Skip to main content
Log in

Generic instabilities in the relativistic Chapman–Enskog heat conduction law

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

Abstract

We address the well-posedness of the Cauchy problem corresponding to the relativistic first-order fluid equations, coupled with the Chapman–Enskog heat-flux constitutive relation. We show that the system of equations that results by considering linear perturbations with respect to a generic time direction is non-hyperbolic, since there are modes that may arbitrarily grow as wave-number increases. Then, using a result provided by Strang (J Differ Equ 2:107–114, 1966), we conclude that the full non-linear first-order theory is also non-hyperbolic, thus admitting an ill-posed initial-value formulation. Unlike Eckart’s theory, these instabilities are not present when the time direction is aligned with the fluid’s direction. However, since in general the fluid velocity is not surface-forming, the instability can only be avoided in the particular case where no rotation is present.

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

Similar content being viewed by others

Notes

  1. In [14], the curl of the spatial components of the momentum balance equation was calculated in order to decouple the transverse mode. Also, in [15], Mountain’s approximate method is used in order to examine the longitudinal modes.

  2. In the context of DTT, tensor \(I^{\mu \nu }\) contains all dissipative fluxes, while the divergence of \(A^{\mu \nu \sigma }\) has information about the driving forces. Since the latter are given in terms of the gradients of state variables, only local equilibrium quantities may appear in \(A^{\mu \nu \sigma }\) when considering first-order theories.

References

  1. Chapman, S., Cowling, T.G.: The Mathematical Theory of Non-uniform Gases. Cambridge University Press, Cambridge (1970)

    MATH  Google Scholar 

  2. Cercignani, C., Medeiros Kremer, G.: The Relativistic Boltzmann Equation: Theory and Applications. Cambridge University Press, Cambridge (1991)

    MATH  Google Scholar 

  3. Cercignani, C.: The Boltzmann Equation and Its Applications. Applied Mathematical Sciences, vol. 67. Springer, Berlin (1988)

  4. Gabbana, A., Mendoza, M., Succi, S., Tripiccione, R.: Kinetic approach to relativistic dissipation. Phys. Rev. E 96, 023305 (2017). Aug

    Article  ADS  Google Scholar 

  5. Gabbana, A., Simeoni, D., Succi, S., Tripiccione, R.: Relativistic dissipation obeys Chapman–Enskog asymptotics: analytical and numerical evidence as a basis for accurate kinetic simulations. Phys. Rev. E 99, 052126 (2019)

    Article  ADS  MathSciNet  Google Scholar 

  6. Eckart, C.: The thermodynamics of irreversible processes. iii. Relativistic theory of the simple fluid. Phys. Rev. 58, 919–924 (1940)

    Article  ADS  Google Scholar 

  7. Hiscock, W.A., Lindblom, L.: Generic instabilities in first-order dissipative relativistic fluid theories. Phys. Rev. D 31, 725 (1985)

    Article  ADS  MathSciNet  Google Scholar 

  8. Landau, L.D., Lifshitz, E.M.: Fluid Mechanics. Number v. 6 in Course on Theoretical Physics. Elsevier, Amsterdam (2013)

  9. Bemfica, F.S., Disconzi, M.M., Noronha, J.: Nonlinear causality of general first-order relativistic viscous hydrodynamics. Phys. Rev. D 100, 104020 (2019)

    Article  ADS  MathSciNet  Google Scholar 

  10. Kovtun, P.: First-order relativistic hydrodynamics is stable. JHEP 10, 034 (2019)

    Article  ADS  MathSciNet  Google Scholar 

  11. Tsumura, K., Kunihiro, T.: First-principles derivation of stable first-order generic-frame relativistic dissipative hydrodynamic equations from kinetic theory by renormalization-group method. Prog. Theor. Phys. 126(5), 761–809 (2011)

    Article  ADS  Google Scholar 

  12. Denicol, G.S., Kodama, T., Koide, T., Mota, P.: Stability and causality in relativistic dissipative hydrodynamics. J. Phys. G: Nuclear Part. Phys. 35(11), 115102 (2008)

    Article  ADS  Google Scholar 

  13. Shi, P., Koide, T., Rischke, D.H.: Does stability of relativistic dissipative fluid dynamics imply causality? Phys. Rev. D 81, 114039 (2010)

    Article  ADS  Google Scholar 

  14. Garcia-Colin, L.S., Sandoval-Villalbazo, A., García-Perciante, A.L.: Relativistic transport theory for simple fluids to first order in the gradients. Physica A 388, 3765–3770 (2009)

    Article  ADS  Google Scholar 

  15. Garcia-Perciante, A.L., Garcia-Colin, L.S., Sandoval-Villalbazo, A.: Rayleigh-brillouin spectrum in special relativistic hydrodynamics. Phys. Rev. E 79, 066310 (2009)

    Article  ADS  Google Scholar 

  16. Geroch, R., Lindblom, L.: Dissipative relativistic fluid theories of divergence type. Phys. Rev. D 41, 1855–1861 (1990)

    Article  ADS  MathSciNet  Google Scholar 

  17. Geroch, R.: Relativistic theories of dissipative fluids. J. Math. Phys. 36(8), 4226–4241 (1995)

    Article  ADS  MathSciNet  Google Scholar 

  18. Geroch, R., Lindblom, L.: Causal theories of dissipative relativistic fluids. Ann. Phys. 207(2), 394–416 (1991)

    Article  ADS  MathSciNet  Google Scholar 

  19. Geroch, R.: On hyperbolic “theories” of relativistic dissipative fluids. arXiv preprint gr-qc/0103112 (2001)

  20. Geroch, R.P.: Partial differential equations of physics. In: General relativity. Proceedings, 46th Scottish Universities Summer School in Physics, NATO Advanced Study Institute, Aberdeen, UK, 16–29 July 1995 (1996)

  21. Strang, G.: Necessary and insufficient conditions for well-posed cauchy problems. J. Differ. Equ. 2, 107–114 (1966)

    Article  ADS  MathSciNet  Google Scholar 

  22. Jüttner, F.: Das maxwellsche gesetz der geschwindigkeitsverteilung in der relativtheorie. Annalen der Physik 339(5), 856–882 (1911)

    Article  ADS  Google Scholar 

  23. Chacón-Acosta, G., Dagdug, L., Morales-Técotl, H.A.: Manifestly covariant jüttner distribution and equipartition theorem. Phys. Rev. E 81, 021126 (2010)

    Article  ADS  Google Scholar 

  24. Mazur, P., De, S.R.: Non-equilibrium Thermodynamics. Dover, New York (2011)

    MATH  Google Scholar 

  25. Méndez, A.R., García-Perciante, A.L.: Heat conduction in relativistic neutral gases revisited. Gen. Relat. Gravit. 43, 2257–2275 (2011)

    Article  ADS  MathSciNet  Google Scholar 

  26. Hurwitz, A.: Ueber die bedingungen, unter welchen eine gleichung nur wurzeln mit negativen reellen theilen besitzt. Mathematische Annalen 46(2), 273–284 (1895)

    Article  MathSciNet  Google Scholar 

  27. García-Perciante, A.L., Reula, O.: On the illposedness and stability of the relativistic heat equation. arXiv 1909.08271 (2019)

Download references

Acknowledgements

We would like to thank CONICET and SECyT-UNC for partial support. M.E.R is a postdoctoral fellow from CONICET, Argentina.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Marcelo E. Rubio.

Additional information

Communicated by Shin-ichi Sasa.

Publisher's Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

A

A

In this appendix we give a brief review about divergence-type fluid theories, and then we compute the most general constitutive tensor for first-order DTT, considering the Jüttner distribution function. Finally, we get an expression of the heat-flux constitutive relation predicted from this formalism by taking the appropriate projection for the divergence of the constitutive tensor.

1.1 A.1 Brief detour on DTT

Roughly speaking, a set of dynamic equations is considered a divergence- type theory (DTT) if it can be written as a set of equations on the divergence of the corresponding dynamical variables. Any fluid theory which is governed by a set of conservation laws (particle number density, energy and momentum densities, etc) constitutes, indeed, a divergence-type theory. However, constitutive equations, which play a major role in the assessment of causality and hyperbolicity through Geroch’s criterion need to be also expressed in divergence form. In other words, an additional tensor containing the thermodynamic forces is needed, and the equation for its divergence should lead to the corresponding constitutive equations.

Formally speaking, divergence-type fluid theories are the type of theories which satisfy the following three conditions:

(i):

The dynamical variables are given by the energy-momentum tensor \(T^{\mu \nu }\) of the fluid and the particle number current, \(N^{\mu }\);

(ii):

The dynamical equations are given by

$$\begin{aligned} \nabla _{\mu }N^{\mu }= & {} 0 \end{aligned}$$
(A.1)
$$\begin{aligned} \nabla _{\mu }T^{\mu \nu }= & {} 0 \end{aligned}$$
(A.2)
$$\begin{aligned} \nabla _{\mu }A^{\mu \nu \sigma }= & {} I^{\nu \sigma } \end{aligned}$$
(A.3)

Here, both the constitutive tensor \(A^{\mu \nu \sigma }\) and the source tensor \(I^{\nu \sigma }\) are algebraic functions of the dynamical variables \(N^{\mu }\) and \(T^{\mu \nu }\), and \(I^{\nu \sigma }\) is symmetric and traceless.

(iii):

There exists a four-current \(S^{\alpha }\), which is also a local algebraic function of \(T^{\mu \nu }\) and \(N^{\mu }\) that satisfies, as a consequence of the dynamical equations,

$$\begin{aligned} \nabla _{\alpha }S^{\alpha }=\sigma , \end{aligned}$$

with \(\sigma \ge 0\) an algebraic function of \(T^{\mu \nu }\) and \(N^{\mu }\) and \(\sigma = 0\) if and only if \(I^{\nu \sigma }\equiv 0\).

In order to close the system and to have the same quantity of variables and equations, condition \(g_{\mu \nu }A^{\alpha \mu \nu } = 0\) is required. Conditions (ii) and (iii) imply the existence of a generating function \(\chi (\xi ,\xi _{\mu },\xi _{\mu \nu })\) which contains all the information of the theory. In particular, particle flux and energy-momentum tensor can be obtained by taking derivatives of \(\chi \) with respect to the corresponding variables, as it is shown in Eqs. (2.6) and (2.7). The study of hyperbolicity properties of the theory in terms of this new formulation is much more direct. In fact, by introducing a collective abstract variable \(\xi ^A := (\xi , \xi _{\mu },\; \xi _{\mu \nu })\), Eqs. (A.1), (A.2) and (A.3) can be set into the form

$$\begin{aligned} \mathcal {K}^{\mu }{}_{AB}\nabla _{\mu } \xi ^B = J_A, \end{aligned}$$
(A.4)

where

$$\begin{aligned} \mathcal {K}^{\mu }{}_{AB} := \frac{\partial ^3 \chi }{\partial \xi _{\mu } \partial \xi ^A \partial \xi ^B} \end{aligned}$$
(A.5)

is the principal symbol of (A.4) (which by construction is symmetric in the capital indices) and \(J_A := (0,0,\;I_{\mu \nu })\). Then, following Geroch’s formalism [20], we say that the system is symmetric-hyperbolic is there exists a time-like vector field \(t^a\) such that the symmetric form \(h_{AB} = t_a\mathcal {K}^a{}_{AB}\) is a norm; that is, it is positive-definite.

1.2 A.2 Details of calculations of Section 2

We now compute the most general constitutive tensor field that can be constructed from first-order divergence-type fluid theories, using a Jüttner distribution function in local equilibrium, \(f^{(0)}\). As was pointed out in Section 2, such a constitutive tensor is made up by means of the third moment of \(f^{(0)}\).

From the macroscopic point of view, the constitutive tensor field must be an algebraic function of the dynamical variables \(N^{\mu }\) and \(T^{\mu \nu }\). On the other hand, since we are considering first-order theories and \(A^{\mu \alpha \beta }\) includes, by definition, first derivatives with respect to the dissipative tensor, we conclude that it must be of zeroth-order (for that reason we are considering just the Jüttner distribution function). Up to this order, both \(N^{\mu }\) and \(T^{\mu \nu }\) are made up in terms of the fluid four-velocity \(u^{\mu }\) and the background metric \(g_{\mu \nu }\). It is rather straightforward to see that the most general tensor field that satisfies these requirements, and has the symmetries imposed by the theory, is the one given in Eq. (2.8). In fact, recalling that the possible \(p^{\mu }\) are those restricted to the mass-shell \(p^{\mu }p_{\mu }=-1\) (where the mass of each fluid component is normalized to \(m = 1\)), we get

$$\begin{aligned} A_o= & {} - \frac{2}{3} A^{\mu \alpha \beta }\left( 2u_{\mu }u_{\alpha } + h_{\mu \alpha }\right) u_{\beta } \nonumber \\= & {} 2 \int {f^{(0)}\left( -u_{\alpha }p^{\alpha }\right) \left[ \left( -u_{\alpha }p^{\alpha }\right) ^2 - \frac{1}{2}\right] \,d\Omega }. \end{aligned}$$
(A.6)

By the change of variables

$$\begin{aligned} \varepsilon = -u_{\alpha }p^{\alpha }, \end{aligned}$$

we get

$$\begin{aligned} f^{(0)} = \frac{n}{4\pi z K_2\left( 1/z\right) }\,e^{-\varepsilon /z}, \end{aligned}$$

and \(d\Omega = \sqrt{\varepsilon ^2 - 1}\, d\varepsilon \, dS_2\), where \(dS_2\) is the area element in the unit sphere of momentum directions. Thus, integral (A.6) reduces to

$$\begin{aligned} A_o= & {} \frac{2n}{zK_2\left( 1/z\right) }\int _1^{\infty }{e^{-\varepsilon /z}\varepsilon \left( \varepsilon ^2 - \frac{1}{2}\right) \sqrt{\varepsilon ^2 - 1}\;d\varepsilon }\nonumber \\= & {} \frac{2n}{zK_2\left( 1/z\right) }\left( \frac{1}{2}\mathcal {I}_1(z) + \mathcal {I}_2(z)\right) , \nonumber \end{aligned}$$

where, for \(\ell \in \mathbb {N}\),

$$\begin{aligned} \mathcal {I}_{\ell }(z) := \int _{1}^{\infty }{e^{-\varepsilon /z}\varepsilon \left( \varepsilon ^2 - 1\right) ^{\ell -1/2}d\varepsilon }. \end{aligned}$$
(A.7)

The integral in Eq. (A.7) can be easily computed by means of some properties of modified Bessel functions, as shown in the following

Proposition 1

Let \(K_{\ell }\) be the \(\ell \)-th modified Bessel function. Then, the following identities hold for any \(\ell \in \mathbb {N}\):

  1. (i)
    $$\begin{aligned} \int _{1}^{\infty }{e^{-\varepsilon /z}\varepsilon \left( \varepsilon ^2 - 1\right) ^{\ell -1/2}d\varepsilon } = \frac{\Gamma \left( \ell +\frac{1}{2}\right) }{\Gamma \left( \frac{1}{2}\right) }\,\left( 2z\right) ^{\ell }K_{\ell +1}\left( \frac{1}{z}\right) . \end{aligned}$$
    (A.8)
  2. (ii)
    $$\begin{aligned} \frac{dK_{\ell }(x)}{dx} = \frac{\ell K_{\ell }(x)}{x} - K_{\ell +1}(x). \end{aligned}$$
    (A.9)

Proof. Identity (ii) is a direct consequence of the derivative formula

$$\begin{aligned} \frac{d}{dx}\left[ \frac{K_{\ell }(x)}{x^{\ell }}\right] = - \frac{K_{\ell + 1}(x)}{x^{\ell }}. \end{aligned}$$

Identity (i) is also consequence of the above formula and the following important one:

$$\begin{aligned} K_{\ell }(x) = \left( \frac{x}{2}\right) ^{\ell }\frac{\Gamma (1/2)}{\Gamma (\ell +1/2)}\int _{1}^{\infty }{e^{-xy}\left( y^2 - 1\right) ^{\ell -1/2}\,dy}. \end{aligned}$$

                                                                                                                                          \(\Box \)

By applying the proposition above we get, then, \(\mathcal {I}_1 = zK_2(1/z)\), \(\mathcal {I}_2 = 3z^2 K_3(1/z)\) and

$$\begin{aligned} A_o = n\left( 1 + 6z\mathcal {G}(z)\right) . \end{aligned}$$

Analogously, for \(A_1\) we get

$$\begin{aligned} A_1= & {} - \frac{1}{3}A^{\mu \alpha \beta }u_{\mu }u_{\alpha }u_{\beta } \nonumber \\= & {} \frac{n}{3zK_2(1/z)}\int _1^{\infty }{e^{-\varepsilon /z}\varepsilon \left( \varepsilon ^2 - \frac{1}{4}\right) \sqrt{\varepsilon ^2 - 1}\;d\varepsilon } \nonumber \\= & {} \frac{n}{3zK_2\left( 1/z\right) }\left( \frac{3}{4}\mathcal {I}_1(z) + \mathcal {I}_2(z)\right) \nonumber \\= & {} n\left( z\mathcal {G}(z) + \frac{1}{4}\right) . \end{aligned}$$
(A.10)

Now, in order to obtain the Chapman–Enskog constitutive relation, we proceed to project the constitutive tensor in the space perpendicular to \(u^{\mu }\). In order to do so, we find it useful to express the constitutive tensor as a sum of three contributions, namely

$$\begin{aligned} A^{\mu \alpha \beta } = A_1^{\mu \alpha \beta } + A_2^{\mu \alpha \beta } + A_3^{\mu \alpha \beta }, \end{aligned}$$
(A.11)

where

$$\begin{aligned} A_1^{\mu \alpha \beta }= & {} A_o u^{\mu }u^{\alpha }u^{\beta } \nonumber \\ A_2^{\mu \alpha \beta }= & {} A_1 u^{\mu }g^{\alpha \beta } \nonumber \\ A_3^{\mu \alpha \beta }= & {} (A_o - 4A_1)g^{\mu (\alpha }u^{\beta )} \nonumber \end{aligned}$$

Then,

$$\begin{aligned} h_{\alpha }{}^{\gamma }\nabla _{\mu }A_{1}^{\mu \alpha \beta }= & {} h_{\alpha }{}^{\gamma }\left[ u^{\alpha }u^{\beta }\dot{A}_o+A_o\nabla _{\mu }\left( u^{\mu }u^{\alpha }u^{\beta }\right) \right] \nonumber \\= & {} A_o u^{\beta }\dot{u}^{\gamma }, \end{aligned}$$
(A.12)

yielding

$$\begin{aligned} u_{\beta }h_{\alpha }{}^{\gamma }\nabla _{\mu }A_{1}^{\mu \alpha \beta } = -A_o \dot{u}^{\gamma }. \end{aligned}$$

By similar calculations, we get

$$\begin{aligned} h_{\alpha }{}^{\gamma }\nabla _{\mu }A_{2}^{\mu \alpha \beta }=\left( \dot{A}_1 + A_1 \nabla _{\mu }u^{\mu }\right) h^{\gamma \beta }, \end{aligned}$$

which implies \(u_{\beta }h_{\alpha }{}^{\gamma }\nabla _{\mu }A_{2}^{\mu \alpha \beta } = 0\). Finally, defining

$$\begin{aligned} A_3 := \frac{A_o}{2} - 2A_1, \end{aligned}$$

we have

$$\begin{aligned} u_{\beta }h_{\alpha }{}^{\gamma }\nabla _{\mu }A_{3}^{\mu \alpha \beta }= & {} 2u_{\beta }h_{\alpha }{}^{\gamma }\left[ g^{\mu (\alpha }u^{\beta )}\nabla _{\mu } A_3 + A_3\nabla ^{(\alpha }u^{\beta )}\right] \nonumber \\= & {} - h^{\gamma \mu }\left( \nabla _{\mu }A_3\right) + A_3\dot{u}^{\gamma } \nonumber \end{aligned}$$

Plugging all together, we get

$$\begin{aligned} q^{\gamma }\propto & {} u_{\beta }h^{\gamma }{}_{\alpha }\nabla _{\mu }A^{\mu \alpha \beta } \nonumber \\\propto & {} h^{\gamma \mu }\left( \nabla _{\mu }A_3 + A_3\frac{\nabla _{\mu }n}{n}\right) + \left( \frac{A_o}{2} + 2A_1\right) \dot{u}^{\gamma } \nonumber \\= & {} zh^{\gamma \mu }\left( \left( z\mathcal {G}(z)\right) ' \frac{\nabla _{\mu }T}{T} + \mathcal {G}(z)\frac{\nabla _{\mu }n}{n}\right) \nonumber \\+ & {} \left( 1+5z\mathcal {G}(z)\right) \dot{u}^{\gamma } \end{aligned}$$
(A.13)

Then, we use the local expression (2.13) for the acceleration at the leading order, and the following formula for the derivative of \(\mathcal {G}(z)\):

$$\begin{aligned} \mathcal {G}'(z) = -\frac{1}{z^2}\left[ \mathcal {G}^2(z)-5z\mathcal {G}(z)-1\right] , \end{aligned}$$
(A.14)

which implies that

$$\begin{aligned} (z\mathcal {G}(z))'= \mathcal {G}(z)\left[ 1-\frac{\bar{\mathcal {G}}(z)}{z}\right] , \end{aligned}$$
(A.15)

where

$$\begin{aligned} \bar{\mathcal {G}}(z):=\mathcal {G}(z)-\frac{1+5z\mathcal {G}(z)}{\mathcal {G}(z)}. \end{aligned}$$

Finally, recalling the contribution to the heat flux \(\delta q^{\gamma }\) given in Eq. (2.18) by taking into account first-order corrections in the four-acceleration, we get

$$\begin{aligned} q^{\gamma }\propto & {} h^{\gamma \mu } \mathcal {G}(z) \left[ \left( 1 - \frac{\bar{\mathcal {G}}(z)}{z}\right) \frac{\nabla _{\mu }T}{T} + \frac{\nabla _{\mu }n}{n}\right] - \frac{1+5z\mathcal {G}(z)}{\mathcal {G}(z)} h^{\gamma \mu }\left( \frac{\nabla _{\mu }T}{T} + \frac{\nabla _{\mu }n}{n}\right) + \delta q^{\gamma } \nonumber \\= & {} h^{\gamma \mu } \mathcal {G}(z) \left[ \left( 1 - \frac{\bar{\mathcal {G}}(z)}{z}\right) \frac{\nabla _{\mu }T}{T} + \frac{\nabla _{\mu }n}{n}\right] + h^{\gamma \mu }\left[ \bar{\mathcal {G}}(z) - \mathcal {G}(z)\right] \left( \frac{\nabla _{\mu }T}{T} + \frac{\nabla _{\mu }n}{n}\right) + \delta q^{\gamma } \nonumber \\= & {} h^{\gamma \mu }\bar{\mathcal {G}}(z)\left[ \left( 1-\frac{\mathcal {G}(z)}{z}\right) \frac{\nabla _{\mu }T}{T} + \frac{\nabla _{\mu }n}{n}\right] + \delta q^{\gamma } \nonumber \\= & {} -h^{\gamma \mu }\bar{\mathcal {G}}(z)\left( \frac{\kappa }{\lambda }\frac{\nabla _{\mu }T}{T} - \frac{\nabla _{\mu }n}{n}\right) + \delta q^{\gamma }. \end{aligned}$$
(A.16)

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

García-Perciante, A.L., Rubio, M.E. & Reula, O.A. Generic instabilities in the relativistic Chapman–Enskog heat conduction law. J Stat Phys 181, 246–262 (2020). https://doi.org/10.1007/s10955-020-02578-0

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s10955-020-02578-0

Navigation