilde{\psi }_{0,h} = \tilde{Z}_{0,h}^{-1} \begin{pmatrix} 2 e h \\ -m v \\ 1 \end{pmatrix}, \qquad \tilde{\psi }_{\sigma ,h} = \tilde{Z}_{\sigma ,h}^{-1} \begin{pmatrix} 2 e \sigma h' \\ m(c_h - 2 \sigma v h) \\ 2 \sigma h \end{pmatrix}. \end{aligned}$$
(8.21)

Since the interaction potential is either zero or infinite, \(\Upsilon _h = \frac{1}{2} \beta ^{-1} \langle y; y \rangle = -e/h'\), and the normalization constants in (8.16) become

$$\begin{aligned} \tilde{Z}_{0,h} = \sqrt{m e}\,c_h/\sqrt{-h'}, \qquad \tilde{Z}_{\sigma ,h} = 2 \sqrt{m e}\,c_h \,. \end{aligned}$$
(8.22)

Specializing (8.19) and (8.20) to the square-well interaction potential leads to the coupling matrices

$$\begin{aligned} G_h^0 = \sqrt{- h'\,e/m} \begin{pmatrix} -1 &{}\quad 0 &{}\quad 0 \\ 0 &{}\quad 0 &{}\quad 0 \\ 0 &{}\quad 0 &{}\quad 1 \end{pmatrix} \end{aligned}$$
(8.23)

and for \(\sigma = \pm 1\)

$$\begin{aligned} G_h^{\sigma } = \frac{1}{2} \sqrt{e/m} \left[ \frac{1}{2(2 h^2 - h')} \begin{pmatrix} a_3 &{}\quad a_1 &{}\quad -a_3 \\ a_1 &{}\quad a_2 &{}\quad -a_1 \\ -a_3 &{}\quad -a_1 &{}\quad a_3 \end{pmatrix} - 4 h \begin{pmatrix} \frac{1}{2}(1+\sigma ) &{}\quad 0 &{}\quad 0 \\ 0 &{}\quad 0 &{}\quad 0 \\ 0 &{}\quad 0 &{}\quad \frac{1}{2}(1-\sigma ) \end{pmatrix} \right] \nonumber \\ \end{aligned}$$
(8.24)

with

$$\begin{aligned} \begin{array}{ll} a_1 &{}= 2 (-h')^{-1/2} \big ( h h'' - h'^2 - 2 h^2 h' \big ), \\ a_2 &{}= 4 h (-h')^{-1}\big ( h h'' - 2 h'^2 \big ),\\ a_3 &{}= 4 h^3 - 6 h h' + h'' . \end{array} \end{aligned}$$
(8.25)

As above, the signs of some entries in \(G_h^{\sigma }\) are flipped compared to [28], due to different sign conventions for the eigenvectors of A.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Mendl, C.B., Spohn, H. Shocks, Rarefaction Waves, and Current Fluctuations for Anharmonic Chains. J Stat Phys 166, 841–875 (2017). https://doi.org/10.1007/s10955-016-1626-5

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s10955-016-1626-5

Keywords

Navigation