1: \documentclass[11pt]{article}
2: \usepackage{amssymb,amsthm,amsmath,amsfonts}
3: \usepackage{epsfig}
4:
5: \theoremstyle{plain}
6: \begingroup
7: %\theorembodyfont{plain}
8: \newtheorem*{metathm}{Metatheorem}
9: \newtheorem{theorem}{Theorem}
10: \newtheorem{corollary}{Corollary}
11: \newtheorem{lemma}{Lemma}
12: %\newtheorem{prop}[thm]{Proposition}
13: %\newtheorem{metathm}{Metatheorem}[section]
14: %\newtheorem{conj}[thm]{Conjecture}
15: %\newtheorem*{refthm}{Theorem}
16: \endgroup
17:
18:
19: \theoremstyle{definition}
20: \newtheorem{remark}{Remark}
21:
22:
23: \newcommand{\nwc}{\newcommand}
24: \nwc{\bit}{\begin{itemize}}
25: \nwc{\eit}{\end{itemize}}
26: \nwc{\Levy}{L\'evy}
27: \nwc{\LK}{L\'evy-Khintchine}
28: \nwc{\LI}{L\'evy-It\^{o}}
29: \nwc{\CH}{Cole-Hopf}
30: \nwc{\Holder}{H\"{o}lder}
31: \nwc{\cadlag}{c\`{a}dl\`{a}g}
32: \nwc{\be}{\begin{equation}}
33: \nwc{\ee}{\end{equation}}
34: \nwc{\ba}{\begin{eqnarray}}
35: \nwc{\ea}{\end{eqnarray}}
36: \nwc{\la}{\label}
37: \nwc{\nn}{\nonumber}
38: \nwc{\Z}{\mathbb{Z}}
39: \nwc{\C}{\mathbb{C}}
40: \nwc{\E}{\mathbb{E}}
41: \nwc{\R}{\mathbb{R}}
42: \nwc{\N}{\mathbb{N}}
43: \nwc{\prob}{\mathbb{P}}
44: \nwc{\Skor}{\mathbb{D}}
45: \nwc{\attr}{\mathcal{A}}
46: \nwc{\PP}{\mathcal{P}}
47: \nwc{\PPE}{\mathcal{P}(E)}
48: \nwc{\M}{\mathcal{M}}
49: \nwc{\Tt}{T^{(t)}}
50: \nwc{\Ut}{U^{(t)}}
51: \nwc{\Vt}{V^{(t)}}
52: \nwc{\tlambda}{\tilde{\lambda}}
53: \nwc{\tbar}{\bar{t}}
54: \nwc{\gtx}{g^{(t)}_x}
55: \nwc{\order}{\prec}
56: \nwc{\law}{\stackrel{\mathcal{L}}{\rightarrow}}
57: \nwc{\eqd}{\stackrel{\mathcal{L}}{=}}
58: \nwc{\vp}{\varphi}
59: \nwc{\Vp}{\Phi}
60: \nwc{\psilevy}{\Psi}
61: \nwc{\ve}{\varepsilon}
62: \nwc{\veps}{\varepsilon}
63: \nwc{\eps}{\ve}
64: \nwc{\betarsc}{\theta}
65: \nwc{\cl}{c\'{a}dl\'{a}g}
66: \nwc{\qref}[1]{(\ref{#1})}
67: \nwc{\D}{\partial}
68: \nwc{\Ebar}{\bar{E}}
69: \nwc{\mmt}{m}
70: \nwc{\dnto}{\downarrow}
71: \nwc{\fzero}{F_\rho} %{F_{0,\rho}}
72: \nwc{\fone}{M_\rho} %{F_{1,\rho}}
73: \nwc{\ip}[1]{\langle #1 \rangle}
74: \nwc{\ipbig}[1]{\left\langle #1 \right\rangle}
75: \nwc{\Lip}{\mathop{\rm Lip}\nolimits}
76: \nwc{\Tmin}{T_{\min}}
77: \nwc{\Tmax}{T_{\max}}
78: \nwc{\Tgel}{T_{\rm gel}}
79: \nwc{\LL}{\mathcal{L}}
80: \nwc{\mudot}{\mu}
81: \nwc{\nudot}{\nu}
82: \nwc{\rme}{{\rm e}}
83: \nwc{\rmi}{{\rm i}}
84: \nwc{\nsup}{^{(n)}}
85: \nwc{\ksup}{^{(k)}}
86: \nwc{\jsup}{^{(j)}}
87: \nwc{\nksup}{^{(n_k)}}
88: \nwc{\inv}{^{-1}}
89: \nwc{\qxfac}{(1-\rme^{-qx})}
90: \nwc{\one}{\mathbf{1}}
91: \nwc{\psib}{\psi^{(b)}}
92: \nwc{\pss}{\psi_\#}
93: \nwc{\phss}{\Phi_\#}
94: \nwc{\asharp}{\gamma}
95: \nwc{\vfour}{\hat{v}}
96: \nwc{\Pss}{\Psi_\#}
97: \nwc{\pibar}{\Pi}
98: \nwc{\Ltail}{\bar{\Lambda}}
99: \nwc{\vb}{v^{(b)}}
100: \nwc{\spec}{\tilde{\beta}}
101: \nwc{\hv}{\hat{v}}
102: \nwc{\hvl}{\hat{v}_L}
103: \nwc{\lap}{{\,\cal L}}
104: \renewcommand{\Re}{\mathop{\rm Re}\nolimits}
105: \renewcommand{\Im}{\mathop{\rm Im}\nolimits}
106: \begin{document}
107: %
108: \title{Universality classes in Burgers turbulence}
109: \author{Govind Menon\textsuperscript{1} and Robert. L.
110: Pego\textsuperscript{2}}
111:
112: \date{\today}
113:
114: \maketitle
115: %
116: \begin{abstract}
117: We establish necessary and sufficient conditions for the shock statistics
118: to approach self-similar form in Burgers turbulence with \Levy\/
119: process initial data. The proof relies upon an elegant closure theorem
120: of Bertoin %~\cite{B_burgers} %({\em Comm. Math. Phys. {\bf 193}, 1998\/})
121: and Carraro and Duchon %~\cite{Duchon1} %%Can't put citations in abstract
122: that reduces the
123: study of shock statistics to Smoluchowski's coagulation equation with
124: additive kernel, and upon our previous characterization of the
125: domains of attraction
126: of self-similar solutions for this equation. %~\cite{MP1}.
127: %~({\em Comm. Pure. Appl. Math. {\bf 57}, 2004\/}).
128: \end{abstract}
129:
130: \medskip
131: \noindent
132: {\bf Keywords:\/} Burgers turbulence, Smoluchowski's coagulation equation,
133: \Levy\/ processes, dynamic scaling, regular variation,
134: agglomeration, coagulation, coalescence, shock statistics.
135:
136: \medskip
137: \noindent
138: {\bf MSC classification:\/} 60J75, 35R60, 35L67, 82C99
139: %
140: \medskip
141: \footnotetext[1]
142: {Division of Applied Mathematics, Box F, Brown University, Providence, RI 02912.
143: Email: menon@dam.brown.edu}
144: \footnotetext[2]{Center for Nonlinear Analysis, Department of
145: Mathematical Sciences, Carnegie Mellon University, Pittsburgh, PA 15213.
146: Email: rpego@cmu.edu}
147:
148: \pagebreak
149: \section{Introduction}
150: The construction of stochastic processes that are also weak solutions
151: to the equations of fluid mechanics is one approach to rigorous
152: mathematical theories of turbulence. This is poorly understood at
153: present, and we must
154: settle for insights from vastly simplified model problems. We consider
155: the invisicid Burgers equation
156: \begin{equation}
157: \label{E.burgers}
158: \D_t u + \D_x\left( \frac{u^2}{2} \right) =0, \quad t>0,\ x \in \R, \quad
159: u(x,0)=u_0(x),
160: \end{equation}
161: with random initial data $u_0$. The problem is to determine the
162: statistical properties of the \CH\ (entropy) solution $u(x,t)$ to
163: (\ref{E.burgers}), given the statistical properties of $u_0$.
164: There is a large literature on the subject; we refer to
165: Burgers' book~\cite{Burgers} and the
166: more recent survey articles~\cite{ES,Gurbatov2,Woc}.
167: The problem was proposed by Burgers as a
168: model for turbulence in incompressible fluids, but it has several
169: well-known flaws in this regard.
170:
171: Explicit solutions play a special role in the theory.
172: Burgers studied the case when $u_0$ is
173: white noise in his monograph~\cite{Burgers}.
174: His work remains the foundation for several rigorous results, which
175: culminate with the
176: complete solution by Frachebourg and Martin for the velocity and shock
177: statistics (see ~\cite{Frachebourg} and references therein).
178: The case when $u_0$ is a Brownian motion has attracted much attention
179: since the work of
180: She, Aurell and Frisch~\cite{She} and Sinai~\cite{Sinai}.
181: An elegant solution to
182: this problem was obtained by Bertoin~\cite{B_burgers} and Carraro and
183: Duchon~\cite{Duchon1}. More generally, these authors considered
184: initial data that comprise a \Levy\/ process
185: with only downward jumps (i.e., shocks). A \Levy\/ process
186: $X_x$ ($x \geq 0$) is a continuous-time random walk with stationary and
187: independent increments. It is determined completely
188: by its characteristic exponent $\Psi$, satisfying $\E(e^{ikX_x})
189: = e^{-x\Psi(k)}$, via the celebrated \LK\/ formula
190: \be
191: \label{E.LK1}
192: \Psi(k) = i b k +
193: \frac{\sigma^2 k^2}{2} + \int_{\R} \left(1-e^{iks}+
194: i ks\one_{|s|<1} \right) \Pi(ds), \quad k \in \R.
195: \ee
196: The process $X$ is the superposition of three independent
197: processes related to this formula: a Brownian motion
198: with variance $\sigma^2$ and drift $b \in \R$, a compound Poisson
199: process with jump measure $\Pi \one_{|s| \geq 1}$, and a pure
200: jump martingale
201: %\footnote{??check mean drift. 6/1/06: this is OK.}
202: with jump measure $\Pi \one_{|s| <1}$ (see~\cite[Ch.1]{B_book}).
203: The measure $\Pi$ is arbitrary, subject to the condition
204: $\int_\R (1\wedge s^2)\Pi(ds)<\infty$, where $a\wedge b$ means $\min(a,b)$.
205:
206: We say $X$ is {\em spectrally negative\/} if $\Pi$ is
207: concentrated on the half-line $s<0$. In all that follows we assume
208: \be
209: \la{E.initial}
210: u_0(x) = \left\{ \begin{array}{l} 0, \quad x < 0, \\
211: \text{a spectrally negative \Levy\/ process}, \quad x
212: \geq 0, \end{array} \right.
213: \ee
214: As we show below, we can always reduce to the case where $u_0(x)$ has zero mean
215: $\E(u_0(x))$ for all $x$,
216: and $\int_{-\infty}^0(|s|\wedge s^2) \Pi(ds) < \infty$. It is then
217: more convenient to use the Laplace exponent
218: \be
219: \label{E.LKu}
220: \psi(q) = -\Psi(-iq) = \frac{\sigma^2 q^2}{2} + \int_0^\infty \left(
221: e^{-qs} -1 + qs \right) \,\Lambda(ds), \quad q >0,
222: \ee
223: where $\Lambda((s,\infty)) = \Pi((-\infty,-s))$ for every $s>0$, so that
224: $\E(e^{qX_x}) = e^{x\psi(q)}$ and $\E(X_x)=x\psi'(0)=0$.
225:
226: For this class of initial data, Bertoin
227: proved a remarkable closure property for the entropy
228: solution of \qref{E.burgers}, namely:
229: $x\mapsto u(x,t)-u(0,t)$ {\em remains a spectrally negative
230: \Levy\/ process for all $t>0$.}
231: This closure property was first noted by Carraro and
232: Duchon in connection with their notion of statistical
233: solutions to Burgers equation~\cite{Duchon0}. That these
234: statistical solutions agree with the \CH\/ solution for spectrally
235: negative data was shown by Bertoin~\cite[Thm. 2]{B_burgers}.
236: The closure property fails if $u_0$ has positive jumps---These
237: positive jumps open into rarefaction waves for $t>0$, and this is
238: incompatible with the rigidity of sample paths of \Levy\/ processes.
239: An interesting formal analysis of closure properties of Burgers equation
240: is presented in~\cite{Duchon2}.
241:
242: Henceforth, we write $v(x,t)=u(x,t)-u(0,t)$ for brevity.
243: The \LK\ representation now implies that the law of the \Levy\/ process
244: $x\mapsto v(x,t)$ is completely described by a corresponding
245: ``\Levy\/ triplet'' $(b_t,\sigma_t^2, \Lambda_t)$.
246: The mean drift $b_t=\E(v(1,t))$ satisfies $b_t=b_0=0$ for every
247: $t \geq 0$. Moreover, for every $t >0$, $v(\cdot,t)$ is
248: of bounded variation, thus the variance $\sigma_t=0$. Consequently,
249: the law of $v(\cdot,t)$ is completely determined
250: by only the jump measure $\Lambda_t$ which contains the shock statistics.
251:
252: It is a striking fact, implicit in \cite{B_burgers}, that the evolution of
253: $\Lambda_t$ is described by {\em Smoluchowski's coagulation equation with
254: additive kernel}, an equation that arises in entirely different
255: areas such as the analysis of algorithms~\cite{Chassaing}, the kinetics of
256: polymerization~\cite{Ziff}, and cloud formation from
257: droplets~\cite{Golovin} (see~\cite{Aldous} for a review).
258: %% Smoluchowski's coagulation equation in the following theorem.
259: %% \begin{theorem}
260: %% Assume $u_0$ is a spectrally negative \Levy\/ process with \Levy\/ triplet
261: %% $(0, \sigma^2_0, \Lambda_0)$. With probability $1$, $v(\cdot,t)$, $t
262: %% \in [0,\infty)$ are \Levy\/ processes whose \Levy\/
263: %% measures form a continuous semigroup generated by Smoluchowski's coagulation
264: %% equation with rate $x+y$ as described in
265: %% \qref{E.smolburgers}.
266: %% \end{theorem}
267: What this means is that {\em mean-field theory is exact\/} for Burgers
268: equation with
269: initial data of the form \qref{E.initial}, i.e., random one-sided
270: data with stationary and independent increments.
271: We give a precise statement to this effect below in
272: Theorem~\ref{T.meanfield}. Several connections
273: between stochastic models of coalescence and Burgers turbulence are
274: reviewed in~\cite{B_icm}.
275:
276: Here, we use the closure property as a basis for a rigorous study of
277: {\em universality classes for dynamic scaling} in Burgers turbulence.
278: Our motivation is the
279: following. A central theme in studies of homogeneous isotropic turbulence in
280: incompressible fluids is the universality of the Kolmogorov
281: spectrum~\cite{Kolm}. A possible rigorous formulation of such
282: universality involves (a) the construction of stochastic
283: processes that mimic a `typical turbulent flow', and (b) a characterization
284: of the domains of attraction of these processes.
285: For Burgers turbulence,
286: step (a) consists of constructing exact solutions for special initial data, say
287: white noise or Brownian motion. In this article, we
288: carry out step (b) for initial data that satisfy \qref{E.initial}.
289:
290: Domains of attraction are studied in the classical limit theorems in
291: probability (e.g., the central limit theorem), and their process
292: versions (e.g., Donsker's invariance principle). For Smoluchowski's
293: coagulation equation with additive kernel, we characterized
294: all possible domains of attraction in~\cite{MP1},
295: a result akin to the classical limit theorems.
296: In this article we deal with a process version.
297: In all that follows, we consider the processes
298: $x\mapsto v(x,t)$ as elements
299: of the space $\Skor$ of right continuous paths $\R_+\to \R$
300: with left limits (\cadlag\/ paths) equipped with the Skorokhod
301: topology~\cite[Ch. VI]{Jacod}. The shock statistics determine
302: completely the law of this process (a probability measure on
303: $\Skor$). Approach to limiting forms will be phrased in terms of
304: weak convergence of probability measures on $\Skor$.
305:
306: Among the initial data we consider, the stable processes are of particular
307: importance because of their self-similarity. Let $X^\alpha, \alpha \in
308: (1,2]$ denote the stable process with Laplace exponent $q^\alpha$
309: ($\alpha=2$ corresponds to Brownian motion).
310: The corresponding jump measure
311: $\Lambda(ds)=s^{-1-\alpha}\,ds/\Gamma(-\alpha)$ for $\alpha<2$.
312: There is a one-to-one correspondence between (a) these stable
313: processes, (b) statistically self-similar solutions in
314: Burgers turbulence, and (c) self-similar solutions to Smoluchowski's
315: coagulation equation.
316: Precisely, this works as follows. Let $\alpha \in (1,2]$, and let $T^\alpha$
317: denote the first-passage process for $x\mapsto X^\alpha_x+x$, i.e.,
318: \be
319: \label{E.firstpass}
320: T^\alpha_x = \inf \{y \left| X^\alpha_y + y > x \right.\}.
321: \ee
322: Then for the solution to (\ref{E.burgers}) with
323: $u_0(x)=X^\alpha_x$ for $x \geq 0$,
324: $v(x,t)$ is statistically self-similar, with
325: \begin{equation}
326: \label{eq:selfsimilar}
327: v(x,t) \eqd
328: t^{1/\beta-1} V^\alpha_{xt^{-1/\beta}}
329: :=
330: t^{1/\beta-1}\left(xt^{-1/\beta} -
331: T^\alpha_{xt^{-1/\beta}} \right)
332: , \quad \; t,x>0,
333: \end{equation}
334: where $\beta= (\alpha-1)/\alpha$.
335: Here $\eqd$ means both processes define the same measure on $\Skor$.
336: The process $T^\alpha$ is a pure jump \Levy\ process with
337: \Levy\ measure $f_\alpha(s)
338: \,ds$, where $f_\alpha$ is the number density profile
339: of a self-similar solution to Smoluchowski's coagulation
340: equation~\cite[Sec 6]{MP1}:
341: \be
342: \label{E.ndensity}
343: f_\alpha(s) = \frac{1}{\pi} \sum_{k=1}^\infty \frac{(-1)^{k-1}
344: s^{k\beta-2}}{k!} \Gamma(1+k-k \beta) \sin \pi k \beta, \quad \alpha
345: \in (1,2].
346: \ee
347: These solutions are related to classical distributions in probability
348: theory by rescaling. If $p(s;\alpha,2-\alpha)$ denotes the density of
349: a maximally-skewed \Levy\ stable law~\cite[XVII.7]{Feller} we have
350: \cite{B_eternal,MP1}
351: \be
352: \label{E.max_stable}
353: f_\alpha (s) = s^{\beta-2} p(s^\beta;\alpha, 2-\alpha).
354: \ee
355: By the \LI\ decomposition~\cite[Thm 1.1]{B_book} and
356: (\ref{eq:selfsimilar}) we may conclude
357: that the magnitudes of shocks in $u(\cdot,t)$ form a Poisson point
358: process valued in $(0,\infty)$ whose characteristic measure is
359: \begin{equation}
360: \label{eq:shock_intensity}
361: \Lambda^\alpha_t(ds) = t^{1-2/\beta}
362: f_\alpha\left( {s t^{1-1/\beta}} \right) \, ds.
363: \end{equation}
364:
365: For $1<\alpha <2$ the self-similar solutions have algebraic tails,
366: with $f_\alpha(s) \sim s^{-1-\alpha}/\Gamma(-\alpha)$ as $s\to\infty$.
367: The case $\alpha=2$ is particularly important since it corresponds to Brownian
368: initial data. Here we obtain a solution found by Golovin in a model
369: for cloud formation from droplets~\cite{Golovin},
370: \be
371: \label{E.golovin}
372: f_2(s) = (4\pi)^{-1/2} s^{-3/2} e^{-s/4}.
373: \ee
374: For the corresponding solution to \qref{E.burgers}, the
375: law of $v(x,t)$ can be recovered from
376: the law of $T^2_x$, the first-passage time for Brownian motion with
377: unit drift, which is explicitly given as follows (see section 2.6):
378: %and thus $V^2_x$, may then be determined
379: %explicitly by Laplace inversion~\cite[Ch.29]{Abramowitz}, and we have
380: \be
381: \label{E.golovin2}
382: {\prob}(T^2_x \in (y,y+dy) ) =
383: \frac{x \one_{y >0}}{2\sqrt{\pi y^3}}
384: \exp\left(-\frac{(x-y)^2}{4y}\right) \, dy.
385: \ee
386:
387:
388: %% Notice that the typical length scale in $x$ is $t^{1/\beta}$. However, it is
389: %% imprecise to think of this as the typical gap between shocks, since
390: %% the shocks are dense ($\int_0^\infty n(t,s) ds$ is divergent) .
391: %% When $\alpha=2$ these densities may be expressed in the closed form
392: %% \[ n(t,s) ds = \frac{1}{\sqrt{2\pi}} s^{-3/2} e^{-s/2t^2}.\]
393:
394: Considering now arbitrary solutions to \qref{E.burgers} with initial data
395: \qref{E.initial}, we classify solutions that approach self-similar
396: form as $t\to\infty$ as follows. We establish necessary and
397: sufficient conditions for convergence of the laws of the rescaled processes
398: \begin{equation}
399: \label{E.rscV}
400: x\mapsto \Vt_x = \frac{t}{\lambda(t)}v(\lambda(t)x, t)
401: %,\qquad x \geq 0,
402: \end{equation}
403: in the sense of weak convergence of measures on $\Skor$. (Since
404: the shocks coalesce, a rescaling $\lambda(t) \to \infty$ is needed
405: to obtain a non-trivial limit. The amplitude scaling is natural,
406: see Section~\ref{S.proof}.) Convergence
407: to a process $V^*$ is written $\Vt \law V^*$ as in~\cite{Jacod}.
408: Recall that a positive function $L$ is said to be slowly varying at $\infty$
409: if $\lim_{t\to\infty}L(tx)/L(t)=1$ for all $x>0$.
410: %We implicitly exclude the trivial case when $\Lambda_0$ vanishes in
411: %the following theorem.
412: \begin{theorem}
413: \label{T.main}
414: Let $u_0$ be a spectrally negative \Levy\ process with zero mean
415: $\E(u_0(x))$,
416: variance $\sigma_0^2\ge0$, and downward jump measure %$\Lambda_0$
417: satisfying $\int_0^\infty (s \wedge s^2) \Lambda_0(ds)<\infty$.
418: \begin{enumerate}
419: \item Suppose there is a rescaling $\lambda(t) \rightarrow \infty$ as
420: $t\to\infty$ and a \Levy\ process $V^*$ with zero mean $\E(V^*_1)$
421: such that the random variables $\Vt_1$ converge to $V^*_1$ in law.
422: Then there exists $\alpha \in (1,2]$
423: and a function $L$ slowly varying at infinity such that
424: \begin{equation}
425: \label{E.RV}
426: \sigma_0^2+ \int_0^s r^2 \Lambda_0(dr) \sim {s^{2-\alpha} L(s)}
427: \quad\text{as $ s \to \infty.$}
428: \end{equation}
429: \item Conversely, assume that there exists $\alpha \in (1,2]$ and a
430: function $L$ slowly varying at infinity such that (\ref{E.RV})
431: holds. Then there is a strictly increasing rescaling $\lambda(t) \to
432: \infty$ such that $\Vt \law V^{\alpha}$.
433: \end{enumerate}
434: \end{theorem}
435: \begin{remark}
436: %It is not necessary to assume that $\Vt \law V^*$ in order to
437: %obtain \qref{E.RV}.
438: %$\Vt_{x_0}$ for some fixed $x_0 \in (0,\infty)$ suffices.
439: Since $\Vt$ and $V^*$ are \Levy\/ processes, we have
440: $\Vt \law V^*$ if and only if
441: we have convergence in law of the random variables
442: $\Vt_{x_0}$ for some fixed $x_0 \in (0,\infty)$ (see \qref{E.eqv1}-\qref{E.eqv2} in
443: section 3 below). We take $x_0=1$ without loss of generality.
444: Part (2) implies in particular that the only
445: possible limits are statistically self-similar.
446: \end{remark}
447: \begin{remark}
448: \label{R.energy}
449: We say a solution has finite energy if for any finite
450: interval $I \subset \R_+$ we have $\E\left(\int_I|v(x,t)|^2 \, dx\right) <
451: \infty$. The jump measure $\Lambda_t$ for the solution is related
452: to the energy by (see Section~\ref{S.energy})
453: \[ \E\left(\int_I v(x,t)^2 \, dx \right) = \left(\int_0^\infty s^2
454: \Lambda_t(ds) \right) \int_I x\,dx. \]
455: %% Let $(b_t,0, \Lambda_t)$ denote the triplet for the \Levy\/
456: %% process $v(\cdot,t)$. Then (see Section~\ref{S.energy})
457: %% \[ \E\left(\int_I v(x,t)^2 \, dx \right) = b_t^2 \int_I x^2\, dx +
458: %% \left(\int_0^\infty y^2 \Lambda_t(dy) \right) \int_I x\,dx. \]
459: The integral in \qref{E.RV} is thus a measure of the energy in an
460: interval. If it is initially finite, it is conserved for $t>0$,
461: and it remains infinite if it is initially infinite.
462: The only self-similar solution with finite energy corresponds to
463: $\alpha=2$, and Theorem~\ref{T.main} implies it attracts all solutions with
464: initially finite energy. In this sense, one may say that the finite energy
465: solution is universal. However, Theorem~\ref{T.main} also indicates
466: the delicate dependence of the domains of attraction on the tail
467: behavior of $\Lambda_0$. Heavy-tailed solutions seem to us no less interesting
468: than those with finite energy. Finer results on asymptotics, and a
469: compactness theorem for subsequential limits that
470: builds on Bertoin's \LK\/ classification for eternal
471: solutions to Smoluchowski's equation~\cite{B_eternal}, will be
472: developed elsewhere.
473: \end{remark}
474: \begin{remark}
475: The case of zero mean, $b_0=0$, is the most interesting.
476: If $b_0> 0$ or $b_0<0$ we can
477: reduce to this case by a change of variables (see below). If $b_0<0$,
478: the solution is defined only for $0 \leq t <
479: -b_0^{-1}$. Theorem~\ref{T.main} then characterizes the approach to
480: self-similarity at the blow-up time. If $b_0>0$ then the behavior of
481: the solution as $t\to\infty$ is determined by the zero-mean solution
482: with the same $\sigma_0^2$ and $\Lambda_0$ at the finite time $b_0^{-1}$.
483: \end{remark}
484: \begin{remark} The \CH\/ solution is geometric and
485: Theorem~\ref{T.main} may be a viewed as a limit theorem for
486: statistics of minima. The utility of regular variation in such
487: problems is widely known~\cite{Resnick}. If the initial data is white
488: noise, the \CH\/ solution is a study of the parabolic hull of
489: Brownian motion. Groeneboom's work on this problem~\cite{Groen} is the
490: basis for several results on Burgers turbulence~(in
491: particular~\cite{AE,Frachebourg,Giraud}). We have been
492: unable to find a similar reference to the problem we consider in the
493: probability literature (\cite{Bingham2} seems the closest).
494: \end{remark}
495:
496: \begin{remark}
497: There is a growing literature on intermittence, and the asymptotic
498: self-similarity of Burgers turbulence, see for
499: example~\cite{Woc2,Gurbatov}.
500: Numerical simulations and heuristic arguments suggest that this is a
501: subtle problem with several distinct regimes.
502: It is hard to obtain rigorous results for general initial data.
503: Theorem~\ref{T.main} tells us that the approach
504: to self-similarity is at least as complex as in the classical limit
505: theorems of probability.
506: \end{remark}
507:
508: The rest of this article is organized as follows. We explain the
509: mapping from Burgers equation to Smoluchowski's coagulation
510: equation in Section~\ref{S.meanfield}. This is followed by the proof of
511: Theorem~\ref{T.main} in Section~\ref{S.proof}. Finally, in
512: Section~\ref{S.energy} we compute a number of statistics of
513: physical interest: energy and dissipation in solutions, the
514: Fourier-Laplace spectrum, and the multifractal spectrum.
515:
516:
517: \section{Mean field theory for Burgers equation}
518: \label{S.meanfield}
519: In this section we explain the connection between Burgers equation with
520: spectrally negative \Levy\ process data and Smoluchowski's coagulation
521: equation. The main results are due to
522: Bertoin~\cite{B_burgers} and Carraro and Duchon~\cite{Duchon1}.
523: We follow Bertoin's approach, and explain results implicit
524: in~\cite{B_burgers} and \cite{B_eternal}.
525: We think it worthwhile to make this connection widely known
526: in full generality, since the results are of interest to
527: many non-probabilists. Exact solutions of this simplicity are also
528: useful as benchmark problems for numerical calculations.
529: %% To this end,
530: %% we also tabulate some quantities of physical interest: the BV decay,
531: %% dissipation at shocks, the multifractal spectrum, and the power spectrum.
532: %%
533:
534: \subsection{Shock coalescence and Smoluchowski's coagulation equation}
535: Smoluchowski's coagulation equation is a widely used mean-field model of
536: cluster growth~(see~\cite{Aldous,Drake} for introductions).
537: We begin with a heuristic derivation of
538: the coagulation equation as a mean-field model of shock
539: coalescence. First consider the evolution of
540: a single shock of size $s>0$.
541: Let $u_0(x)= -s \one_{x \geq 0}$. Then the solution is
542: \be
543: \la{E.ba1}
544: u(x,t) = -s \one_{x \geq x_1(t)}, \quad x_1(t) =
545: -\frac{s}{2}t.
546: \ee
547: Shock coalescence is nicely seen as follows.
548: Let $u_0(x) = -\sum_{k=1}^N s_k(0)
549: \one_{x \geq x_k(0)}$ where $s_k(0) >0$ for
550: $k=1,\ldots,N$ and $x_1(0) < \ldots x_N(0)$.
551: The solution may be constructed
552: using the method of characteristics and the standard jump condition
553: \[
554: \dot{x} = \frac12(u^- + u^+)
555: \]
556: across a shock at $x=x(t)$, where $u^-$ and $u^+$ denote respectively the left
557: and right limits of $u(\cdot,t)$ at $x$. At any time
558: $t>0$, there are $N(t) \leq N(0)$ shocks at locations $x_1(t) <
559: x_k(t) < x_{N(t)}(t)$ and
560: \be
561: \la{E.ba2}
562: u(x,t) = - \sum_{k=1}^N s_k(t) \one_{x \geq x_k(t)}, \quad
563: \dot{x}_k(t_+) = -\sum_{j=1}^{k-1} s_j(t_+) -
564: \frac{s_k(t_+)}{2}.
565: \ee
566: The shock sizes $s_k(t)$ are constant between collisions,
567: and add upon collision---when shocks $k$ and $k+1$ collide,
568: we set $s_k(t_+) = s_k(t_-)+ s_{k+1}(t_-)$ and relabel.
569: This yields an appealing {\em sticky
570: particle\/} or {\em ballistic aggregation\/} scenario.
571: We say a system of particles with position, mass and
572: velocity $(x_k(t),m_k(t),v_k(t))$ undergoes ballistic aggregation if
573: (a) the particles move with constant mass and velocity between collisions,
574: and (b) at collisions, the colliding particles stick to form a single particle,
575: conserving mass and momentum in the process. We map this shock coalescence
576: problem to a sticky particle system by setting $m_k= s_k$ and
577: $v_k = \dot x_k$. %-\sum_{j=1}^{k-1} s_j - s_k/2$.
578: Suppose particles $k$ and $k+1$ meet at time $t$.
579: Then, with unprimed variables denoting values before collision and
580: primed variables denoting values after,
581: since $v_{k+1}=v_k-(m_k+m_{k+1})/2$ we have
582: \[
583: m_k v_k + m_{k+1} v_{k+1}
584: %= m_k v_k + m_{k+1}\left(v_k - \frac{m_k+m_{k+1}}{2} \right)
585: =(m_k + m_{k+1})\left(v_k - \frac{m_{k+1}}{2} \right)
586: = m_k'v_k'.
587: \]
588:
589: %Then just before the collision
590: %\[ (m_k v_k)(t_-) + (m_{k+1} v_{k+1}(t_-) =
591: %(m_k + m_{k+1})\left(v_k(t_-) -
592: %\frac{s_{k+1}(t_-)}{2} \right). \]
593: %The velocity just after collision is given by conservation of momentum
594: %\[ v_k(t_+)= -\sum_{j=1}^{k-1} s_j
595: %-\frac{s_k(t_-)+s_{k+1}(t_-)}{2} = -\sum_{j=1}^{k-1} s_j
596: %-\frac{s_k(t_+)}{2}, \]
597: %in agreement with \qref{E.ba2}.
598:
599: The calculations so far involve no randomness. Suppose now that the
600: shock sizes $s_j$ are independent and let $f(s,t)\,ds$ denote
601: the expected number of shocks per unit length with size in
602: $[s,s+ ds]$. We derive a mean-field rate equation for $f$ as follows.
603: Let $I$ be an interval of unit length.
604: The number density changes because of the flux of
605: shocks entering and leaving $I$ and because of shock collisions within $I$.
606: On average, the velocity difference across $I$ is
607: \[
608: M_1(t) = \int_0^\infty s f(s,t) ds,
609: \]
610: therefore the average influx is $M_1(t)f(s,t)\,ds$.
611: Next consider the formation of a shock of size $s_1+ s_2$ by a collision
612: of shocks of size $s_1$ and $s_2$ as shown in
613: Figure~\ref{fig:heuristic}. The relative velocity
614: between these shocks is $(s_1+ s_2)/2$
615: (see Figure~\ref{fig:heuristic}).
616: The expected number of neighboring pairs with sizes in
617: $[s_1,s_1+ds_1]$,
618: $[s_2,s_2+ds_2]$
619: %$s_1$ and $s_2$ is $f(s_1,t)f(s_2,t)$.
620: respectively is
621: \[
622: f(s_1,t)f(s_2,t) \,ds_1\,ds_2 .
623: \]
624: The probability that these neighboring shocks are near enough to collide in
625: time $dt$ is $\frac12(s_1+s_2)\,dt$,
626: thus the number of these shocks
627: that collide in time $dt$ is
628: %-----------------------------------------------------------------------
629: \begin{figure}
630: \centerline{\epsfysize=5cm{\epsffile{heuristic2.eps}}}
631: \caption{Binary clustering of shocks \label{fig:heuristic}
632: }
633: \end{figure}
634: %-----------------------------------------------------------------------
635: \be
636: \label{E.rate}
637: % f(m_1,t) f(m_2,t) \frac{\Delta_1+ \Delta_2}{2} dt =
638: f(s_1,t) f(s_2,t) \frac{s_1+ s_2}{2}\, ds_1\,ds_2\,dt.
639: \ee
640: Summing over all collisions that create shocks of size $s=s_1+s_2$,
641: and accounting for the loss of shocks of size $s$ ($=s_1$ or $s_2$)
642: in collisions with other shocks, we obtain the rate equation
643: %For every such collision, we gain one particle of size $m_1+m_2$
644: %and lose two particles. We
645: \[
646: %\nn %\la{E.smol_ba} %\lefteqn
647: {\partial_t f (s,t)= M_1(t) f + Q(f,f)}
648: \] %\ea
649: where
650: $Q(f,f)$ denotes the collision operator given by
651: \[
652: Q(f,f)(s,t) = \frac{1}{2}\int_0^s s f(s_1,t)
653: f(s-s_1, t) \, ds_1
654: - \int_0^\infty (s+ s_1) f(s,t) f(s_1,t) \,ds_1.
655: \]
656: We integrate in $s$ to
657: find $\dot M_1 = M_1^2$, therefore the normalized
658: density $f/M_1$ satisfies the equation
659: \be
660: \la{E.smol_ba}
661: \frac1{M_1} \partial_t \left( \frac{f}{M_1} \right) = Q \left( \frac{f}{M_1},
662: \frac{f}{M_1} \right).
663: \ee
664: Up to a change of time scale, this is a fundamental mean-field model of
665: coalescence: Smoluchowski's coagulation equation with additive kernel.
666: We treat this equation in greater depth below.
667:
668: More precisely, it turns out that the random solution $u(x,t)$ has the
669: structure described in \qref{E.ba2} when the initial data $u_0$ consists of a
670: compound Poisson process with only downward jumps. The mean drift rate at
671: time $t$ is then $-M_1(t)$, and this example shows that the solution
672: blows up at the time $M_1(0)^{-1}$ in this case.
673: We show below (see \qref{E.changedrift}) that one may remove the mean drift by
674: a change of scale and slope, yielding `sawtooth' data with a deterministic
675: upward drift that compensates the random downward jumps.
676: For such data we obtain a global solution. Thus, there is no essential
677: distinction between sawtooth data and the decreasing initial data
678: considered above.
679:
680:
681: %% This heuristic calculation is made precise by Bertoin's theorem.
682: %% Arguably the simplest random initial data is a compound Poisson
683: %% process: that is, the jumps $\Delta_k$ are iid with
684: %% law $\nu_0$ on $(0,\infty)$, and $x_k(0)$ are points of an
685: %% independent Poisson process with rate $r>0$. In this case, the
686: %% \Levy\/ measure of $u_0$ is $\Lambda_0(ds) = r \nu_0(ds)$. This is a
687: %% typical `sawtooth' profile, and remarkably this form is preserved by
688: %% the equation.
689:
690:
691: %% Since
692: %% $\E(u_0(x)) = x\left( c_0 - \int_0^\infty y\Lambda(dy)\right)$, it is
693: %% natural to assume that $M_0 = \int_0^\infty y \Lambda(dy)< \infty$,
694: %% and $c_0=M_0$ in order that the process $u_0$ has no net drift. A particular
695: %% case of Bertoin's theorem is that the solution remains a compound
696: %% Poisson process with no drift for all time, and the statistics of
697: %% $u(x,t)$ can be recovered from \qref{E.smol_ba} with initial data
698: %% $\nu_0$.
699:
700:
701: %% To summarize, what has worked in our favor here is that the rigidity of
702: %% sample paths of \Levy\/ processes is perfectly compatible with
703: %% entropy solutions for Burgers equation. This is clear for the compound
704: %% Poisson processes. Remarkably, it holds even
705: %% when the initial data is not of bounded variation.
706:
707:
708: \subsection{The \CH\ formula}
709: The modern notion of an entropy solution stems from the penetrating
710: analysis by Hopf of the vanishing viscosity limit to
711: \qref{E.burgers}. His work was based on a change of variables
712: (re)discovered independently by Cole
713: and Hopf~\cite{Cole,Hopf}. This solution is obtained via
714: minimization of the \CH\/ function \cite{Cole,Hopf}
715: \begin{equation}
716: \label{E.CH}
717: H(y,t;x) = \frac{(x-y)^2}{2t}
718: + \int_{-\infty}^y u_0(y')dy'.
719: \end{equation}
720: The minimum in $y$ is well-defined for all $t>0$
721: provided $U(y)= \int_0^y u_0(y') dy'$ is
722: lower semicontinuous and $\lim_{x \to \pm \infty} y^{-2} U(y)=0$.
723: This is a mild assumption and holds for the random data we consider
724: provided that the mean drift is zero.
725: We denote the extreme points where $H$ is minimized by
726: \be
727: \la{E.IL}
728: a_-(x,t) = \inf\{z | H(z,t;x)= \min_y H\}, \quad
729: a_+(x,t) = \sup\{z| H(z,t;x)= \min_y H \}.
730: \ee
731: Notice that any $z \in \R$ such that $x= tu_0(z) +z$ is a critical
732: point of $H$, and represents a Lagrangian point that arrives at $x$ at
733: time $t$. Of these $z$, the `correct' Lagrangian points are
734: the minimizers of $H$. If $a_-(x,t)=a_+(x,t)$, this point is unique,
735: and we have
736: \begin{equation}
737: \label{E.hopf}
738: u(x,t) = \frac{x - a_\pm(x,t)}{t}, \quad x \in \R,\ t >0. \quad
739: \end{equation}
740: There is a shock at $(x,t)$ when $a_-(x,t) \neq a_+(x,t)$. In this
741: case, the Lagrangian interval
742: $[a_-(x,t),a_+(x,t)]$ is absorbed into the shock and the velocity of the
743: shock is given by the Rankine-Hugoniot condition (conservation of momentum)
744: \be
745: \la{E.ushock}
746: u(x,t) = \frac{u(x_+,t) + u(x_-,t)}{2} =
747: \frac{1}{a_+(x,t)-a_-(x,t)}\int_{a_-(x,t)}^{a_+(x,t)} u_0(y) \, dy.
748: \ee
749: It will be convenient for us to assume that $u$ is
750: right-continuous and we call $a(x,t) = a_+(x,t)$ the {\em inverse
751: Lagrangian function\/}. Of course, the speed of shocks are still
752: determined by the right-hand side of \qref{E.ushock}.
753:
754: %% A similar calculation holds if the collisions are not binary. We now
755: %% incorporate a drift and consider `sawtooth' initial data of the form
756: %% \be
757: %% \label{E.ba4}
758: %% u^{(c)}_0(x) = cx - \sum_{k=1}^N \Delta_k(0) \one_{x > x_k(0)}.
759: %% \ee
760: In order to deal with non-zero mean drift in initial data,
761: we will use the following interesting
762: invariance of Burgers equation.
763: Assume that $u_0(x)=o(|x|)$ as $|x| \to \infty$, and
764: let $u(x,t)$ be the \CH\/ solution with $u(x,0)=u_0(x)$,
765: defined for all $t \geq 0$. Let $c\in\R$ and
766: define
767: \be
768: \label{E.Tc}
769: u^{(c)}_0(x) = u_0(x) + cx , \qquad
770: T_c = \left\{ \begin{array}{rl} -c^{-1}, &\quad c <0, \\
771: +\infty, &\quad c \geq 0. \end{array} \right.
772: \ee
773: Then the \CH\/ solution with initial data $u^{(c)}_0$ is given by
774: \be
775: \label{E.changedrift}
776: u^{(c)}(x,t) = \frac{1}{1+ct} u\left( \frac{x}{1+ct}, \frac{t}{1+ct} \right)
777: + \frac{cx}{1+ct}, \quad t \in [0,T_c).
778: \ee
779: This is seen as follows. An elementary calculation shows that the
780: \CH\/ functionals for the different data are related by
781: \[ H^{(c)}(y,t;x) = H\left(y, \frac{t}{1+ct};\frac{x}{1+ct}\right) +
782: \frac{cx^2}{2(1+ct)}, \]
783: which implies the inverse Lagrangian functions are related by
784: \be
785: \label{E.changedrift2}
786: a^{(c)}(x,t) = a\left(\frac{x}{1+ct}, \frac{t}{1+ct}\right).
787: \ee
788: We now substitute in \qref{E.hopf} to obtain \qref{E.changedrift}.
789: %% can be verified by differentiation at points where $u(x,t)$ is
790: %% smooth. At discontinuities,
791: %% \qref{E.changedrift} preserves Oleinik's entropy condition $u(x_-,t) >
792: %% u(x_+,t)$ which is sufficient to guarantee uniqueness.
793: %% For sawtooth initial data, we now see from \qref{E.ba2} and
794: %% \qref{E.changedrift} that the shock speeds are constant between
795: %% collisions. Thus, the solution is generated by sticky
796: %% particle dynamics as before. If initially
797: %% \be
798: %% \label{E.ba6}
799: %% m_k = \Delta_k(0), \quad v_k(0) = c_0 x_k(0) - \sum_{j=1}^{k-1}
800: %% \Delta_j(0) - \frac{\Delta_k(0)}{2},
801: %% \ee
802: %% then the solution to Burgers equation is reconstructed by
803: %% \be
804: %% \label{E.ba5}
805: %% u^{(c)}(x,t) = c(t) x - \sum_{k=1}^{N(t)} \Delta_k(t) \one_{x \geq x_k(t)}.
806: %% \ee
807: %% where
808: %% \be
809: %% \label{E.ba7}
810: %% c(t) = \frac{c_0}{1+c_0t}, \quad \Delta_j(t) = \frac{m_j(t)}{1+c_0t}.
811: %% \ee
812: %% %% In particular, $u(x,t)$ is of the same form as the initial
813: %% %% data. Moreover, $c(t)$ and
814: %% %% $\Delta_1(t)$ reflect the typical decay of BV data.
815: %% %% The
816: %% %% crucial assumption here is that the slope $c_0$ is the same to
817: %% the left and
818: %% %% the right of the shock. This ensures that while $u_\pm(x_1(t))\neq
819: %% %% u_\pm(x_1(0)$, it is still true that the shock speed
820: %% %% \[ \frac{u_-(x_1(t)) + u_+(x_1(t))}{2} = - \frac{\Delta}{2},
821: %% %% \quad t \geq 0. \]
822: %% %% Thus, the shock path is linear and it is appealing to think of the shock
823: %% %% as a particle of mass $\Delta(0)$. This is the basis for the widely
824: %% %% known mapping from Burgers equation to {\em ballistic aggregation\/}
825: %% %% or {\em sticky particles\/}. Consider $N$ particles with mass $m_k$
826: %% %% and initial velocity $v_k(0)$ located at $x_1(0) < x_2 \ldots <
827: %% %% x_N(0)$. The particles move with constant velocity between
828: %% %% collisions. At collisions they stick, conserving mass and momentum.
829: %% %% This yields an intuitive picture for Burgers equation with
830: %% %% `sawtooth' initial data with slope $c_0 >0$, and $N$
831: %% %% downward jumps $\Delta_k(0) >0$ at points $x_1(0) < x_2(0) < \ldots <
832: %% %% x_N(0)$:
833: %% %% We define an associated system of sticky particles at $x_k(0)$, with
834: %% %% %k=1, \ldots N$ with mass and velocity
835: %% %% %% Following
836: %% %% %% It is clear that at least for a short time, the shocks cannot
837: %% %% %% collide, and we may piece together the solution using \qref{E.ba3}.
838: %% %% %% The extension to collisions is simple. If two shocks meet, say
839: %% %% %% $x_k(t)=x_{k+1}(t)$, then they stick, and the new velocity is
840: %% %% %% determined by conservation of momentum
841: %% %% %% \[ v_k(t_+) = \frac{ \Delta_k(t_-) v_k(t_-) + \Delta_{k+1}(t_-)
842: %% %% %% v_{k+1}(t_-)}{\Delta_k(t_-) + \Delta_{k+1}(t_-)}. \]
843: %% %% %
844: %% %% %What we have described is the basic and widely known mapping from
845: %% %Burgers equation to {\em ballistic aggregation\/}.
846: %% The heuristic
847: %% calculation above, can be made precise for the class of \Levy\/
848: %% process initial data with only downward jumps.
849: %% The main idea is as
850: %% follows.
851:
852: \subsection{Solutions with \Levy\/ process initial data}\label{S.sol}
853: Here we describe how the solution of \qref{E.burgers},
854: with initial data of the form \qref{E.initial},
855: is determined in terms of Laplace exponents,
856: essentially following Bertoin's treatment in \cite{B_burgers}.
857:
858: Suppose $x\mapsto u_0(x)$ is an arbitrary spectrally
859: negative \Levy\/ process for $x\ge0$, with Laplace exponent $\psi_0$
860: having downward jump measure $\Lambda_0$.
861: We first show that we may assume without loss of generality that
862: $\int_0^\infty (s\wedge s^2)\Lambda_0(ds)<\infty$. Indeed,
863: if $\int_1^\infty s\Lambda_0(ds)=\infty$, then
864: $u_0(x)/x \to-\infty$ almost surely as $x\to\infty$. (This follows from
865: the fact that for the compound Poisson process $X_x$ with jump measure
866: $\Lambda_0(ds) \one_{|s| \geq 1}$, one has $X_x/x\to\infty$ as $x\to\infty$
867: by the law of large numbers.)
868: In this case the \CH\/ function $H(y,t;x)$ has no minimum for any $t>0$,
869: and equation \qref{E.burgers} has no finite entropy solution
870: for any positive time.
871: Hence, we may suppose that $\int_1^\infty s\Lambda_0(ds)<\infty$.
872:
873: Next, we show that one may assume the mean drift $b_0=\E(u_0(1))$ is
874: zero. If $b_0$ is nonzero, we have $\lim_{x \to \infty} u_0(x)/x =
875: b_0$ a.s.\, by the strong law of large numbers.
876: If $b_0<0$, then by comparison to compression-wave solutions
877: with initial data $A+b \max(x,0)$, we find using the maximum
878: principle that a.s.\, the solution blows up exactly
879: at time $-b_0^{-1}$. If $b_0>0$ there is a global solution.
880: In either case, we may use the transformation
881: \qref{E.changedrift} with $c=b_0$ to reduce to the case $b_0=0$, replacing
882: $u_0(x)$ by $u_0(x)-b_0x\one_{x>0}$. More precisely, we apply
883: \qref{E.changedrift} for $x \geq 0$ noting that $a(0,t) \geq 0$,
884: thus $a(x,t) \geq a(0,t) \geq 0$ for $x \geq 0$, so that
885: \qref{E.changedrift2} holds for $x \geq 0$. %\marginpar{must check: ok}
886: We have:
887: \begin{lemma}
888: If $u^{(c)}_0$ is a spectrally negative \Levy\/ process with
889: \Levy\/ triplet $(c, \sigma^2_0, \Lambda_0)$, the Cole-Hopf
890: solution $u^{(c)}(x,t)$ is determined
891: via \qref{E.changedrift} for $x\ge0$ and $t \in [0,T_c)$,
892: in terms of a solution $u(x,t)$ having zero mean drift and
893: defined for all $t\ge0$.
894: \end{lemma}
895:
896: With these reductions, we may restrict ourselves to Laplace
897: exponents $\psi_0$ of the form
898: \begin{equation}\label{psi0}
899: \psi_0(q) = \frac{\sigma_0^2q^2}2 + \int_0^\infty
900: \left(e^{-qs} - 1 + qs \right)
901: \Lambda_0(ds), \quad q\ge0.
902: \end{equation}
903: We will always assume that $a$
904: and $u$ are right continuous (i.e., $a(x,t)= a(x_+,t)$, compare with
905: \qref{E.ushock}). This ensures $a$ is an element of the
906: Skorokhod space $\Skor$, so that we may use the
907: standard Skorokhod topology to study limiting behavior.
908: For brevity, we write
909: \[
910: v(x,t)=u(x,t)-u(0,t),\qquad
911: l(x,t)=a(x,t)-a(0,t),
912: \]
913: and rewrite \qref{E.hopf} as
914: \be
915: \la{E.vl}
916: v(x,t) = \frac{x-l(x,t)}{t}, \quad x \geq 0,\ t > 0.
917: \ee
918: %It will be convenient to break with convention and suppose \qref{E.vl}
919: %holds everywhere; that is $v(x,t)=v(x_+,t)$.
920: Bertoin has shown that for all $t>0$, $x\mapsto l(x,t)$ is an increasing
921: \Levy\ process (a subordinator) with the same law as the first passage
922: process for $tu_0(x)+x$.
923: We denote the Laplace exponents of $l$ and $v$ by $\Phi$ and
924: $\psi$ respectively:
925: \be
926: \la{E.LElv}
927: \E\left(e^{-ql(x,t)}\right) = e^{-x\Phi(q,t)}, \quad \E
928: \left(e^{qv(x,t)}\right) = e^{x \psi(q,t)}, \quad x,q,t \geq 0.
929: \ee
930: We combine (\ref{E.vl}) and \qref{E.LElv} to obtain
931: \begin{equation}
932: \label{E.psi_t}
933: \psi(q,t) =\frac{q}{t}-\Phi\left(\frac{q}{t},t\right).
934: \end{equation}
935: Since $l$ is a subordinator, it has the simpler \LK\/ representation
936: \begin{equation}
937: \label{E.phi}
938: \Phi(q,t) = d_tq + \int_0^\infty (1-e^{-qs})\mu_t(ds), \quad q >0,
939: \end{equation}
940: where $d_t \geq 0$ supplies the deterministic part of the drift,
941: and $\mu_t$ is the \Levy\ measure of $l(\cdot,t)$,
942: which now must satisfy $\int_0^\infty (1\wedge s)\mu_t(ds)<\infty$ \cite{B_book}.
943: We see from \qref{E.vl} that
944: $v(\cdot,t)$ is a \Levy\ process with no Gaussian component,
945: and thus has a \LK\/ representation
946: \be
947: \la{E.defpsi}
948: \psi(q,t) = b_t q + \int_0^\infty\left(e^{-qs} - 1 + qs \right)
949: \Lambda_t(ds), \quad q\ge0,\ t>0,
950: \ee
951: related to \qref{E.phi} by
952: \be
953: \la{E.mulambda}
954: \Lambda_t( ds) = \mu_t(t\,ds) ,
955: \qquad
956: b_t + \int_0^\infty s\Lambda_t(ds) = \frac{1-d_t}{t}.
957: \ee
958:
959: Due to the result that $l$ and the first passage process
960: of $tu_0(x)+x$ have the same law, a simple functional relation holds between
961: $\psi_0$ and $\Phi(q,t)$ \cite[Thm.~2]{B_burgers}:
962: \begin{equation}
963: \label{E.main}
964: \psi_0(t\Phi(q,t)) + \Phi(q,t)= q, \quad q \geq 0,\ t >0.
965: \end{equation}
966: %[[Implies $t\D_t\Phi-\Phi\D_q\Phi=-\Phi$.]]
967: The evolution takes a remarkably simple form when
968: we combine equations (\ref{E.psi_t}) and (\ref{E.main}) to obtain
969: \begin{equation}
970: \label{E.psi_soln}
971: \psi(q,t) = \psi_0(q -t\psi(q,t)), \quad q \geq 0,\ t>0.
972: \end{equation}
973: But then $\psi(q,t)$ solves the inviscid Burgers equation (in $q$ and $t$!)
974: \begin{equation}
975: \label{E.psi_burgers}
976: \partial_t \psi + \psi \partial_q \psi =0, \quad \psi(0,q) = \psi_0(q).
977: \end{equation}
978: The solution to (\ref{E.psi_burgers}) may be constructed
979: by the method of characteristics and takes the form
980: (\ref{E.psi_soln}). The remarkable fact that the Laplace exponent is
981: also a solution to Burgers equation was first observed by Carraro and
982: Duchon~\cite[Thm.~2]{Duchon1}.
983:
984: Since $\psi_0$ is analytic and strictly convex,
985: the solution (\ref{E.psi_soln}) is analytic for all time and unique,
986: and the condition $\partial_q \psi(0,t)=0$ is preserved for all $t>0$.
987: By \qref{E.psi_t}--\qref{E.mulambda}, we have
988: %Thus, $b_t=0$, % Since $\partial_q\psi(0,t)=0$ for $t>0$,
989: %and equation (\ref{E.psi_t}) implies
990: \begin{equation}
991: \label{E.phi_mass}
992: b_t = 0, \qquad d_t = 1 - t \int_0^\infty
993: s\Lambda_t(ds), \quad t>0.
994: %\partial_q \Phi(q,t)\left|_{q=0} \right.= d_t + \int_0^\infty x
995: %\mu_t(ds) = 1.
996: \end{equation}
997: %Thus, $m(t) : = \int_0^\infty x \mu_t(ds) \leq 1$, and $\lim_{q \to \infty}
998: %\partial_q \Phi(q,t) =d_t$.
999: %We now use \qref{E.psi_t} to see that $\mu_t$
1000: %and $\Lambda_t$ are related by the simple rescaling
1001: %% Moreover, (\ref{E.main}) implies $\lim_{q \to
1002: %% \infty} \Phi(q,t)=\infty$.
1003: Let
1004: \begin{equation}\la{E.m0}
1005: M_0= \lim_{q \to \infty} \psi'_0(q) , \qquad
1006: %= \int_0^\infty y\Lambda_0(dy) .
1007: \end{equation}
1008: %(argh, if $\sigma^2=0$)
1009: We find $\Phi(q,t)\to\infty$ as $q\to\infty$ from (\ref{E.main}),
1010: and differentiate to obtain
1011: \begin{equation}
1012: \label{E.drift}
1013: d_t = \lim_{q \to \infty} \partial_q \Phi(q,t) = \lim_{q\to \infty}
1014: \frac{1}{1+t\psi_0'\left(t \Phi(q,t)\right)} = \frac{1}{1+tM_0}, \quad t >0,
1015: \end{equation}
1016: with the understanding that $d_t=0$ if $M_0=+\infty$.
1017: Then
1018: \begin{equation}\la{E.m}
1019: M(t):= \lim_{q\to\infty} \D_q\psi(q,t) = \int_0^\infty s\Lambda_t(ds)
1020: = \frac{M_0}{1+tM_0}, \quad t>0,
1021: \end{equation}
1022: with the understanding that $M(t)=t^{-1}$ when $M_0=\infty$.
1023: Note $M'=-M^2$. Below we will characterize the evolution
1024: of $\Lambda_t$ differently.
1025: %
1026: %
1027: %
1028:
1029: \subsection{BV regularity}
1030: It is clear from the \CH\ formula that
1031: $u$ is locally of bounded variation for every $t>0$. We derive a
1032: decay estimate that quantifies this.
1033: The sample paths of $u_0$ have unbounded variation if and only
1034: if~\cite[p.15]{B_book}
1035: \be
1036: \label{E.sigma1}
1037: \sigma_0^2 >0 \quad \mbox{or} \quad \int_0^\infty s \Lambda_0(ds) =
1038: \infty.
1039: \ee
1040: Heuristically, this corresponds to the presence of many small
1041: jumps (`dust'). This is reflected in the Laplace exponent as
1042: $M_0=\lim_{q \to \infty} \psi_0'(q) = +\infty$ in this case.
1043: On the other hand, $M_0$ is finite if and only if
1044: $u_0$ is BV, in which case $\sigma_0=0$ and
1045: $M_0 = \int_0^\infty s \Lambda_0(ds) < \infty$.
1046:
1047: The analytic
1048: formula (\ref{E.phi}) has the following probabilistic meaning.
1049: If we take a Poisson point process $x\mapsto m^t_x$ (masses of clusters) with
1050: jump measure $\mu_t$ we have the representation~\cite[p.16]{B_book}
1051: \begin{equation}
1052: \label{E.a_repn}
1053: l(x,t) = d_tx + \sum_{0 \leq y \leq x} m^{t}_y.
1054: \end{equation}
1055: The velocity field, and a point process of shock strengths
1056: $s^t_y= t^{-1}m^t_y$ are
1057: determined from (\ref{E.hopf}), (\ref{E.drift}) and (\ref{E.a_repn}) by
1058: \begin{equation}
1059: \label{E.BV}
1060: v(x,t) = M(t) x - \frac{1}{t}\sum_{0 \leq y \leq
1061: x} m^{t}_y = M(t)x - \sum_{0 \leq y} s^t_y,
1062: %:= \frac{M_0}{1+tM_0}x - \sum_{0 \leq y \leq x} \Delta^{t}_y,
1063: \end{equation}
1064:
1065: For every $t>0$, $v(x,t)$ is the difference of
1066: two increasing functions: a linear drift and a pure jump process.
1067: Thus, it is of bounded variation, and by \qref{E.phi_mass} and
1068: \qref{E.a_repn} we have
1069: \begin{equation}
1070: \label{E.m1_finite}
1071: \E \left( \int_0^x |\partial_y v(y,t)| dy \right) =
1072: 2M(t)x ,
1073: %= 2 x \int_0^\infty x \Lambda_t (ds) =: 2x M(t),
1074: \quad x,t>0,
1075: \end{equation}
1076: because $\E \left( \sum_{0 \leq y \leq x} m^t_y\right) =
1077: xM(t)=x(1-d_t)$.
1078: %% Let us also note that \qref{E.BV} implies
1079: %% $v(\cdot,t)$ has the characteristic sawtooth profile seen in
1080: %% simulations for every $t>0$. If the shocks are discrete ($\int_0^\infty
1081: %% \mu_t(ds) < \infty$), this implies the shock paths are straight lines
1082: %% with speed given by \qref{E.ushock}. As $t$ increases, the shock
1083: %% structure coarsens purely because of collisions. Remarkably, this is
1084: %% exactly described by the following model.
1085: %% Let us also note that \qref{E.BV} implies
1086: %% $v(\cdot,t)$ has the characteristic sawtooth profile seen in
1087: %% simulations for every $t>0$. If the shocks are discrete ($\int_0^\infty
1088: %% \mu_t(ds) < \infty$), this implies the shock paths are straight lines
1089: %% with speed given by \qref{E.ushock}. As $t$ increases, the shock
1090: %% structure coarsens purely because of collisions. Remarkably, this is
1091: %% exactly described by the following model.
1092: %% %% This is the basic BV estimate. It will appear below
1093: %% %% in the time change for the mapping to Smoluchowski's equation.
1094:
1095: %
1096: \subsection{Relation to Smoluchowski's coagulation equation}
1097:
1098: We consider a positive measure $\nu_\tau(ds)$ interpreted as the
1099: number of clusters of mass or size $s$ per unit
1100: volume at time $\tau$. Clusters of mass $r$ and $s$ coalesce by
1101: binary collisions at a rate governed by a symmetric kernel $K(r,s)$.
1102: A weak formulation of Smoluchowski's coagulation equation
1103: can be based on a general moment identity
1104: for suitable test functions $\zeta$ (see~\cite{MP1}):
1105: \begin{align}
1106: \label{E.smol}
1107: &\D_\tau \int_0^\infty \zeta(s)\,\nu_\tau(ds)
1108: = \nonumber\\ &\qquad\qquad
1109: \frac{1}{2} \int_0^\infty \!\! \int_0^\infty
1110: \left( \zeta(r+s)-\zeta(r)-\zeta(s) \right) K(r,s)\,\nu_\tau(dr) \,\nu_\tau(ds).
1111: \end{align}
1112: We consider only the {\em additive kernel \/}$K(r,s)=r+s$. It is classical that
1113: (\ref{E.smol}) can then be solved by the Laplace
1114: transform~\cite{Drake}. We denote the
1115: initial time by $\tau_0$ (to be chosen
1116: below). The minimal (and natural) hypothesis on initial data
1117: $\nu_{\tau_0}$ is that the mass $\int_0^\infty s \nu_{\tau_0}(ds)$ is
1118: finite. We scale the initial data such that $\int_0^\infty s \nu_{\tau_0}
1119: =1$. The Laplace exponent
1120: \begin{equation}\label{E.nu}
1121: \vp(q,\tau) = \int_0^\infty
1122: (1-e^{-qs}) \nu_\tau(ds)
1123: \end{equation}
1124: then satisfies
1125: \begin{equation}
1126: \label{E.vp1}
1127: \quad \partial_\tau \vp - \vp \partial_q \vp = -\vp, \quad \tau >\tau_0.
1128: \end{equation}
1129: We showed in~\cite{MP1} that (\ref{E.vp1}) may be used to define
1130: unique, global, mass-preserving solutions to (\ref{E.smol}).
1131: In particular, a map $\tau\mapsto \nu_\tau$ from $[\tau_0,\infty)$
1132: to the space of positive Radon measures on $(0,\infty)$, such that
1133: $\int_0^\infty s\nu_\tau(ds)=1$ for all $\tau\ge\tau_0$, is a solution
1134: of Smoluchowski's equation with $K(r,s)=r+s$ (in an appropriate weak
1135: sense detailed in \cite{MP1}) if and only if $\vp$ satisfies \qref{E.vp1}.
1136:
1137: We now connect solutions of the inviscid Burgers equation \qref{E.burgers}
1138: with \Levy\/ process initial data to solutions of Smoluchowski's
1139: equation through a change of scale.
1140: Let $u_0$ satisfy \qref{E.initial} and assume as in subsection~\ref{S.sol}
1141: that the corresponding downward jump measure $\Lambda_0$ satisfies
1142: $\int_0^\infty (s\wedge s^2)\Lambda_0(ds)<\infty$ and the mean drift is zero.
1143: Let $\Lambda_t$ be the jump measure of the \CH\/ solution.
1144: With $M_0$ and $M(t)$ as in \qref{E.m0} and \qref{E.m},
1145: let $\tau_0=-\log M_0$ if $u_0$ is
1146: of bounded variation, and $\tau_0= -\infty$ otherwise, and set
1147: \be
1148: \la{E.smolburgers}
1149: \tau = -\log M(t), \quad
1150: \nu_\tau(ds) = \Lambda_t(M(t)ds).
1151: %\qquad[[=\mu_t(tM(t)\,ds)]]
1152: %\tau = \log\left(t + e^{\tau_0} \right), \quad \nu_\tau(ds) =
1153: %\mu_{t} \left( \frac{t\,ds}{t+e^{\tau_0}} \right) =
1154: %\Lambda_t
1155: %\left(\frac{ds}{t+e^{\tau_0}} \right).
1156: \ee
1157: From \qref{E.m} it follows $\int_0^\infty s \nu_\tau(ds)=1$,
1158: and by \qref{E.nu} and \qref{E.defpsi} we find
1159: \be
1160: \label{E.changevars}
1161: \vp(q,\tau) = q - \psi(qe^\tau,t) .
1162: \ee
1163: %\[
1164: %[[= q-\psi(q/M(t),t) =
1165: %q-\left(\frac{q}{tM(t)}-\Phi(\frac{q}{tM(t)},t)\right)
1166: % = \Phi(\frac{q}{tM(t)},t)-\frac{q}{tM_0}. ]]
1167: %\]
1168: We see that $\psi$ solves (\ref{E.psi_burgers}) if and only if $\vp$
1169: solves \qref{E.vp1}.
1170: Therefore, the rescaled \Levy\/ measure of $v(\cdot,t)$
1171: evolves according to Smoluchowski's equation.
1172: Conversely, given any solution of Smoluchowski's equation with initial
1173: data $\nu_0$ at a finite $\tau_0$, we can
1174: construct a corresponding solution of \qref{E.burgers} by choosing
1175: $u_0$ to be a spectrally negative \Levy\/ process with jump measure
1176: $\Lambda_{t_0}$ via \qref{E.smolburgers}
1177:
1178: Initial data $u_0$ with unbounded variation are of particular
1179: interest. Here we have {\em eternal solutions\/} $\nu_\tau$ to
1180: (\ref{E.smol}) defined for all $\tau \in \R$.
1181: We see that eternal solutions are in
1182: one-to-one correspondence with initial data $u_0$ of unbounded
1183: variation via \qref{E.changevars}. A finer correspondence
1184: mapping the clustering of shocks to the additive coalescent is found
1185: in~\cite{B_cluster}.
1186: %% When the initial data have bounded
1187: %% variation, the change of time scale is required for the following
1188: %% reason. Here shocks grow because of coalescence, and also because they
1189: %% absorb mass from intervals that have not participated in shocks ({\em
1190: %% Lagrangian regular intervals\/}).
1191:
1192: To summarize, we have the following correspondence.
1193:
1194: \begin{theorem}
1195: \label{T.meanfield}
1196: Assume $u_0$ is a spectrally negative \Levy\/ process with
1197: \Levy\/ triplet $(0, \sigma_0^2, \Lambda_0)$, with the same assumptions
1198: as in Theorem~\ref{T.main}.
1199: Then for all $t>0$, $v(\cdot,t)$ and $l(\cdot,t)$ are
1200: \Levy\/ processes with triplet $(0,0,\Lambda_t)$, whose jump
1201: measures $\Lambda_t$ determine a solution $\nu_\tau(ds)$ to Smoluchowski's coagulation
1202: equation with rate kernel $K(r,s)=r+s$ as described in
1203: \qref{E.smolburgers}.
1204: \end{theorem}
1205:
1206: %(b) If $u^{(c)}_0$ is a spectrally negative \Levy\/ process with
1207: % \Levy\/ triplet $(c, \sigma^2_0, \Lambda_0)$, the Cole-Hopf
1208: % solution $u^{(c)}(x,t)$ is in one-to-one correspondence with a
1209: % solution $u(x,t), t \in [0,T_c)$ with zero drift, via
1210: % \qref{E.changedrift}.
1211: %[[Shouldn't this statement be in sec. 2.3??]]\marginpar{move??}
1212:
1213: %% Moreover, the maps $t \mapsto v(\cdot,t)$ and
1214: %% $t \mapsto l(\cdot,t)$ are continuous in the Skorokhod topology at
1215: %% all but a countable set of points corresponding to the jumps of
1216: %% $u(0,t)$ (a shock at $x=0$).
1217: %
1218: %
1219: %
1220: %
1221: %
1222: %
1223: %
1224: %
1225: \subsection{Self-similar solutions}
1226: Bertoin's characterization of eternal solutions is the analogue of the
1227: \LK\ characterization of infinitely divisible
1228: distributions~\cite{B_book,Feller}. Among the latter, the
1229: stable distributions are of particular interest, and their analogues
1230: for Smoluchowski's equations are obtained by choosing the Laplace
1231: exponent $\psi_0(q) = q^\alpha$, $\alpha \in (1,2]$. For $\alpha \in
1232: (1,2)$ the corresponding \Levy\/ measures are
1233: \[
1234: \Lambda(ds) =
1235: \frac{ s^{-(1+\alpha)}}
1236: {\Gamma(-\alpha)}
1237: \,ds.
1238: \]
1239: The Laplace exponent $q^2$
1240: corresponds to an atom at the origin. We thereby
1241: obtain for $\alpha \in (1,2]$ a family of self-similar
1242: solutions to Smoluchowski's equation with Laplace exponent of the form
1243: $\vp(\tau,q) = e^{-\beta \tau} \vp_\alpha(q e^{\beta \tau})$ where
1244: $\vp_\alpha$ solves
1245: \begin{equation}
1246: \label{E.add_scaling}
1247: %\frac{1}{\alpha}
1248: \vp_\alpha(q)^\alpha + \vp_\alpha(q) =
1249: q, \quad q>0.
1250: \end{equation}
1251: The self-similar solutions to Smoluchowski's coagulation
1252: equation are
1253: \begin{equation}
1254: \label{E.nform}
1255: \nu_\tau(ds) = e^{-2\tau/\beta} f_{\alpha}(e^{-\tau/\beta}s)\,ds, \quad
1256: \beta=\frac{\alpha-1}{\alpha}, \quad \alpha \in (1,2],
1257: \end{equation}
1258: where $f_\alpha$ has been defined in \qref{E.ndensity}.
1259: An analytic proof that these are the only self-similar solutions to
1260: Smoluchowski's equation may be found in~\cite{MP1}. Each of these
1261: solutions corresponds to a self-similar process. Precisely,
1262: let $X^\alpha$ denote the stable process with Laplace exponent
1263: $q^\alpha$, and $T^\alpha$ and $V^\alpha$ denote the processes
1264: \begin{equation}
1265: \label{E.first_pass}
1266: T^\alpha_x=\inf \{y\geq 0: X^\alpha_y +y > x \}, \qquad
1267: %\mathrm{and}\quad
1268: V^\alpha_{x} = x - T^\alpha_x.
1269: \end{equation}
1270: %
1271: %\quad \E(e^{-qT^\alpha_x}) = e^{-x\vp_\alpha(q)}, \quad q >0.
1272: %\end{equation}
1273: We have $M_0=+\infty$ and $M(t)=t^{-1}=e^{-\tau}$ in this case, and
1274: the Laplace exponent of the process $l(\cdot,t)$ is of the self-similar form
1275: %\marginpar{check??}
1276: \begin{equation}
1277: \label{E.ss3}
1278: \Phi(q,t) = \vp(q,\tau)= t^{-1/\beta} \vp_\alpha \left( qt^{1/\beta}\right), \;\;
1279: t >0.
1280: \end{equation}
1281: The solution processes have the scaling property
1282: \begin{equation}
1283: \label{E.shock_scaling}
1284: l(x,t) \eqd t^{1/\beta} T^\alpha_{xt^{-1/\beta}}, %\quad \mbox{and}
1285: \qquad v(x,t) \eqd t^{1/\beta-1} V^\alpha_{xt^{-1/\beta}}.
1286: \end{equation}
1287: The corresponding \Levy\/ measures are obtained from
1288: \qref{E.mulambda}, \qref{E.smolburgers} and \qref{E.nform}:
1289: \be
1290: \mu^\alpha_t(ds)= t^{-2/\beta} f_{\alpha}(t^{-1/\beta}s) \, ds, \quad
1291: \Lambda^{\alpha}_t(ds) = t^{1-2/\beta} f_{\alpha} \left(
1292: t^{1-1/\beta} s \right) \, ds.
1293: \ee
1294: In the important case $\alpha=2$, we have
1295: $1/\beta=2$ and $\vp_2(q)=-\frac12+\sqrt{\frac14+q}$,
1296: and by Laplace inversion~\cite[Ch.29]{Abramowitz}
1297: we obtain the explicit expression in \qref{E.golovin2}
1298: for the distribution of $T^2_x$.
1299:
1300: \section{The convergence theorem}
1301: \label{S.proof}
1302: In \cite{MP1} we proved the following theorem characterizing solutions
1303: that approach self-similar form in Smoluchowski's coagulation
1304: equation with additive kernel. To every solution $\nu_\tau$ of
1305: \qref{E.smol} with
1306: $\int_0^\infty s \nu_\tau(ds)=1$ we associate the probability
1307: distribution function
1308: \be\label{E.Fsd}
1309: F(s,\tau) = \int_{(0,s]} r\nu_\tau(dr).
1310: \ee
1311: To a self-similar solution $f_{\alpha}$, $\alpha \in (1,2]$ with
1312: $\beta = (\alpha-1)/\alpha$ we associate
1313: \begin{equation}
1314: \label{E.Fdef}
1315: F_{\alpha}(s) = \int_0^s rf_{\alpha}(r)\,dr =
1316: \sum_{k=1}^\infty \frac{(-1)^{k-1}s^{k\beta}}{k!}
1317: \Gamma(1+k-k\beta)\frac{\sin \pi k\beta}{\pi k\beta}.
1318: \end{equation}
1319: A probability distribution function $F^*$ is called nontrivial if
1320: $F^*(s) <1 $ for some $s>0$; this means the distribution is proper
1321: ($\lim_{s \to \infty}F^*(s)=1$) and not concentrated at $0$.
1322: \begin{theorem}
1323: \label{T.add}
1324: Suppose $\tau_1 \in \R$ and $\nu_{\tau}$, $\tau \in [\tau_1,\infty)$, is
1325: a solution to Smoluchowski's
1326: coagulation equation with additive kernel such that $\int_0^\infty s
1327: \nu_{\tau_1}(ds) =1$.
1328: \begin{enumerate}
1329: \item Suppose there is a rescaling function
1330: $\tlambda(\tau) \rightarrow \infty$ as $\tau\to\infty$ and a
1331: nontrivial probability distribution function $F^*$ such that
1332: \begin{equation}
1333: \label{E.add5}
1334: \lim_{\tau \rightarrow \infty} F(\tlambda(\tau)s,\tau) = F^*(s)
1335: \end{equation}
1336: at all points of continuity of $F^*$. Then there exists $\alpha \in (1,2]$
1337: and a function $L$ slowly varying at infinity such that
1338: \begin{equation}
1339: \label{E.add_M2}
1340: \int_0^s r^2 \nu_{\tau_1}(dr) \sim {s^{2-\alpha} L(s)}
1341: \quad\text{as $s \to \infty.$}
1342: \end{equation}
1343: \item Conversely, assume that there exists $\alpha \in (1,2]$ and a
1344: function $L$ slowly varying at infinity such that (\ref{E.add_M2})
1345: holds. Then there is a strictly increasing rescaling $\tlambda(\tau) \to
1346: \infty$ such that
1347: \[ \lim_{\tau\to\infty} F(\tlambda(\tau) s,\tau) = F_{\alpha}(s), \quad 0
1348: \leq s < \infty,\]
1349: where $F_{\alpha}$ is a distribution function for a self
1350: similar solution as in \qref{E.Fdef}.
1351: %whose Laplace transform $u_\rho$ solves (\ref{eq:add_scaling}).
1352: \end{enumerate}
1353: \end{theorem}
1354:
1355: We now prove Theorem~\ref{T.main}. Let $u_0$ be a spectrally
1356: negative \Levy\/ process with zero mean drift and $\int_0^\infty (s \wedge
1357: s^2) \Lambda_0(ds) < \infty$. To the solution increment
1358: $v(x,t)=u(x,t)-u(0,t)$ with downward jump measure $\Lambda_t$,
1359: associate a solution $\nu_\tau$ of Smoluchowski's coagulation equation
1360: \qref{E.smol} as in Theorem~\ref{T.meanfield}
1361: with Laplace exponent $\vp(q,\tau)$ given by \qref{E.changevars}.
1362: Let $\tau_1=\tau_0=-\log M_0$ if $M_0<\infty$, and
1363: let $\tau_1=0$ if $M_0=+\infty$ and $\tau_0=-\infty$.
1364:
1365: We deduce Theorem~\ref{T.main} from Theorem~\ref{T.add} by
1366: establishing two equivalences:
1367: \begin{itemize}
1368: \item[(a)]
1369: There is a rescaling $\lambda(t) \rightarrow \infty$ as $t\to\infty$
1370: and a \Levy\ process $V^*$ with zero mean drift $\E(V^*)$ such that
1371: % the marginal distributions of $\Vt$ converge to those of $V^*$
1372: $\Vt \law V^*$
1373: %\marginpar{which? 6/1: convergence in law seems better here; we note
1374: % the equivalence with convergence of marginals in the proof}
1375: if and only if
1376: there is a rescaling $\tlambda(\tau)\to\infty$ as $\tau\to\infty$
1377: and a nontrivial probability distribution function $F^*$
1378: such that \qref{E.add5} holds.
1379: \item[(b)]
1380: $ \int_0^\infty s^2 \nu_{\tau_1}(ds)< \infty$ if and only if
1381: $\int_0^\infty s^2 \Lambda_0 (ds) < \infty$.
1382: Moreover,
1383: $\int_0^s r^2 \nu_{\tau_1}(dr) \sim s^{2-\alpha}L(s)$ as $s \to \infty$
1384: if and only if
1385: $\int_0^s r^2 \Lambda_0(dr) \sim s^{2-\alpha}L(s)$ as $s \to \infty$.
1386: \end{itemize}
1387:
1388: %\begin{proof}[Proof of (1)]
1389: {\em Proof of (a).}
1390: We prove claim (a) by showing each part equivalent
1391: to a corresponding convergence statement for rescaled Laplace
1392: exponents.
1393: First, convergence in law in
1394: $\Skor$ for processes with independent
1395: increments can be reduced to the convergence of characteristic
1396: exponents~\cite[Cor. VII.4.43, p.440]{Jacod}.
1397: In particular, suppose $\lambda(t)\to\infty$ as $t\to\infty$.
1398: Then we have
1399: \be
1400: \la{E.eqv1}
1401: \Vt \law V^*, \quad\mbox{with \ $\E(V^*_1)=0$,}
1402: \ee
1403: if and only if
1404: $\E(e^{ik\Vt_x}) \to \E(e^{ikV^*_x})$ for all
1405: {$k \in \R$, uniformly for $x$ in compact sets, and $\E(V^*_1)=0$}.
1406: But since we are working with \Levy\/ processes, the \LK\/ formula
1407: shows the dependence on $x$ is trivial, and thus \qref{E.eqv1} is equivalent to
1408: \be
1409: \la{E.eqv2}
1410: \E(e^{ik\Vt_1}) \to \E(e^{ikV^*_1}) \quad\mbox{for all $k \in \R$,
1411: \quad and\ \ $\E(V^*_1)=0$}.
1412: \ee
1413: But pointwise convergence of characteristic functions is equivalent to
1414: convergence in distribution of the random variables
1415: $\Vt_1$~\cite[XV.3.2]{Feller}, and since $\Vt_1=1-\Tt_1
1416: \leq 1$, \qref{E.eqv2} is equivalent to convergence of the Laplace
1417: transforms~\cite[XIII.1.2]{Feller}:
1418: \be
1419: \la{E.eqv4}
1420: \E(e^{q\Vt_1}) \to \E(e^{qV^*_1}) \quad \mbox{for all $q >0$},
1421: \quad\mbox{and}\ \ \E(V^*_1)=0.
1422: \ee
1423: Taking logarithms and using \qref{E.rscV} and \qref{E.LElv},
1424: \qref{E.eqv4} is equivalent to
1425: \be \la{C.psi}
1426: \lambda \psi\left({qt}/{\lambda},t\right) \to \psi^*(q)
1427: \quad\mbox{for all $q>0$},
1428: \quad \mbox{and}\ \ \D_q\psi^*(0)=0,
1429: %\lambda \Phi(q \lambda^{-1},t) \to \Phi^*(q), \quad -\partial_q
1430: %\Phi^*(0)=1,
1431: \ee
1432: where $\E(e^{qV^*_x})=e^{x\psi^*(q)}$.
1433: This expresses the convergence of $\Vt$ in terms of convergence
1434: of rescaled Laplace exponents.
1435:
1436: Now suppose $\tlambda(\tau)\to\infty$ as $\tau\to\infty$.
1437: Using~\cite[XIII.1.2]{Feller} again, the (proper) convergence in \qref{E.add5}
1438: is equivalent to pointwise convergence of Laplace transforms:
1439: \be \la{C.eta}
1440: \eta(q,\tau) \to \eta^*(q) \quad\mbox{for all $q>0$},
1441: \quad\mbox{with}\ \ \eta^*(0)=1.
1442: \ee
1443: where $\eta(q,\tau) := \int_0^\infty e^{-qs}
1444: F(\tlambda(\tau)\,ds,\tau)$,
1445: $ \eta^*(q) := \int_0^\infty e^{-qs} F^*(ds)$.
1446: By \qref{E.Fsd} and \qref{E.nu}, we have
1447: \be
1448: \eta(q,\tau)= (\D_q\vp)(q/\tlambda,\tau), \qquad
1449: \int_0^q \eta(r,\tau)\,dr = \tlambda \vp(q/\tlambda,\tau).
1450: \ee
1451: We claim that \qref{C.eta} is equivalent to the statement that
1452: (with $\vp^*(q)=\int_0^q \eta^*(r)\,dr$)
1453: \be \label{C.vp}
1454: \tlambda\vp(q/\tlambda,\tau) \to \vp^*(q)
1455: \mbox{\ \ for all $q>0$}, \quad\mbox{and}\quad
1456: \D_q\vp^*(0)=1.
1457: \ee
1458: Clearly, since $\eta(\cdot,\tau)$ is completely monotone
1459: and bounded, \qref{C.eta} implies \qref{C.vp}. In the other
1460: direction, assume \qref{C.vp}. For any sequence $\tau_j\to\infty$
1461: there is a subsequence along which $\eta(q,\tau_j)$ converges
1462: for all (rational, hence real) $q>0$, to some limit whose integral
1463: must be $\vp^*$. Thus \qref{C.eta} follows.
1464:
1465: We now finish the proof of claim (a) by observing that due
1466: to \qref{E.changevars}, we have
1467: \be
1468: \tlambda\vp(q/\tlambda,\tau)= q -
1469: \tlambda\psi({q e^\tau}/{\tlambda},t).
1470: \ee
1471: Hence the convergence in \qref{C.psi}
1472: is equivalent to that in \qref{C.vp} provided we have
1473: \be \la{E.lrel}
1474: \lambda(t)/t = \tlambda(\tau)/e^\tau,
1475: %\qquad\mbox{or}\quad
1476: \ee
1477: or $\lambda(t) = tM(t) \tlambda(\tau)$, since $tM(t)\to1$ as $t\to\infty$.
1478: (Note $tM(t)=1$ if $M_0=\infty$.)
1479:
1480:
1481: %
1482: %
1483: %
1484: %\begin{proof}[Proof of (2)]
1485: {\em Proof of (b).}
1486: It is only the case $M_0=\infty$ that requires some work. Indeed, if
1487: $M_0 < \infty$ we see from \qref{E.smolburgers} that $\nu_{\tau_1}(ds)
1488: = \Lambda_0(M_0 \,ds)$. In what follows, we suppose that $M_0=\infty$.
1489: We then have an eternal solution to Smoluchowski's equation, and $t=e^\tau$.
1490: We shall compare the tails of
1491: $\nu_{0}$ ($\tau=0$) with that of $\Lambda_0$ ($t=0$).
1492:
1493: Claim (b) is a purely analytic fact that follows from Karamata's Tauberian
1494: theorem~\cite{Feller}. We first reformulate it as a statement about
1495: Laplace transforms. Let $\vp_0(q)=\vp(q,0)$, $\psi_0(q)=\psi(q,0)$.
1496: For every $\alpha \in (1,2]$ we have
1497: \begin{equation}
1498: \label{eq:tauber1}
1499: \int_0^s r^2 \nu_0(dr) \sim s^{2-\alpha}L(s) \iff 1-\vp_0'(q) \sim
1500: q^{\alpha-1} L\left(\frac{1}{q}\right) \frac{\Gamma(3-\alpha)}{\alpha-1}
1501: \end{equation}
1502: as $s \to \infty$ and $q \to 0$ respectively (see~\cite[eq. 7.4]{MP1}).
1503: By the same argument,
1504: \begin{equation}
1505: \label{eq:tauber2}
1506: \int_0^s r^2 \Lambda_0(dr) \sim s^{2-\alpha}L(s) \iff \psi_0'(q) \sim
1507: q^{\alpha-1} L\left(\frac{1}{q}\right) \frac{\Gamma(3-\alpha)}{\alpha-1},
1508: \end{equation}
1509: with the following caveat when $\alpha=2$.
1510: If $\int_0^\infty r^2 \Lambda(dr)=\infty$ then
1511: (\ref{eq:tauber2}) holds. On the other hand, if $\int_0^\infty r^2
1512: \Lambda_0(dr) <\infty$ then we must modify the second condition
1513: in (\ref{eq:tauber2}) to
1514: %this is $\lim_{x \to \infty} L(x)$. If
1515: %$\sigma^2 > 0$, then we have
1516: \[\psi_0'(q) \sim \left( \sigma^2 + \int_0^\infty r^2 \Lambda(dr)\right) q,
1517: \quad q \rightarrow 0.\]
1518: We set $t=1$ in (\ref{E.changevars}) and differentiate
1519: \qref{E.psi_soln} with respect to $q$ to obtain
1520: \begin{equation}
1521: \label{eq:psiphi}
1522: \psi_0'(\vp_0(q)) = \frac{1- \vp_0'(q)}{\vp_0'(q)}
1523: = \frac1{\vp_0'(q)}-1.
1524: \end{equation}
1525: The functions $\psi_0'$, $\vp_0$, and $1/\vp'$
1526: are strictly increasing. Since $\vp_0'(0)=1$
1527: we also have $\vp_0(q)= q(1+o(1))$ as $q \to 0$. A sandwich
1528: argument as in~\cite{Feller} may now be used to deduce claim (b).
1529: First suppose that (\ref{eq:tauber1}) holds. Fix $b,\ve>0$.
1530: Then for $q$ sufficiently small we use monotonicity
1531: and (\ref{eq:psiphi}) to obtain
1532: \[
1533: \frac{1- \vp_0'(bq(1-\ve))}{1-\vp_0'(q(1+\ve))}
1534: \frac{\vp_0'(q(1+\ve))}{\vp_0'(bq(1-\ve))}
1535: <
1536: \frac{\psi_0'(bq)}{\psi_0'(q)} <
1537: \frac{1- \vp_0'(bq(1+\ve))}{1-\vp_0'(q(1-\ve))}
1538: \frac{\vp_0'(q(1-\ve))}{\vp_0'(bq(1+\ve))}.
1539: \]
1540: Letting first $q$ and then $\ve \to 0$, we obtain
1541: \[ \lim_{q \to 0} \frac{\psi_0'(bq)}{\psi_0'(q)} = b^{\alpha-1}.\]
1542: Thus, $\psi_0'$ is regularly varying with exponent $\alpha-1$.
1543: Similarly, if we assume that (\ref{eq:tauber2}) holds, we sandwich
1544: \[
1545: \frac{\psi_0'(b(1-\ve)q)}{\psi_0'((1+\ve)q)} <
1546: \frac{1-\vp_0'(bq)}{1-\vp_0'(q)}
1547: \frac{\vp_0'(q)}{\vp_0'(bq)}.
1548: <
1549: \frac{\psi_0'(b(1+\ve)q)}{\psi_0'((1-\ve)q)},
1550: \]
1551: to deduce that $1-\vp_0'$ is regularly varying with exponent
1552: $\alpha-1$. Finally, since $\vp_0'(0)=1$ it follows from (\ref{eq:psiphi}) that
1553: $\lim_{q \to 0} {\psi_0'(q)}/(1-\vp_0'(q))=1.$
1554: This finishes the proof of Theorem~1.
1555:
1556: \section{Energy, dissipation and spectra}
1557: \label{S.energy}
1558: In this section, we compute several statistics of physical
1559: interest for the solution increments: mean energy and dissipation,
1560: the law of the Fourier-Laplace transform, and the multifractal spectrum.
1561: While the computations are routine, some
1562: interesting features emerge, namely (i) conservation of energy
1563: despite dissipation at shocks, (ii) a simple evolution rule for the
1564: Fourier-Laplace spectrum, and (iii) a multifractal spectrum in sharp
1565: variance with %[[??]]
1566: that of fully developed turbulence. Simple proofs of fine regularity
1567: properties (e.g., Hausdorff dimension of the set of Lagrangian regular
1568: points) may be found in~\cite{B_burgers}.
1569:
1570: \subsection{Energy and dissipation}
1571: The energy in any finite interval $I \subset \R_+$is computed using
1572: the \LK\/ formula \qref{E.defpsi} and Fubini's theorem as follows.
1573: \ba
1574: \nn
1575: && \E\left(\int_I v(x,t)^2 \, dx \right) =
1576: \int_I \E \left( v(x,t)^2\right) \, dx =
1577: \int_I
1578: \left( \partial_q^2\left. \E \left( e^{q v(x,t)}\right)
1579: \right|_{q=0} \right) dx \\
1580: \nn
1581: &&
1582: = \int_I \left. \partial_q^2 e^{x \psi(q,t)} \right|_{q=0} \, dx
1583: = \int_I \left( x^2 \partial_q \psi (0,t)^2 + x \partial_q^2 \psi(0,t)
1584: \right) \, dx \\
1585: \label{E.energy}
1586: &&
1587: = b_t^2 \int_I x^2\, dx + \left(\int_0^\infty y^2 \Lambda_t(dy)
1588: \right) \int_I x\,dx.
1589: \ea
1590: Let us restrict attention to solutions of mean zero, that is
1591: $b_t=0$. Then we have conservation of energy in the sense that
1592: \be
1593: \label{E.cons2}
1594: \E \left(\int_I v(x,t)^2 \, dx\right)= \E \left( \int_I v(x,0)^2 \, dx
1595: \right), \quad t \geq 0.
1596: \ee
1597: Indeed, by \qref{E.energy}, we see that \qref{E.cons2} is equivalent to
1598: \be
1599: \label{E.cons3}
1600: \partial_q^2 \psi(0,t) = \partial_q^2 \psi_0 = \sigma_0^2 +
1601: \int_0^\infty s^2 \Lambda_0(ds) =: M_2,
1602: \ee
1603: with the understanding that $\partial_q^2\psi(0,t) =\infty$ if
1604: $\int_0^\infty s^2 \Lambda_0(ds)$ is divergent. It is only necessary
1605: to differentiate \qref{E.psi_soln} to obtain
1606: \[ \partial_q \psi(q,t) = \frac{\psi_0'(q-t\psi)}{1+ t \psi_0'(q-t\psi)}, \quad
1607: \partial_q^2 \psi(q,t) = \frac{1}{\left(1+
1608: t\psi_0'(q-t\psi)\right)^3} \psi_0''(q-t\psi), \]
1609: and then take the limit $q\to 0$ to obtain \qref{E.cons2}.
1610:
1611: The dissipation at a shock with left and right limits $u_\pm$ is
1612: % defined to \be \la{E.diss1}
1613: %\frac{(u_- -u_+)^3}{3}. \ee
1614: obtained as follows. The decay
1615: of the $L^2$ norm for solutions to Burgers equations with viscosity
1616: $\eps$, $u_t + uu_x = \eps u_{xx}$, is given by
1617: \[ \frac{d}{dt} \int_{\R} u^2 \, dx = 2 \eps \int_{\R} u_x^2 \, dx. \]
1618: The right hand side may be evaluated exactly for traveling waves (viscous
1619: shocks) of the form $u(x,t)=u^\eps(x-ct)$. It is easily seen that for
1620: any $\eps>0$ a traveling wave profile connecting the states $u_- >
1621: u_+$ at $\mp \infty$ is of the form $u^\eps(x-ct)= w((x-ct)/\eps)$
1622: where $w$ satisfies the ordinary differential equation
1623: \[ -c\left(w-u_-\right) + \frac{1}{2}\left(w^2 -u_-^2 \right) =
1624: \frac{dw}{d\xi},
1625: \quad c = \frac{u_-+u_+}{2}.\]
1626: We therefore have
1627: \ba
1628: \nn
1629: && 2 \eps \int_\R u_x^2 \, dx = 2 \int_\R (w')^2 \, d\xi =2
1630: \int_{u_-}^{u_+} \left[-c(w-u_-) + \frac{1}{2}\left(w^2 -u_-^2
1631: \right) \, \right] dw\\
1632: \nn
1633: &&
1634: = 2(u_--u_+)^3 \int_0^1 w(1-w) \, dw = \frac{(u_--u_+)^3}{3}.
1635: \ea
1636: %multiply both sides by $u^\eps$, integrate, and simplify to find
1637: %\[ 2 \eps \int_{-\infty}^\infty (u^\eps_x)^2 \, dx = (u_--u_+)^3
1638: %\int_0^1 s(1-s) \, ds = \frac{(u_--u_+)^3}{6}.\]
1639: %The decay is well-defined, even if the traveling waves are not in $L^2$.
1640: The right hand side is independent of $\eps$ and captures the
1641: dissipation of the entropy solution in the limit $\eps \to 0$.
1642: The dissipation at shocks in any finite interval $I \subset \R_+$ may now be
1643: computed by summing over all shocks in $I$ using \qref{E.a_repn} and
1644: \qref{E.BV}:
1645: \be
1646: \label{E.diss}
1647: \frac{1}{3}\E \left( \sum_{y \in I} (v(y_-,t)- v(y_+,t))^3\right)
1648: = \frac{1}{3} \E \left( \sum_{y \in I} (s_y^t)^3 \right)
1649: = \frac{|I|}{3} \int_0^\infty s^3 \Lambda_t(ds),
1650: \ee
1651: where $|I|$ is the length of $I$.
1652:
1653: Conservation of energy in the sense described in \qref{E.cons2} is
1654: rather surprising in
1655: view of the dissipation at shocks. In particular, there are
1656: solutions with finite energy ($\int_0^\infty s^2 \Lambda_t(ds) < \infty$),
1657: but infinite dissipation ($\int_0^\infty s^3 \Lambda_t(ds) =
1658: \infty$). However, there is no contradiction, since \qref{E.cons2}
1659: refers to the expected value of the energy in any finite interval $I$, and
1660: the energy dissipated in shocks is compensated by energy input
1661: from the endpoints of $I$.
1662: %Indeed, for any an interval $I=(0,x)$ we have the iden
1663: % On any interval $(0,x)$ we have the identity
1664: %\[ 0 = \frac{d}{dt} \E\left( \int_0^x v(y,t)^2\, dy \right) =
1665: % -\frac{\E(v(y,t)^3)}{3} - \]
1666: %% This is the basic BV estimate. It will appear below
1667: %% in the time change for the mapping to Smoluchowski's equation.
1668:
1669: \subsection{The Fourier-Laplace spectrum}
1670: We show that the law of the Fourier transform $\hat{v}(k,t)$
1671: of paths $x\mapsto v(x,t)$, is determined by a \Levy\/ process
1672: with jump measure $s^{-1}\Ltail_t(s)\, ds$, where
1673: $\Ltail_t(s)= \int_s^\infty \Lambda_t(ds)$.
1674: Here $\Lambda_t(ds)$ denotes the jump measure of $v(x,t)$
1675: (see Theorem~\ref{T.spectrum} below).
1676: The assertion $\hat{v} \sim k^{-1}$ as $k \to \infty$
1677: for white noise initial data is common in the Burgers turbulence
1678: literature (e.g., see \cite{Frisch-Parisi,Woc}).
1679: For the present case of \Levy\/ process initial data,
1680: we show that $\hat{v}(k,t) \sim -iM(t)k^{-2}$ as $k \to \infty$.
1681: In addition, we find precise corrections under additional assumptions on
1682: $\Lambda_t$ (for example, for self-similar solutions).
1683:
1684: These computations with the laws of the Fourier-Laplace
1685: transform should be contrasted with the conventional notion of
1686: the power spectrum.
1687: %, defined to be the measure with density $\E\left(|\hv (k,t)|^2\right)$.
1688: Despite its widespread use for wide-sense stationary processes,
1689: the power spectrum is of limited utility for the present problem
1690: involving stationary increments, as we now show.
1691: Fix $L>0$ and consider the interval $[0,L]$. Almost every sample path
1692: $v(x,t)$ is bounded on $[0,L]$ and we may define the truncated Fourier
1693: transform
1694: \be
1695: \label{E.Fourier1}
1696: \hvl (k, t) = %\frac{1}{L}
1697: \int_0^L e^{-ikx} v(x,t) \, dx.
1698: %\quad k = \frac{2 n \pi}{L},\ n \in \Z.
1699: \ee
1700: %More precisely, these are the Fourier coefficients of the periodic
1701: %extension of $v(x,t)$ restricted to $[0,L]$.
1702: If the energy is finite ($M_2 < \infty$ in \qref{E.cons3}), we may
1703: compute a truncated power spectral density $S_L(k)$ as follows. We have
1704: %$\E(|\hvl(k,t)|^2)$
1705: \be
1706: \label{E.Fourier2}
1707: \frac1L |\hvl(k,t)|^2 = \frac{1}{L} \int_0^L \int_0^L e^{-ik(x-y)} v(x,t)
1708: v(y,t) \, dx \, dy.
1709: \ee
1710: Since $v(x,t)$ is a \Levy\/ process with mean zero,
1711: the autocorrelation is
1712: \be
1713: \mathbb{E}(v(x,t) v(y,t)) = (x\wedge y)M_2.
1714: \ee
1715: We take expectations in \qref{E.Fourier2} to find
1716: \be
1717: \label{E.Fourier3}
1718: % \E\left(|\hvl(0,t)|^2\right) = \frac{M_2 L}{3},
1719: %\qquad
1720: S_L(k):= \frac1L \E\left(|\hvl(k,t)|^2\right) = \frac{2 M_2}{k^2}
1721: \left(1-\frac{\sin kL}{kL}\right),\quad k \neq 0.
1722: \ee
1723: The power spectrum $S(k)=\lim_{L\to\infty}S_L(k)=2M_2/k^2$ is now seen to be
1724: well-defined, but is unsuitable for distinguishing solutions
1725: because all solutions with
1726: the same energy (possibly infinite) have identical power spectrum.
1727:
1728: A well-defined spectrum that distinguishes solutions may be obtained by
1729: taking the Fourier-Laplace transform of process paths.
1730: For fixed $p>0$ we define the random variable
1731: \be
1732: \label{E.Fourier4}
1733: \lap{v}(p,t) = \int_0^\infty e^{-px} v(x,t) \,dx =
1734: \frac{1}{p}\int_0^{\infty} e^{-px} v(dx,t).
1735: \ee
1736: The integrals are well-defined because $\lim_{x\to \infty}
1737: v(x,t)/x=0$ a.s.\ by the strong law of large numbers, and $v(x,t)$ is of
1738: bounded variation. If $s^t$ denotes a point process of shock strengths
1739: as in \qref{E.BV} we have
1740: \be
1741: \label{E.Fourier_CM}
1742: p\lap{v}(p,t) = \frac{M(t)}{p} - \sum_{0\le x} e^{-px}s^t_x.
1743: %\int_0^\infty e^{-px} s^t(dx).
1744: \ee
1745: We determine the law of $p\lap{v}(p,t)$ by computing its Laplace transform
1746: via the `infinitesimal' \LK\/ formula $\E\left( e^{q v(dx,t)} \right) =
1747: e^{\psi(q,t) \, dx}$. We then have
1748: \ba
1749: \nn
1750: \lefteqn{\E \left(e^{q p \lap{v}(p,t)}\right) =
1751: \exp \left( \int_0^\infty \psi( q e^{-px}, t) \, dx \right)} \\
1752: \label{E.Fourier5}
1753: && = \exp \left( \frac{1}{p} \int_0^q \frac{\psi(q',t)}{q'} \,
1754: dq'\right) =: \exp \left( {\frac{1}{p}\pss(q,t)}\right), \quad p,q>0,
1755: \ea
1756: after the change of variables $q'=q e^{-px}$. We now observe that
1757: $\pss$ determines a Laplace exponent as follows.
1758: Let $\Ltail_t(s) = \int_s^\infty \Lambda_t(ds)$ denote the tail of the
1759: \Levy\ measure $\Lambda_t$. Since $\int_0^\infty (s\wedge s^2)
1760: \Lambda_t(ds) < \infty$ we have the bounds
1761: \[ s\Ltail_t(s) \leq \int_s^\infty r \Lambda_t(dr),
1762: %\quad\mathrm{and}
1763: \qquad s^2 \Ltail_t(s) \leq \int_0^\eps r^2 \Lambda_t(dr)
1764: + s^2 \Ltail_t(\eps), \quad s \in (0,\eps). \]
1765: Therefore,
1766: \[
1767: \lim_{s \to \infty} s \Ltail_t(s) = 0,
1768: \qquad \lim_{s \to 0} s^2 \Ltail_t(s)=0,
1769: \]
1770: and we may integrate by parts in \qref{E.defpsi} to obtain
1771: \be
1772: \la{E.pss1}
1773: \frac{\psi(q',t)}{q'}
1774: = \int_0^\infty (1-e^{-q's}) \Ltail_t(s) \, ds.
1775: \ee
1776: Integrating once more in $q'$ we find
1777: \be
1778: \label{E.Fourier6}
1779: \pss(q,t) =
1780: \int_0^q \frac{\psi(q',t)}{q'} \,dq' =
1781: \int_0^\infty (e^{-qs}-1 + qs) \frac{\Ltail_t(s)}{s} \,
1782: ds. %\quad \Ltail_t(s) = \Lambda_t((s,\infty)),
1783: \ee
1784: We integrate by parts in \qref{E.m} to see that
1785: \be
1786: \label{E.Lam1}
1787: \int_0^\infty \Ltail_t(s) \, ds= \int_0^\infty s
1788: \Lambda_t(ds) = M(t) <\infty.
1789: \ee
1790: This enables us to write
1791: \be
1792: \label{E.phss1}
1793: \pss(q,t) = M(t)q - \phss(q,t), \quad
1794: \phss(q,t) = \int_0^\infty (1-e^{-qs}) \frac{\Ltail_t(s)}{s} \, ds.
1795: \ee
1796: Since \qref{E.Lam1} ensures $s^{-1}\Ltail_t(s) \,ds$ satisfies
1797: the finiteness conditions for a jump measure, $\pss$ is a
1798: Laplace exponent for a \Levy\/ process with zero mean drift that we denote
1799: by $Z^t$. Similarly, $\phss$ is the Laplace exponent for a subordinator that
1800: we denote $Y^t$. We summarize our calculations in the identities
1801: \be
1802: \label{E.phss2}
1803: Z^t_r = M(t)r -Y^t_r, \;\;
1804: \E\left(e^{qZ^t_r} \right) = e^{r \pss(q,t)}, \;\; \E\left(e^{-qY^t_r}
1805: \right) = e^{-r \phss(q,t)},\;\; r,q, t >0.
1806: \ee
1807:
1808: The result is that the Laplace spectrum of the solution increments is determined by
1809: \be
1810: \label{E.Fourier7}
1811: \E \left(e^{q p \lap{v}(p,t)}\right) = \E \left( e^{ q Z^t_{1/p}} \right),
1812: \quad q>0,\ p >0,
1813: \ee
1814: which implies that $\lap{v}(p,t)$ has the same law as $p^{-1} Z_{1/p}^t$
1815: for fixed $p>0$. Note that $\lap{v}(p,t)$ is not a \Levy\/ process
1816: in $p$. In fact, for a fixed realization, $\lap{v}(p,t)$ is analytic
1817: in $p$. Nevertheless, its law is determined by the \Levy\/ process $Z^t$.
1818:
1819: We extend this computation to the Fourier spectrum ($p=ik$) as
1820: follows. The calculations leading to \qref{E.Fourier5} hold for
1821: complex $q$ with $\mathrm{Re}(q) \geq 0$, and in particular for $q =
1822: i\xi$, $\xi \in \R$. Moreover, $\lap{v}(p,t)$ is a well-defined random
1823: variable for every $p$ with $\mathrm{Re}(p)>0$. Thus, we may
1824: analytically continue the identity
1825: $E(e^{qp \lap{v}(p,t)}) = \exp( p^{-1}\pss(q,t))$
1826: to all $p$ with $\mathrm{Re}(p)>0$, and
1827: $q=i\xi$. As in \qref{E.LKu}, let $\Pss(\xi,t)=-\pss(i\xi,t)$ define
1828: the characteristic exponent corresponding to the \Levy\/ process $Z^t$.
1829: For $\eps, k>0$ we set $p= \eps +ik$, $\hat{v}(k-i\eps,t)=\lap{v}(\eps+ik,t)$
1830: and pass to the limit $\eps \to 0$
1831: on both sides of \qref{E.Fourier5} to obtain
1832: \ba
1833: \lim_{\eps \dnto 0} \lefteqn{\E \left(e^{i\xi (ik \hat{v}(k-i\eps,t))}
1834: \right) = \exp \left( \frac{1}{ik}\pss(i\xi,t) \right) } \\
1835: && = \exp \left( \frac{i}{k}\Pss(\xi,t) \right) = \E \left( e^{ i\xi
1836: Z^t_{1/k}} \right), \quad \xi \in \R, k>0.
1837: \ea
1838: Thus, for fixed $k >0$, as $\eps\dnto 0$
1839: the random variables $ik\hat{v}(k-i\eps,t)$ converge in law
1840: to the (real) random variable $Z^t_{1/k}$.
1841: We denote this limit by $ik\vfour(k,t)$.
1842: As before we do not assert that
1843: the processes $ik\hat{v}(k-i\eps,t)$ converge in law to the process
1844: $Z^t_{1/k}$, simply the convergence of random variables for fixed
1845: $k$. We summarize our conclusions as follows.
1846: \begin{theorem}
1847: \label{T.spectrum}
1848: Let $Y^t$ be a subordinator with Laplace exponent $\phss(q,t)$ from
1849: \qref{E.phss1}, and let $Z^t$ the \Levy\/ process defined by \qref{E.phss2}.
1850: Then for every fixed $p>0$ and $k>0$ the random variables $p\lap{v}(p,t)$
1851: and $ik \vfour(k,t)$ have the same law as $Z^t_{1/p}$ and $Z^t_{1/k}$,
1852: respectively.
1853: \end{theorem}
1854:
1855: Due to this result and \qref{E.phss2},
1856: we always have the upper bound $ik^2\vfour(k,t) \leq M(t)$ a.s.
1857: This crude bound may be refined as $k \to \infty$ using information
1858: related to the sample path
1859: behavior of subordinators (see~\cite[Ch.~III.4]{B_book}).
1860: %To avoid repetition, we restrict attention to $\vfour$.
1861: %The same conclusions hold for $\lap{v}(p)$.
1862: \begin{corollary}
1863: \label{C.asympt1}
1864: For every $t>0$, $\lim_{k \to \infty} ik^2 \vfour(k,t) = M(t)$ in probability.
1865: %and $\lim_{p \to \infty} p^2 \hat{v}(p,t) =M(t)$
1866: \end{corollary}
1867: \begin{proof}
1868: This follows from the fact that
1869: $\lim_{r \dnto 0}Y^t_r/r=0$ in probability, proved as
1870: follows. By \qref{E.phss2} we have
1871: $\E(e^{-q Y^t_r/r})= e^{-r\Phi_\#(q/r,t)}$, and since
1872: $1-e^{-s}\le 1\wedge s$, by \qref{E.phss1} we have
1873: that the Laplace exponent
1874: \[
1875: r\Phi_\#(q/r,t) = r\int_0^\infty(1-e^{-qs/r})
1876: \frac{\Ltail_t(s)}s\,ds
1877: \le \int_0^\infty (r\wedge qs)
1878: \frac{\Ltail_t(s)}s\,ds \to 0
1879: \]
1880: as $r\dnto0$ for each $q>0$.
1881: Hence $Y^t_r/r\to 0$ in law.
1882: \end{proof}
1883:
1884: A similar conclusion holds for the Laplace spectrum as $p\to\infty$.
1885: Actually, for the subordinator $Y^t_r$, the sample paths have the stronger
1886: property that $\lim_{r\dnto0} Y^t_r/r\to 0$ a.s.~\cite[III.4.8]{B_book}.
1887: Under a mild assumption on the integrability of the small jumps,
1888: we can strengthen convergence in probability to almost-sure
1889: convergence of the Laplace spectrum.
1890:
1891: \begin{corollary}
1892: \label{C.asympt2}
1893: For every $t>0$, $\lim_{p\to\infty} p^2\lap{v}(p,t)=M(t)$ in
1894: probability. If we also assume $\int_0^1 |\log s| \Ltail_t(s)\,ds < \infty$,
1895: then $\lim_{p \to \infty} p^2 \lap{v}(p,t)=M(t)$ a.s.
1896: \end{corollary}
1897: \begin{proof}
1898: For notational convenience, we suppress the dependence on $t$ in the proof.
1899: Fix $\eps >0$, % choose $a$ such that $1 < a< 1 + \eps/M$
1900: and let $p_m= 2^m$ for positive integers $m$. We will show that
1901: $\lim_{m \to \infty} p_m^2\lap{v}(p_m) =M$ a.s. That is, for every
1902: $\eps>0$, we claim
1903: \be
1904: \label{E.bc1}
1905: \prob\left( \left| p_m^2 \lap{v}(p_m) - M \right| > \eps
1906: \ \mbox{infinitely often}\right) =0.
1907: \ee
1908: This is sufficient to establish $\lim_{p\to \infty}p^2 \lap{v}(p)=M$
1909: a.s. Indeed, since $M/p- p\lap{v}(p)$ is completely monotone by
1910: \qref{E.Fourier_CM}, for $p\in (p_m,p_{m+1})$ we have the bounds
1911: \[
1912: 0 < M(t) - p^2\lap{v}(p)
1913: <
1914: \frac{p}{p_m} \left( M(t) - p^2_{m}\lap{v}(p_{m}) \right)
1915: <
1916: 2 \left( M- p_m^2 \lap{v}(p_m)\right) ,
1917: \]
1918: and therefore
1919: \be
1920: \{M-p^2\lap{v}(p) > 2\eps\} \subset \{ M - p_m^2 \lap{v}(p_m) > \eps \}.
1921: %\subset \{M-p_m^2 \lap{v}(p_m) > 2\eps\}
1922: \ee
1923: %
1924: %\{M- p_{m+1}^2 \lap{v}(p_{m+1}) > 2\eps,
1925: % % -(a-1)M \}} \\
1926: % &&
1927: % \nn
1928: % \subset \{M- p_{m+1}^2 \lap{v}(p_{m+1}) > \eps \}
1929: % \ea
1930: % by the choice of $a$.
1931: %Now \qref{E.bc1} implies $\lim_{p\to \infty}p^2 \lap{v}(p)=M$ a.s.
1932:
1933: In order to prove \qref{E.bc1} we use the elementary estimate
1934: \ba
1935: \nn
1936: && \prob \left( \left| p_m^2 \lap{v}(p) - M \right| > \eps \right)
1937: = \prob\left( p_m Y_{1/p_m} > \eps \right)
1938: \leq \frac{e}{e-1}\E\left(1 - \exp \left(-\frac{p_m}{\eps}
1939: Y_{1/p_m}\right) \right) \\
1940: \nn
1941: && = \frac{e}{e-1} \left(1-
1942: \exp\left(-\frac{1}{p_m}\phss(\frac{p_m}{\eps}) \right) \right)
1943: \leq \frac{e}{e-1}
1944: \frac{1}{p_m}\phss(\frac{p_m}{\eps}).
1945: \ea
1946: We will show that $\sum_{m=1}^\infty p_m^{-1} \phss(p_m/\eps) <
1947: \infty$. The first Borel-Cantelli lemma then implies \qref{E.bc1}.
1948: For clarity, we suppose $\eps=1$. This causes no essential
1949: difference and reveals the main computation.
1950:
1951: %In the following, $\pibar$ denotes the integrated tail
1952: %measure for the subordinator $Y^t$:[[move into proof??]]
1953: Denote the integrated tail of the \Levy\ measure for $Y^t$ by
1954: \be
1955: \label{E.pibar}
1956: \pibar_t(s) = \int_s^\infty \frac{\Ltail_t(s')}{s'} \, ds'.
1957: \ee
1958: We integrate by parts and use Tonelli's theorem to find
1959: \be
1960: \label{E.lil3}
1961: \sum_{m=1}^\infty p_m^{-1} \phss(p_m) = \int_0^\infty
1962: \sum_{m=1}^\infty e^{-p_m s} \pibar_t(s) \,ds.
1963: \ee
1964: It is only necessary to check that the integral over $s\in (0,1)$ is
1965: finite. Here we use the elementary estimate
1966: \[
1967: \sum_{m=1}^\infty e^{-2^m s} \leq \int_0^\infty \exp(-e^{x\log 2}s)
1968: \,dx = \frac{1}{\log 2} \int_{s}^\infty e^{-y} \frac{dy}{y} \leq
1969: \frac{|\log s|+1}{\log 2},
1970: \]
1971: so that
1972: \[
1973: \int_0^1 \sum_{m=1}^\infty e^{-p_m s} \pibar_t(s) \,ds \leq
1974: \frac{1}{\log 2} \int_0^1 (1+|\log s|) \pibar_t(s) \,ds.
1975: \]
1976: By the definition of $\pibar_t(s)$ in \qref{E.pibar}, the last integral is
1977: \ba
1978: \nn
1979: && {\int_0^1 |\log s| \int_s^\infty \frac{\Ltail_t(r)}{r} \, dr \, ds
1980: = \int_0^\infty \frac{\Ltail_t(r)}{r} \, dr \int_0^{1\wedge r}
1981: |\log s| \, ds} \\
1982: \nn
1983: &&\qquad \leq \int_0^1 |\log r| \Ltail_t(r)\, dr + \int_0^\infty
1984: \Ltail_t(r)\, dr,
1985: \ea
1986: which is finite by assumption.
1987: \end{proof}
1988:
1989: Corrections to the bound $ik^2\hat{v}(k,t)\le M(t)$
1990: involve the law of the iterated logarithm~\cite[III.4]{B_book}.
1991: The following corollary holds for initial data that is not
1992: BV (so $M_0=+\infty$ and $M(t)=1/t$) with suitably regular small
1993: jumps (`dust').
1994: %We define a rate function
1995: %\be
1996: %\la{E.LIL1}
1997: %h(k,t) = \frac{k \log \log k}{\psi_0(t k \log \log k)+ k \log\log k},
1998: %\quad e < k,
1999: %\ee
2000: %using the map $r \mapsto g_0(r):=\psi_0(tr)+r$ which is the inverse of
2001: %$q\mapsto \Phi(q,t)$ (see \qref{E.main}).
2002: %[[?? Isn't $\Phi_\#$ the relevant Laplace exponent for $Y^t$?]]
2003: % $\phss^{-1}(\phss(q,t),t)=q$, $q >0$. The following corollary holds if
2004: % the small jumps are suitably regular. In
2005: % particular, it holds for the self-similar solutions.
2006: % Recall that $\sigma_0^2$ and $\Lambda_0$ denote the Gaussian
2007: %part and jump measure of the initial data.
2008: \begin{corollary}
2009: \label{C.LIL}
2010: %\marginpar{Restate in terms of $\Lambda_t(s)$ near $0$?}
2011: Assume $\sigma_0\ne0$ and $\alpha=2$, or assume
2012: $\sigma_0=0$ and $\Ltail_0(s)=\int_s^\infty \Lambda(dr)$ is regularly varying at
2013: zero with exponent $-\alpha$ where $\alpha \in (1,2)$.
2014: Then for every $c>0$ and $t>0$ we have
2015: \be
2016: \label{E.LIL7}
2017: -\log \prob\left( \frac{t^{-1}- ik^2\vfour(k,t)}
2018: {h(k\log\log k)} \leq c \right) \sim
2019: \frac %{\alpha-1}
2020: {\log\log k}
2021: {\gamma c^\gamma t^{1+2\gamma}}
2022: %\frac{1-\alpha^{-1}}{(c\alpha )^{1/(\alpha-1)}} % C_{c,\alpha}
2023: , \quad k \to \infty,
2024: \ee
2025: where $\gamma=1/(\alpha-1)$ and $h(k)=k/\psi_0(k)$.
2026: %\be \label{E.hdef}
2027: %h(k) = %\frac{\alpha^\alpha}{t^{1+\alpha}}
2028: %\frac{k \log \log k}{\psi_0( k \log \log k)}.
2029: %\ee
2030: %with $C_{c,\alpha} = (1-1/{\alpha}) (\alpha c)^{1/(1-\alpha)}$.
2031: \end{corollary}
2032:
2033: This corollary is a consequence of~\cite[Lemma III.12]{B_book} and is
2034: associated with the following lemma of independent interest which shows that
2035: the evolution preserves the regularity of the dust.
2036: \begin{lemma}
2037: \label{L.LIL}
2038: (a) Assume that $\sigma_0=0$ and $\Ltail_0(s)$ is regularly varying at
2039: zero with exponent $-\alpha$, $\alpha \in (1,2)$. Then $\Ltail_t(s)$ is
2040: regularly varying at zero with exponent $-1/\alpha$ for every $t>0$.
2041:
2042: (b) If $\sigma_0\neq 0$, then $\Ltail_t(s) \sim (\sigma_0 t)^{-1}
2043: \sqrt{2/(\pi s)}$ as $s \to 0$, for every $t>0$.
2044: \end{lemma}
2045: \begin{proof}
2046: Recall that $t^{-1}-ik^2\hat{v}(k,t)$ agrees in law with
2047: $kY^t_{1/k}$.
2048: Combining \qref{E.psi_t} with \qref{E.Fourier6} we find
2049: that the Laplace exponent of the subordinator $Y^t$ satisfies
2050: \be\label{E.Phish}
2051: \Phi_\#(q,t) = \int_0^q \Phi\left(\frac{q'}{t},t\right) \frac{dq'}{q'}
2052: = \int_0^{q/t} \Phi\left({q'},t\right) \frac{dq'}{q'}.
2053: \ee
2054: We claim that $\Phi(\cdot,t)$, and hence $\Phi_\#(\cdot,t)$, is
2055: regularly varying at $\infty$ with exponent
2056: $\hat\alpha=1/\alpha\in[\frac12,1)$.
2057:
2058: To prove the claim, we integrate by parts in \qref{psi0} to obtain
2059: \[ \frac{\psi_0(q)}{q^2} = \frac{\sigma_0^2}{2}+ \int_0^\infty e^{-qs}
2060: \left(\int_s^\infty \Ltail_0(r) \, dr \right) \, ds. \]
2061: First assume $\sigma_0=0$ and $\Ltail_0$ is regularly varying at
2062: zero with exponent $-\alpha$.
2063: Then $\psi_0$ is regularly varying at infinity with exponent $\alpha$.
2064: This follows from \cite[XIII.5.3]{Feller}, or
2065: may be proved directly. If $\sigma_0 \neq 0$, we have
2066: $\lim_{q \to \infty} \psi_0(q)/q^2=\sigma_0^2/2$.
2067: The Laplace exponent $\Phi(q,t)$ is determined via
2068: the functional relation \qref{E.main}. Since $\alpha >1$, the map
2069: $\Phi \mapsto g_0(\Phi):= \psi_0(t\Phi)+\Phi$ is regularly varying
2070: (in $\Phi)$ at $\infty$ with
2071: exponent $\alpha$. Therefore, the inverse function $\Phi(q,t)$ is
2072: regularly varying (in $q$) at $\infty$ with exponent $1/\alpha$.
2073:
2074: Now let $g_\#(\cdot,t)$ be the inverse function to $\Phi_\#(\cdot,t)$.
2075: Then by \cite[Lemma III.4.12]{B_book}
2076: we infer that for every $\hat c>0$,
2077: \be
2078: \label{E.LIL8}
2079: -\log \prob\left( \frac{k Y^t_{1/k}} {h_\#(k\log\log k,t)} \leq \hat{c} \right) \sim
2080: (1-\hat\alpha)(\hat\alpha/\hat{c})^{{\hat\alpha}/(1-\hat\alpha)}
2081: \log\log k, \quad k \to \infty,
2082: \ee
2083: where
2084: \[
2085: h_\#(k,t) = \frac{ k}{g_\#(k,t)}.
2086: \]
2087: By \qref{E.Phish} and regular variation we have
2088: $\Phi_\#(q,t)\sim \Phi(q/t,t)/\hat\alpha$ as $q\to\infty$,
2089: and thus by \qref{E.main} we find that as $q\to\infty$,
2090: \[
2091: g_\#(q,t)\sim t g_0(\hat\alpha q)\sim t\psi_0(\hat\alpha tq)
2092: \sim t^{1+\alpha}\alpha^{-\alpha} \psi_0(q).
2093: \]
2094: Substituting $\hat c=ct^{1+\alpha}\alpha^{-\alpha}$ into \qref{E.LIL8}
2095: yields Corollary~\ref{C.LIL}.
2096:
2097: Karamata's Tauberian theorem and the monotone density theorem now imply that
2098: $\Ltail_t(s)$ is regularly varying at zero with exponent
2099: $-1/\alpha$. If $\sigma_0 \neq 0$, we find $\Phi(q,t) \sim (\sigma_0
2100: t)^{-1} \sqrt{2q}$ as $q \to \infty$. Assertion (b) of the Lemma then follows from
2101: the Tauberian theorem.
2102: \end{proof}
2103:
2104: For the self-similar solutions, $\psi_0(q)=q^\alpha$ with $\alpha \in (1,2]$,
2105: $\Ltail_0(s)=s^{-\alpha}/(\alpha\Gamma(-\alpha))$ for $\alpha\in(1,2)$, and we have
2106: $\Ltail_t(s) \sim t^{-1}(ts)^{-1/\alpha}/\Gamma(1-1/\alpha)$ as $s \to 0$.
2107:
2108: \subsection{The multifractal spectrum}
2109: The notion of a multifractal spectrum was introduced by Frisch and
2110: Parisi to describe the intermittency of velocity fields in fully developed
2111: turbulence~\cite{Frisch-Parisi}. The multifractal spectrum $d(h)$ measures the
2112: dimension of the set $S_h$ where the velocity field has singularities of
2113: order $h$. There are different mathematical formulations of
2114: multifractality, corresponding to different notions of what one means
2115: by singularities of order $h$. Here we follow the treatment by
2116: Jaffard, which yields $d(h)$ rather easily~\cite{Jaffard} (the
2117: notation has been changed slightly for consistency with this
2118: article).
2119:
2120: We say a function
2121: $v:\R_+ \to \R$, is $C^r(x_0)$ for a point $x_0 \in \R_+$ if there is
2122: a polynomial $P_{x_0}$ of degree at most $[r]$ such that
2123: \[ |v(x)-P_{x_0}(x)| \leq C|x-x_0|^r, \]
2124: in a neighborhood of $x_0$. The \Holder\/ exponent of $v$ at $x_0$ is
2125: defined as
2126: \[ h_v(x_0)= \sup\{r\left| v \in C^r(x_0)\right. \}. \]
2127: We define $S_h$ to be the set of points where $v$ is of \Holder\/
2128: exponent $h$. The multifractal spectrum $d(h)$ is the Hausdorff
2129: dimension of $S_h$. If $S_h$ is empty, the convention is
2130: $d(h)=-\infty$. As an example, let us compute the multifractal
2131: spectrum when the initial data is of bounded variation. Then $M_0< \infty$
2132: in~\qref{E.BV} and there is a finite number of
2133: shocks $s^t_y$ in a finite interval $[0,x]$ with probability
2134: $1$. Suppose $x_0$ is not a shock location for $v(\cdot,t)$. Then~\qref{E.BV}
2135: shows $v$ is analytic near $x_0$ and $h_v(x_0) = \infty$. If $x_0$ is a shock
2136: location, then $v \in C^{-\eps}(x_0)$ for every $\eps >0$, so that
2137: $h_v(x_0)=0$. Thus, we have
2138: simply $d(0)=0$ and $d(h)= -\infty$ for every $h \neq 0$.
2139:
2140: The multifractal spectrum is more interesting for initial data of
2141: unbounded variation, that is, when \qref{E.sigma1} holds. In this case, the
2142: jumps in $v$ are dense. Following Jaffard~\cite{Jaffard},
2143: the multifractal spectrum is computed as follows. We define
2144: %If $\Lambda_t$ denotes the \Levy\/ measure for $v(\cdot,t)$,
2145: \be
2146: \label{E.multfrac}
2147: C_j(t) = \int_{2^{-j-1}}^{2^{-j}} \Lambda_t(ds), \quad
2148: \spec_t = \max \left(0, \limsup_{j \to \infty} \frac{\log C_j(t)}{j
2149: \log 2} \right).
2150: \ee
2151: For any $t >0$, $v(\cdot, t)$ has no Brownian component. It then
2152: follows from~\cite[Thm. 1]{Jaffard} that
2153: \be
2154: \label{E.multfrac2}
2155: d_t(h) = \left\{ \begin{array}{ll} \spec_th, &\quad h \in [0,
2156: 1/\spec_t], \\ -\infty, &\quad \mathrm{else}. \end{array} \right.
2157: \ee
2158: Experiments suggest that the multifractal spectrum for fully developed
2159: three-dimensional turbulence is a concave
2160: curve~\cite[Fig. 2]{Meneveau-Sreenivasan}. This is in clear contrast
2161: with \qref{E.multfrac2}.
2162:
2163:
2164: For example, let us compute the multifractal spectrum for the
2165: self-similar process $V^\alpha$ of index
2166: $\alpha \in (1,2]$. Since $\Lambda^\alpha_t(s)$ is a
2167: scaled copy of $\Lambda^\alpha_1(s)$, $d_t(h)$ is independent of $t$.
2168: We use \qref{E.ndensity} to obtain the asymptotics as $s \to
2169: 0$:
2170: \[ \Lambda^\alpha_1(ds) = f_\alpha(s) \, ds \sim \frac{\sin \pi
2171: \beta}{\pi} s^{\beta-2}
2172: \Gamma(2-\beta) \, ds , \quad \beta = \frac{\alpha-1}{\alpha}. \]
2173: We then have $\spec_t= \alpha^{-1}, t >0$ and
2174: \be
2175: \label{E.multfrac3}
2176: d(h) = \left\{ \begin{array}{ll} h/\alpha, &\quad h \in [0, \alpha],
2177: \\ -\infty, &\quad \mathrm{else}. \end{array} \right.
2178: \ee
2179: In particular, \qref{E.multfrac3} implies that $d(\alpha)=1$, that is
2180: $v(x,t)$ is $C^{\alpha}(x)$ for a.e $x \in \R_+$. For this set a finer
2181: characterization of the local variation of $v(\cdot,t)$ may be
2182: obtained by using the Fristedt-Pruitt law of the iterated
2183: logarithm (see~\cite[Cor.~1]{B_burgers}). However, the
2184: multifractal spectrum, also describes sets
2185: $S_h$, $0<h< \alpha$, that are not covered by the Fristedt-Pruitt law.
2186:
2187:
2188: \section{Acknowledgement}
2189: This material is based upon work supported by the National Science Foundation
2190: under grants DMS 03-05985 and DMS 04-05343. G.M. thanks the University
2191: of Crete for hospitality during part of this work.
2192: %(Center for Nonlinear
2193: %Analysis)
2194:
2195: \bibliographystyle{siam}
2196: \bibliography{mp4}
2197: \end{document}
2198:
2199:
2200:
2201:
2202:
2203:
2204:
2205:
2206:
2207: