1: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2: % Last modif by B. on Dec 21
3:
4: \documentclass [11pt,a4paper]{article}
5:
6: \usepackage{amsmath}
7:
8: \usepackage{graphics}
9: \usepackage{latexsym}
10: \usepackage{amssymb}
11: \usepackage{enumerate}
12: %\usepackage{euscript}
13:
14: \newtheorem {theorem}{Theorem}
15: \newtheorem {thm}[theorem]{Theorem}
16: \newtheorem {lemma}[theorem]{Lemma}
17: \newtheorem {corollary}[theorem]{Corollary}
18: \newtheorem {prop}[theorem]{Proposition}
19: \newtheorem {definition}[theorem]{Definition}
20:
21: %%%%Some macros
22: \def \N {\mathbb N}
23: \def \Z {\mathbb Z}
24: \def \Q {\mathbb Q}
25: \def \R {\mathbb R}
26: \def \C {\mathbb C}
27: \def \ind{1\!\!1}
28: \def \eps {\varepsilon}
29: \def \s {{\sigma}}
30: \def \sigm {{u}}
31: \def \rh {{\rho}}
32:
33: \def \prob {\ensuremath{\mathbf{P}}}
34: \def \expect {\ensuremath{\mathbf{E}}}
35: \def \var {\ensuremath{\mathbf{Var}}}
36: \def \cov {\ensuremath{\mathbf{Cov}}}
37:
38: \def\pt{\partial_t}
39: \def\px{\partial_x}
40:
41: \def\us{\mathbf{s}}
42: \def\uzero{\mathbf{0}}
43: \def\uone{\mathbf{1}}
44: \def\vareps{\varepsilon}
45:
46: \newcommand {\sgn}{\mathop{\mathrm{sgn}}}
47: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
48: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
49:
50: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
51: \title {Hydrodynamic Equation for a Deposition Model}
52: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
53: \author {B\'alint T\'oth
54: %\thanks{Institute of Mathematics, TU Budapest}
55: \and
56: Wendelin Werner
57: %\thanks{D\'epartment de Math\'ematiques, Universit\'e Paris-Sud, Orsay}
58: }
59: \date {Technical University Budapest and Universit\'e Paris-Sud}
60: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
61:
62: \begin {document}
63:
64: \setlength{\baselineskip}{1.23\baselineskip}
65:
66: \maketitle
67: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
68:
69: \begin {abstract}
70: We show that the two-component
71: system of hyperbolic
72: conservation laws $\partial_t \rho + \partial_x (\rho u)
73: =0 = \partial_t u + \partial_x \rho$
74: appears naturally in the formally
75: computed hydrodynamic limit of
76: some randomly growing interface models,
77: and we study some properties of this system.
78: \end {abstract}
79: \vskip 3mm
80:
81: {\bf Key Words:} Hyperbolic conservation laws, KPZ equation
82:
83: {\bf MSC-class.:} 35L65, 82C41, 60K35
84: %
85: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
86:
87: \section{Introduction}
88:
89: The macroscopic behaviour of physical systems can
90: often be described in terms of non-linear partial
91: differential equations. In many cases, it had been
92: shown that functionals of microscopic models
93: from statistical physics converge in the hydrodynamic limit
94: towards certain solutions of these partial differential equations.
95:
96: Studying the partial differential equation
97: (or the system of partial differential equations)
98: can turn out to be a very hard challenge in itself:
99: Appearance of
100: singularities in finite time, shocks etc.
101: The so-called hyperbolic conservation laws have in particular
102: received a lot of interest. Even in one space dimension,
103: these PDEs proved to be extremely interesting and challenging
104: both mathematically and phenomenologically.
105: These are partial differential equations of the type
106: %
107: \begin{equation*}
108: \partial_t u + \partial_x J(u) = 0
109: \end{equation*}
110: %
111: where $u= u(t,x)$
112: takes its value in $\R^n$ and $J$ is a non-linear function
113: from $\R^n$ into $\R^n$.
114:
115: The best known and most investigated examples are the following.
116: (See e.g. \cite{hormander, serre, smoller} for a comprehensive
117: introduction and survey of the subject.)
118: \begin{enumerate}[(1)]
119: \item
120: Burgers' equation (with no viscosity): $n=1$ and
121: %
122: \begin{equation*}
123: \label{burgers}
124: \pt u+ \px(u^2/2) = 0.
125: \end{equation*}
126: %
127: \item
128: The isentropic gas dynamics equation in one space dimension:
129: $n=2$, the components are the density field $\rho(t,x)$ and momentum
130: field $m(t,x)$
131: %
132: \begin{equation}
133: \label{igd}
134: \left\{
135: \begin{array}{l}
136: \pt \rho +\px m=0
137: \\[5pt]
138: \pt m + \px \big( m^2/\rho +p(\rho)\big)=0
139: \end{array}
140: \right.
141: \end{equation}
142: %
143: where $p(\rho)$ is the pressure, depending on density only.
144: \item
145: The so-called p-system, which is an alternative formulation of
146: the dynamics of one-dimensional gas. The two components
147: are the velocity field $u(t,x)$ and the specific volume (=
148: inverse density) field $v(t,x)$:
149: %
150: \begin{equation}
151: \label{psystem}
152: \left\{
153: \begin{array}{l}
154: \pt v - \px u=0
155: \\[5pt]
156: \pt u + \px p(v)=0.
157: \end{array}
158: \right.
159: \end{equation}
160: %
161: Here $p(v)$ denotes the pressure, as a function of specific
162: volume.
163: \item
164: The shallow water equation is another two component system:
165: $h(t,x)$ denotes the height of the (shallow) layer of water,
166: $u(t,x)$ is the velocity field:
167: %
168: \begin{equation}
169: \label{shallowwater}
170: \left\{
171: \begin{array}{l}
172: \pt h + \px(h u) =0
173: \\[5pt]
174: \pt u + \px \big(u^2/2 +h\big)=0.
175: \end{array}
176: \right.
177: \end{equation}
178: %
179: \end{enumerate}
180:
181: Since Riemann, a considerable amount of knowledge and technology
182: (more recently, for instance, entropy solutions, compensated
183: compactness method) has been derived that give a better
184: understanding of the physically relevant solutions to these
185: equations.
186:
187: In the present paper, we will be considering a
188: particular two-component
189: (i.e. $n=2$) system of hyperbolic conservation laws
190: that arises in the context of surface growth
191: (or more precisely growing interfaces, since the surface is one-dimensional).
192: In other words, at each time
193: $t \ge 0$, one sees a landscape $
194: x \mapsto h(t, x)$ where $x \in \R$.
195: The function $h$ is increasing in time. The rough phenomenological
196: description of the phenomena we are interested in corresponds to the
197: case where the surface is growing in the normal direction to
198: its boundary, but there exists a `tension' that tends to
199: keep the surface together, in the sense that it will
200: fill in holes quickly.
201: In the physics literature, a famous equation has been
202: proposed by Kardar, Parisi and Zhang (the KPZ equation)
203: as a model for such situations, cf. \cite{kardarparisizhang}.
204: It is (in the mathematical
205: jargon) an ill-posed non-linear partial differential
206: equation with a stochastic term:
207: %
208: \begin{equation*}
209: \label{kpzequ}
210: \pt h = \Delta h - (\px h)^2 + W
211: \end{equation*}
212: %
213: where $W=W(t,x)$ denotes a space-time white noise.
214: We do not want to give a review of the huge physics
215: literature on this equation, but we briefly stress two
216: aspects.
217: (See \cite{barabasistanley} for a
218: state-of-the-art survey of the physics literature on the subject
219: and an exhaustive list of references up to 1995.)
220: First, there exists to our knowledge no completely satisfactory
221: (see however \cite {bertinigiacomin}) derivation of
222: this equation from a microscopic model.
223: Second, it is predicted that
224: `the' solution to this equation has a special scaling
225: behaviour at late times. More precisely, it is believed
226: that when $\alpha, t, x$ are very large,
227: $h^{(\alpha)} (t,x) =
228: \alpha^{-1/3} h( \alpha t , \alpha^{2/3} x)$
229: is also a solution to the KPZ equation.
230: The exponents $1/3$ and $2/3$ should be related
231: to various conjectures and recent rigorous results
232: concerning the fluctuations of highest eigenvalues of
233: random matrices, of first passage percolation
234: paths, of longest increasing sequences etc etc.
235:
236: One way to define one-dimensional
237: interfaces $h(t,x)$ in terms of
238: particle systems goes as follows:
239: Start with a (finite or infinite) system of particles
240: that evolve randomly in the potential $h(t,x)$ (or in
241: some potential defined in terms of $h$)
242: and that all contribute to increase the potential
243: in the sense that $h(t ,x)$ corresponds to the joint
244: local time (i.e. cummulated occupation time density)
245: of the particles at time $t$ and site $x$. In other
246: words, $h(t,x)$ increases locally at $x$ if there is
247: a particle at $x$ and time $t$.
248: Note that this leads naturally to a two-component system
249: in the (formally computed) hydrodynamical limit:
250: the first component is the
251: density of particles, and the second component is the
252: gradient of the profile of the potential.
253:
254: In \cite{tothwerner}, we constructed a continuous stochastic
255: process, corresponding on a heuristic level to
256: the case where there is exactly (and only) one particle (its location at
257: time $t$ is denoted by $X_t$)
258: which is driven by
259: %
260: \begin{equation*}
261: \label{tsrm1}
262: dX_t = - \px h (t, X_t) dt
263: \end{equation*}
264: %
265: and $h(t,x)$ is the local time of $X$ at $x$ and time $t$,
266: so that
267: %
268: \begin{equation*}
269: \label{tsrm2}
270: \pt h(t, x) = \delta(X_t -x).
271: \end{equation*}
272: %
273: For details concerning the construction and primary properties of
274: this process and a rigorous version of
275: these equations, see \cite{tothwerner}. Let us just emphasize
276: a couple of features: The process $(X_t, t \ge 0)$ is a random
277: process, even though the previous `differential equations'
278: look very deterministic.
279: One reason is that (in the stationary regime), the function
280: $x \mapsto h(t,x)$ is not regular; in fact, it is a
281: Brownian motion in the space variable (for fixed $t$).
282: Second, $X_t$ is not a usual stochastic process (it is
283: not solution of a stochastic differential equation
284: for instance), it
285: has the $2/3$ scaling: $(\alpha^{-2/3} X_{\alpha t}, t\ge 0)$ has
286: the same law as $(X_t, t \ge 0)$. In particular,
287: $(\alpha^{-1/3} h(\alpha^{2/3}x, \alpha t), t \ge 0, x \in \R)$
288: has the same law as $(h(x,t), t \ge 0,
289: x \in \R)$ so that $h$ has the same scaling property as
290: the asymptotic scaling conjectured for the
291: KPZ equation.
292:
293: The process $(X_t, t \ge 0)$ can be viewed as the scaling limit
294: of a discrete negatively reinforced (i.e. self-repellent)
295: random walk $(S_n , n \ge 0)$ on $\Z$ called the
296: `true self-avoiding walk' in the physics literature. This is
297: a nearest-neighbour
298: walk on $\Z$ that decides at each step to jump
299: to the left or to the right according to a probability
300: depending on how many times it has visited the neighbouring
301: sites (or edges) before. Suppose for instance that after
302: $n$ steps $S_n=x$ and that the discrete walk
303: $(S_i)_{i \le n}$ has jumped already $l$ (resp. $r$) times on the
304: edge immediately to the left (resp. to the right) of $x$.
305: Then, $S_{n+1} = x+1$ with probability
306: %
307: \begin{equation*}
308: \label{tsaw}
309: \prob \big( S_{n+1} = x+1 \big |
310: l, r, S_n=x \big)
311: = \frac {e^{-\beta l}} {e^{-\beta l} + e^{-\beta r}}
312: \end{equation*}
313: %
314: where $\beta>0$ is some fixed constant.
315: In other words, the walk will
316: prefer to go along the edge it has visited less often in the
317: past.
318: Note also that the probability in fact
319: depends only on the
320: difference $l-r$ (which depends on all the past trajectory).
321: The distribution of the rescaled position of the random walker,
322: $S_n/n^{2/3}$, converges to (a multiple of) the one-dimensional
323: marginal
324: distribution of the continuous process $X_t$ described above,
325: \cite{toth}.
326:
327: It seems natural to consider the case where this
328: one particle is replaced by many particles performing the same
329: kind of self-repelling walk on $\Z$, with a \emph{joint
330: cumulated local time of all particles}. Or, in the continuous
331: space-time setting: a continuously distributed cloud of particles
332: (that all contribute to the same local time), which is the subject of
333: the present paper. As we shall see, this leads in the (formally
334: computed) hydrodynamic limit to the following system of
335: hyperbolic conservation laws:
336: %
337: \begin{equation}
338: \left\{
339: \begin{array}{l}
340: \pt \rho+\px (\rho u) =0
341: \\[5pt]
342: \pt u + \px \rho = 0
343: \end{array}
344: \right.
345: \label{ourpde}
346: \end{equation}
347: %
348: where $\rho$ corresponds to the density of particles at
349: $x$ and time $t$, and $u (x,t) = -\px h$ corresponds to
350: the negative gradient of the interface.
351: It seems, that although this system looks very natural,
352: it has not been considered in the literature.
353: We should emphasize that in spite of some formal similarities
354: with the p-system (\ref{psystem}) and the shallow water equation
355: (\ref{shallowwater}), the system (\ref{ourpde}) shows very
356: different behaviour and describes a quite different phenomenon.
357: We hope that its study may lead to improved understanding
358: of some aspects of `growing interfaces' in general.
359: In particular, this equation could shed some light on some
360: of the conjectured properties of the KPZ equation.
361: \emph{The goal of the present paper is not to present a complete
362: treatment of this system of partial differential equation,
363: but rather to initiate it as an alternative approach to 1-d
364: domain growth and deposition phenomena}.
365:
366:
367:
368: \section{The PDE: phenomenological derivation}
369:
370: We define a deposition model in the following terms.
371: The actual state of the system is described by two functions:
372: %
373: \begin{equation*}
374: \rho:\R_+\times\R\to\R_+
375: \quad\mbox{ and }\quad
376: h:\R_+\times\R\to \R.
377: \end{equation*}
378: %
379: $\rho(t,x)$ is the density of the population performing the
380: deposition, while $h(t,x)$ is the deposition height at time $t$
381: and space coordinate $x$.
382: The rules governing the time evolution of the system are the following
383: \begin{enumerate}[(1)]
384: %
385: \item
386: The total population is conserved, so
387: that the continuity equation
388: %
389: \begin{equation*}
390: \pt\rho+\px(\rho u)=0
391: \end{equation*}
392: %
393: is valid, where $u(t,x)$ is the velocity field, to be specified
394: by the dynamical rules.
395: %
396: \item
397: The deposition rate is proportional to the density of the
398: population, i.e.
399: %
400: \begin{equation}
401: \pt h=c_1\rho,
402: \label{depositionrate}
403: \end{equation}
404: %
405: %
406: where $c_1$ is a positive constant.
407: %
408: \item
409: The population is driven by a velocity field proportional to
410: the negative gradient of height
411: %
412: \begin{equation}
413: u=-c_2\px h,
414: \label{velocityfield}
415: \end{equation}
416: %
417: where $c_2$ is another positive constant.
418: This rule corresponds to the self-repellence mechanism
419: described in the introductory section.
420: \end{enumerate}
421: %
422: From (\ref{depositionrate}) and (\ref{velocityfield}) we
423: readily get
424: %
425: \begin{equation*}
426: \pt u+c_1c_2\px \rho=0
427: \end{equation*}
428: %
429: Without loss of generality, we can
430: choose $c_1c_2=1$ and get the two component system of hyperbolic
431: conservation laws
432: %
433: \begin{equation}
434: \left\{
435: \begin{array}{l}
436: \pt \rho+\px (\rho u) =0
437: \\[5pt]
438: \pt u + \px \rho = 0
439: \end{array}
440: \right.
441: \label{ourpde2}
442: \end{equation}
443: %
444: This system of PDEs with initial conditions
445: %
446: \begin{equation}
447: \rho(0,x)=\rho^{(0)}(x),
448: \qquad
449: u(0,x)=u^{(0)}(x)
450: \label{ouric}
451: \end{equation}
452: %
453: is the main object of the present paper.
454: %We
455: %propose it as a hydrodynamic description of deposition
456: %phenomena, possible alternative of the stochastic KPZ approach.
457:
458: As a first remark we mention here the scale invariance of
459: (\ref{ourpde2}). Let $\nu\in \R$ be fixed. Given the functions
460: $(t,x)\mapsto \rho(t,x)$ and $(t,x)\mapsto u(t,x)$ and a positive
461: fixed number $\alpha$, define the rescaled functions
462: %
463: \begin{align*}
464: \rho^{(\alpha)}(t,x)
465: &
466: :=
467: \alpha^{2(1-\nu)}\rho(\alpha t,\alpha^{\nu}x),
468: \\
469: u^{(\alpha)}(t,x)
470: &
471: :=
472: \alpha^{1-\nu} u(\alpha t,\alpha^{\nu}x).
473: \end{align*}
474: %
475: One can easily check that if $(\rho,u)$ is solution of
476: (\ref{ourpde2}), then
477: $(\rho^{(\alpha)}, u^{(\alpha)})$ is also a solution,
478: for any $\alpha>0$. The choice $\nu=1$ yields the hyperbolic
479: scale invariance valid for any hyperbolic conservation law. More
480: interesting is for our purposes the choice $\nu=2/3$. This is
481: the physically relevant scale invariance, since the density
482: changes covariantly under this scaling, i.e., the total mass
483: $\int\rho^{(\alpha)}dx$ is unchanged.
484:
485: With this choice of $\nu$ the following scale invariance of the
486: deposition height follows:
487: %
488: \begin{equation*}
489: h^{(\alpha)}(t,x)
490: :=
491: \alpha^{-1/3} h(\alpha t,\alpha^{2/3}x).
492: \end{equation*}
493: %
494: Recall that this is exactly the \emph{conjectured asymptotic scale
495: invariance of the one-dimensional KPZ equation}.
496:
497:
498:
499:
500: \section{Bricklayers}
501:
502: We define a system of interacting particles living on $\Z$,
503: with \emph{two conserved} quantities, whose hydrodynamic modes
504: are governed by a two-component system of hyperbolic
505: conservation laws which, after taking another limit (low
506: density/late time), transforms into our system (\ref{ourpde2}).
507: The computations of the present section are somewhat
508: formal. Working out all technical details (e.g. proving
509: uniqueness of the equilibrium Gibbs measures or technical details
510: of Yau's hydrodynamic limit) needs more effort.
511: The present section serves as microscopic motivation of the
512: PDE proposed above.
513:
514: \subsection{The particle system}
515:
516: The Great Wall of China is being built by a brigade of
517: bricklayers. The wall consists of columns of unit-size bricks,
518: piled above the edges of the lattice $\Z$. The height of the
519: column piled above the edge $(j,j+1)$
520: (i.e., number of bricks in this column) is $h_j$. In the
521: dynamics of the system the discrete negative gradients
522: $z_j:=h_{j-1}-h_{j}\in\Z$ will be relevant.
523: The bricklayers occupy
524: the sites of the lattice. At each site $j\in\Z$ there might be
525: an unlimited number $n_j\in\N$ of bricklayers. Bricklayers jump
526: to neighbouring sites and at each jump $j\to j\pm1$ a brick is
527: added to the respective column of bricks.
528:
529: In more technical
530: terms: particles (= bricklayers) perform continuous time nearest
531: neighbour walk on the lattice $\Z$ and $h_j$ measures the
532: cumulated (discrete) local time on the lattice edge $(j,j+1)$.
533:
534:
535: \begin{figure}[ht]
536: \centering
537: %\vspace {-2cm}
538: \resizebox{0.95 \textwidth}{!} {\includegraphics {gwc.eps}}
539: %\vspace {-1cm}
540: \caption{
541: The great wall being built}
542: \end{figure}
543:
544:
545: About the dynamics: the jump rates are chosen so that the
546: following conditions hold:
547: \begin{enumerate}[(1)]
548: %
549: \item
550: the bricklayers' jumps are driven
551: by the local shape of the wall so that they try to reduce the
552: differences $z_j$ (i.e. to keep the height of the wall even),
553: %
554: \item
555: conditionally on the actual shape of the wall the
556: bricklayers jump independently.
557: \end{enumerate}
558: % NEW
559: This is done as follows. The instantaneous rate of jump from
560: site $j$ to site $j\pm1$ (for each bricklayer sitting at site $j$)
561: is equal to $r(\pm z_j)$, where $r:\Z\to (0,\infty)$ is a fixed monotone
562: increasing function which defines the model. In order to be able
563: to compute explicitely
564: the stationary measures (see subsection 3.2) we impose
565: that $r(1-z) r(z)$ is a positive constant (this is
566: for instance the case if $r(z) = \exp (\beta z)$), and
567: multiplying time by a constant term, we can
568: in fact restrict ourselves
569: to the case where
570: % END NEW
571: \begin{equation}
572: \label{rcond}
573: r(z)r(-z+1)=1,
574: \quad
575: \text{ for all }
576: z\in\Z.
577: \end{equation}
578: Thus, the following
579: changes of configuration may occur:
580: %
581: \begin{equation*}
582: (n_j,z_j),(n_{j+1},z_{j+1})
583: \to
584: (n_j-1,z_j-1),(n_{j+1}+1,z_{j+1}+1)
585: \end{equation*}
586: %
587: with rate $n_j r(z_j)$, and
588: %
589: \begin{equation*}
590: (n_j,z_j),(n_{j-1},z_{j-1})
591: \to
592: (n_j-1,z_j+1),(n_{j-1}+1,z_{j-1}-1)
593: \end{equation*}
594: %
595: with rate $n_j r(-z_j)$.
596: % TO HERE
597:
598: Clearly, $\sum_j n_j$ and $\sum_j z_j$ are formally conserved
599: quantities of the dynamics. It is also clear that besides these
600: globally conserved quantities the parity of $n_j+z_j$ is also
601: conserved on each lattice site $j\in\Z$.
602:
603: Now we give a more formal description of our interacting
604: particle system. For $s\in\{0,1\}$ let
605: %
606: \begin{equation*}
607: \big(\N\times\Z\big)_s:=
608: \{(n,z)\in\N\times\Z: n+z=s \mbox{ mod } 2\}.
609: \end{equation*}
610: %
611: Given the sequence $\us=(s_j)_{j\in\Z}\in\{0,1\}^{\Z}$ we define
612: the state space of our system as
613: %
614: \begin{equation*}
615: \Omega_{\us}:=
616: \prod_{j\in\Z}\big(\N\times\Z\big)_{s_j}.
617: \end{equation*}
618: %
619: Elements of $\Omega_{\us}$ will be denoted by $\mathbf{\omega}$, i.e.
620: $\mathbf{\omega} = (\omega_j)_{j\in\Z}$ with
621: $\omega_j = (n_j,z_j) \in \big(\N\times\Z\big)_{s_j}$.
622: The (formal) infinitisimal generator of the Markov process
623: described verbally in the first paragraph of this section, is:
624: % FROM HERE
625: %
626: \begin{equation*}
627: Lf(\omega) =
628: \sum_{j\in\Z} n_j r(z_j)
629: \big(f(\Theta_{j+}\omega) - f(\omega)\big)
630: +
631: \sum_{j\in\Z} n_j r(-z_j)
632: \big(f(\Theta_{j-}\omega) - f(\omega)\big),
633: \end{equation*}
634: %
635: % TO HERE
636: where the maps $\Theta_{j+}$ and $\Theta_{j-}$ act on the subsets
637: $\{\mathbf{\omega}\in\Omega_{\us}: n_j\ge1\}$ as
638: %
639: \begin{equation*}
640: \big(\Theta_{j+}\mathbf{\omega}\big)_i:=
641: \left\{
642: \begin{array}{lcl}
643: (n_i,z_i)
644: &\quad \mbox{ if } \quad&
645: i\not=j, j+1
646: \\[5pt]
647: (n_i-1, z_i- 1)
648: &\quad \mbox{ if } \quad&
649: i=j
650: \\[5pt]
651: (n_{i}+1, z_{i}+ 1)
652: &\quad \mbox{ if } \quad&
653: i=j+1
654: \end{array}
655: \right.
656: \end{equation*}
657: %
658: respectively
659: %
660: \begin{equation*}
661: \big(\Theta_{j-}\mathbf{\omega}\big)_i:=
662: \left\{
663: \begin{array}{lcl}
664: (n_i,z_i)
665: &\quad \mbox{ if } \quad&
666: i\not=j, j-1
667: \\[5pt]
668: (n_i-1, z_i+ 1)
669: &\quad \mbox{ if } \quad&
670: i=j
671: \\[5pt]
672: (n_{i}+1, z_{i}- 1)
673: &\quad \mbox{ if } \quad&
674: i=j -1
675: \end{array}
676: \right.
677: \end{equation*}
678: %
679:
680: \subsection{Equilibrium Gibbs measures}
681:
682: % NEW
683: For $k\ge0$ denote
684: %
685: \begin{equation*}
686: R(z):=\prod_{k=1}^{|z|} r(k)
687: \end{equation*}
688: %
689: and
690: %
691: \begin{equation*}
692: \theta^*:=\lim_{k\to\infty} r(k)\in(1,\infty].
693: \end{equation*}
694: Note that (\ref {rcond}) implies that for all $z \in \Z$,
695: \begin {equation}
696: \label {rprop}
697: R(-z)=
698: R(z) = R(z-1) r(z)
699: = R(z+1) r (-z).
700: \end {equation}
701: %
702: Fix the parameters $s\in\{0,1\}$, $\lambda>0$,
703: $\theta\in(1/\theta^*, \theta^*)$ and
704: define the probability measure $\mu_{s,\lambda,\theta}$ on
705: $\big(\N\times\Z\big)_s$ as follows:
706: %
707: \begin{equation*}
708: \mu_{s,\lambda,\theta}(n,z):=
709: \frac{1}{Z_s(\lambda,\theta)}
710: \frac{\lambda^n}{n!}\frac{\theta^z}{R(z)},
711: \end{equation*}
712: %
713: where
714: %
715: \begin{equation*}
716: Z_s(\lambda,\theta):=
717: \sum_{(n,z)\in(\N\times\Z)_s}
718: \frac{\lambda^n}{n!}\frac{\theta^z}{R(z)}
719: \end{equation*}
720: %
721: is the normalizing factor (partition function). The measure
722: $\mu_{s,\lambda,\theta}$ is a product measure on $\N\times\Z$
723: restricted to the subset $n+z=s\mbox{ mod }2$.
724: It is worth noting that
725: \begin{equation}
726: \label{symmetries}
727: Z_s(\lambda,\theta)=
728: Z_s(\lambda,\theta^{-1})
729: \quad
730: \text{ and }
731: \quad
732: \mu_{s,\lambda,\theta}(n,z)
733: =
734: \mu_{s,\lambda,\theta^{-1}}(n,-z).
735: \end{equation}
736:
737: For a fixed sequence $\us\in\{0,1\}^\Z$ and fixed parameters
738: $\lambda>0$, $\theta\in(1/\theta^*,\theta^*)$
739: we define on $\Omega_{\us}$ the
740: probability measure
741: %
742: \begin{equation*}
743: \mu_{\us,\lambda,\theta}:=
744: \prod_{j\in\Z} \mu_{s_j, \lambda, \theta}.
745: \end{equation*}
746: %
747:
748: By direct computations, one can check using (\ref {rprop})
749: that
750: for any function $f$
751: that depends only on the value
752: of finitely many ${\mathbf {\omega}}_k$'s, for
753: any fixed $s_j$, $s_{j+1}$ and
754: $({\mathbf {\omega}}_i)_{i \not= j, j+1}$,
755: \begin {eqnarray*}
756: \lefteqn {\sum_{{\mathbf {\omega}}_j, {\mathbf {\omega}}_{j+1}}
757: n_j r(z_j) \mu_{s_j, \lambda, \theta } ({\mathbf {\omega}}_j)
758: \mu_{s_{j+1}, \lambda, \theta } ({\mathbf {\omega}}_{j+1} )
759: f ( \Theta_{j+} ( {\mathbf {\omega}} ))
760: }\\
761: &=&
762: \sum_{{\mathbf {\omega}}_j, {\mathbf {\omega}}_{j+1}}
763: n_{j+1} r(z_{j+1}) \mu_{s_j, \lambda, \theta } ({\mathbf {\omega}}_j)
764: \mu_{s_{j+1}, \lambda, \theta } ({\mathbf {\omega}}_{j+1})
765: f ( {\mathbf {\omega}} )
766: \end {eqnarray*}
767: and a similar dentity holds for the jumps to the left.
768: It follows that given the
769: local parities $n_j+z_j=s_j\mbox{ mod }2$, the probability
770: measures $\mu_{\us,\lambda,\theta}$ are stationary for the dynamics.
771: These are the \emph{equilibrium Gibbs measures} of our system.
772: For a similar computation in the context of a simpler
773: one-component domain growth model see also \cite{balazs}.
774: % TO HERE
775:
776: Invariance under spatial translations is unfortunately lost
777: in this very general setup. In order to impose it, we restrict
778: ourselves to one of the following two choices: either $\us=\uzero$
779: or $\us=\uone$.
780:
781: \subsection{The hydrodynamic equations}
782:
783: For the rest of this section we fix either
784: $\us=\uzero$ or $\us=\uone$ and we do not denote any more the
785: dependence on $\us$.
786:
787: As we have mentioned already the globally conserved quantities
788: of our system are $\sum_j n_j$ and $\sum_j z_j$. In the
789: equilibrium regime $\mu_{\lambda,\theta}$ the averages of these
790: quantities are
791: % FROM HERE
792: %
793: \begin{equation*}
794: \rho
795: :=
796: \langle n_j \rangle_{\lambda,\theta}
797: =
798: \lambda\frac{\partial \log Z(\lambda,\theta)}{\partial \lambda},
799: \quad
800: \sigm
801: :=
802: \langle z_j \rangle_{\lambda,\theta}
803: =
804: \theta\frac{\partial \log Z(\lambda,\theta)}{\partial \theta}.
805: \end{equation*}
806: % TO HERE
807: %
808: These are the particle density (per site) and the average slope
809: of the height of the wall, in equilibrium.
810: It is easy to see that the map
811: $\R_+\times (1/\theta^*,\theta^*)
812: \ni (\lambda,\theta) \mapsto (\rho,\sigm) \in
813: \R_+\times\R$
814: is globally invertible. Indeed,
815: % FROM HERE
816: %
817: \begin{equation}
818: \left(
819: \begin{array}{cc}
820: {\partial \rho}/{\partial \lambda}
821: &
822: {\partial \rho}/{\partial \theta}
823: \\[3pt]
824: {\partial \sigm}/{\partial \lambda}
825: &
826: {\partial \sigm}/{\partial \theta}
827: \end{array}
828: \right)
829: =
830: \left(
831: \begin{array}{cc}
832: \var(n)
833: &
834: \cov(n,z)
835: \\[3pt]
836: \cov(n,z)
837: &
838: \var (z)
839: \end{array}
840: \right)
841: \left(
842: \begin{array}{cc}
843: \lambda^{-1}
844: &
845: 0
846: \\[3pt]
847: 0
848: &
849: \theta^{-1}
850: \end{array}
851: \right).
852: \label{invert}
853: \end{equation}
854: %
855: % TO HERE
856: %
857: So the gradient matrix on the left hand side of (\ref{invert}) is
858: everywhere invertible and this implies global invertibility of
859: the map $(\lambda, \theta)\mapsto(\rho, \sigm)$.
860: With slight abuse of notation we denote the components of the
861: inverse function $\lambda=\lambda(\rho,\sigm)$ and
862: $\theta=\theta(\rho,\sigm)$. From (\ref{symmetries}) it
863: follows that
864: \begin{equation}
865: \label{symmetries2}
866: \lambda(\rho,-\sigm)=\lambda(\rho,\sigm)
867: \quad\text{ and }\quad
868: \theta(\rho,-\sigm)=1/\theta(\rho,\sigm).
869: \end{equation}
870:
871: In order to guess the system of hydrodynamic equations we have
872: to see first how the infinitisimal generator acts on the
873: conserved quantities. An easy computation shows:
874: %
875: % FROM HERE
876: %
877: \begin{align*}
878: &
879: Ln_j
880: =
881: \big( n_{j-1} r(z_{j-1}) - n_{j} r(- z_{j}) \big)
882: -
883: \big( n_{j} r( z_{j}) - n_{j+1} r(-z_{j+1}) \big)
884: \\
885: &
886: Lz_j
887: =
888: \big( n_{j-1} r(z_{j-1}) + n_{j} r(-z_{j}) \big)
889: -
890: \big( n_{j} r(z_{j}) + n_{j+1} r(-z_{j+1}) \big)
891: \end{align*}
892: %
893: % TO HERE
894: %
895: On the right hand side of these equations we see \emph{discrete
896: gradients of fluxes}. This fact helps us guessing the
897: hydrodynamic equations. Applying the standard formal
898: manipulations to our gradient system (see e.g. \cite{fritz},
899: \cite{kipnislandim}) and using the straightforward identities
900: %
901: % FROM HERE
902: %
903: \begin{equation*}
904: \langle n_j r(\pm z_j ) \rangle_{\lambda,\theta}
905: =
906: \lambda\theta^{\pm1}
907: \end{equation*}
908: %
909: in the hydrodynamic limit taken with \emph{hyperbolic (Eulerian)
910: scaling} of space and time, we arrive at the system of PDEs
911: %
912: \begin{equation}
913: \left\{
914: \begin{array}{l}
915: \pt \rho +
916: \px \big( \lambda(\rho,\sigm)
917: (\theta(\rho,\sigm)-\theta(\rho,\sigm)^{-1})
918: \big)
919: =0,
920: \\[5pt]
921: \pt \sigm +
922: \px \big( \lambda(\rho,\sigm)
923: (\theta(\rho,\sigm)+\theta(\rho,\sigm)^{-1})
924: \big)
925: =0.
926: \end{array}
927: \right.
928: \label{hydrodyneq}
929: \end{equation}
930: %
931:
932: Under growth conditions on the rate function $r(z)$, as
933: $z\to\infty$,
934: Yau's `relative entropy method' (see e.g. \cite{yau},
935: %\cite{ollavaradhanyau},
936: \cite{fritz}, \cite{kipnislandim})
937: in principle can be applied to our system of interacting
938: particles, resulting in the validity of the above system of PDEs in
939: the hydrodynamic limit, \emph{as long as the solutions are
940: smooth}.
941: %
942: %
943: % WORK IN PROGRESS
944: %
945: % TO HERE
946:
947: From the system (\ref{hydrodyneq}) we can derive the system
948: (\ref{ourpde2}) by taking a second limit:
949: We replace
950: $\rho(t,x)$ by $\alpha^{2/3}\rho(\alpha t,\alpha^{2/3}x)$
951: and
952: $\sigm(t,x)$ by $\alpha^{1/3} \sigm(\alpha t,\alpha^{2/3}x)$
953: We note that for small values of the variables $\rho$ and
954: $\sigm$,
955: %
956: \begin{equation*}
957: \lambda(\rho, \sigm)=\rho +o(\rho),
958: \qquad
959: \theta(\rho,\sigm)=1+c \sigm + o(\sigm).
960: \end{equation*}
961: %
962: where
963: %
964: \begin{equation*}
965: c=\left(\left.\frac{\partial^2 Z}{\partial
966: \theta^2}\right|_{\lambda=0,\theta=1} \right)^{-1}\in(0,\infty).
967: \end{equation*}
968: %
969: Letting now $\alpha\to0$, we arrive at (\ref{ourpde2}).
970: We should emphasize here that this scaling limit does not depend
971: much on the details of microscopic system. Also,
972: from any conservation law of the form
973: %
974: \begin{equation*}
975: \left\{
976: \begin{array}{l}
977: \pt\rho + \px J(\rho,u) =0
978: \\[5pt]
979: \pt u + \px K(\rho,u) =0
980: \end{array}
981: \right.
982: \end{equation*}
983: %
984: we would get (\ref{ourpde2}) under the same limiting procedure,
985: provided that
986: %
987: \begin{equation*}
988: J(\rho,u)=\rho u +o(\rho u),
989: \quad
990: K(\rho,u)=\rho + o(\rho),
991: \quad
992: \text{ as }
993: \rho,u\to 0.
994: \end{equation*}
995: %
996: This indicates that (\ref{ourpde2}) is valid for a wider class
997: of microscopic systems.
998:
999:
1000:
1001: \section{Analysis of the PDE}
1002:
1003: We are now going to see how the methods developed in the PDE
1004: literature (see \cite{hormander, serre, smoller}) can be applied
1005: to our system. In order to put things into perspective, we
1006: briefly recall general
1007: results and see how they can be applied in the context of our
1008: system (\ref{ourpde}).
1009:
1010: \subsection{Two-component systems of hyperbolic conservation laws}
1011:
1012: For a generic two-component system we shall use the notation
1013: $v=v(t,x)=(v_1(t,x),v_2(t,x))^T$. (The superscript $^T$ will
1014: denote transposition of vectors/matrices.) The generic
1015: two-component system is
1016: %
1017: \begin{equation}
1018: \pt v + \px J (v) = 0,
1019: \label{hcl}
1020: \end{equation}
1021: %
1022: where $v\mapsto J(v)=(J_1(v),J_2(v))^T$ is a smooth vector field
1023: over $\R\times\R$. $J$ is the flux of the flow of
1024: the conserved quantity $v$. The initial conditions are specified by
1025: %
1026: \begin{equation}
1027: v(0,x)=v^{(0)}(x).
1028: \label{incond}
1029: \end{equation}
1030: %
1031:
1032: For a (possibly vector- or matrix valued)
1033: function $f=f(v)$ we denote the gradient with respect to the
1034: $v$-variables
1035: $\nabla f =
1036: ({\partial f}/{\partial v_1},{\partial f}/{\partial v_2})$.
1037: For classical \emph{smooth} solutions $v(t,x)$, (\ref{hcl}) is
1038: equivalent to
1039: %
1040: \begin{equation}
1041: \pt v + \nabla J \cdot \px v = 0
1042: \label{smoothhcl}
1043: \end{equation}
1044: %
1045: (we use $\cdot$ to indicate products of matrices).
1046:
1047: As a technical device one usually also considers the so-called
1048: viscous equations
1049: %
1050: \begin{equation}
1051: \pt v + \nabla J \cdot \px v = \vareps \px^2 v.
1052: \label{visc}
1053: \end{equation}
1054: %
1055: Existence and unicity of smooth solution $v^{(\vareps)}(t,x)$ of
1056: (\ref{visc}), for any bounded and smooth initial conditions
1057: (\ref{incond}) is
1058: guaranteed by the smoothening effect of the \emph{artificial
1059: viscosity term} on the right hand side. One hopes that physically
1060: acceptable (stable) solutions of the original system (\ref{hcl})
1061: can be obtained as a \emph{strong} limit of the viscous solution
1062: $v^{(\vareps)}(t,x)$, as $\vareps\to 0$. The existence of this
1063: strong limit is a very difficult problem and is
1064: a main object of investigation in the context of hyperbolic
1065: conservation laws.
1066:
1067: In our case (\ref{ourpde2}) the two components are
1068: $
1069: v_1=\rho$,
1070: $v_2=u$,
1071: and the corresponding fluxes are
1072: $
1073: J_1(\rho,u)=\rho u$,
1074: $
1075: J_2(\rho,u)=\rho
1076: $.
1077: The inviscid system is (\ref{ourpde2}). The (artificially)
1078: viscous system is
1079: %
1080: \begin{equation}
1081: \left\{
1082: \begin{array}{l}
1083: \pt \rho+\px (\rho u) =\vareps \px^2 \rho
1084: \\[5pt]
1085: \pt u + \px \rho = \vareps \px^2 u.
1086: \end{array}
1087: \right.
1088: \label{ourvisc}
1089: \end{equation}
1090: %
1091: The viscous solutions (which do exist and are unique) will be
1092: denoted by $\big(\rho^{(\vareps)}(t,x),
1093: u^{(\vareps)}(t,x)\big)$.
1094:
1095:
1096:
1097:
1098: \subsection{Hyperbolicity}
1099:
1100: One has to check that the matrix $\nabla J$ has two distinct
1101: real eigenvalues
1102: $\mu < \lambda$.
1103: The domain where
1104: this
1105: holds will be denoted
1106: \begin{equation*}
1107: {\cal D}_{\mathrm{hyp}}:=\{v\in\R\times\R: \mu(v)<\lambda(v)\}.
1108: \end{equation*}
1109:
1110: The corresponding left (row) and right (column) eigenvectors
1111: will be denoted by
1112: $l$ and $r$, respectively, $m$ and $s$. That is:
1113: %
1114: \begin{align}
1115: &l \cdot\nabla J = \lambda l,
1116: \qquad
1117: &
1118: \nabla J \cdot r = \lambda r,
1119: \label{eveqlambda}
1120: \\
1121: &m \cdot \nabla J = \mu m,
1122: \qquad
1123: &
1124: \nabla J \cdot s = \mu s.
1125: \label{eveqmu}
1126: \end{align}
1127: %
1128:
1129: For our system we find:
1130: %
1131: \begin{equation*}
1132: \nabla J=
1133: \left(
1134: \begin{array}{cc}
1135: u & \rho
1136: \\
1137: 1 & 0
1138: \end{array}
1139: \right),
1140: \end{equation*}
1141: %
1142: and
1143: %
1144: \begin{align}
1145: &
1146: \lambda=+\frac12(\sqrt{u^2+4\rho}+u),
1147: \qquad
1148: \phantom{m}
1149: l=(\lambda,\rho),
1150: \qquad
1151: \phantom{s}
1152: r=(\lambda,1)^T,
1153: \label{ourevlambda}
1154: \\
1155: &
1156: \mu=-\frac12(\sqrt{u^2+4\rho}-u),
1157: \qquad
1158: \phantom{l}
1159: m=(\mu,\rho),
1160: \qquad
1161: \phantom{r}
1162: s=(\mu,1)^T.
1163: \label{ourevmu}
1164: \end{align}
1165: %
1166: Note that
1167: $ l \cdot s = m \cdot r = 0$,
1168: as it should be.
1169:
1170: We conclude that for our system,
1171: \begin{equation*}
1172: {\cal D}_{\mathrm{hyp}}=
1173: \{(\rho,u)\in\R\times\R:u^2+4\rho>0\}.
1174: \end{equation*}
1175: %
1176: Note that in the physically relevant
1177: domain with non-negative densities
1178: \begin{equation*}
1179: {\cal D}_{\mathrm{ph}}:=
1180: \{(\rho,u)\in\R\times\R:\rho\ge0\},
1181: \end{equation*}
1182: there is one single point
1183: where strict hyperbolicity is lost, namely $(\rho,u)=(0,0)$.
1184: On the other hand, we found that the system is still hyperbolic
1185: in the physically meaningless domain
1186: ${\cal D}_{\mathrm {hyp}} \setminus
1187: {\cal D}_{\mathrm {ph}}
1188: = \{(\rho,u)\in{\cal D}_{\mathrm{hyp}}: \rho<0\}
1189: \not=\emptyset$.
1190: At the moment nothing seems to prevent solutions to flow into
1191: this domain.
1192: Later we
1193: shall see that Lax's maximum principle (valid for stable entropy
1194: solutions) takes care of this problem.
1195:
1196: \subsection{Riemann invariants, characteristics}
1197:
1198: In the generic two-component case, we are looking for scalar
1199: functions
1200: ${\cal D}_{\mathrm{hyp}}\ni v\mapsto w(v)\in \R$
1201: and space-time trajectories
1202: $\R_+\ni t\mapsto\xi(t)\in \R$ such that
1203: for smooth solutions of (\ref{hcl}) (or, equivalently, of
1204: (\ref{smoothhcl}))
1205: $w$ is conserved along the trajectory $\xi(t)$, i.e.
1206: %
1207: \begin{equation*}
1208: \frac{d}{dt} w\big( v(t,\xi(t) \big) = 0.
1209: \end{equation*}
1210: %
1211: Using (\ref{smoothhcl}) we find:
1212: %
1213: \begin{equation}
1214: \frac{d \xi}{d t}=
1215: \frac{(\nabla w \cdot \nabla J)\cdot \px v}
1216: { \nabla w \cdot \px v }.
1217: \label{wcons}
1218: \end{equation}
1219: %
1220:
1221: In order to solve (\ref{wcons}), $\nabla w$ must be a left
1222: eigenvector of
1223: the matrix $\nabla J$. It follows that this relation admits two
1224: solutions: one for each eigenvalue of $\nabla J$.
1225: We denote the two solutions by $w$
1226: (corresponding to the eigenvalue $\lambda$), respectively, by $z$
1227: (corresponding to the eigenvalue $\mu$). The gradients $\nabla
1228: w$, respectively $\nabla z$, are parallel to the row vectors $l$,
1229: respectively $m$, defined in (\ref{eveqlambda}), respectively
1230: (\ref{eveqmu}). In other words,
1231: % (see (\ref{leftdotright})),
1232: %
1233: \begin{align*}
1234: \nabla w \cdot s =0,
1235: \quad
1236: &
1237: \frac{ d \xi }{d t} = \lambda,
1238: \\
1239: \nabla z \cdot r =0,
1240: \quad
1241: &
1242: \frac{ d \xi }{d t} = \mu.
1243: \end{align*}
1244: %
1245: These equations, of course, do not determine uniquely the
1246: functions $w(v)$ and $z(v)$. Given two smooth, monotone maps
1247: $f,g:\R\to\R$, the transformation $\hat w:=f(w)$, $\hat z:=g(z)$
1248: leaves the above equations invariant. The functions $w$ and $z$
1249: are called the {\it Riemann invariants}, or {\it characteristic
1250: coordinates} of the problem.
1251:
1252: In our case the most convenient choice of the Riemann invariants
1253: $w$ and $z$ is the following: let
1254: %
1255: \begin{align*}
1256: &
1257: {\cal D}_w:=
1258: \{ (\rho,u) \in {\cal D}_{\mathrm{hyp}} : \sqrt{u^2+4\rho}-u\ge0\},
1259: \\
1260: &
1261: {\cal D}_z:=
1262: \{ (\rho,u) \in {\cal D}_{\mathrm{hyp}} : \sqrt{u^2+4\rho}+u\ge0\},
1263: \end{align*}
1264: %
1265: and define $w:{\cal D}_w\to\R$, $z:{\cal D}_z\to \R$ by the
1266: formulas:
1267: %
1268: \begin{align*}
1269: &
1270: w(\rho,u)
1271: =
1272: -\sqrt{\sqrt{u^2+4\rho}-u}
1273: \left(\sqrt{u^2+4\rho}+2u\right),
1274: \\
1275: &
1276: z(\rho,u)
1277: =
1278: -\sqrt{\sqrt{u^2+4\rho}+u}
1279: \left(\sqrt{u^2+4\rho}-2u\right).
1280: \end{align*}
1281: %
1282:
1283: \begin{figure}[ht]
1284: \centering
1285: %\vspace {-2cm}
1286: \resizebox{0.95 \textwidth}{!} {\includegraphics {zw.eps}}
1287: %\vspace {-1cm}
1288: \caption{
1289: Level lines of the Riemann invariants: $z=cst$ and $w=cst$}
1290: \end{figure}
1291:
1292:
1293: Note that
1294: ${\cal D}_w\cap{\cal D}_z={\cal D}_{\mathrm{ph}}$,
1295: so that both Riemann invariants are
1296: defined in the physically relevant subdomain.
1297:
1298: It is straightforward to check that both Riemann invariants $w$
1299: and $z$ defined above are \emph{convex} functions of the
1300: variables $(\rho,u)$. This fact will have crucial importance in
1301: later analysis.
1302:
1303: \subsection{Genuine nonlinearity}
1304:
1305: In plain words, genuine nonlinearity of a two-component system
1306: of hyperbolic conservation laws means that on the level curves
1307: $w(v)=\mbox{const.}$,
1308: respectively $z(v)=\mbox{const.}$, the characteristic speed
1309: $\mu$, respectively $\lambda$, varies strictly monotonically.
1310: Formally:
1311: %
1312: \begin{equation*}
1313: \left.
1314: \frac{\partial \lambda}{\partial w} \right|_z
1315: \not=0\not=
1316: \left. \frac{\partial \mu}{\partial z} \right|_w.
1317: \end{equation*}
1318: %
1319: Performing straightforward computations this turns out to be
1320: equivalent to
1321: %
1322: \begin{equation*}
1323: \nabla \lambda \cdot r
1324: \not=0\not=
1325: \nabla \mu \cdot s.
1326: \end{equation*}
1327: %
1328: That is: the characteristic speeds $\lambda$ and $\mu$
1329: vary strictly monotonically in the direction of their corresponding
1330: right eigenvectors.
1331:
1332: In our case, given the formulas (\ref{ourevlambda}) and
1333: (\ref{ourevmu}) we easily get
1334: %
1335: \begin{equation*}
1336: \nabla \lambda \cdot r
1337: = \frac{2\lambda}{\lambda-\mu},
1338: \qquad
1339: \nabla \mu \cdot s
1340: = \frac{2\mu}{\mu-\lambda}.
1341: \end{equation*}
1342: %
1343: Recall from (\ref{ourevlambda}), (\ref{ourevmu}) that
1344: on ${\cal D}_{\mathrm{ph}}$ we have $\mu\le0\le\lambda$,
1345: with strict inequalities for $\rho>0$.
1346: We conclude that our system is genuinely nonlinear in the
1347: interior of the physically relevant domain
1348: ${\cal D}_{\mathrm{ph}}$.
1349: On the half lines $\rho=0$,
1350: $u\le0$, respectively, $\rho=0$, $u\ge0$ (on the boundary
1351: of ${\cal D}_{\mathrm{ph}}$) genuine nonlinearity
1352: of the first, respectively, of the second, characteristic speed
1353: is lost.
1354:
1355:
1356: \subsection{Weak solutions, shocks, Rankine-Hugoniot conditions}
1357:
1358: As it is well-known, a nonlinear system of hyperbolic conservation laws
1359: (\ref{hcl}) can develop
1360: singularities (e.g. discontinuities), irrespectively of the
1361: smoothness of the initial conditions. A
1362: \emph{generalized} or \emph{weak} solution of (\ref{hcl}), (\ref{incond})
1363: in a space-time domain is a
1364: bounded, measurable function $(t,x)\mapsto v(t,x)$ satisfying
1365: %
1366: \begin{align}
1367: \label{weaksln}
1368: \int_{-\infty}^\infty\int_0^\infty
1369: \big\{\pt\phi(t,x)
1370: \cdot v(t,x)
1371: +
1372: \px\phi(t,x) \cdot
1373: &
1374: J(v(t,x))\big\} dtdx
1375: +
1376: \\
1377: &
1378: +
1379: \int_{-\infty}^\infty \phi(0,x)\cdot v^{(0)}(x) dx
1380: =
1381: 0
1382: \notag
1383: \end{align}
1384: %
1385: for any row vector valued test function $\phi=(\phi_1,\phi_2)$
1386: with compact support in the respective space-time domain. This
1387: last equation is
1388: obtained by a formal integration by parts. It is easily
1389: seen that a strong (smooth) solution is also a weak solution.
1390:
1391: Assuming a (locally) piecewise $C^1$ solution with a spatially
1392: isolated jump discontinuity at some space-time position
1393: $(t,x)\in\R_+\times\R$, one derives the
1394: Rankine-Hugoniot conditions which relate the left- and right
1395: limits of the function $x\mapsto v(t,x)$ at the discontinuity and the
1396: propagation speed of the discontinuity:
1397: %
1398: \begin{equation}
1399: \frac{J_1(v(t,x^+)) - J_1(v(t,x^-)) }{v_1(t,x^+) - v_1(t,x^-) }
1400: =\s=
1401: \frac{J_2(v(t,x^+)) - J_2(v(t,x^-)) }{v_2(t,x^+) - v_2(t,x^-) },
1402: \label{rankine}
1403: \end{equation}
1404: %
1405: where $\s$ is the propagation speed of the discontinuity, i.e.
1406: the slope in space-time of the line of discontinuity. (\ref{rankine})
1407: is derived from (\ref{weaksln}) by an elementary local
1408: argument, using the
1409: divergence theorem (in space-time). Given the two independent
1410: relations in (\ref{rankine}), any three of the five values
1411: $v_1(t,x^-)$, $v_1(t,x^+)$, $v_2(t,x^-)$, $v_2(t,x^+)$, $\s$
1412: determine the
1413: other two. This imposes a serious restriction on the possible
1414: jump discontinuities of weak solutions. Note that the conditions
1415: are left-right symmetric.
1416:
1417: We turn now to our system (\ref{ourpde2}). We denote by
1418: $(\rho^{\text{left}},u^{\text{left}})$,
1419: respectively
1420: $(\rho^{\text{right}},u^{\text{right}})$,
1421: the values of the component functions
1422: at the two sides of the presumed discontinuity.
1423: The Rankine-Hugoniot conditions are:
1424: %
1425: \begin{equation}
1426: \label{ourrankine1}
1427: \frac
1428: {u^{\text{right}}\rho^{\text{right}}-u^{\text{left}}\rho^{\text{left}}}
1429: {\rho^{\text{right}}-\rho^{\text{left}}}
1430: =\s=
1431: \frac
1432: {\rho^{\text{right}}-\rho^{\text{left}}}
1433: {u^{\text{right}}-u^{\text{left}}}.
1434: \end{equation}
1435: %
1436: Given the value at one side of the discontinuity, the value at
1437: the other side as function of propagation speed is expressed as
1438: follows:
1439: %
1440: \begin{equation}
1441: \label{ourrankine2}
1442: \rho^{\text{right}}=\s^2-\s u^{\text{left}},
1443: \qquad
1444: u^{\text{right}}=\s-\frac{\rho^{\text{left}}}{\s}
1445: \end{equation}
1446: %
1447: Note that $\rho^{\text{right}}$, respectively,
1448: $u^{\text{right}}$,
1449: is expressed as function of
1450: $\s$ and $u^{\text{left}}$, respectively, as function of
1451: $\s$ and $\rho^{\text{left}}$,
1452: only. (In principle, both should be expressed as functions of
1453: $\s$, $\rho^{\text{left}}$ and $u^{\text{left}}$.)
1454: This is a special feature of our system.
1455:
1456: The propagation speed, as function of the values of the
1457: components on both sides of the discontinuity, is expressed as:
1458: %
1459: \begin{equation*}
1460: \s_{\pm}=
1461: \pm\frac12
1462: \big\{
1463: \sqrt{(u^{\text{right}})^2+4\rho^{\text{left}}}\pm u^{\text{right}}
1464: \big\}=
1465: \mp\frac12
1466: \big\{
1467: \sqrt{(u^{\text{left}})^2+4\rho^{\text{right}}}\mp u^{\text{left}}
1468: \big\}.
1469: \end{equation*}
1470: %
1471:
1472: Lax's condition of stability for Rankine-Hugoniot
1473: discontinuities, \cite{lax1}, specified for two-component
1474: systems reads as follows:
1475: Assume that the weak solution (\ref{weaksln}) of the
1476: two-component system (\ref{hcl}) is piecewise smooth, with a
1477: spatially
1478: isolated discontinuity with values $v^{\text{left}}$,
1479: respectively, $v^{\text{right}}$ on the two sides, propagating
1480: according to the Rankine-Hugoniot conditions (\ref{rankine}). The
1481: discontinuity is a stable \emph{back shock}, respectively,
1482: \emph{front shock}, according whether
1483: %
1484: \begin{equation}
1485: \label{backshock}
1486: \mu(v^{\text{right}})
1487: <\s<
1488: \min\{\mu(v^{\text{left}}),\lambda(v^{\text{right}})\},
1489: \end{equation}
1490: %
1491: or
1492: %
1493: \begin{equation}
1494: \label{frontshock}
1495: \max\{\lambda (v^{\text{right}}),\mu (v^{\text{left}})\}
1496: <\s<
1497: \lambda(v^{\text{left}}).
1498: \end{equation}
1499: %
1500: Rankine-Hugoniot discontinuities which do not obey either one of
1501: the conditions (\ref{backshock}) or (\ref{frontshock}), are
1502: unstable, physically not realisable.
1503:
1504: Tedious (but, in principle straightforward) computations show,
1505: that in the case of our system (\ref{ourpde}) the discontinuities
1506: propagating according to (\ref{ourrankine1}), or equivalently
1507: (\ref{ourrankine2}) are stable back shocks if $\s<0$ and stable
1508: front shocks if $\s>0$.
1509:
1510:
1511: \subsection{Entropies}
1512:
1513: Given the two-component system of conservation laws (\ref{hcl}),
1514: we look for \emph{additional} conserved quantities, i.e., for pairs
1515: of functions
1516: ${\cal D}_{\mathrm{hyp}}\ni v\mapsto (S(v),F(v))\in\R\times\R$
1517: which satisfy
1518: %
1519: \begin{equation}
1520: \pt S(v) + \px F(v)=0
1521: \label{entropycons}
1522: \end{equation}
1523: %
1524: for \emph{smooth solutions} of the original problem (\ref{hcl})
1525: (or, equivalently: for smooth solutions of (\ref{smoothhcl})).
1526: Indeed, (\ref{entropycons}) means, that $S(v(t,x))$ is globally
1527: conserved quantity, with flux $F(v(t,x))$.
1528: The pair of functions $(S,F)$
1529: is called \emph{entropy/flux} pair. Using the form
1530: (\ref{smoothhcl}), valid for smooth solutions of (\ref{hcl}),
1531: one finds the system of
1532: PDEs defining an entropy/flux pair:
1533: %
1534: \begin{equation}
1535: \nabla F = \nabla S\cdot \nabla J,
1536: \label{entropyeq1}
1537: \end{equation}
1538: %
1539: or, in extended form:
1540: %
1541: \begin{equation*}
1542: \label{entropyeq2}
1543: \frac{\partial F}{\partial v_k}
1544: =
1545: \sum_{l=1}^2
1546: \frac{\partial S}{\partial v_l}
1547: \frac{\partial J_l}{\partial v_k},
1548: \qquad
1549: k=1,2.
1550: \end{equation*}
1551: %
1552: This is a two-component linear hyperbolic system of PDEs for the
1553: two unknown functions $S$ and $F$ -- just well determined. There
1554: are various alternative equivalent ways of writing it. E.g.,
1555: eliminating the function $F$ we get a second order hyperbolic
1556: PDE (a wave equation with variable coefficients) for $S$:
1557: %
1558: \begin{equation*}
1559: \label{entropyeq3}
1560: \frac{\partial J_1}{\partial v_2}
1561: \frac{\partial^2 S}{\partial v_1^2}
1562: +
1563: \left(
1564: \frac{\partial J_2}{\partial v_2}-
1565: \frac{\partial J_1}{\partial v_1}
1566: \right)
1567: \frac{\partial^2 S}{\partial v_1 \partial v_2}
1568: +
1569: \frac{\partial J_2}{\partial v_1}
1570: \frac{\partial^2 S}{\partial v_2^2}
1571: =0.
1572: \end{equation*}
1573: %
1574: Or, changing variables to the characteristic coordinates $(w,z)$:
1575: %
1576: \begin{equation*}
1577: \label{entropyeq4}
1578: \frac{\partial F}{\partial w} =
1579: \lambda \frac{\partial S}{\partial z},
1580: \qquad
1581: \frac{\partial F}{\partial z} =
1582: \nu \frac{\partial S}{\partial w}.
1583: \end{equation*}
1584: %
1585: Or, eliminating $F$ between these two equations:
1586: %
1587: \begin{equation*}
1588: \label{entropyeq5}
1589: \frac{\partial^2 S}{\partial w\partial z}
1590: =
1591: \frac{1}{\lambda-\mu}
1592: \left(
1593: \frac{\partial \mu}{\partial w}
1594: \frac{\partial S}{\partial z}
1595: -
1596: \frac{\partial \lambda}{\partial z}
1597: \frac{\partial S}{\partial w}
1598: \right).
1599: \end{equation*}
1600: %
1601: These last two forms explicitly show the wave-character of the
1602: entropy equations (\ref{entropyeq1}).
1603: Of particular importance are those entropy/flux pairs for
1604: which the function $v\mapsto S(v)$ is {\it convex}.
1605: Such pairs will
1606: be simply called (with slight abuse of terminology)
1607: \emph{convex entropy/flux pairs}.
1608:
1609: In the case of our system (\ref{ourpde2}) the entropy equations,
1610: written in terms of the physical variables $\rho$ and $u$, are:
1611: %
1612: \begin{equation*}
1613: \label{ourentropyeq1}
1614: \frac{\partial F}{\partial \rho}=
1615: u\frac{\partial S}{\partial \rho}+
1616: \frac{\partial S}{\partial u},
1617: \qquad
1618: \frac{\partial F}{\partial u}=
1619: \rho \frac{\partial S}{\partial \rho}
1620: \end{equation*}
1621: %
1622: Or, eliminating $F$:
1623: %
1624: \begin{equation}
1625: \label{ourentropyeq2}
1626: \rho \frac{\partial^2 S}{\partial\rho^2}
1627: -u \frac{\partial^2 S}{\partial\rho\partial u}
1628: -\frac{\partial^2 S}{\partial u^2}
1629: =0
1630: \end{equation}
1631: %
1632:
1633: The existence of a strictly convex entropy/flux pair,
1634: \emph{globally defined} on
1635: ${\cal D}_{\mathrm{ph}}=\{(\rho,u): \rho\ge 0, \ u\in\R\}$ and
1636: with $S$ bounded from below is very important, since the
1637: applicability of Lax's Maximum Principle cited in the next
1638: subsection relies on it.
1639: Here it is:
1640: %
1641: \begin{equation}
1642: \label{ourfirstentropy}
1643: S(\rho,u)=\rho\log\rho+\frac{u^2}{2},
1644: \qquad
1645: F(\rho,u)=u\rho(\log\rho+1).
1646: \end{equation}
1647: %
1648:
1649: Lax's `entropy wave construction' (cf. \cite{lax2})
1650: applies
1651: also to our system (\ref{ourpde}). Since these
1652: computations are rather involved, we do not reproduce them
1653: here. Let us just point out, that this robust method ensures the
1654: existence of a sufficiently rich family of convex entropy/flux
1655: pairs in any fixed subdomain compactly contained in ${\cal
1656: D}_{\mathrm{ph}}$.
1657:
1658: There are also other (more ad hoc) methods of constructing
1659: entropy/flux pairs. Following, e.g., the ideas of
1660: \cite{lionsperthametadmor} we may try to find so called
1661: similarity solutions of the entropy equation
1662: (\ref{ourentropyeq2})
1663: of the form:
1664: %
1665: \begin{equation}
1666: \label{similarity}
1667: S(\rho,u)=\rho^{\alpha}\phi(\rho^\beta u)
1668: \end{equation}
1669: %
1670: Elementary
1671: computations show that $\beta=-1/2$ is the only choice
1672: consistent with (\ref{ourentropyeq2}).
1673: Inserting (\ref{similarity}), with $\beta=-1/2$
1674: into (\ref{ourentropyeq2})
1675: we find the following ordinary differential equation for the
1676: function $\phi:\R\to\R$:
1677: %
1678: \begin{equation}
1679: \label{similarityequ}
1680: 3(y^2-4/3)\phi''(y)+
1681: (5-8\alpha)y\phi'(y)+
1682: 4\alpha(\alpha-1)\phi(y)=0.
1683: \end{equation}
1684: %
1685: Any solution of (\ref{similarityequ}), with any $\alpha\in\R$
1686: fixed provides an entropy of our system, via (\ref{similarity}).
1687: So, we are able to construct a sufficiently rich family of
1688: entropy/flux pairs to our system (\ref{ourpde2}).
1689:
1690: \subsection{Entropy solutions}
1691:
1692: A weak solution $(t,x)\mapsto v(t,x)$ of the generic system
1693: (\ref{hcl})
1694: is called \emph{entropy solution} if for any convex entropy/flux pair
1695: $(S,F)$ we have
1696: %
1697: \begin{equation}
1698: \pt S +\px F \le 0
1699: \label{entropycond}
1700: \end{equation}
1701: %
1702: in the sense of distributions, i.e., for any \emph{positive}
1703: test function $(t,x)\mapsto \phi(t,x)$
1704: %
1705: \begin{equation*}
1706: \int_{-\infty}^\infty\int_0^\infty
1707: \left\{\pt \phi(t,x)S(v(t,x)) + \px \phi(t,x) F(v(t,x))\right\}
1708: dtdx \ge 0.
1709: \end{equation*}
1710: %
1711:
1712: Entropy solutions are the only physically admissible, stable
1713: ones among the weak solutions. Strong limits of all convergent
1714: approximation schemes (such as vanishing viscosity or various
1715: convergent
1716: finite difference schemes) result in entropy solutions. It is
1717: also expected that convergent hydrodynamic limits of interacting
1718: particle systems result in entropy solution of the corresponding
1719: hyperbolic conservation laws.
1720: For piecewise smooth weak solutions, Lax's stability
1721: condition for the shocks mentioned in a previous paragraph is
1722: equivalent with the entropy conditions (\ref{entropycond}).
1723:
1724: Of particular interest is the following Maximum Principle, due to
1725: P. Lax, see e.g. \cite{lax2}.
1726:
1727: \smallskip
1728:
1729: \noindent
1730: {\bf Maximum Principle for Entropy Solutions.}
1731: \emph{
1732: Assume that the following two conditions hold
1733: \begin{enumerate}[(i)]
1734: \item
1735: The Riemann invariants $v\mapsto w(v)$ and $v\mapsto z(v)$
1736: of the system of hyperbolic conservation laws (\ref{hcl})
1737: are (globally) convex functions of $v$.
1738: \item
1739: There exists a globally defined convex entropy/flux pair, with
1740: entropy function bounded from below.
1741: \end{enumerate}
1742: Then, starting with bounded initial data,
1743: $\sup_{-\infty<x<\infty}|v^{(0)}(x)|<\infty$, along entropy
1744: solutions $(t,x)\mapsto v(t,x)$ the maximum values of the Riemann
1745: invariants,
1746: $\sup_{-\infty<x<\infty}w(v(t,x))$
1747: and
1748: $\sup_{-\infty<x<\infty}z(v(t,x))$
1749: do not increase with $t$.
1750: }
1751:
1752: \smallskip
1753:
1754: \noindent{\sl Remark:}
1755: The same statement applies for solutions $v^{(\vareps)}(t,x)$ of
1756: the viscous system (\ref{visc}) --- this follows from the
1757: classical maximum principle. If $v^{(\vareps)}$ converges
1758: strongly as $\vareps\to 0$, then the limiting $v$ is in fact an
1759: entropy solution of the inviscid system (\ref{hcl}) and forcibly
1760: it obeys Lax's Maximum Principle. It is not clear whether all
1761: entropy solutions arise as limits of viscous solutions, with
1762: vanishing viscosity. A general proof of the Maximum Principle for
1763: entropy solutions can be found in \cite{lax2}.
1764:
1765:
1766: Applying this theorem to our system we find that if we start with
1767: bounded initial data $x\mapsto(\rho^{(0)}(x),u^{(0)}(x))\in{\cal
1768: D}_{\mathrm{ph}}$
1769: (that is: with non-negative initial density) then \emph{entropy
1770: solutions will stay in the physical domain, i.e., for any $t\ge0$
1771: $x\mapsto(\rho(t,x),u(t,x))\in{\cal D}_{\mathrm{ph}}$.}
1772: (See Fig. 2 for graphical representation of the level curves
1773: $w(\rho,u)=\text{const.}$ and $z(\rho,u)=\text{const}$.) This is
1774: a very important consequence of the Maximum Principle: as we
1775: already mentioned, a priori we could not see any reason banning a
1776: (physically relevant) solution from flowing out into the
1777: physically meaningless domain with $\rho<0$.
1778:
1779: In the case of isentropic gas dynamics, (\ref{igd}), choosing
1780: convex versions of the Riemann invariants $w$ and $z$, for any
1781: $w_{\text{max}}\in\R$, $z_{\text{max}}\in\R$, the domains
1782: %
1783: \begin{equation*}
1784: \{(\rho,m)\in\R_+\times\R: w(\rho,m)\le w_{\text{max}},
1785: z(\rho,m)\le z_{\text{max}}\}
1786: \end{equation*}
1787: %
1788: are compact. So starting with
1789: bounded initial data global boundedness of (viscous and) entropy
1790: solutions is guaranteed by the Maximum Principle. This is
1791: unfortunately not the case for our system. The domains
1792: %
1793: \begin{equation*}
1794: \{(\rho,u)\in {\cal D}_{\mathrm{ph}}:
1795: w(\rho,u)\le w_{\text{max}},
1796: z(\rho,u)\le z_{\text{max}}\}
1797: \end{equation*}
1798: %
1799: are \emph{not} compact, see Fig. 2. So here is an open question:
1800: \emph{Is it the case, that if the initial data (\ref{ouric}) are
1801: bounded then the solutions
1802: $\big(\rho^{(\vareps)}(t,x), u^{(\vareps)}(t,x)\big)$
1803: of the viscous equation (\ref{ourvisc}) stay bounded for ever?
1804: Similarly: is it the case that entropy solutions of
1805: (\ref{ourpde2}) with bounded initial data stay bounded?}
1806: We guess that the answer to these questions are affirmative, but
1807: we could not prove this yet.
1808:
1809: \subsection{Vanishing viscosity, existence of entropy solutions}
1810:
1811: The \emph{existence of entropy solutions} for a two-component
1812: syetem of hyperbolic conservationlaws (\ref{hcl}) is a
1813: notoriously difficult question. The most powerful approach seems
1814: to be the program initiated by R. DiPerna in \cite{diperna1},
1815: completed for the case of isentropic gas dynamics (\ref{igd})
1816: in \cite{diperna2}, then refined and extended in
1817: Lions et al. \cite{lionsperthametadmor} and in several other papers.
1818:
1819: In \cite{diperna1}, DiPerna proves the following result:
1820:
1821: \smallskip
1822:
1823: \noindent{\bf DiPerna's Theorem.}
1824: \emph{
1825: Consider the two-component system of hyperbolic
1826: conservation laws (\ref{hcl}) and the corresponding viscous
1827: system (\ref{visc}). Assume that
1828: \begin{enumerate}[(i)]
1829: \item
1830: The Riemann invariants $v\mapsto w(v)$ and $v\mapsto z(v)$ are
1831: convex. (More precisely: there are convex choices of the Riemann
1832: invariants. See subsection 4.3.)
1833: \item
1834: The system is genuinely nonlinear. (See subsection 4.4.)
1835: \end{enumerate}
1836: Let $\cal C$ be a domain compactly
1837: contained in ${\cal D}_{\mathrm{hyp}}$ and assume that the
1838: sequence of solutions $v^{(\vareps)}(t,x)$, $t\in[0,T]$,
1839: $x\in\R$, $\vareps\to 0$, of the viscous systems (\ref{visc}),
1840: (\ref{incond})
1841: takes values from $\cal C$. Then there is a subsequence
1842: $v^{(\vareps^{'})}(t,x)$ which converges strongly in
1843: $L^1_{\mathrm{loc}}([0,T]\times\R)$. The limit $v(t,x)$ is
1844: entropy solution of the system (\ref{hcl}).
1845: }
1846:
1847: \smallskip
1848:
1849: \noindent
1850: {\sl Some Remarks:}
1851: \begin{enumerate}[(1)]
1852: \item
1853: The proof relies on the construction of Lax's `entropy waves',
1854: hinted at in subsection 4.6 and essentially on the so-called
1855: compensated compactness method developed by Murat and Tartar. We
1856: do not have a chance to reproduce here any technical part of the
1857: proof.
1858: \item
1859: It is assumed
1860: that the viscous solutions stay in the domain $\cal
1861: C$.
1862: However, even in this form the
1863: theorem is technically very-very difficult. Extra difficulties
1864: arise by relaxing this condition and imposing conditions {\em only on
1865: the initial data}: in the isentropic gas dynamics and in our case
1866: too, the solution data will typically flow to the boundary of the
1867: domain of hyperbolicity and genuine nonlinearity, $\rho=0$, where
1868: this theorem is not any more valid.
1869: \item
1870: For extensions, physically more satisfactory formulations and
1871: enormous further technical difficulties see e.g. \cite{diperna2},
1872: \cite{lionsperthametadmor}, etc.
1873: \end{enumerate}
1874:
1875:
1876: This theorem can be applied in a straightforward way for domains
1877: $\cal C$, compactly contained in ${\cal D}_{\mathrm{ph}}$. We can
1878: add to this that if initially
1879: \begin{equation*}
1880: \max_x w(\rho^{(0)}(x), u^{(0)}(x))<0
1881: \quad\mathrm{or}\quad
1882: \max_x z(\rho^{(0)}(x), u^{(0)}(x))<0
1883: \end{equation*}
1884: then, due to the Maximum Principle, the viscous solutions
1885: $\rho^{(\vareps)}, u^{(\vareps)}$ are kept away from
1886: the `dangerous' vacuum line $\rho=0$, see Fig. 2. So, in this
1887: case one has to care only about the boundedness of the
1888: solutions.
1889:
1890: \bigskip
1891: \noindent
1892: {\bf Acknowledegments.}
1893: BT thanks illuminating discussions with
1894: M\'arton Bal\'azs, J\'ozsef Fritz and Benedek Valk\'o.
1895: We also thank Sophie Lemaire for kindly helping us producing Figure 2.
1896: Cooperation between the authors is partially supported by the
1897: French-Hungarian joint scientific research grant `Balaton'.
1898:
1899: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1900: \begin{thebibliography}{99}
1901:
1902: \bibitem{balazs}
1903: M. Bal\'azs:
1904: Microscopic structure of the shock in a domain growth model.
1905: {\sl Preprint} (2000)
1906:
1907: \bibitem{barabasistanley}
1908: A.L. Barab\'asi, E.H. Stanley:
1909: {\sl Fractal Concepts in Surface Growth.\/}
1910: Cambridge Univ. Press, 1995.
1911:
1912: \bibitem{diperna1}
1913: R. DiPerna:
1914: Convergence of approximate solutions to conservation laws.
1915: {\sl Arch. Rat. Mech. Anal.\/} {\bf 82}: 27-70 (1983)
1916:
1917: \bibitem{diperna2}
1918: R. DiPerna:
1919: Convergence of the viscosity method for isentropic gas dynamics.
1920: {\sl Commun. Math. Phys.} {\bf 91}: 1-30 (1983)
1921:
1922: \bibitem{fritz}
1923: J. Fritz:
1924: {\sl An Introduction to the Theory of Hydrodynamic Limits.\/}
1925: Graduate School of Mathematics, Univ. Tokyo, 2000.
1926:
1927: \bibitem {bertinigiacomin}
1928: L. Bertini, G. Giacomin:
1929: Stochastic Burgers and KPZ equations from particle systems,
1930: {\sl Comm. Math. Phys.} {\bf 183}: 571-607 (1997)
1931:
1932: \bibitem{hormander}
1933: L. H\"ormander:
1934: {\sl Lectures on Non-Linear Hyperbolic Differential Equations.\/}
1935: Springer, 1997.
1936:
1937: \bibitem{kardarparisizhang}
1938: M. Kardar, G. Parisi, Y.-C. Zhang:
1939: Dynamic scaling of growing interfaces.
1940: {\sl Phys. Rev. Lett.\/} {\bf 56}: 889-892 (1986)
1941:
1942: \bibitem{kipnislandim}
1943: C. Kipnis, C. Landim:
1944: {\sl Scaling Limits of Interacting Particle Systems.\/}
1945: Springer, 1999.
1946:
1947: \bibitem{lax1}
1948: P. Lax:
1949: Hyperbolic systems of conservation laws II.
1950: {\sl Commun. Pure Appl. Math.\/} {\bf 10}: 537-566 (1957)
1951:
1952: \bibitem{lax2}
1953: P. Lax:
1954: Shock waves and entropy.
1955: In: {\sl Contributions to Nonlinear Functional Analysis,\/}
1956: ed: E.A. Zarantonello. Academic Press, 1971.
1957:
1958: \bibitem{lax3}
1959: P. Lax:
1960: Shock waves, increase of entropy and loss of information.
1961: In: {\sl Seminar on Nonlinear PDEs, Berkeley, California 1983.\/}
1962: Springer, 1984.
1963:
1964: \bibitem{lionsperthametadmor}
1965: P.L. Lions, B. Perthame, E. Tadmor:
1966: Kinetic formulation of the isentropic
1967: gas dynamics and $p$-systems,
1968: {\sl Commun. Math. Phys.\/} {\bf 163}: 415-431 (1995)
1969:
1970: \bibitem{serre}
1971: D. Serre:
1972: {\sl Syst\`emes de lois de conservation, vol. 1 and 2},
1973: Diderot Editeur, 1996.
1974:
1975: \bibitem{smoller}
1976: J. Smoller:
1977: {\sl Shock Waves and Reaction-Diffusion equations.\/}
1978: (second edition) Springer, 1994.
1979:
1980: \bibitem{toth}
1981: B. T\'oth:
1982: The `true' self-avoiding walk with bond repulsion on $\Z$: limit
1983: theorems.
1984: {\sl Ann. Probab.\/} {\bf 23}: 1523-1556 (1995)
1985:
1986: \bibitem{tothwerner}
1987: B. T\'oth, W. Werner:
1988: The true self-repelling motion.
1989: {\sl Probab. Theory Rel. Fields\/} {\bf 111}: 375-452 (1998)
1990:
1991: \bibitem{yau}
1992: H.T. Yau:
1993: Relative entropy and hydrodynamics of Ginzburg-Landau models.
1994: {\sl Lett. math. Phys.\/} {\bf 22}: 63-80 (1991)
1995:
1996:
1997: \end{thebibliography}
1998: -----------------------------------------------------------------
1999: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2000:
2001: \vskip1cm
2002:
2003:
2004: \hbox{\sc
2005: \vbox{\noindent
2006: \hsize66mm
2007: B\'alint T\'oth\\
2008: Institute of Mathematics\\
2009: Technical University Budapest\\
2010: Egry J\'oszef u. 1.\\
2011: H-1111 Budapest, Hungary\\
2012: {\tt balint@math.bme.hu}
2013: }
2014: \hskip5mm
2015: \vbox{\noindent
2016: \hsize66mm
2017: Wendelin Werner\\
2018: D\'ept. de Math\'ematiques\\
2019: Universit\'e Paris-Sud\\
2020: B\^at. 425\\
2021: 91405 Orsay cedex, France\\
2022: {\tt wendelin.werner@math.u-psud.fr}
2023: }
2024: }
2025:
2026: \end {document}
2027:
2028:
2029:
2030:
2031:
2032:
2033:
2034:
2035:
2036:
2037:
2038:
2039:
2040:
2041:
2042:
2043:
2044:
2045: