1: \documentclass[11pt]{article}
2:
3:
4: \usepackage{amssymb,amsmath,amsfonts,amsthm}
5: \usepackage{graphicx,psfrag}
6:
7:
8: \newcommand{\N}{{\mathbb N}}
9: \newcommand{\R}{{\mathbb R}}
10: \newcommand{\SP}{{\mathbb S}}
11: \newcommand{\W}{{\mathbb W}}
12:
13:
14: \newtheorem{thh}{Theorem}
15: \newtheorem{lemm}{Lemma}
16: \newtheorem{prop}{Proposition}
17: \newtheorem{definition}{Definition}
18: \newtheorem{remarque}{Remark}
19:
20:
21: \setlength{\textwidth}{16cm}
22: \setlength{\textheight}{26cm}
23: \setlength{\oddsidemargin}{0cm}
24: \setlength{\evensidemargin}{0cm}
25: \setlength{\topmargin}{-1in}
26:
27:
28:
29:
30:
31:
32: \begin{document}
33:
34:
35:
36: \title{Shock Profiles for Non Equilibrium Radiating Gases}
37:
38: \author{Chunjin {\sc Lin}$^\dag$, Jean-Fran\c{c}ois {\sc Coulombel}$^{\dag \ddag}$,
39: Thierry {\sc Goudon}$^{\dag \ddag}$\\
40: \\
41: {\small $\dag$ Team SIMPAF--INRIA Futurs \& Universit\'e Lille 1,
42: Laboratoire Paul Painlev\'e, UMR CNRS 8524}\\
43: {\small Cit\'e scientifique, 59655 VILLENEUVE D'ASCQ Cedex, France}\\
44: {\small $\ddag$ CNRS}\\
45: {\small E-mails: {\tt chunjin.lin@math.univ-lille1.fr},
46: {\tt jfcoulom@math.univ-lille1.fr}},\\
47: {\small \tt thierry.goudon@math.univ-lille1.fr}}
48: \date{\today}
49: \maketitle
50:
51:
52: \begin{abstract}
53: We study a model of radiating gases that describes the interaction of an
54: inviscid gas with photons. We show the existence of smooth traveling waves
55: called 'shock profiles', when the strength of the shock is small. Moreover,
56: we prove that the regularity of the traveling wave increases when the
57: strength of the shock tends to zero.
58: \end{abstract}
59:
60:
61:
62:
63:
64:
65: \section{Introduction and main results}
66:
67:
68:
69: We are interested in a system of PDEs describing astrophysical flows, where
70: a gas interacts with radiation through energy exchanges. Similar questions
71: arise in the modeling of reentry problems, or high temperature combustion
72: phenomena. The gas is described by its density $\rho >0$, its bulk velocity
73: $u \in \R$, and its specific total energy $E=e+u^2/2$, where $e$ stands for
74: the specific internal energy. (Our analysis is restricted to a one-dimensional
75: framework, but this is not a loss of generality, as shown below.) We consider
76: a situation where the gas is not in thermodynamical equilibrium with the
77: radiations, which are thus described by their own energy $n$. The evolution
78: of the gas flow is governed by the system:
79: \begin{equation}
80: \label{eulerevol}
81: \begin{cases}
82: \partial_t \rho +\partial_x (\rho \, u)=0 \, ,& \\
83: \partial_t (\rho \, u) +\partial_x (\rho \, u^2+P)=0\, ,& \\
84: \partial_t (\rho \, E) +\partial_x (\rho \, E \, u+P \, u)=n-\theta^4 \, ,&
85: \end{cases}
86: \end{equation}
87: where the right-hand side in the last equation accounts for energy exchanges
88: with the radiations, $P$ being the pressure of the gas, and $\theta$ its
89: temperature. Throughout the paper, we always assume that the gas obeys the
90: perfect gas pressure law:
91: \begin{equation}
92: \label{pressure}
93: P=R \, \rho \, \theta =(\gamma-1) \, \rho \, e \, ,
94: \end{equation}
95: where $R$ is the perfect gas constant, and $\gamma>1$ is the ratio of the
96: specific heats at constant pressure, and volume. This assumption yields many
97: algebraic simplifications, but we believe that our results still hold for a
98: general pressure law satisfying the usual requirements of thermodynamics.
99: System \eqref{eulerevol} is completed by considering that radiations are
100: described by a stationary diffusion regime that reads:
101: \begin{equation}
102: \label{diffn}
103: -\partial_{xx} n =\theta^4 -n \, .
104: \end{equation}
105: We detail in Appendix \ref{model} how the system \eqref{eulerevol}, \eqref{diffn}
106: can be formally derived by asymptotics arguments, starting from a more complete
107: system involving a kinetic equation for the specific intensity of radiation.
108:
109:
110: As a matter of fact, the operator $(1-\partial_{xx})$ can be explicitly
111: inverted, and \eqref{diffn} can be recast as a convolution:
112: \begin{equation}
113: \label{diffn2}
114: n(t,x) =\dfrac{1}{2} \, \displaystyle \int_{\R}
115: {\rm e}^{-|x-y|} \, \theta (t,y)^4 \, dy \, .
116: \end{equation}
117: Let us introduce the quantity:
118: \begin{equation}
119: \label{defq}
120: q(t,x) :=-\partial_x n \, (t,x)=\dfrac{1}{2} \displaystyle \int_{\R}
121: {\rm e}^{-|x-y|} \, {\rm sgn}(x-y) \, \theta (t,y)^4 \, dy \, ,
122: \end{equation}
123: where sgn is the sign function:
124: \begin{equation*}
125: {\rm sgn} (x)=\begin{cases}
126: 1 &\text{if $x>0$,} \\
127: 0 &\text{if $x=0$,} \\
128: -1 &\text{if $x<0$.}
129: \end{cases}
130: \end{equation*}
131: The quantity $q$ can be interpreted as the radiative heat flux. Then, we
132: can rewrite \eqref{eulerevol}, \eqref{diffn} as follows:
133: \begin{equation}
134: \label{Euler2}
135: \begin{cases}
136: \partial_t \rho +\partial_x (\rho \, u) =0\, ,& \\
137: \partial_t (\rho \, u) +\partial_x (\rho \, u^2+P)=0 \, ,& \\
138: \partial_t (\rho \, E) +\partial_x (\rho \, E \, u +P \, u +q) =0 \, ,&
139: \end{cases}
140: \end{equation}
141: with $q$ given by \eqref{defq}. Recall that $E=e+u^2/2$, and $P$ is given
142: by \eqref{pressure}.
143:
144:
145: In this paper, we address the question of the influence of the energy exchanges
146: on the structure of shock waves. More precisely, let us consider given states at
147: infinity $(\rho_\pm, u_\pm,e_\pm)$, and let us asume that:
148: \begin{equation}
149: \label{shock}
150: (\rho,u,e)(t,x) =\begin{cases}
151: (\rho_-,u_-,e_-) &\text{if $x<\sigma \, t$,} \\
152: (\rho_+,u_+,e_+) &\text{if $x>\sigma \, t$,}
153: \end{cases}
154: \end{equation}
155: is a shock wave, with speed $\sigma$, solution to the standard Euler
156: equations (that is, system \eqref{Euler2} with $q \equiv 0$). We refer to
157: \cite{lax,serrelivre,smoller} for a detailed study of shock waves for the
158: Euler equations. The question we ask is the following: does there exist a
159: traveling wave $(\rho,u,e)(x-\sigma t)$ solution to \eqref{Euler2}, with $q$
160: given by \eqref{defq}, that satisfies the asymptotic conditions:
161: \begin{equation}
162: \label{asym-cond}
163: \lim_{\xi \rightarrow \pm\infty} (\rho,u,e)(\xi) =(\rho_\pm,u_\pm,e_\pm) \, .
164: \end{equation}
165: In other words, we are concerned with the existence of a shock profile, and a
166: natural expectation (at least for shocks of small amplitude) is that the step
167: shock \eqref{shock} is smoothed into a continuous profile, due to the dissipation
168: introduced by \eqref{diffn}. The analogous problem for the compressible Navier-Stokes
169: system has been treated a long time ago, see \cite{gilbarg}, without any smallness
170: assumption on the shock wave. Concerning radiative transfer, a formal analysis
171: of shock profiles has been performed in \cite{HB}, together with rough numerical
172: simulations. (We refer also to \cite{ZR,MM} for the physical background.) The
173: main purpose of this work is to make the analysis of \cite{HB} rigorous. Since
174: we are only concerned in this paper with the existence of shock profiles, and not
175: with their stability, the problem is purely one-dimensional (due to the Galilean
176: invariance of the Euler equations). This is why we have directly restricted to
177: the one-dimensional case. However, the formal derivation of Appendix \ref{model}
178: is made in several space dimensions.
179:
180:
181: Before stating our main results, let us mention that a simplified version
182: of \eqref{Euler2}, \eqref{defq} has been introduced, and studied in \cite{SchoTad} and later in
183: \cite{kawanishi,kawanishi2}. This 'baby-model' consists in a Burgers
184: type equation:
185: \begin{equation*}
186: \partial_t u + \partial_x \Big( \dfrac{u^2}{2} \Big)=-\partial_x q \, ,
187: \end{equation*}
188: coupled to the diffusion equation:
189: \begin{equation*}
190: -\partial_{xx} q+q=-\partial_x u \, .
191: \end{equation*}
192: These two equations can be seen as a scalar version of \eqref{Euler2},
193: \eqref{defq} since they can be recast as:
194: \begin{equation}
195: \label{Kburgers}
196: \partial_t u +\partial_x \Big( \dfrac{u^2}{2} \Big) =Ku-u \, ,
197: \end{equation}
198: where $K$ is the integral operator already arising in \eqref{defq}:
199: \[
200: Ku (t,x)=\dfrac{1}{2} \displaystyle \int_{\R} {\rm e}^{-|x-y|} \, u(t,y) \, dy \, .
201: \]
202: The thorough study of \eqref{Kburgers} has motivated a lot of works; we mention
203: in particular \cite{nishi,marcati,tadmor,serre}. Clearly \eqref{Kburgers} can be
204: seen as a prototype for discussing \eqref{Euler2}, \eqref{defq}; nevertheless,
205: replacing \eqref{Euler2}, and \eqref{defq} by \eqref{Kburgers} has two important
206: consequences: the equation becomes scalar, and the 'diffusion' $K-1$ applies to
207: the unique unknown (while in \eqref{Euler2}, the 'diffusion' appears through the
208: radiative heat flux $q$ only in the third equation). Our work is a first attempt
209: to extend the known results for \eqref{Kburgers} to the more physical model
210: \eqref{Euler2}, \eqref{defq}.
211:
212:
213: Let us now state our main results. The first result deals with the existence
214: of smooth shock profiles when the strength of the shock is small:
215:
216:
217: \begin{thh}
218: \label{exis}
219: Let $\gamma$ satisfy
220: \begin{equation*}
221: 1< \gamma <\dfrac{\sqrt{7}+1}{\sqrt{7}-1} \simeq 2.215 \, ,
222: \end{equation*}
223: and let $(\rho_-,u_-,e_-)$ be fixed. Then there exists a positive constant
224: $\delta$ (that depends on $(\rho_-,u_-,e_-)$, and $\gamma$) such that, for
225: all state $(\rho_+,u_+,e_+)$ verifying:
226: \begin{itemize}
227: \item $\|(\rho_+,u_+,e_+)-(\rho_-,u_-,e_-)\| \le \delta$,
228:
229: \item the function \eqref{shock} is a shock wave, with speed $\sigma$, for
230: the (standard) Euler equations,
231: \end{itemize}
232: then there exists a $C^2$ traveling wave $(\rho,u,e)(x-\sigma t)$ solution
233: to \eqref{Euler2}, \eqref{defq}, \eqref{asym-cond}.
234: \end{thh}
235:
236:
237: As in the study of the 'baby-model' \eqref{Kburgers}, the existence of
238: a smooth shock profile is linked to a smallness assumption on the shock
239: strength, see \cite{kawanishi}. Here the smallness parameter $\delta$
240: may depend on the state $(\rho_-,u_-,e_-)$, while for \eqref{Kburgers},
241: the smallness parameter is uniform (and even explicit!).
242:
243:
244: The restriction on the adiabatic constant $\gamma$ might be unnecessary,
245: but it simplifies the proof, and it covers the main physical cases
246: $1<\gamma \le 2$.
247:
248:
249: Our second result is also in the spirit of \cite{kawanishi}, and deals with
250: the smoothness of the shock profile constructed in the previous Theorem:
251:
252:
253: \begin{thh}
254: \label{smooth}
255: Let $\gamma$ satisfy
256: \begin{equation*}
257: 1< \gamma < \dfrac{\sqrt{7}+1}{\sqrt{7}-1} \simeq 2.215 \, ,
258: \end{equation*}
259: and let $(\rho_-,u_-,e_-)$ be fixed. Then there exists a decreasing
260: sequence of positive numbers $(\delta_n)_{n \in \N}$ (the sequence depends
261: on $(\rho_-,u_-,e_-)$, and $\gamma$) such that, for all $n\in \N$, and for
262: all state $(\rho_+,u_+,e_+)$ verifying:
263: \begin{itemize}
264: \item $\|(\rho_+,u_+,e_+)-(\rho_-,u_-,e_-)\| \le \delta_n$,
265:
266: \item the function \eqref{shock} is a shock wave, with speed $\sigma$, for
267: the (standard) Euler equations,
268: \end{itemize}
269: then there exists a $C^{n+2}$ traveling wave $(\rho,u,e)(x-\sigma t)$ solution
270: to \eqref{Euler2}, \eqref{defq}, \eqref{asym-cond}.
271: \end{thh}
272:
273:
274: To a large extent, our analysis follows the arguments of \cite{HB}, \cite{SchoTad} and
275: \cite{kawanishi}. The proof of Theorem \ref{exis} is presented in Section
276: \ref{Proof}, while Section \ref{moresmooth} is devoted to the proof of
277: Theorem \ref{smooth}. The investigation of strong shocks, as well as
278: stability issues will be addressed in a forthcoming work.
279:
280:
281:
282:
283:
284:
285: \section{Existence of smooth shock profiles}
286: \label{Proof}
287:
288:
289:
290: In this section, we prove Theorem \ref{exis}. We first recall some basic facts
291: on shock waves for the Euler equations. Then, we make some transformations on
292: the traveling wave equation. Eventually, we prove Theorem \ref{exis} by using
293: an auxiliary system of Ordinary Differential Equations, that is introduced and
294: studied in the last paragraph of this section.
295:
296:
297:
298:
299:
300:
301: \subsection{Shock wave solutions to the Euler equations}
302:
303:
304:
305: In this paragraph, we recall some basic facts about the (entropic) shock
306: wave solutions to the Euler equations:
307: \begin{equation*}
308: \begin{cases}
309: \partial_t \rho +\partial_x (\rho \, u) =0 \, ,\\
310: \partial_t (\rho \, u) +\partial_x (\rho \, u^2 +P) =0 \, ,\\
311: \partial_t (\rho \, E) +\partial_x (\rho \, E \, u+ P \, u) = 0 \, ,
312: \end{cases}
313: \end{equation*}
314: where $P$, and $E$ are given as in the introduction. We refer to
315: \cite{lax,serrelivre,smoller} for all the details, and omit the calculations.
316: In all what follows, we only consider shock waves that satisfy Lax shock
317: inequalities. We shall thus speak of $1$-shock waves, or $3$-shock waves.
318:
319:
320: We consider a fixed 'left' state $(\rho_-,u_-,e_-)$. Then the 'right' states
321: $(\rho_+,u_+,e_+)$ such that $(\rho_\pm,u_\pm,e_\pm)$ define a $1$-shock
322: wave, with speed $\sigma$, is a half-curve initiating at $(\rho_-,u_-,e_-)$.
323: Introducing the notation $v_\pm=u_\pm -\sigma$, the Rankine-Hugoniot jump
324: conditions can be rewritten as:
325: \begin{equation}
326: \label{defjC1C2}
327: \begin{cases}
328: \rho_+ \, v_+ =\rho_- \, v_- =:j \, ,\\
329: \rho_+ \, v_+^2 +P_+ =\rho_- \, v_-^2 +P_- =:j \, C_1 \, ,\\
330: \rho_+ \, v_+ \big( e_+ +\dfrac{v_+^2}{2} \big) +P_+ \, v_+
331: =\rho_- \, v_- \big( e_- +\dfrac{v_-^2}{2} \big) +P_- \, v_- =:j \, C_2 \, .
332: \end{cases}
333: \end{equation}
334: Observe that $v_-$ does not only depend on the 'left' state $(\rho_-,u_-,e_-)$,
335: but also on $(\rho_+,u_+,e_+)$, because $v_-$ is defined with the help of the
336: shock speed $\sigma$. Consequently, the constants $j$, $C_1$, and $C_2$ depend
337: on both $(\rho_-,u_-,e_-)$, and $(\rho_+,u_+,e_+)$.
338:
339:
340: For $1$-shocks, that is when the inequalities:
341: \begin{equation*}
342: u_+ -c_+ < \sigma <u_+ \, ,\quad \sigma < u_- -c_- \, ,
343: \end{equation*}
344: are satisfied ($c$ denotes the sound speed), all quantities $j$, $C_1$, and
345: $C_2$ are positive. Moreover, when the strength of the shock tends to zero,
346: that is when $(\rho_+,u_+,e_+)$ tends to $(\rho_-,u_-,e_-)$, one has
347: \begin{equation}
348: \label{asymp}
349: \begin{pmatrix}
350: \sigma \\
351: j \\
352: C_1 \\
353: C_2 \end{pmatrix} \longrightarrow \begin{pmatrix}
354: u_- - c_- \\
355: \rho_- \, c_- \\
356: c_- +(\gamma-1) e_-/c_- \\
357: \gamma \, e_- +c_-^2/2 \end{pmatrix} \, .
358: \end{equation}
359: Consequently, when the strength of the shock is small, all quantities
360: $j$, $C_1$, $C_2$ are bounded away from zero. Recall also that $1$-shocks
361: are compressive, in the sense that $\rho_+ >\rho_-$. This inequality
362: immediately implies that $0<v_+ <v_-$. Eventually, the strength of the
363: shock tends to zero if, and only if $u_+$ tends to $u_-$ (in that case,
364: we also have $\rho_+ \rightarrow \rho_-$, and $e_+ \rightarrow e_-$
365: because of the Rankine-Hugoniot jump conditions).
366:
367:
368: In all what follows, we limit our discussion to the case of $1$-shocks
369: for simplicity, but the extension to $3$-shocks is immediate.
370:
371:
372:
373:
374:
375:
376: \subsection{Reduction of the traveling wave equation}
377:
378:
379:
380: In this paragraph, we derive, and transform the equation satisfied by
381: traveling wave solutions to \eqref{Euler2}, \eqref{defq}. A traveling
382: wave solution to \eqref{Euler2}, \eqref{defq} with speed $\sigma$ is a
383: solution $(\rho,u,e)(x-\sigma t)$. For such solutions, the radiative
384: heat flux $q$ also depends on the sole variable $x-\sigma t$:
385: \begin{equation*}
386: q(x-\sigma t)=\dfrac{1}{2} \displaystyle \int_{\R}
387: {\rm e}^{-|x-\sigma t-y|} \, {\rm sgn}(x-\sigma t-y) \, \theta (y)^4 \, dy \, ,
388: \end{equation*}
389: and \eqref{Euler2} reads:
390: \begin{equation*}
391: \begin{cases}
392: (\rho\, (u-\sigma))' =0 \, ,\\
393: (\rho \, u \, (u-\sigma)+(\gamma-1) \, \rho \, e)' =0 \, ,\\
394: (\rho \, (e+\frac{u^2}{2}) \, (u-\sigma)+(\gamma-1) \, \rho \, e \, u+q)' =0 \, ,
395: \end{cases}
396: \end{equation*}
397: where $'$ denotes differentiation with respect to the variable
398: $\xi=x-\sigma t$. Introducing the new unknown $v=u-\sigma$, the
399: above system is easily seen to be equivalent to:
400: \begin{equation}
401: \label{Euler-xi}
402: \begin{cases}
403: (\rho \, v)'=0 \, ,\\
404: (\rho \, v^2+(\gamma-1)\, \rho \, e)' =0 \, ,\\
405: (\rho \, v \, (e+\frac{v^2}{2})+(\gamma-1) \, \rho \, v \, e+q)' =0 \, .
406: \end{cases}
407: \end{equation}
408: Since we are looking for a shock profile, the traveling wave solution
409: should also satisfy:
410: \begin{equation}
411: \label{asym-cond2}
412: \lim_{\xi\to \pm\infty} (\rho,v,e) =(\rho_{\pm},v_{\pm},e_{\pm}) \, ,
413: \end{equation}
414: where $v_{\pm}=u_{\pm}-\sigma$, and $(\rho_\pm,u_\pm,e_\pm)$ defines a
415: $1$-shock wave with speed $\sigma$ for the Euler equations. Notice that
416: the quantity $a=|u_+-u_-|/2$, that measures the strength of the shock,
417: is invariant with respect to our change of velocity, that is
418: $a=|u_+-u_-|/2=|v_+-v_-|/2$. Recall also that for $1$-shocks,
419: there holds $v_->v_+>0$.
420:
421:
422: Observing that we have:
423: \begin{equation*}
424: q(\xi)=\dfrac{1}{2} \displaystyle \int_0^{+\infty} {\rm e}^{-y} \,
425: \big( \theta (\xi-y)^4 -\theta (\xi+y)^4 \big) \, dy \, ,
426: \end{equation*}
427: we conclude that $q$ tends to zero at $\pm \infty$ by Lebesgue's Theorem
428: (because $\theta$ is necessarily bounded since it has finite limits at
429: $\pm \infty$). We can thus integrate the system
430: \eqref{Euler-xi}-\eqref{asym-cond2} once, and \eqref{Euler-xi} reads
431: equivalently:
432: \begin{equation}
433: \label{system}
434: \begin{cases}
435: (\rho \, v)(\xi)=j \, ,\\
436: (\rho \, v^2+(\gamma-1) \, \rho \, e)(\xi)=j \, C_1 \, ,\\
437: (\rho \, v \, (e+\frac{v^2}{2})+(\gamma-1) \, \rho \, v \, e+q)(\xi)=j \, C_2 \, ,
438: \end{cases}
439: \end{equation}
440: where the constants $j,$ $C_1$ and $C_2$ are given by the Rankine-Hugoniot
441: conditions \eqref{defjC1C2}. For small shocks, the positive constants $j$,
442: $C_1$, $C_2$ have the asymptotic behavior \eqref{asymp}.
443:
444:
445: From the two first equations of \eqref{system}, we derive the relations:
446: \begin{equation*}
447: \rho (\xi)=\dfrac{j}{v(\xi)} \, ,\quad
448: e(\xi)=\dfrac{(C_1-v(\xi)) \, v(\xi)}{\gamma-1} \, .
449: \end{equation*}
450: The third equation of \eqref{system} thus reduces to:
451: \begin{equation}
452: \label{v-q}
453: v(\xi)^2 -\dfrac{2 \, \gamma \, C_1}{\gamma+1} \, v(\xi)
454: +\dfrac{2 \, (\gamma-1) \, C_2}{\gamma+1}
455: =\dfrac{2 \, (\gamma-1)}{j \, (\gamma+1)} \, q(\xi) \, .
456: \end{equation}
457: Using the equation of state \eqref{pressure}, as well as the second equation
458: of \eqref{system}, we get:
459: \begin{equation*}
460: \theta (\xi)=\dfrac{(\gamma-1) \, e(\xi)}{R} =\dfrac{(C_1-v(\xi)) \, v(\xi)}{R} \, .
461: \end{equation*}
462: Consequently, \eqref{v-q} can be recast as an integral equation with a single
463: unknown function $v$:
464: \begin{equation}
465: \label{eqv}
466: v(\xi)^2 -\dfrac{2 \, \gamma \, C_1}{\gamma+1} \, v(\xi)
467: +\dfrac{2 \, (\gamma-1) \, C_2}{\gamma+1}
468: =\dfrac{(\gamma-1)}{j \, (\gamma+1) \, R^4} \int_\R {\rm e}^{-|\xi-y|} \,
469: {\rm sgn} (\xi-y) \, v(y)^4 (C_1-v(y))^4 \, dy \, .
470: \end{equation}
471: We are searching for a solution $v$ to \eqref{eqv}, that satisfies the asymptotic
472: conditions $v(\xi)\rightarrow v_\pm$, as $\xi \rightarrow \pm \infty$.
473:
474:
475: \begin{remarque}
476: If we find a $C^2$ solution $v$ to \eqref{eqv} that does not vanish, and that
477: satisfies $v(\xi) \to v_{\pm}$ as $\xi \to \pm \infty$, then we obtain a $C^2$
478: shock profile $(\rho,u,e)$ by simply setting:
479: \begin{equation*}
480: \rho(\xi)=\dfrac{j}{v(\xi)} \, ,\quad u(\xi)=v(\xi)+\sigma \, ,\quad
481: e(\xi)=\dfrac{(C_1-v(\xi))\, v(\xi)}{\gamma-1} \, .
482: \end{equation*}
483: In particular, if $v(\xi) \in [v_+,v_-]$ for all $\xi$, then $v$ does not vanish.
484: \end{remarque}
485:
486:
487: \begin{remarque}
488: Since the heat flux $q$ vanishes at $\pm \infty$, \eqref{eqv} can be also
489: rewritten as:
490: \begin{equation*}
491: (v(\xi)-v_-)(v(\xi)-v_+) =
492: \dfrac{(\gamma-1)}{j \, (\gamma+1) \, R^4} \int_\R {\rm e}^{-|\xi-y|} \,
493: {\rm sgn} (\xi-y) \, v(y)^4 \, (C_1-v(y))^4 \, dy \, .
494: \end{equation*}
495: \end{remarque}
496:
497:
498: We are going to rewrite \eqref{eqv} as a second order differential equation,
499: that will be easier to study than the integral equation \eqref{eqv}. Indeed,
500: assuming that $v$ is a $C^2$ function of $\xi$, and differentiating twice
501: \eqref{eqv} with respect to $\xi$, we get (see \cite{HB} for the details of
502: the computations):
503: \begin{equation}
504: \label{eqv1}
505: (v-\dfrac{\gamma \, C_1}{\gamma+1}) \, v''+(v')^2
506: -\dfrac{4 \, (\gamma-1)}{j \, (\gamma+1) \, R^4} \, (C_1-v)^3 \, v^3
507: \, (C_1-2v) \, v' -\dfrac{1}{2} \, (v-v_-) \, (v-v_+) =0 \, .
508: \end{equation}
509: Conversely, if $v$ is a $C^2$ solution to \eqref{eqv1} that satisfies
510: $v(\xi) \to v_{\pm}$ as $\xi \to \pm \infty$, then $v$ is also a solution
511: to \eqref{eqv}. If in addition $v$ takes its values in the interval
512: $[v_+,v_-]$, then we can construct a $C^2$ shock profile, and thus prove
513: Theorem \ref{exis}.
514:
515:
516: The differential equation \eqref{eqv1} can be simplified by introducing the
517: new unknown function $\hat{v}=v-(v_- +v_+)/2$, and by rewriting the second
518: order differential equation as a first order system:
519: \begin{equation}
520: \label{ode}
521: \left\{
522: \begin{array}{lll}
523: \hat{v}' &=& w \, ,\\
524: \hat{v} \, w' &=& -w^2-f(\hat{v}) \, w+\dfrac{\hat{v}^2-a^2}{2} \, ,
525: \end{array}
526: \right.
527: \end{equation}
528: where $f$ is the following polynomial function:
529: \begin{equation}
530: \label{fun f}
531: f(\hat{v})=\dfrac{4 \, (\gamma-1)}{j \, R^4 \, (\gamma+1)} \,
532: \left( \dfrac{C_1}{\gamma+1}-\hat{v} \right)^3 \,
533: \left( \hat{v}+\dfrac{\gamma\, C_1}{\gamma+1} \right)^3 \,
534: \left( 2\hat{v}+\dfrac{(\gamma-1) \, C_1}{\gamma+1} \right) \, .
535: \end{equation}
536: We recall that $a=|v_- -v_+|/2$, and that $a$ measures the strength of the
537: shock.
538:
539:
540: \begin{remarque}
541: \label{f0}
542: The asymptotic behavior \eqref{asymp} of $j$, $C_1$, and $C_2$ shows that when
543: the strength of the shock tends to zero ($a \to 0^+$), the limit of $f(0)$ is
544: given by:
545: \begin{equation*}
546: f(0) \rightarrow \dfrac{4 \, \gamma^3 \, (\gamma -1)^2}{R^4 \, (\gamma +1)^8}
547: \, \dfrac{(c_- +(\gamma-1) \, \frac{e_-}{c_-})^7}{\rho_- \, c_-} >0 \, .
548: \end{equation*}
549: \end{remarque}
550:
551:
552: Since $v_+ <v_-$ for a $1$-shock, we are searching for a solution to \eqref{ode}
553: that is defined on all $\R$, and that satisfies:
554: \begin{equation}
555: \label{asy-ode}
556: \lim_{\xi \to -\infty} (\hat{v},w)(\xi) = (a,0) \, ,\quad
557: \lim_{\xi \to +\infty} (\hat{v},w)(\xi) = (-a,0) \, .
558: \end{equation}
559: To prove Theorem \ref{exis}, we are thus reduced to showing the existence
560: of a heteroclinic orbit for \eqref{ode} that connects the stationary solutions
561: $(\pm a,0)$. Due to the previous transformation $\hat{v}=v-(v_-+v_+)/2$, if
562: $\hat{v}$ takes its values in $[-a,a]$, then $v=\hat{v}+(v_-+v_+)/2$ will
563: take its values in the interval $[v_+,v_-]$, and therefore will not vanish.
564:
565:
566: \begin{remarque}
567: The system \eqref{ode} is 'singular' at $\hat{v}=0$. Nevertheless, we are
568: searching for a smooth solution connecting $(\pm a,0)$, so that $\hat{v}$
569: vanishes in at least one point. Because $w'=\hat{v}''$ should also have a
570: limit at this point, a $C^2$ shock profile can exist only if the equation:
571: \begin{equation*}
572: w^2 +f(0) \, w +\dfrac{a^2}{2} =0 \, ,
573: \end{equation*}
574: has real roots. The corresponding discriminant condition turns out to be
575: much less simple than the one found in \cite{kawanishi} for the 'baby model'
576: \eqref{Kburgers}. (In particular, $f(0)$ depends on the shock through the
577: constants $j$, and $C_1$). This is a first 'nonexplicit' restriction on
578: the shock strength to derive the existence of a smooth shock profile.
579: \end{remarque}
580:
581:
582: Due to the singular nature of the system \eqref{ode} at $\hat{v}=0$, it is
583: more convenient to work on an auxiliary system of ODEs, where the singularity
584: has been eliminated (at least formally) thanks to a change of variables. This
585: procedure was already used in \cite{kawanishi}. In the next paragraph, we
586: shall introduce this auxiliary system, and complete the proof of Theorem
587: \ref{exis}.
588:
589:
590:
591:
592:
593:
594: \subsection{Existence of a heteroclinic orbit}
595:
596:
597:
598: We begin with a result on an auxiliary system of ODEs, where the singularity
599: at $\hat{v}=0$ has been eliminated:
600:
601:
602: \begin{prop}
603: \label{lem}
604: Assume that $\gamma$ satisfies $1<\gamma<(\sqrt{7}+1)/(\sqrt{7}-1)$, and
605: consider the following system of ODEs:
606: \begin{equation}
607: \label{ode-ref}
608: \left\{
609: \begin{array}{rcl}
610: V' & = & V\, W \, ,\\
611: W' & = & -W^2-f(V) \, W+\dfrac{(V^2-a^2)}{2} \, .
612: \end{array}
613: \right.
614: \end{equation}
615: There exists a positive constant $a_0$, that depends only on
616: $(\rho_-,u_-,e_-)$, and $\gamma$ such that if the shock strength $a$
617: satisfies $a\in (0,a_0]$, the following properties hold:
618: \begin{itemize}
619: \item $f(0)^2-2a^2>0$, and we define $w_0 := \big(
620: -f(0)+\sqrt{f(0)^2-2a^2} \big) /2 <0$.
621:
622: \item There exists a solution $(V_\flat,W_\flat)$ to \eqref{ode-ref} that
623: is defined on all $\R$, and that satisfies
624: \begin{equation*}
625: \lim_{\eta \rightarrow -\infty} (V_\flat,W_\flat)(\eta) =(a,0) \, ,\quad
626: \lim_{\eta \rightarrow +\infty} (V_\flat,W_\flat)(\eta) =(0,w_0) \, .
627: \end{equation*}
628: Furthermore, $V_\flat$ is decreasing, and the convergence of $V_\flat$ to
629: $0$ as $\eta \rightarrow +\infty$ is exponential.
630:
631:
632: \item There exists a solution $(V_\sharp,W_\sharp)$ to \eqref{ode-ref} that
633: is defined on all $\R$, and that satisfies
634: \begin{equation*}
635: \lim_{\eta \rightarrow -\infty} (V_\sharp,W_\sharp)(\eta) =(-a,0) \, ,\quad
636: \lim_{\eta \rightarrow +\infty} (V_\sharp,W_\sharp)(\eta) =(0,w_0) \, .
637: \end{equation*}
638: Furthermore, $V_\sharp$ is increasing, and the convergence of $V_\sharp$ to
639: $0$ as $\eta \rightarrow +\infty$ is exponential.
640: \end{itemize}
641: \end{prop}
642:
643:
644: Assuming that the result of Proposition \ref{lem} holds, the existence of
645: a heteroclinic orbit for \eqref{ode} connecting $(\pm a,0)$ can be derived
646: by following the analysis of \cite{SchoTad, kawanishi}. We briefly recall the method.
647: Using the solution $(V_\flat,W_\flat)$, we introduce the change of variable:
648: $$
649: \Xi_\flat (\eta) = -\int_{\eta}^{+\infty} V_\flat (\zeta) \, d\zeta \, .
650: $$
651: Since $V_\flat$ tends to $0$ exponentially as $\eta$ tends to $+\infty$,
652: $\Xi_\flat$ is well-defined, and it is an increasing $C^\infty$ diffeomorphism
653: from $\R$ to $(-\infty,0)$. Then $(\hat{v}_\flat,w_\flat) :=(V_\flat,W_\flat)
654: \circ \Xi_\flat^{-1}$ is a $C^\infty$ solution to \eqref{ode} on the interval
655: $(-\infty,0)$, that satisfies:
656: \begin{equation*}
657: \lim_{\xi \rightarrow -\infty} (\hat{v}_\flat,w_\flat)(\xi) =(a,0) \, ,\quad
658: \lim_{\xi \rightarrow 0^-} (\hat{v}_\flat,w_\flat)(\xi) =(0,w_0) \, .
659: \end{equation*}
660: Similarly, with the help of the solution $(V_\sharp,W_\sharp)$ we can construct
661: a $C^\infty$, decreasing diffeomorphism $\Xi_\sharp$ from $\R$ to $(0,+\infty)$,
662: and a $C^\infty$ solution $(\hat{v}_\sharp,w_\sharp)$ to \eqref{ode} on the interval
663: $(0,+\infty)$. This solution $(\hat{v}_\sharp,w_\sharp)$ connects $(0,w_0)$ and
664: $(-a,0)$, as $\xi$ varies from $0^+$ to $+\infty$. Let us now 'glue' the solutions
665: $(\hat{v}_\flat,w_\flat)$, and $(\hat{v}_\sharp,w_\sharp)$, by defining:
666: \begin{equation}
667: \label{defsolution}
668: (\hat{v},w) (\xi) :=\begin{cases}
669: (\hat{v}_\flat,w_\flat) (\xi) &\text{if $\xi<0$,}\\
670: (\hat{v}_\sharp,w_\sharp) (\xi) &\text{if $\xi>0$,}
671: \end{cases}
672: \end{equation}
673: and extend the functions $\hat{v}$, and $w$ at $0$ by setting
674: $(\hat{v},w)(0)=(0,w_0)$. In this way, $\hat{v}$, and $w$ are continuous
675: on $\R$, and $C^\infty$ on $\R \setminus \{ 0\}$. It remains to show that
676: $\hat{v} \in C^2(\R)$, that $(\hat{v},w)$ solves \eqref{ode} on $\R$, and
677: that $\hat{v}$ takes its values in $(-a,a)$.
678:
679:
680: Observe first of all that $\hat{v}$ is a decreasing function, because of
681: the monotonicity properties of $V_\flat,V_\sharp,\Xi_\flat,\Xi_\sharp$. Using
682: the asymptotic behavior of $V_\flat$, $V_\sharp$ at $-\infty$, we get that
683: $\hat{v}(\xi) \in (-a,a)$ for all $\xi \in \R$.
684:
685:
686: Let us now note that the above construction of $(\hat{v},w)$ shows that
687: $(\hat{v},w)$ is a solution to \eqref{ode} on $\R \setminus \{ 0\}$. In
688: particular, $\hat{v}' (\xi)=w(\xi)$ if $\xi \neq 0$. Moreover, $w$ is
689: continuous on $\R$, so $\hat{v} \in C^1(\R)$, and $\hat{v}'(0)=w(0)=w_0$.
690: To prove that $\hat{v}\in C^2(\R)$, it is sufficient to show that $w\in
691: C^1(\R)$, which is equivalent to showing that $w'$ has a limit at $0$
692: (because we already know that $w$ is $C^\infty$ on $\R \setminus \{ 0\}$).
693: To prove that $w'$ has a limit at $0$, we are going to study the asymptotic
694: behavior of $(V_\flat,W_\flat)$, and $(V_\sharp,W_\sharp)$ at $+\infty$.
695: More precisely, let us denote $U(V,W)$ the vector field associated to the
696: ODE \eqref{ode-ref}:
697: \begin{equation}
698: \label{defU}
699: U(V,W)=\begin{pmatrix}
700: V \, W \\
701: -W^2- f(V) \, W+\dfrac{(V^2-a^2)}{2} \end{pmatrix} \, ,
702: \end{equation}
703: where $f$ is given by \eqref{fun f}. The Jacobian matrix of $U$ at
704: $(0,w_0)$ is:
705: \begin{equation*}
706: \begin{pmatrix}
707: w_0 & 0 \\
708: -f'(0) \, w_0 & -2\, w_0-f(0) \end{pmatrix} = \begin{pmatrix}
709: \lambda_1^{(0)} & 0 \\
710: b_0 & \lambda_2^{(0)} \end{pmatrix} \, .
711: \end{equation*}
712: For $a$ sufficiently small, one checks that $\lambda_2^{(0)}<\lambda_1^{(0)}<0$
713: (see Proposition \ref{lem} for the definition of $w_0$). The eigenvectors
714: corresponding to the eigenvalues $\lambda_1^{(0)}$ and $\lambda_2^{(0)}$ are:
715: \begin{equation*}
716: e_1^{(0)}=\begin{pmatrix}
717: f(0)+3 \, w_0\\
718: b_0 \end{pmatrix} \, ,\quad e_2^{(0)}=\begin{pmatrix}
719: 0\\
720: 1 \end{pmatrix} \, .
721: \end{equation*}
722: The standard theory of autonomous ODEs, see e.g. \cite{pontriaguine}, shows that
723: there are exactly two solutions to \eqref{ode-ref} that tend to $(0,w_0)$ as
724: $\eta$ tends to $+\infty$, and that are tangent to the straight line $(0,w_0)
725: +\R \, e_2^{(0)}$. Moreover, all the other solutions to \eqref{ode-ref} that
726: tend to $(0,w_0)$ as $\eta$ tends to $+\infty$ are tangent to the straight line
727: $(0,w_0) +\R \, e_1^{(0)}$. Now, it is rather simple to see that the two solutions
728: to \eqref{ode-ref} that tend to $(0,w_0)$ as $\eta$ tends to $+\infty$, and that
729: are tangent to the straight line $(0,w_0) +\R \, e_2^{(0)}$, satisfy
730: $V \equiv 0$, and:
731: \begin{equation*}
732: W'=-W^2-f(0) \, W -\dfrac{a^2}{2} \, .
733: \end{equation*}
734: Because the solutions $(V_\flat,W_\flat)$, and $(V_\sharp,W_\sharp)$ given by
735: Proposition \ref{lem} cannot satisfy $V_\flat \equiv 0$, and $V_\sharp \equiv 0$,
736: we can conclude that the solutions $(V_\flat,W_\flat)$, and $(V_\sharp,W_\sharp)$
737: are tangent to $(0,w_0) +\R \, e_1^{(0)}$ as $\eta$ tends to $+\infty$. In
738: particular, this yields:
739: \begin{equation}
740: \label{asympderiv}
741: \lim_{\eta \rightarrow +\infty} \dfrac{W_\flat' (\eta)}{V_\flat' (\eta)}
742: =\lim_{\eta \rightarrow +\infty} \dfrac{W_\sharp' (\eta)}{V_\sharp' (\eta)}
743: =\dfrac{-f'(0) \, w_0}{f(0)+3 \, w_0} \, .
744: \end{equation}
745: (A quick verification shows that $f(0)+3 \, w_0>0$ for small enough $a$.)
746: From the construction of the solutions $(\hat{v}_\flat,w_\flat)$, and
747: $(\hat{v}_\sharp,w_\sharp)$, we get:
748: \begin{equation*}
749: \lim_{\xi \rightarrow 0^-} w_\flat' (\xi)
750: =\lim_{\xi \rightarrow 0^+} w_\sharp' (\xi)
751: =\dfrac{-f'(0) \, w_0^2}{f(0)+3 \, w_0} \, .
752: \end{equation*}
753: As a consequence, when $a$ is small enough, $w \in C^1(\R)$, and therefore
754: $\hat{v} \in C^2(\R)$. Moreover, $(\hat{v},w)$ solves \eqref{ode} on
755: $\R \setminus \{ 0\}$, so by continuity, it solves \eqref{ode} on $\R$.
756: This completes the proof of Theorem \ref{exis}, provided that the result
757: of Proposition \ref{lem} holds.
758:
759:
760:
761:
762:
763:
764: \subsection{Proof of Proposition \ref{lem}}
765:
766:
767:
768: In this paragraph, we prove Proposition \ref{lem}, which will complete the
769: proof of Theorem \ref{exis}. At first, we define the set:
770: $$
771: P = \Big\{ (V,W) | V \in [-a,a], W^2+f(V) \, W-\dfrac{V^2-a^2}{2}=0 \Big\} \, ,
772: $$
773: so that the points $(\pm a,0)$ belong to $P$. The following Lemma gives a
774: description of $P$ for $a>0$ small enough. We refer to figure \ref{K} for
775: a schematic picture.
776:
777:
778: \begin{figure}
779: \centering
780: \psfrag{A}{$a$}
781: \psfrag{B}{$-a$}
782: \psfrag{C}{$V$}
783: \psfrag{D}{$W$}
784: \psfrag{E}{$\overline{V}$}
785: \psfrag{P1}{$P_1$}
786: \psfrag{P2}{$P_2$}
787: \psfrag{w0}{$w_0$}
788: \includegraphics{EnsembleP.eps}
789: \caption{The set $P =P_1 \cup P_2$.}
790: \label{K}
791: \end{figure}
792:
793:
794: \begin{lemm}
795: \label{geo}
796: Assume that $1 < \gamma < (\sqrt{7}+1)/(\sqrt{7}-1)$. Then there exists
797: a constant $a_0>0$, that only depends on $(\rho_-,u_-,e_-)$ and $\gamma$
798: such that if the shock strength $a$ satisfies $a\in (0,a_0]$, we have the
799: following results:
800: \begin{itemize}
801: \item For all $V \in [-a,a]$, $f(V)^2+2\, (V^2-a^2)>0$. We can
802: thus define
803: \begin{align*}
804: \forall \, V \in [-a,a] \, ,\quad
805: \W_1(V) &= \dfrac{-f(V)+\sqrt{f(V)^2 +2\, (V^2-a^2)}}{2} \, ,\\
806: \W_2(V) &= \dfrac{-f(V)-\sqrt{f(V)^2 +2\, (V^2-a^2)}}{2} \, .
807: \end{align*}
808:
809: \item $P = P_1\cup P_2$, where $P_1$ and $P_2$ are two curves defined by
810: $$
811: P_1 =\big\{ (V,\W_1(V)) | V \in [-a,a] \big\} \, , \quad
812: P_2 =\big\{ (V,\W_2(V)) | V \in [-a,a] \big\} \, ,
813: $$
814: so that the points $(\pm a,0)$, and $(0,w_0)$ belong to $P_1$.
815:
816: \item There exists a unique point $\overline{V} \in \, (-a,0)$ such that
817: $\W_1$ is increasing on the interval $[\overline{V},a]$, and $\W_1$ is
818: decreasing on the interval $[-a,\overline{V}]$.
819:
820: \item For all $V \in [\overline{V},0]$, one has
821: $\W_2(V) < \W_1(\overline{V})$.
822: \end{itemize}
823: \end{lemm}
824:
825:
826: \begin{proof}
827: Let us first define a function $\Delta$ by setting:
828: \begin{equation*}
829: \Delta (V) := f(V)^2 +2 \, (V^2-a^2) \, .
830: \end{equation*}
831: Using \eqref{asymp}, for $a$ small enough, we have:
832: \begin{equation*}
833: \dfrac{C_1}{\gamma +1} -a \ge \kappa >0 \, ,\quad
834: \dfrac{\gamma \, C_1}{\gamma +1} -a \ge \kappa >0 \, ,\quad
835: \dfrac{(\gamma -1)\, C_1}{\gamma +1} -2a \ge \kappa >0 \, ,
836: \end{equation*}
837: where $\kappa$ is a positive constant that only depends on $(\rho_-,u_-,e_-)$
838: and $\gamma$. Moreover, \eqref{asymp} also shows that $j \ge \kappa$ for
839: $a \in (0,a_0]$, up to restricting $\kappa$. Consequently, there exist
840: $a_0>0$, and $\kappa>0$ such that for $a \in (0,a_0]$, we have $f(V) \ge \kappa$,
841: and $\Delta (V) \ge \kappa$ for all $V\in [-a,a]$. This directly shows that
842: the set $P$ is the union of the two curves $P_1$, and $P_2$. It is rather
843: clear from the definition of $P_1$ that $(\pm a,0)$, and $(0,w_0)$ belong
844: to $P_1$ (recall that $w_0$ is defined in Proposition \ref{lem}). Observe
845: also that $\W_2(V) <\W_1 (V) \le 0$ for $V \in [-a,a]$, and $\W_1(V)<0$ if
846: $V \in (-a,a)$.
847:
848:
849: The functions $\W_1$, and $\W_2$ are $C^\infty$ on $[-a,a]$. Moreover, we
850: compute the relation:
851: \begin{equation}
852: \label{derivW1}
853: \forall \, V \in [-a,a] \, ,\quad
854: \sqrt{\Delta (V)} \, \W_1'(V) =V-\W_1(V) \, f'(V) \, ,
855: \end{equation}
856: and from \eqref{fun f}, we also compute
857: \begin{multline}
858: \label{derivf}
859: f'(V)=\dfrac{14 \, (\gamma-1)}{j \, R^4 \, (\gamma+1)} \\
860: \left( \dfrac{C_1}{\gamma+1}-\hat{v} \right)^2 \,
861: \left( \hat{v}+\dfrac{\gamma\, C_1}{\gamma+1} \right)^2 \,
862: \left( 2\hat{v}+\dfrac{(\gamma-1) \, C_1}{\gamma+1} +\dfrac{C_1}{\sqrt{7}} \right)
863: \left( -2\hat{v}-\dfrac{(\gamma-1) \, C_1}{\gamma+1} +\dfrac{C_1}{\sqrt{7}} \right) \, .
864: \end{multline}
865: As we have done for $f$, and $\Delta$, a careful analysis shows that for
866: $1<\gamma<(\sqrt{7}+1)/(\sqrt{7}-1)$, and for $a$ small enough, one has
867: $f'(V) \ge \kappa>0$ for all $V \in [-a,a]$, because each term in the product
868: \eqref{derivf} is positive. Using this information in \eqref{derivW1}, we can
869: already conclude that $\W_1$ is increasing on the interval $[0,a]$ (see figure
870: \ref{K}). Moreover, the relation \eqref{derivW1} also shows that $\W'_1(0)>0$,
871: and $\W'_1(-a)<0$. Consequently, there exists some $\overline{V} \in (-a,0)$
872: such that $\W_1'(\overline{V})=0$. Let us prove that $\overline{V}$ is the only
873: zero of $\W_1'$. We claim that it is sufficient to show the following property:
874: \begin{equation}
875: \label{property}
876: \W_1' (V) =0 \Longrightarrow \W_1'' (V) >0 \, .
877: \end{equation}
878: Indeed, if the property \eqref{property} holds true, then any point where $\W_1'$
879: vanishes is a local strict minimum. If there existed two such local strict minima
880: $-a<\overline{V}_1<\overline{V}_2<a$, then $\W_1$ would admit a local maximum
881: $\overline{V}_3 \in (\overline{V}_1,\overline{V}_2)$, which is obviously
882: impossible. Therefore let us prove that the property \eqref{property} holds
883: true.
884:
885:
886: Differentiating \eqref{derivW1} with respect to $V$, we obtain that if
887: $\W_1'(\overline{V})=0$, then
888: $$
889: \sqrt{\Delta (\overline{V})} \, \W_1''(\overline{V}) = 1 -f''(\overline{V})
890: \, \W_1(\overline{V}) \, .
891: $$
892: Observing that
893: \begin{equation*}
894: |f''(\overline{V}) \, \W_1(\overline{V})| \le C \, |\W_1(\overline{V})|
895: \le C \, \dfrac{a^2 -\overline{V}^2}{f(\overline{V})} \le
896: \dfrac{C \, a^2}{\kappa} \, ,
897: \end{equation*}
898: for suitable positive constants $C$, and $\kappa$ (that are independent
899: of $a \in (0,a_0]$), we can conclude that $\W_1''(\overline{V})>0$, provided
900: that $a$ is small enough. This completes the proof that $\W_1'$ has a unique
901: zero $\overline{V} \in (-a,0)$, and therefore $\W_1$ is decreasing on
902: $[-a,\overline{V}]$, and is increasing on $[\overline{V},a]$.
903:
904:
905: For the last point of the lemma, we use the relation:
906: $$
907: \W_1'(V) +\W_2'(V)=-f'(V)<0 \, .
908: $$
909: Because $\W_1'(V)\ge 0$ for $V \in [\overline{V},0]$, we get $\W_2'(V)<0$
910: for $V \in [\overline{V},0]$. Thus for $V \in [\overline{V},0]$, we have
911: $\W_2(V) \le \W_2 (\overline{V}) <\W_1 (\overline{V})$, and the proof of
912: the Lemma is complete.
913: \end{proof}
914:
915:
916: Using Lemma \ref{geo}, we are going to prove Proposition \ref{lem}.
917: The analysis follows \cite{gilbarg}.
918:
919:
920: As we have already seen in the preceeding paragraph, the point $(w_0,0)$ is
921: a stable node of \eqref{ode-ref}. We now study the nature of the equilibrium
922: points $(\pm a,0)$. Recall that the vector field associated to \eqref{ode-ref}
923: is denoted $U$, see \eqref{defU}. The Jacobian matrix of $U$ at $(a,0)$ is:
924: \begin{equation*}
925: \begin{pmatrix}
926: 0 & a \\
927: a & f(a) \end{pmatrix} \, ,
928: \end{equation*}
929: so it has exactly one negative eigenvalue $\mu_1$, and one positive eigenvalue
930: $\mu_2$ (the equilibrium point $(a,0)$ is a saddle point):
931: \begin{equation*}
932: \mu_1 =\dfrac{-f(a)-\sqrt{f(a)^2+4 \, a^2}}{2} \, ,\quad
933: \mu_2 =\dfrac{-f(a)+\sqrt{f(a)^2+4 \, a^2}}{2} \, .
934: \end{equation*}
935: An eigenvector associated to $\mu_2$, and is $r_2=(a,\mu_2)$. Moreover,
936: using the relation \eqref{derivW1}, we can check that for $a$ small enough,
937: the following inequality holds:
938: \begin{equation}
939: \label{ineg1}
940: 0 <\dfrac{\mu_2}{a} < \dfrac{a}{f(a)} =\W_1'(a) \, ,
941: \end{equation}
942: where the function $\W_1$ is defined in Lemma \ref{geo}. Let us now define a
943: compact set $K_1$ by:
944: \begin{equation*}
945: K_1 := \Big\{ (V,W) \in [0,a] \times \R \, | \, \W_1 (V) \le W \le 0 \Big\} \, ,
946: \end{equation*}
947: Then the inequalities \eqref{ineg1} show that for $s<0$ small enough, the
948: point $(a,0) + s \, r_2$ belongs to the interior of $K_1$. We refer to
949: figure \ref{compactK1} for a detailed picture of the situation.
950:
951:
952: \begin{figure}
953: \centering
954: \psfrag{$W$}{$W$}
955: \psfrag{$V$}{$V$}
956: \psfrag{$w_0$}{$w_0$}
957: \psfrag{$a$}{$a$}
958: \psfrag{$K_1$}{$K_1$}
959: \psfrag{A}{$(a,0)+{\mathbb R} \, r_2$}
960: \includegraphics{compactK1.eps}
961: \caption{The compact set $K_1$.}
962: \label{compactK1}
963: \end{figure}
964:
965:
966: From the standard theory of autonomous ODEs, see e.g. \cite{pontriaguine}, we
967: know that there exists a maximal solution $(V_\flat,W_\flat)$ to \eqref{ode-ref}
968: that tends to the saddle point $(a,0)$ as $\eta$ tends to $-\infty$, and that
969: is tangent to the half-straight line $(a,0) +\R^- \, r_2$. This solution is
970: defined on an open interval $(-\infty,\eta_*)$ (with possibly $\eta_*=+\infty$).
971: For large negative $\eta$, the preceeding analysis shows that
972: $(V_\flat,W_\flat) (\eta)$ belongs to the interior of $K_1$. Moreover,
973: $(V_\flat,W_\flat)$ cannot reach the boundary of $K_1$. Indeed $V_\flat$ cannot
974: identically vanish so $(V_\flat,W_\flat)(\eta) \not \in \partial K_1 \cap
975: \{ V=0\}$. Similarly, we have $(V_\flat,W_\flat)(\eta) \neq (a,0)$. Eventually,
976: on the set:
977: \begin{equation*}
978: \Big\{ (V,0) \, | \, V \in (0,a) \Big\} \cup
979: \Big\{ (V,\W_1(V)) \, | \, V \in \, (0,a) \Big\} \, ,
980: \end{equation*}
981: the vector field $U$ is not zero, and is directed towards the interior of
982: $K_1$. Therefore the solution $(V_\flat,W_\flat)$ cannot reach $\partial K_1$,
983: so it takes its values in the compact set $K_1$. The maximal solution
984: $(V_\flat,W_\flat)$ is thus defined on $\R$. It cannot reach the boundary
985: of $K_1$, so $W_\flat$ takes negative values, which means that $V_\flat$ is
986: decreasing (because $V_\flat'=V_\flat \, W_\flat$). Because $(V_\flat,W_\flat)$
987: takes values in the interior of $K_1$, the function $W_\flat$ is also decreasing.
988: This shows that $(V_\flat,W_\flat)(\eta)$ has a limit as $\eta$ tends to $+\infty$,
989: and this limit is necessarily be a stationary solution of \eqref{ode-ref}. The
990: only possibility is that $(V_\flat,W_\flat)(\eta)$ tends to $(0,w_0)$ as $\eta$
991: tends to $+\infty$. The convergence is necessarily exponential, because the
992: Jacobian matrix of $U$ at $(0,w_0)$ has two negative eigenvalues, see e.g.
993: \cite{pontriaguine}.
994:
995:
996: To construct the other solution $(V_\sharp,W_\sharp)$, we argue similarly by defining
997: a compact set $K_2$:
998: \begin{equation*}
999: K_2 := \Big\{ (V,W) \in [-a,\overline{V}] \times \R | \W_1 (V) \le W \le 0 \Big\}
1000: \cup \Big\{
1001: (V,W) \in [\overline{V},0] \times \R | \W_1 (\overline{V}) \le W \le 0 \Big\} \, ,
1002: \end{equation*}
1003: see figure \ref{compactK2}. The Jacobian matrix of $U$ at $(-a,0)$ has one
1004: negative eigenvalue $\nu_1$, and one positive eigenvalue $\nu_2$, with:
1005: \begin{equation*}
1006: \nu_2 = \dfrac{-f(-a)+\sqrt{f(-a)^2+4 \, a^2}}{2} \, .
1007: \end{equation*}
1008: An eigenvector associated to the eigenvalue $\nu_2$ is $R_2=(-a,\nu_2)$.
1009: As was done earlier, we check that the inequalities:
1010: \begin{equation*}
1011: \W_1'(-a)=\dfrac{-a}{f(-a)} < \dfrac{\nu_2}{-a} <0 \, ,
1012: \end{equation*}
1013: hold true. Therefore, one can reproduce the above analysis, and show that there
1014: exists a solution $(V_\sharp,W_\sharp)$ to \eqref{ode-ref} that takes its values
1015: in $K_2$ (and is thus defined on $\R$), and that tends to $(-a,0)$ at $-\infty$.
1016: Moreover, $(V_\sharp,W_\sharp)$ can not reach the boundary of $K_2$, so $V_\sharp$
1017: is increasing. It only remains to study the monotonicity of $W_\sharp$. This is
1018: slightly more complicated than for $W_\flat$. Observe that $K_2$ is the union of
1019: the sets:
1020: \begin{align*}
1021: K_2^1 &:=\Big\{ (V,W) \in [-a,0] \times \R \, | \, \W_1 (V) \le W \le 0 \Big\} \, ,\\
1022: K_2^2 &:=\Big\{ (V,W) \in [\overline{V},0] \times \R \, | \,
1023: \W_1 (\overline{V}) \le W \le \W_1(V) \Big\} \, .
1024: \end{align*}
1025: When $(V_\sharp,W_\sharp)$ takes its values in the interior of $K_2^1$, the function
1026: $W_\sharp$ is decreasing (this is the case for large negative $\eta$). At the opposite,
1027: when $(V_\sharp,W_\sharp)$ takes its values in the interior of $K_2^2$, the function
1028: $W_\sharp$ is increasing, because thanks to Lemma \ref{geo}, we have:
1029: \begin{align*}
1030: W_\sharp'(\eta)
1031: &=-W_\sharp(\eta)^2 -f(V_\sharp(\eta)) \, W_\sharp(\eta)
1032: +\dfrac{V_\sharp(\eta)^2 -a^2}{2} \\
1033: &=\big(
1034: \W_1 (V_\sharp(\eta)) -W_\sharp (\eta) \big) \,
1035: \big( W_\sharp(\eta) -\W_2 (V_\sharp(\eta)) \big) \\
1036: &\ge \big( \W_1 (V_\sharp(\eta)) -W_\sharp (\eta) \big) \,
1037: \big( \W_1(\overline{V}) -\W_2 (V_\sharp(\eta)) \big) >0 \, .
1038: \end{align*}
1039: Moreover, if $(V_\sharp,W_\sharp) (\eta_0)$ belongs to the interior of $K_2^2$ for
1040: some $\eta_0 \in \R$, then $(V_\sharp,W_\sharp) (\eta)$ belongs to the interior of
1041: $K_2^2$ for all $\eta \ge \eta_0$ (because it cannot reach the boundary of $K_2^2$
1042: for $\eta \ge \eta_0$). Summing up, either $(V_\sharp,W_\sharp)(\eta)$ belongs to
1043: $K_2^1$ for all $\eta$, and $W_\sharp$ is monotonic on $\R$, either
1044: $(V_\sharp,W_\sharp)(\eta)$ belongs to $K_2^2$ for all $\eta$ greater than some
1045: $\eta_0$, and $W_\sharp$ is monotonic on $[\eta_0,+\infty)$. In any case, the function
1046: $W_\sharp$ is monotonic on a neighborhood of $+\infty$, and thus has a limit at
1047: $+\infty$. This shows that $(V_\sharp,W_\sharp)(\eta)$ tends to $(0,w_0)$ as $\eta$
1048: tends to $+\infty$, and the convergence is exponential. As a matter of fact, we
1049: have seen in the preceeding paragraph that $(V_\sharp,W_\sharp)$ is tangent to the
1050: straight line $(0,w_0)+\R \, e_1^{(0)}$ as $\eta$ tends to $+\infty$, so one can
1051: check that $(V_\sharp,W_\sharp) (\eta)$ belongs to the interior of $K_2^2$ for
1052: large positive $\eta$. This means that $W_\sharp$ is decreasing on some interval
1053: $(-\infty,\eta_0)$, and increasing on $[\eta_0,+\infty)$. The proof of
1054: Proposition \ref{lem} is now complete.
1055:
1056:
1057: \begin{figure}
1058: \centering
1059: \psfrag{$W$}{$W$}
1060: \psfrag{$V$}{$V$}
1061: \psfrag{$K_2$}{$K_2$}
1062: \psfrag{$-a$}{$-a$}
1063: \psfrag{A}{$\overline{V}$}
1064: \psfrag{B}{$W_1(\overline{V})$}
1065: \psfrag{C}{$(-a,0)+{\mathbb R} \, R_2$}
1066: \includegraphics{compactK2.eps}
1067: \caption{The compact set $K_2$.}
1068: \label{compactK2}
1069: \end{figure}
1070:
1071:
1072:
1073:
1074:
1075:
1076: \section{Additional regularity of shock profiles}
1077: \label{moresmooth}
1078:
1079:
1080:
1081: As should be clear from the preceeding section, the key point in the construction
1082: of a shock profile is Proposition \ref{lem} that gives the existence of two
1083: heteroclinic orbits for the system \eqref{ode-ref}. To prove Theorem \ref{smooth},
1084: we are going to study the behavior of the derivatives of $(V_\flat,W_\flat)$, and
1085: $(V_\sharp,W_\sharp)$ near $+\infty$. The proof of Theorem \ref{smooth} follows
1086: from an induction argument. To make the arguments clear, we deal with the first
1087: case separately. In all what follows, $(V_\flat,W_\flat)$, and $(V_\sharp,W_\sharp)$
1088: are the solutions to \eqref{ode-ref} that are defined in Proposition \ref{lem},
1089: and $(\hat{v},w)$ denotes the solution to \eqref{ode} that is defined by
1090: \eqref{defsolution}. We have the following:
1091:
1092:
1093: \begin{prop}
1094: \label{n=3}
1095: Under the assumptions of Proposition \ref{lem}, there exists a positive constant
1096: $a_1 \le a_0$ (that depends on $(\rho_-,u_-,e_-)$, and $\gamma$), such that for
1097: all $a \in (0,a_1]$, one has $w \in C^2 (\R)$, $\hat{v} \in C^3 (\R)$, and:
1098: \begin{equation*}
1099: w(\xi)=w_0 +w_1 \, \hat{v} (\xi) +w_2 \, \hat{v} (\xi)^2 +o(\hat{v} (\xi)^2) \, ,\quad
1100: \text{\rm as $\xi \rightarrow 0$,}
1101: \end{equation*}
1102: for some suitable constants $w_1,w_2$ ($w_0$ has already been defined in Proposition
1103: \ref{lem}).
1104: \end{prop}
1105:
1106:
1107: \begin{proof}
1108: Recall that $V_b$, and $V_\sharp$ do not vanish on $\R$, so we can introduce
1109: some $C^\infty$ functions $W_{\flat,1}$, and $W_{\sharp,1}$ that are defined by:
1110: \begin{equation*}
1111: W_\flat = w_0 + V_\flat \, W_{\flat,1} \, ,\quad
1112: W_\sharp = w_0 + V_\sharp \, W_{\sharp,1} \, .
1113: \end{equation*}
1114: Substituting in \eqref{ode-ref} shows that $(V_\flat,W_{\flat,1})$, and
1115: $(V_\sharp,W_{\sharp,1})$ are solutions to the system:
1116: \begin{equation}
1117: \label{w_1}
1118: \begin{cases}
1119: V' = V (w_0+V \, W_1) \, , &\\
1120: W_1' = -w_0 \, \dfrac{f(V)-f(0)}{V} -2 \, V \, W_1^2 -3 \, w_0 \, W_1
1121: -f(V) \, W_1 +\dfrac{V}{2} \, . &
1122: \end{cases}
1123: \end{equation}
1124: Moreover, we already know from Proposition \ref{lem}, and \eqref{asympderiv}
1125: that:
1126: \begin{equation*}
1127: \lim_{\eta \rightarrow +\infty} (V_\flat,W_{\flat,1}) (\eta)
1128: =\lim_{\eta \rightarrow +\infty} (V_\sharp,W_{\sharp,1}) (\eta)
1129: =\Big( 0,\dfrac{-f'(0) \, w_0}{f(0)+3 \, w_0} \Big) \, .
1130: \end{equation*}
1131:
1132:
1133: We denote $U_1 (V,W_1)$ the vector field associated with \eqref{w_1}:
1134: \begin{equation*}
1135: U_1 (V,W_1) := \begin{pmatrix}
1136: V (w_0+V \, W_1) \\
1137: -w_0 \, \dfrac{f(V)-f(0)}{V} -2 \, V \, W_1^2 -3 \, w_0 \, W_1
1138: -f(V) \, W_1 +\dfrac{V}{2} \end{pmatrix} \, .
1139: \end{equation*}
1140: Recall that $f$ is a polynomial function of degree $7$, see \eqref{fun f},
1141: thus $F(V) :=(f(V)-f(0))/V$ is a polynomial function of degree $6$, and we have
1142: $F(0)=f'(0)$, $F'(0)=f''(0)/2$. Obviously the system of ODEs \eqref{w_1} admits
1143: the equilibrium point $(0,w_1)$, where:
1144: \begin{equation*}
1145: w_1 :=\dfrac{-F(0) \, w_0}{f(0)+3\, w_0} =\dfrac{-f'(0) \, w_0}{f(0)+3\, w_0} \, .
1146: \end{equation*}
1147: We are now going to study the nature of the equilibrium point $(0,w_1)$, and show
1148: that for $a$ small enough, this equilibrium point is a stable node for \eqref{w_1}.
1149: Then we shall show that $w \in C^2 (\R)$, and $\hat{v} \in C^3 (\R)$. In the end,
1150: we shall derive the asymptotic expansion near $\xi=0$.
1151:
1152:
1153: \underline{Step 1}: the Jacobian matrix of $U_1$ at $(0,w_1)$ is:
1154: \begin{equation*}
1155: \begin{pmatrix}
1156: w_0 & 0\\
1157: \dfrac{1}{2} -2\, w_1^{2} -f'(0)\, w_1 -\dfrac{f''(0)}{2} \, w_0 & -f(0)-3\, w_0
1158: \end{pmatrix} =\begin{pmatrix}
1159: \lambda_1^{(1)} & 0 \\
1160: b_1 & \lambda_2^{(1)} \end{pmatrix} \, .
1161: \end{equation*}
1162: Using Remak \ref{f0}, we can conclude that for sufficiently small $a$, that
1163: is $a \in (0,a_1]$ for some positive number $a_1$ less than $a_0$, one has
1164: $\lambda_2^{(1)} <\lambda_1^{(1)}<0$, that is, $f(0)+4\, w_0>0$. Moreover,
1165: the eigenvectors corresponding to the eigenvalues $\lambda_1^{(1)}$, and
1166: $\lambda_2^{(1)}$ are:
1167: \begin{equation*}
1168: e_1^{(1)} =\begin{pmatrix}
1169: f(0) +4\, w_0\\
1170: b_1 \end{pmatrix} \, ,\quad
1171: e_2^{(1)} =\begin{pmatrix}
1172: 0\\
1173: 1 \end{pmatrix} \, .
1174: \end{equation*}
1175: Consequently, $(0,w_1)$ is a stable node of \eqref{w_1}, and there are exactly
1176: two solutions to \eqref{w_1} that tend to $(0,w_1)$ as $\eta$ tends to $+\infty$,
1177: and that are tangent to the straight line $(0,w_1) +\R \, e_2^{(1)}$. All the other
1178: solutions to \eqref{w_1} that tend to $(0,w_1)$ as $\eta$ tends to $+\infty$ are
1179: tangent to the straight line $(0,w_1) +\R \, e_1^{(1)}$. As in the preceeding
1180: section, we can thus conclude that:
1181: \begin{equation}
1182: \label{eq:0}
1183: \lim_{\eta \rightarrow +\infty} \dfrac{W_{\flat,1}'(\eta)}{V_\flat'(\eta)}
1184: =\lim_{\eta \rightarrow +\infty} \dfrac{W_{\sharp,1}'(\eta)}{V_\sharp'(\eta)}
1185: =\dfrac{b_1}{f(0)+4\, w_0} \, .
1186: \end{equation}
1187:
1188:
1189: \underline{Step 2}: if we let $g_1$ denote the second coordinate of
1190: the vector field $U_1$, we have $W_{\flat,1}'=g_1(V_\flat,W_{\flat,1})$, and
1191: $W_{\sharp,1}'=g_1(V_\sharp,W_{\sharp,1})$. Differentiating once with respect
1192: to $\eta$, and using \eqref{eq:0}, we end up with:
1193: \begin{equation}
1194: \label{eq:1}
1195: \lim_{\eta \rightarrow +\infty} \dfrac{W_{\flat,1}''(\eta)}{V_\flat'(\eta)}
1196: =\lim_{\eta \rightarrow +\infty} \dfrac{W_{\sharp,1}''(\eta)}{V_\sharp'(\eta)}
1197: = \ell_1 \, ,
1198: \end{equation}
1199: where the real number $\ell_1$ can be explicitely computed (but its exact expression
1200: is of no use). Following the analysis of the preceeding section, we define some
1201: functions $w_{\flat,1} := W_{\flat,1} \circ \Xi_\flat^{-1}$, and $w_{\sharp,1}
1202: := W_{\sharp,1} \circ \Xi_\sharp^{-1}$. First of all, \eqref{eq:0} yields:
1203: \begin{equation}
1204: \label{eq:2}
1205: \lim_{\xi \rightarrow 0^-} w_{\flat,1}' (\xi)
1206: =\lim_{\xi \rightarrow 0^+} w_{\sharp,1}' (\xi)
1207: =\dfrac{b_1 \, w_0}{f(0)+4\, w_0} \, .
1208: \end{equation}
1209: Observe now that we have the relations:
1210: \begin{equation*}
1211: \hat{v}_\flat \, w_{\flat,1}' =W_{\flat,1}' \circ \Xi_\flat^{-1} \, ,\quad
1212: \hat{v}_\sharp \, w_{\sharp,1}' =W_{\sharp,1}' \circ \Xi_\sharp^{-1} \, ,
1213: \end{equation*}
1214: and combining with \eqref{eq:1}, we get:
1215: \begin{equation}
1216: \label{eq:3}
1217: \begin{split}
1218: \lim_{\xi \rightarrow 0^-} (\hat{v}_\flat \, w_{\flat,1}')' (\xi)
1219: =\lim_{\eta \rightarrow +\infty} \dfrac{W_{\flat,1}''(\eta)}{V_\flat (\eta)}
1220: =\lim_{\eta \rightarrow +\infty}
1221: \dfrac{W_{\flat,1}''(\eta) \, W_\flat (\eta)}{V_\flat' (\eta)} =\ell_1 \, w_0 \, ,\\
1222: \lim_{\xi \rightarrow 0^+} (\hat{v}_\sharp \, w_{\sharp,1}')' (\xi)
1223: =\lim_{\eta \rightarrow +\infty} \dfrac{W_{\sharp,1}''(\eta)}{V_\sharp (\eta)}
1224: =\lim_{\eta \rightarrow +\infty}
1225: \dfrac{W_{\sharp,1}''(\eta) \, W_\sharp (\eta)}{V_\sharp' (\eta)} =\ell_1 \, w_0 \, .
1226: \end{split}
1227: \end{equation}
1228: Differentiating twice the relations $w_\flat =w_0 +\hat{v}_\flat \, w_{\flat,1}$,
1229: and $w_\sharp =w_0 +\hat{v}_\sharp \, w_{\sharp,1}$, we obtain:
1230: \begin{gather*}
1231: w_\flat'' =\hat{v}_\flat'' \, w_{\flat,1} +\hat{v}_\flat' \, w_{\flat,1}'
1232: +(\hat{v}_\flat \, w_{\flat,1}')' =w_\flat' \, w_{\flat,1}
1233: +w_\flat \, w_{\flat,1}' +(\hat{v}_\flat \, w_{\flat,1}')' \, ,\\
1234: w_\sharp'' =w_\sharp' \, w_{\sharp,1}
1235: +w_\sharp \, w_{\sharp,1}' +(\hat{v}_\sharp \, w_{\sharp,1}')' \, .
1236: \end{gather*}
1237: Using \eqref{eq:2}, and \eqref{eq:3}, we get $w_\flat'' (0^-)=w_\sharp'' (0^+)$.
1238: Using the definition \eqref{defsolution}, this shows that $w \in C^2 (\R)$, and
1239: using $\hat{v}'=w$, we obtain $\hat{v} \in C^3 (\R)$.
1240:
1241:
1242: \underline{Step 3}: note that we have the following expansions near $\xi=0$:
1243: \begin{align*}
1244: &w(\xi) = w(0) +w'(0) \, \xi +\dfrac{w''(0)}{2} \, \xi^2 +o(\xi^2) \, ,\\
1245: &\hat{v}(\xi) = w(0) \, \xi +\dfrac{w'(0)}{2} \, \xi^2 +o(\xi^2) \, ,
1246: \end{align*}
1247: with $w(0)=w_0 <0$. We can thus combine these expansions, and derive:
1248: \begin{equation*}
1249: w(\xi) = w_0 +\alpha \, \hat{v}(\xi) +\beta \, \hat{v}(\xi)^2
1250: +o(\hat{v}(\xi)^2) \, ,
1251: \end{equation*}
1252: for some appropriate real numbers $\alpha$, and $\beta$, that we are going
1253: to determine. From the relation $w_\flat (\xi)=w_0+\hat{v}_\flat(\xi) \,
1254: w_{\flat,1} (\xi)$, and using that $w_{\flat,1} (\xi)$ tends to $w_1$ as
1255: $\xi$ tends to $0^-$, we first get $\alpha=w_1$. Then from \eqref{eq:2},
1256: and from the relation $\hat{v}_\flat'(0^-)=w_0$, we can obtain:
1257: \begin{equation*}
1258: w_{\flat,1} (\xi)=w_1 +\dfrac{b_1}{f(0)+4\, w_0} \, \hat{v}_\flat (\xi)
1259: +o(\hat{v}_\flat (\xi)) \, ,\quad \text{as $\xi \rightarrow 0^-$.}
1260: \end{equation*}
1261: We thus obtain $\beta=b_1/(f(0)+4\, w_0)$, which yields:
1262: \begin{equation*}
1263: w(\xi) = w_0 +w_1 \, \hat{v}(\xi) +w_2 \, \hat{v}(\xi)^2 +o(v(\xi)^2) \, ,
1264: \end{equation*}
1265: where $w_2 :=b_1/(f(0)+4\, w_0)$. This latter expansion will be generalized
1266: to any order in what follows.
1267: \end{proof}
1268:
1269:
1270: We now turn to the proof of Theorem \ref{smooth}. More precisely, we are going
1271: to prove the following result, that is a refined version of Theorem \ref{smooth}:
1272:
1273:
1274: \begin{thh}
1275: \label{smooth2}
1276: Let the assumptions of Proposition \ref{lem} be satisfied. Then there exists a
1277: nonincreasing sequence of positive numbers $(a_n)_{n \in \N}$ such that, for
1278: all integer $n$, if $a \in (0,a_n]$, then $w \in C^{n+1} (\R)$, and $\hat{v}
1279: \in C^{n+2} (\R)$. Moreover, $w$ admits the following asymptotic expansion
1280: near $\xi=0$:
1281: \begin{equation}
1282: \label{expansion}
1283: w(\xi) =w_0+w_1 \, \hat{v}(\xi) +\cdots +w_{n+1} \, \hat{v} (\xi)^{n+1}
1284: +o(\hat{v} (\xi)^{n+1}) \, ,
1285: \end{equation}
1286: where the real numbers $w_0,\dots,w_{n+1}$ are defined by:
1287: \begin{equation*}
1288: \begin{cases}
1289: w_0 =\dfrac{-f(0)+\sqrt{f(0)^2 -2\, a^2}}{2} \, ,& \\
1290: w_k =\dfrac{b_{k-1}}{f(0) +(k+2)\, w_0} \, ,&\text{\rm for $k=1,\dots,n+1$,}
1291: \end{cases}
1292: \end{equation*}
1293: and the real numbers $b_0,\dots,b_n$ are given by:
1294: \begin{equation*}
1295: \begin{cases}
1296: b_0 = -f'(0) \, w_0 \, ,& \\
1297: b_1 =\dfrac{1}{2}-2 \, w_1^2 -f'(0) \, w_1 -\dfrac{f''(0)}{2} \, w_0 \, ,& \\
1298: b_k =-\displaystyle \sum_{i=1}^{k+1} \dfrac{f^{(i)}(0)}{i!} \, w_{k+1-i}
1299: -\displaystyle \sum_{i=1}^k (i+1) \, w_i \, w_{k+1-i} \, ,&
1300: \text{\rm for $k=2,\dots,n$.}
1301: \end{cases}
1302: \end{equation*}
1303: \end{thh}
1304:
1305:
1306: \begin{proof}
1307: The case $n=0$ has been proved in the preceeding section, while the case $n=1$
1308: is proved in Proposition \ref{n=3}. (The reader can check that the definition of
1309: $w_0$, $w_1$, $w_2$, $b_0$, and $b_1$ coincide with our previous notations.) We
1310: prove the general case by using an induction with respect to $n$, and we thus
1311: assume that the result of Theorem \ref{smooth2} holds up to the order $n \ge 1$.
1312: We are going to construct $a_{n+1}$ so that the conclusion of Theorem \ref{smooth2}
1313: holds for $a \in (0,a_{n+1}]$. In particular, the real numbers $w_0,\dots,w_{n+1}$,
1314: and $b_0,\dots,b_n$ are given as in Theorem \ref{smooth2}, and we can already define
1315: the real number $b_{n+1}$ by the formula:
1316: \begin{equation*}
1317: b_{n+1} :=-\displaystyle \sum_{i=1}^{n+2} \dfrac{f^{(i)}(0)}{i!} \, w_{n+2-i}
1318: -\displaystyle \sum_{i=1}^{n+1} (i+1) \, w_i \, w_{n+2-i} \, .
1319: \end{equation*}
1320: (Observe indeed that this definition only involves $w_0,\dots,w_{n+1}$, and not
1321: $w_{n+2}$.)
1322:
1323:
1324: \underline{Step 1}: because $V_\flat$, and $V_\sharp$ do not vanish, we can
1325: introduce some functions $W_{\flat,n+1}$, and $W_{\sharp,n+1}$ by the relations:
1326: \begin{equation*}
1327: W_\flat =w_0 +w_1 \, V_\flat +\cdots +w_n \, V_\flat^n
1328: +W_{\flat,n+1} \, V_\flat^{n+1} \, ,\quad
1329: W_\sharp =w_0 +w_1 \, V_\sharp +\cdots +w_n \, V_\sharp^n
1330: +W_{\sharp,n+1} \, V_\sharp^{n+1} \, .
1331: \end{equation*}
1332: Thanks to Taylor's formula, we can write the polynomial function $f$ as:
1333: \begin{equation*}
1334: f(V)=f(0)+f'(0) \, V+\dfrac{f''(0)}{2} \, V^2 +\cdots
1335: +\dfrac{f^{(n)}(0)}{n!} \, V^n +V^{n+1} \, F_{n+1} (V) \, ,
1336: \end{equation*}
1337: where $F_{n+1}$ is a polynomial function such that:
1338: \begin{equation*}
1339: F_{n+1} (0)=\dfrac{f^{(n+1)}(0)}{(n+1)!} \, ,\quad
1340: F_{n+1}' (0)=\dfrac{f^{(n+2)}(0)}{(n+2)!} \, .
1341: \end{equation*}
1342: Substituting the expression of $W_\flat$, and $W_\sharp$ in \eqref{ode-ref}
1343: shows (after a tedious computation!) that $(V_\flat,W_{\flat,n+1})$, and
1344: $(V_\sharp,W_{\sharp,n+1})$ are solutions to the following system of ODEs:
1345: \begin{equation}
1346: \label{eq:k}
1347: \begin{cases}
1348: V' = V \, (w_0+w_1 \, V +\cdots +w_n \, V^n +W_{n+1} \, V^{n+1}) \, ,& \\
1349: W_{n+1}' =g_{n+1} (V,W_{n+1}) \, ,&
1350: \end{cases}
1351: \end{equation}
1352: where the function $g_{n+1}$ is given by:
1353: \begin{align}
1354: g_{n+1} (V,W_{n+1}) := &-(n+2) \, W_{n+1} \left( \sum_{k=0}^n w_k \, V^k
1355: +V^{n+1} \, W_{n+1} \right) -W_{n+1} \, \sum_{k=0}^n (k+1) \, w_k \, V^k \notag\\
1356: &-W_{n+1} \, f(V) -F_{n+1} (V) \, \sum_{k=0}^n w_k \, V^k +b_n
1357: +\dfrac{f^{(n+1)}(0)}{(n+1)!} +V \, Q_{n+1}(V) \, ,\label{defgn+1}
1358: \end{align}
1359: and $Q_{n+1}$ is a polynomial function that satisfies:
1360: \begin{equation*}
1361: Q_{n+1}(0)=b_{n+1} +(n+4) \, w_1 \, w_{n+1} +f'(0) \, w_{n+1}
1362: +\dfrac{f^{(n+1)}(0)}{(n+1)!} \, w_1 +\dfrac{f^{(n+2)}(0)}{(n+2)!} \, w_0 \, .
1363: \end{equation*}
1364: When $n=1$, one has $Q_2 \equiv Q_2(0)=0$ (see the above definition for $b_2$).
1365: Using the expansion \eqref{expansion}, which is part of the induction assumption,
1366: we also know that:
1367: \begin{equation*}
1368: \lim_{\eta \rightarrow +\infty} (V_\flat,W_{\flat,n+1}) (\eta)
1369: =\lim_{\eta \rightarrow +\infty} (V_\sharp,W_{\sharp,n+1}) (\eta)
1370: =(0,w_{n+1}) =\Big( 0,\dfrac{b_n}{f(0)+(n+2)\, w_0} \Big) \, .
1371: \end{equation*}
1372:
1373:
1374: With the above definitions for $g_{n+1}$, and $Q_{n+1}$, we can check that
1375: $(0,w_{n+1})$ is a stationary solution to \eqref{eq:k}. (Recall that $w_{n+1}$
1376: is defined as in Theorem \ref{smooth2} by the induction assumption.) We can
1377: also evaluate the Jacobian matrix of the vector field associated with the
1378: system of ODEs \eqref{eq:k}:
1379: $$
1380: \begin{pmatrix}
1381: w_0 & 0 \\
1382: b_{n+1} & -f(0)-(n+3) \, w_0 \end{pmatrix} =
1383: \begin{pmatrix}
1384: \lambda_1^{(n+1)} & 0 \\
1385: b_{n+1} & \lambda_2^{(n+2)} \end{pmatrix} \, .
1386: $$
1387: There exists a positive number $a_{n+1} \le a_n$ such that for all $a \in
1388: (0,a_{n+1}]$, one has $\lambda_2^{(n+2)}<\lambda_1^{(n+2)}<0$, or equivalently
1389: $f(0)+(n+4) \, w_0>0$. In that case, the eigenvectors corresponding to the
1390: eigenvalues $\lambda_1^{(n+1)}$ and $\lambda_2^{(n+1)}$ are:
1391: $$
1392: e_1^{(n+1)} =\begin{pmatrix}
1393: f(0)+(n+4) \, w_0 \\
1394: b_{n+1} \end{pmatrix} \, ,\quad e_2^{(n+1)}=\begin{pmatrix}
1395: 0 \\
1396: 1 \end{pmatrix} \, .
1397: $$
1398: Using the same argument as in the proof of Proposition \ref{n=3}, we can
1399: conclude that the solutions $(V_\flat,W_{\flat,n+1})$, and
1400: $(V_\sharp,W_{\sharp,n+1})$ of \eqref{eq:k} are tangent to the straight line
1401: $(0,w_{n+1}) +\R \, e_1^{(n+1)}$ as $\eta$ tends to $+\infty$. In particular,
1402: this yields:
1403: \begin{equation}
1404: \label{lim0}
1405: \lim_{\eta \rightarrow +\infty} \dfrac{W_{\flat,n+1}'(\eta)}{V_\flat'(\eta)}
1406: =\lim_{\eta \rightarrow +\infty} \dfrac{W_{\sharp,n+1}'(\eta)}{V_\sharp'(\eta)}
1407: =\dfrac{b_{n+1}}{f(0)+(n+4)\, w_0} =:w_{n+2} \, .
1408: \end{equation}
1409:
1410:
1411: \underline{Step 2}: let us define the function $\widetilde{w}_{n+1}$ by the formula:
1412: \begin{equation*}
1413: \widetilde{w}_{n+1} (\xi) :=\begin{cases}
1414: W_{\flat,n+1} \circ \Xi_\flat^{-1} (\xi) &\text{if $\xi<0$,}\\
1415: w_{n+1} &\text{if $\xi=0$,}\\
1416: W_{\sharp,n+1} \circ \Xi_\sharp^{-1} (\xi) &\text{if $\xi>0$.}
1417: \end{cases}
1418: \end{equation*}
1419: With this definition, $\widetilde{w}_{n+1}$ is continuous, and we have the relation:
1420: \begin{equation}
1421: \label{recurrence0}
1422: w=w_0 +w_1 \, \hat{v} +\dots+w_n \, \hat{v}^n
1423: +\widetilde{w}_{n+1} \, \hat{v}^{n+1} \, .
1424: \end{equation}
1425: Moreover, using \eqref{lim0}, we obtain:
1426: \begin{equation}
1427: \label{recurrence4}
1428: \lim_{\xi \rightarrow 0^-} \dfrac{\widetilde{w}_{n+1}' (\xi)}{\hat{v}'(\xi)}
1429: =\lim_{\xi \rightarrow 0^+} \dfrac{\widetilde{w}_{n+1}' (\xi)}{\hat{v}'(\xi)}
1430: =w_{n+2} \, ,
1431: \end{equation}
1432: which yields $\widetilde{w}_{n+1}'(0^+)=\widetilde{w}_{n+1}'(0^-)$. Therefore, we
1433: have $\widetilde{w}_{n+1} \in C^1 (\R)$. Moreover, using \eqref{eq:k}, we can compute:
1434: \begin{equation}
1435: \label{recurrence1}
1436: \widetilde{w}_{n+1}' \, \hat{v} =g_{n+1} (\hat{v},\widetilde{w}_{n+1}) \, ,
1437: \end{equation}
1438: so we get $\widetilde{w}_{n+1}' \, \hat{v} \in C^1(\R)$.
1439:
1440:
1441: \underline{Step 3}: we use an induction argument to show that $w \in C^{n+2} (\R)$
1442: (which will imply immediately $\hat{v} \in C^{n+3} (\R)$). More precisely, we assume
1443: that for some $k \in \{ 0,\dots,n \}$, we have:
1444: \begin{equation}
1445: \label{recurrence2}
1446: \widetilde{w}_{n+1} \, \hat{v}^k \in C^{k+1}(\R) \, ,\quad
1447: w \in C^{k+1} (\R) \, ,\quad
1448: \widetilde{w}_{n+1}' \, \hat{v}^{k+1} \in C^{k+1} (\R) \, .
1449: \end{equation}
1450: We are going to show that this property implies the same property with $k$ replaced
1451: by $k+1$. (Observe that step 2 above shows that the property \eqref{recurrence2} holds
1452: for $k=0$.)
1453:
1454:
1455: We note that $\hat{v} \in C^{k+2} (\R)$, because $\hat{v}'=w \in C^{k+1} (\R)$. Moreover,
1456: we have $\widetilde{w}_{n+1} \, \hat{v}^{k+1} =(\widetilde{w}_{n+1} \, \hat{v}^k) \,
1457: \hat{v} \in C^{k+1} (\R)$, and we also have:
1458: \begin{equation*}
1459: (\widetilde{w}_{n+1} \, \hat{v}^{k+1})'=\widetilde{w}_{n+1}' \, \hat{v}^{k+1}
1460: +(k+1) \, (\widetilde{w}_{n+1} \, \hat{v}^k) \, w \in C^{k+1} (\R) \, .
1461: \end{equation*}
1462: Therefore, we get $\widetilde{w}_{n+1} \, \hat{v}^{k+1} \in C^{k+2} (\R)$.
1463:
1464:
1465: Using the relation \eqref{recurrence0}, we immediately obtain $w \in C^{k+2}(\R)$.
1466:
1467:
1468: We have $\widetilde{w}_{n+1}' \, \hat{v}^{k+2}=(\widetilde{w}_{n+1}' \, \hat{v}^{k+1})
1469: \, \hat{v} \in C^{k+1} (\R)$, and using \eqref{recurrence1}, we derive:
1470: \begin{multline}
1471: \label{recurrence3}
1472: \big( \widetilde{w}_{n+1}' \, \hat{v}^{k+2} \big)' =\big(
1473: g_{n+1} (\hat{v},\widetilde{w}_{n+1}) \, \hat{v}^{k+1} \big)' \\
1474: = (\partial_1 g_{n+1}) (\hat{v},\widetilde{w}_{n+1}) \, \hat{v}^{k+1} \, w
1475: +(\partial_2 g_{n+1}) (\hat{v},\widetilde{w}_{n+1}) \, \widetilde{w}_{n+1}' \, \hat{v}^{k+1}
1476: +(k+1) \, g_{n+1} (\hat{v},\widetilde{w}_{n+1}) \, \hat{v}^k \, w \, ,
1477: \end{multline}
1478: where $\partial_1 g_{n+1}$ (resp. $\partial_2 g_{n+1}$) denotes the partial derivative
1479: of $g_{n+1}$ with respect to its first (resp. second) variable. From the definition
1480: \eqref{defgn+1}, we see that $g_{n+1} (\hat{v},\widetilde{w}_{n+1})$ can be decomposed
1481: as follows:
1482: \begin{equation*}
1483: g_{n+1} (\hat{v},\widetilde{w}_{n+1})=-(n+2) \, \widetilde{w}_{n+1}^2 \, \hat{v}^{n+1}
1484: +\widetilde{w}_{n+1} \, P_1(\hat{v}) +P_0(\hat{v}) \, ,
1485: \end{equation*}
1486: where $P_0$, and $P_1$ are polynomial functions. Using this decomposition, and
1487: the induction assumption \eqref{recurrence2}, we can show that each term of the
1488: sum in the right-hand side of \eqref{recurrence3} belongs to $C^{k+1} (\R)$.
1489: Consequently $\widetilde{w}_{n+1}' \, \hat{v}^{k+2}$ belongs to $C^{k+2} (\R)$,
1490: and \eqref{recurrence2} holds with $k$ replaced by $k+1$. Because \eqref{recurrence2}
1491: holds for $k=0$, we get that \eqref{recurrence2} holds for $k=n+1$, so we have
1492: proved $w \in C^{n+2} (\R)$, and $\hat{v} \in C^{n+3}(\R)$.
1493:
1494:
1495: \underline{Step 4}: it remains to show that $w$ satisfies the asymptotic expansion
1496: \eqref{expansion} at the order $n+1$. Using \eqref{recurrence4}, and
1497: $\widetilde{w}_{n+1} \in C^1 (\R)$, we obtain:
1498: \begin{equation*}
1499: \widetilde{w}_{n+1} (\xi)-w_{n+1}=w_{n+2} \, \hat{v} (\xi) +o(\hat{v} (\xi))
1500: \, ,\quad \text{as $\xi \rightarrow 0$.}
1501: \end{equation*}
1502: Plugging this expansion in \eqref{recurrence0}, we obtain \eqref{expansion} at the
1503: order $n+1$, so the proof of the induction is complete.
1504: \end{proof}
1505:
1506:
1507: Once we know that the function $\hat{v}$ belongs to $C^{n+2} (\R)$, for
1508: $a \in (0,a_n]$, then $v=\hat{v}+(v_-+v_+)/2$ also belongs to $C^{n+2} (\R)$,
1509: and we have already seen in the previous section that $v$ does not vanish
1510: because $v(\xi)>v_+>0$ for all $\xi$. Moreover, the components $(\rho,u,e)$
1511: of the shock profile are given by:
1512: \begin{equation*}
1513: \rho(\xi)=\dfrac{j}{v(\xi)} \, ,\quad u(\xi)=v(\xi)+\sigma \, ,\quad
1514: e(\xi)=\dfrac{(C_1-v(\xi))\, v(\xi)}{\gamma-1} \, ,
1515: \end{equation*}
1516: so one has $(\rho,u,e) \in C^{n+2} (\R)$, and the proof of Theorem \ref{smooth}
1517: is complete. (Recall that the strength of the shock tends to zero if, and only
1518: if $a=|u_+-u_-|/2$ tends to zero.)
1519:
1520:
1521:
1522:
1523:
1524:
1525: \appendix
1526: \section{Formal derivation of the model}
1527: \label{model}
1528:
1529:
1530:
1531: It is worth describing how the model \eqref{eulerevol}, \eqref{diffn} can be
1532: obtained from a more complete physical system. The derivation we propose below
1533: remains formal -- a rigorous proof being certainly delicate and beyond the scope
1534: of this work -- and we refer to \cite{BD, GL, LHM, MM} for further details. Let
1535: us introduce the specific intensity of radiation $f(t,x,v)$, that depends on
1536: a time variable $t\ge 0$, a space variable $x\in \R^N$, and a direction $v\in
1537: \SP^{N-1}$. We make the 'grey assumption', which means that the frequency
1538: dependence is ignored (all photons have the same frequency). Photons are
1539: subject to two main interaction phenomena:
1540: \begin{itemize}
1541: \item scattering produces changes in the direction of the photons,
1542:
1543: \item absorption/emission where photons are lost/produced through a transfer
1544: mecanism with the surrounding gas.
1545: \end{itemize}
1546: The scattering phenomenon is described by the operator:
1547: \[
1548: Q_s(f)(t,x,v)=\sigma_s \, \Big(
1549: \displaystyle \int_{\SP^{N-1}} f(t,x,v') \, dv'-f(t,x,v) \Big) \, ,
1550: \]
1551: (with $dv$ the normalized Lebesgue measure on $\SP^{N-1}$), and the
1552: absorption/emission phenomenon is described by the operator:
1553: \[
1554: Q_a(f)(t,x,v)=\sigma_a \, \Big(
1555: \dfrac{\sigma}{\pi} \, \theta (t,x)^4 -f(t,x,v) \Big) \, ,
1556: \]
1557: where we used the Stefan-Boltzmann emission law, $\theta$ being the temperature
1558: of the gas, and $\sigma$ the Stefan-Boltzmann constant. In these definitions,
1559: the coefficients $\sigma_{s,a}$ are given positive quantities. These phenomena
1560: are both characterized by a typical mean free path, denoted $\ell_s,\ell_a$
1561: respectively. Therefore, the evolution of the specific intensity is driven by:
1562: \begin{equation}
1563: \label{kindim}
1564: \dfrac{1}{c} \, \partial_t f +v \cdot \nabla_x f =\dfrac{1}{\ell_s} \, Q_s(f)
1565: +\dfrac{1}{\ell_a} \, Q_a(f)=Q(f) \, ,
1566: \end{equation}
1567: where $c$ stands for the speed of light. The equation \eqref{kindim} is
1568: coupled to the Euler system describing the evolution of the fluid:
1569: \begin{equation}
1570: \label{eulerevolbis}
1571: \begin{cases}
1572: \partial_t \rho +\nabla_x \cdot (\rho \, u) =0 \, ,& \\
1573: \partial_t (\rho \, u) +\nabla_x \cdot (\rho \, u\otimes u) +\nabla_x P
1574: =-\dfrac{1}{c} \displaystyle \int_{\SP^{N-1}} v \, Q(f) \, dv \, ,& \\
1575: \partial_t (\rho \, E) +\nabla_x \cdot (\rho \, E \, u+P\, u)
1576: =-\displaystyle \int_{\SP^{N-1}} Q(f) \, dv \, .&
1577: \end{cases}
1578: \end{equation}
1579: The equations \eqref{kindim}, \eqref{eulerevolbis} are thus coupled by the
1580: exchanges of both momentum and energy, and by the Stefan-Boltzmann emission
1581: law. Observe that only the emission/absorption operator enters into the energy
1582: equation since the scattering operator is conservative (this would be different
1583: if Doppler corrections were taken into account). Note also that the total energy:
1584: \[
1585: \dfrac{1}{c} \displaystyle \int_{\R^N} \int_{\SP^{N-1}} f \, dv \, dx
1586: +\displaystyle \int_{\R^N} \rho \, E \, dx \, ,
1587: \]
1588: is (formally) conserved. Writing the system \eqref{eulerevolbis}, and the
1589: kinetic equation \eqref{kindim} in the dimensionless form, we can make four
1590: dimensionless parameters appear:
1591: \begin{description}
1592: \item - $\mathcal C$, the ratio of the speed of light over the typical sound
1593: speed of the gas,
1594:
1595: \item - $\mathcal L_s$, the Knudsen number associated to the scattering,
1596:
1597: \item - $\mathcal L_a$, the Knudsen number associated to the absorption/emission,
1598:
1599: \item - $\mathcal P$, which compares the typical energy of radiation and the
1600: typical energy of the gas.
1601: \end{description}
1602: We thus obtain the rescaled equations:
1603: \begin{equation}
1604: \label{adim}
1605: \begin{cases}
1606: \dfrac{1}{\mathcal C} \, \partial_t f + v \cdot \nabla_x f
1607: =\dfrac{1}{\mathcal L_s} \, Q_s(f)+\dfrac{1}{\mathcal L_a} \, Q_a(f) \, ,& \\
1608: \partial_t \rho +\nabla_x \cdot (\rho \, u)=0 \, ,& \\
1609: \partial_t (\rho \, u) +\nabla_x \cdot (\rho \, u \otimes u) +\nabla_x P
1610: =\dfrac{\mathcal P}{\mathcal L_s} \, \sigma_s \,
1611: \displaystyle \int_{\SP^{N-1}} v \, f(v) \, dv \, ,& \\
1612: \partial_t (\rho \, E) +\nabla_x \cdot (\rho \, E \, u+P\, u)
1613: =-\dfrac{\mathcal P}{\mathcal L_a} \, \sigma_a \, \Big(
1614: \theta^4 -\displaystyle \int_{\SP^{N-1}} f (v) \, dv \Big) \, .&
1615: \end{cases}
1616: \end{equation}
1617: System \eqref{eulerevol}, \eqref{diffn} is then obtained in two steps. First of
1618: all, we assume $\mathcal C \gg 1$. Next, we keep $\mathcal P$ of order 1, and we
1619: are concerned here with a regime where scattering is the leading phenomenon: the
1620: mean free paths are rescaled according to:
1621: \[
1622: \mathcal L_s \simeq \dfrac{1}{\mathcal C} \, ,\qquad
1623: \mathcal L_a \simeq \mathcal C \, .
1624: \]
1625: The asymptotics can be readily understood by means of the Hilbert expansion:
1626: \[
1627: f=f^{(0)} +\dfrac{1}{\mathcal C} \, f^{(1)} +\dfrac{1}{\mathcal C^2} \, f^{(2)}+\dots
1628: \]
1629: Identifying the terms arising with the same power of $1/\mathcal C$, we get:
1630: \begin{description}
1631: \item - at the leading order, $f^{(0)}$ belongs to the kernel of the scattering
1632: operator, so that is does not depend on the microscopic variable $v$:
1633: $f^{(0)}(t,x,v)=n(t,x)$,
1634:
1635: \item - the relation $Q_s(f^{(1)})=v \cdot \nabla_x f^{(0)}$ then leads to:
1636: $f^{(1)}(t,x,v)=-\frac{1}{\sigma_s}\, v \cdot \nabla_x n(t,x)$,
1637:
1638: \item - integrating the equation for $f^{(2)}$ over the sphere yields:
1639: \[
1640: \partial_t n -\dfrac{1}{N \, \sigma_s} \, \Delta_x n
1641: =\sigma_a \, ( \theta^4 -n) \, .
1642: \]
1643: \end{description}
1644: Note also that in the momentum equation, we have:
1645: \[
1646: \dfrac{\sigma_s}{\mathcal L_s} \displaystyle \int_{\SP^{N-1}} v \, f(v) \, dv
1647: \simeq \sigma_s \, \displaystyle \int_{\SP^{N-1}} v \, f^{(1)}(v) \, dv
1648: =-\dfrac{1}{N} \, \nabla_x n \, .
1649: \]
1650: Finally, we obtain the limit system:
1651: \begin{equation}
1652: \label{lim1}
1653: \begin{cases}
1654: \partial_t \rho +\nabla_x \cdot (\rho \, u)=0\, ,& \\
1655: \partial_t (\rho \, u) +\nabla_x \cdot (\rho \, u \otimes u) +\nabla_x P
1656: =-\dfrac{\mathcal P}{N} \, \nabla_x n \, ,& \\
1657: \partial_t (\rho \, E) +\nabla_x \cdot (\rho \, E \, u +P\, u)
1658: =-\mathcal P \, \sigma_a \, ( \theta^4-n ) \, ,& \\
1659: \partial_t n -\dfrac{1}{N \, \sigma_s} \, \Delta_x n =\sigma_a \,
1660: ( \theta^4-n ) \, .
1661: \end{cases}
1662: \end{equation}
1663: The system \eqref{lim1} describes a nonequilibrium regime, where the material
1664: and the radiations have different temperatures ($\theta\neq n^{1/4}$); the
1665: equilibrium regime would correspond to assuming that the emission/absorption
1666: is the leading contribution.
1667:
1668:
1669: After this first asymptotics, we perform a second asymptotics where we set:
1670: \[
1671: \mathcal P \ll 1 \, ,\qquad \mathcal P \, \sigma_a=1 \, ,\qquad
1672: N\, \sigma_s=1/\sigma_a \, .
1673: \]
1674: This leads to \eqref{eulerevol}, \eqref{diffn}. Of course, one might wonder
1675: how this second approximation modifies the shock profiles compared to
1676: \eqref{lim1}, in particular when we get rid of the radiative pressure
1677: in the momentum equation. We refer to \cite[page 579]{MM} for some aspects
1678: of this problem.
1679:
1680:
1681:
1682:
1683:
1684:
1685: \bibliographystyle{alpha}
1686: \bibliography{lcg}
1687: \end{document}
1688: