1: \documentclass[reqno,11pt]{article}
2: \usepackage{amssymb,amsthm,amsmath,amsfonts}
3: \usepackage{epsfig}
4:
5: \pagestyle{myheadings}
6: \markboth{G. Menon and R. L. Pego}{G. Menon and R. L. Pego}
7: \bibliographystyle{siam}
8: \theoremstyle{plain}
9: \begingroup
10: \newtheorem*{metathm}{Metatheorem}
11: \newtheorem{thm}{Theorem}[section]
12: \newtheorem{cor}[thm]{Corollary}
13: \newtheorem{lemma}[thm]{Lemma}
14: \newtheorem{prop}[thm]{Proposition}
15: \newtheorem{conj}[thm]{Conjecture}
16: \newtheorem*{refthm}{Theorem}
17: \endgroup
18:
19: \newcommand{\nwc}{\newcommand}
20: \nwc{\bit}{\begin{itemize}}
21: \nwc{\eit}{\end{itemize}}
22: \nwc{\Levy}{L\'evy}
23: \nwc{\LK}{L\'evy-Khintchine}
24: \nwc{\LI}{L\'evy-It\^{o}}
25: \nwc{\be}{\begin{equation}}
26: \nwc{\ee}{\end{equation}}
27: \nwc{\ba}{\begin{eqnarray}}
28: \nwc{\ea}{\end{eqnarray}}
29: \nwc{\la}{\label}
30: \nwc{\nn}{\nonumber}
31: \nwc{\Z}{\mathbb{Z}}
32: \nwc{\C}{\mathbb{C}}
33: \nwc{\E}{\mathbb{E}}
34: \nwc{\R}{\mathbb{R}}
35: \nwc{\N}{\mathbb{N}}
36: \nwc{\attr}{\mathcal{A}_{\rm p}}
37: \nwc{\attrx}{\mathcal{A}}
38: \nwc{\PP}{\mathcal{P}}
39: \nwc{\PPE}{\mathcal{P}(E)}
40: \nwc{\M}{\mathcal{M}}
41: \nwc{\Tt}{T^{(t)}}
42: \nwc{\Ut}{U^{(t)}}
43: \nwc{\gtx}{g^{(t)}_x}
44: \nwc{\law}{\stackrel{\mathcal{L}}{\rightarrow}}
45: \nwc{\eqd}{\stackrel{d}{=}}
46: \nwc{\vp}{\varphi}
47: \nwc{\Vp}{\Phi}
48: \nwc{\psilevy}{\Psi}
49: \nwc{\ve}{\varepsilon}
50: \nwc{\veps}{\varepsilon}
51: \nwc{\eps}{\ve}
52: \nwc{\betarsc}{\beta}
53: \nwc{\cl}{c\'{a}dl\'{a}g}
54: \nwc{\qref}[1]{(\ref{#1})}
55: \nwc{\D}{\partial}
56: \nwc{\Ebar}{{\bar{E}}}
57: \nwc{\ebar}{[0,\infty)}
58: \nwc{\mmt}{m}
59: \nwc{\fzero}{F_\rho} %{F_{0,\rho}}
60: \nwc{\fone}{M_\rho} %{F_{1,\rho}}
61: \nwc{\ip}[1]{\langle #1 \rangle}
62: \nwc{\ipbig}[1]{\left\langle #1 \right\rangle}
63: \nwc{\Lip}{\mathop{\rm Lip}\nolimits}
64: \nwc{\Tmin}{T_{\min}}
65: \nwc{\Tmax}{T_{\max}}
66: \nwc{\Tgel}{T_{\rm gel}}
67: \nwc{\LL}{\mathcal{L}}
68: \nwc{\mudot}{\mu}
69: \nwc{\nudot}{\nu}
70: \nwc{\rme}{{\rm e}}
71: \nwc{\rmi}{{\rm i}}
72: \nwc{\nsup}{^{(n)}}
73: \nwc{\nsupj}{^{(n_j)}}
74: \nwc{\ksup}{^{(k)}}
75: \nwc{\jsup}{^{(j)}}
76: \nwc{\tsup}{^{(t)}}
77: \nwc{\nksup}{^{(n_k)}}
78: \nwc{\inv}{^{-1}}
79: \nwc{\qxfac}{(1-\rme^{-qx})}
80: \nwc{\Mcan}{\mathcal{S}_d}
81: \nwc{\Mcanx}{\mathcal{S}}
82: \nwc{\sm}{G}
83: \nwc{\dsm}{H}
84: \nwc{\Hmap}{{\mathfrak S}_p}
85: \nwc{\Hmapx}{{\mathfrak S}}
86: \nwc{\canmap}{f}
87: \nwc{\canonical}{\mathcal{M}}
88: \nwc{\dnto}{\downarrow}
89: \nwc{\upto}{\uparrow}
90: \nwc{\cpair}{c} % For Levy pair
91: \nwc{\intR}{\int_0^\infty}
92: \nwc{\tnu}{\tilde\nu}
93: \nwc{\smeas}{s-measure} %{$s$-measure}
94: \nwc{\smeass}{s-measures} %{$s$-measures}
95: \nwc{\barsmeas}{$\overline{\mbox{s}}$-measure} %{$\bar s$-measure}
96: \nwc{\barsmeass}{$\overline{\mbox{s}}$-measures} %{$\bar s$-measures}
97: \nwc{\nubar}{\bar{\nu}}
98: \nwc{\mubar}{\bar{\mu}}
99: \nwc{\dust}{a_0}
100: \nwc{\bdust}{b_0}
101: \nwc{\gel}{a_\infty}
102: \nwc{\bgel}{b_\infty}
103: \nwc{\ggel}{g_\infty}
104: \nwc{\Fbar}{\bar{F}}
105: \nwc{\Fcheck}{\check{F}}
106: \nwc{\nucheck}{\tilde{\nu}}
107: \nwc{\hvp}{\hat{\vp}}
108: \nwc{\hpsi}{\hat{\psi}}
109: \nwc{\htt}{\hat{t}}
110: \nwc{\hq}{\hat{q}}
111: \nwc{\hs}{\hat{s}}
112: \nwc{\dist}{\mathop{\rm dist}\nolimits}
113: \nwc{\extsol}{\Fbar}
114: \nwc{\specialF}{\hat\Fbar}
115: \nwc{\esm}{\bar{\sm}}
116: \nwc{\tlam}{\tilde\lambda}
117: \nwc{\lgr}{a}
118: \nwc{\Fgr}{F_{\rho,\gamma}}
119: \nwc{\vpgr}{\vp_{\rho,\gamma}}
120: \renewcommand{\Re}{\mathop{\rm Re}\nolimits}
121: \renewcommand{\Im}{\mathop{\rm Im}\nolimits}
122:
123: \theoremstyle{definition}
124: \newtheorem{defn}{Definition}[section]
125: \newtheorem{rem}[defn]{Remark}
126: \newtheorem{counterex}[defn]{Counterexample}
127: \newtheorem{ex}[defn]{Example}
128: \theoremstyle{remark}
129: \newtheorem{hyp}{Assumption }
130: \newtheorem{notation}{Notation}
131: \renewcommand{\thenotation}{} % To make notation environment unnumbered
132: \newtheorem{terminology}{Terminology}
133: \renewcommand{\theterminology}{} % To make terminology environment unnumbered
134: %%%%----------------------------------------------------------------------
135: %%%%----------------------------------------------------------------------
136: %% The following causes equations to be numbered within sections:
137: \numberwithin{equation}{section}
138: \numberwithin{figure}{section}
139: %%%%----------------------------------------------------------------------
140: %%%%----------------------------------------------------------------------
141:
142: \begin{document}
143: \title{The scaling attractor and ultimate dynamics for Smoluchowski's
144: coagulation equations}
145: \author{Govind Menon\textsuperscript{1} and Robert. L.
146: Pego\textsuperscript{2}}
147:
148: \date{\today}
149:
150: \maketitle
151:
152: \begin{abstract}
153: We describe a basic framework for studying dynamic scaling
154: that has roots in dynamical systems and probability theory.
155: Within this framework, we study Smoluchowski's coagulation equation
156: for the three simplest rate kernels $K(x,y)=2$, $x+y$ and $xy$.
157: In another work,
158: we classified all self-similar solutions and all universality
159: classes (domains of attraction) for scaling limits under weak
160: convergence (Comm.\ Pure Appl.\ Math 57 (2004)1197-1232). Here
161: we add to this a complete description of the set of all limit
162: points of solutions modulo scaling (the {\it scaling attractor})
163: and the dynamics on this limit set (the {\em ultimate
164: dynamics\/}). The main tool is Bertoin's \LK\/ representation
165: formula for eternal solutions of Smoluchowski's equation (Adv.\
166: Appl.\ Prob.\ 12 (2002) 547--64). This representation
167: {\em linearizes} the dynamics on the scaling attractor,
168: revealing these dynamics to be
169: conjugate to a continuous dilation, and chaotic in a classical sense.
170: Furthermore, our study of scaling limits explains how Smoluchowski dynamics
171: ``compactifies'' in a natural way that accounts for clusters
172: of zero and infinite size (dust and gel).
173: \end{abstract}
174:
175: \noindent
176: {\bf Keywords:\/} dynamic scaling,
177: agglomeration, coagulation, coalescence,
178: infinite divisibility, \Levy\ processes, \LK\/ formula,
179: stable laws, universal laws, semi-stable laws, Doeblin solution
180:
181: \medskip
182: \noindent
183: {\bf MSC classification:\/} 82C22, 44A10, 45K05, 60F05
184:
185: \medskip
186:
187: \footnotetext[1]
188: {Division of Applied Mathematics, Box F, Brown University, Providence, RI 02912.
189: Email: menon@dam.brown.edu}
190: \footnotetext[2]{Department of Mathematical Sciences,
191: Carnegie Mellon University, Pittsburgh, PA 15213.
192: Email: rpego@cmu.edu}
193:
194: \pagebreak
195: \section{Introduction}
196: \label{sec:intro}
197: Smoluchowski's coagulation equation is a fundamental mean-field model
198: of clustering processes. The merging of clusters of mass $x$ and mass
199: $y$ to produce clusters of mass $x+y$
200: occurs at a mass-action rate modulated by a symmetric rate kernel
201: $K(x,y)$. Formally, the evolution equation for the density $n(t,x)$
202: of the size distribution reads
203: \begin{eqnarray}
204: {\partial_t n}(t,x) &=&
205: \frac{1}{2} \int_0^x K(x-y,y) n(t,x-y) n(t,y) dy
206: \nonumber\\
207: &&\ -\ \int_0^\infty K(x,y) n(t,x) n(t,y) dy,
208: \label{eq:smol1}
209: \end{eqnarray}
210: Many kernels arising in applications are homogeneous, that is, there
211: is $\gamma$ such that $K(\alpha x, \alpha y) = \alpha^\gamma K(x,y)$
212: for every $\alpha,x,y >0$. We restrict attention to the ``solvable''
213: kernels $K(x,y)=2$, $x+y$ and $xy$ ($\gamma=0$, 1 and 2 respectively)
214: which may be
215: studied via the Laplace transform. These special kernels arise in a
216: variety of applications,
217: including the aggregation of colloids \cite{Chandra}($K=2$),
218: droplet formation in clouds \cite{Golovin},
219: %gravitational clustering \cite{},
220: a phase transition for parking \cite{Chassaing},
221: Burgers' model of turbulence \cite{B_burgers} ($K=x+y$),
222: random graphs~\cite{Aldous} ($K=xy$), and kinetics of
223: polymerization~\cite{Ziff} (all kernels).
224:
225: As time proceeds, the typical cluster size grows, and
226: an issue of relevance for homogeneous kernels is whether and how the size
227: distribution develops toward self-similar form. In the
228: physics literature, this is called the {\em dynamic scaling\/} problem.
229: This work continues our study of dynamic scaling for Smoluchowski's
230: equation for the solvable kernels. We lay out a
231: general framework for the analysis of dynamic scaling that is
232: inspired by elements from both dynamical systems and probability
233: theory. The main issues may be set out as a list of basic and
234: general questions:
235: \bit
236: \item What scaling solutions exist?
237: Here we seek {\em self-similar solutions\/},
238: or fixed points of the dynamics modulo scaling.
239: \item What are the domains of attraction of these scaling solutions?
240: These comprise the {\em universality classes\/} for dynamic scaling.
241: \item What limit points are possible under scaling dynamics in general?
242: We call the set of such points the {\it scaling attractor} of the system.
243: \item How can we describe the dynamics on the scaling attractor?
244: We call this the {\it ultimate dynamics} of the system.
245: \item How complicated can the ultimate dynamics be?
246: \eit
247:
248: While evidently stated in dynamical terms, strong motivation for this
249: framework comes from the classical limit theorems of probability theory
250: developed by the pioneers of the subject in the 1930s.
251: These theorems concern limits of scaled sums of independent and
252: identically distributed random variables $X_1$, $X_2$, $\ldots$ with
253: common distribution $F$.
254: The distribution of $Y_n=\sum_{j=1}^n X_j$ is the $n$-fold convolution
255: of $F$ with itself and this can be regarded as a discrete evolution
256: law for $F$. The theory as laid out in the marvelous book of Feller
257: \cite{Feller} provides complete answers to the questions above:
258: \bit
259: \item {\em Scaling solutions.}
260: The normal distribution is the unique invariant distribution of finite
261: variance. More generally, all
262: invariant distributions (including those with heavy tails)
263: comprise a two-parameter family, the {\em stable laws\/}
264: first characterized completely by L\'{e}vy (see~\cite{Loeve} for a
265: historical account).
266: %The densities for these distributions may be
267: %written as $p(x;\alpha,\theta)$, where $x \in \R$, $0 < \alpha \leq
268: %2$, $|\theta| \leq \min\{\alpha, 2-\alpha\}$. The parameter $\alpha$
269: %measures the algebraic decay of $p(x;\alpha,\theta)$ as $|x| \to
270: %\infty$. At the endpoint $\alpha=2$ we recover the normal distribution.
271: \item {\em Domains of attraction.}
272: The central limit theorem states that the normal law
273: attracts all distributions with finite variance. The domains of
274: attraction of the stable laws are completely
275: classified in terms of the power-law
276: behavior (more precisely, {\em regular variation\/}) and skewness of
277: the 2nd-moment distribution function. There
278: are no other invariant distributions or domains of attraction in
279: the limit $n\to\infty$.
280: \item {\em Infinite divisibility.}
281: The most general distributions that can arise
282: as rescaled limits for some subsequence $n_j\to\infty$ are the
283: {\em infinitely divisible distributions\/}, characterized
284: by the famous \LK\/ formula.
285: The characteristic function $\varphi(k)
286: = \mathbb{E}(e^{ikX})$ of an infinitely divisible random variable $X$
287: is of the form $\varphi(k) = \exp\Psi(k)$ where
288: \[\Psi(k) =
289: \int_\mathbb{R} \frac{(e^{ikx}-
290: 1 -ik\sin x)}{x^2} M(dx) + ibk
291: \]
292: with $\int_{\R} (1\wedge x^{-2}) M(dx) < \infty$ and $b\in\R$.
293: The measure $M$ is called the {\em canonical measure}.
294: \item {\em Ultimate dynamics.}
295: Under appropriate rescaling, nondegenerate limit points for the
296: discrete evolution $n\mapsto S_n$ correspond one-to-one to
297: {\it \Levy\/ processes} (continuous-time random walks with stationary
298: independent increments).
299: These are stochastic processes $X_t$ ($t>0$)
300: with right continuous paths with left limits,
301: obeying the semigroup formula
302: $\varphi_t(k) = \mathbb{E}(e^{ikX_t}) = e^{t\Psi(k)}$.
303: Rescaling amplitude and $t$ (scaling dynamics) corresponds to
304: linear tranformations (dilation, stretching, shifts) of the canonical measure.
305: %[Stop here? If not, strictly
306: % speaking, one should describe the full \LK\/ representation
307: % including the Gaussian coefficient and mean and not just the jump
308: % measure] in terms of the \Levy\
309: %measure, which is the jump-size distribution for a continuous-time random walk
310: %whose time-one increments follow the given infinitely divisible
311: %law.
312: \item {\em Chaos.}
313: There exist distributions ({\it Doeblin's universal laws})
314: for which {\it every} infinitely divisible distribution appears as
315: some limit point for rescaled self-convolutions.
316: %Thus the ultimate dynamics among the infinitely divisible laws is chaotic
317: %in a sense.
318: \eit
319:
320: In an earlier article \cite{MP1}, we characterized the approach to
321: self-similarity for Smoluchowski's equation as $t\to\infty$ for $K=2$
322: and $x+y$, and the approach to self-similar blow-up as
323: $t$ approaches the {\em gelation time} for $K=xy$.
324: By way of addressing the first two issues in the framework above we found:
325: \bit
326: \item
327: For each solvable kernel, with degree of homogeneity $\gamma\in\{0,1,2\}$, there
328: is a (classically known) self-similar solution with finite
329: $\gamma+1$st
330: moment, unique up to normalization. This solution has
331: exponential decay as $x \to \infty$.
332: But more generally, there is a
333: one-parameter family of self-similar solutions, corresponding to
334: distribution functions
335: written $F_{\rho,\gamma}$ for $\rho \in (0,1], \gamma \in \{0,1,2\}$.
336: We recover the classical self-similar solution at the endpoint $\rho=1$.
337: For $0 < \rho < 1$, the $F_{\rho,\gamma}$ have
338: infinite $\gamma+1$st moment and are directly related
339: to important heavy-tailed distributions of probability theory ---
340: Mittag-Leffler distributions for $K=2$, and stable laws of maximum skewness
341: for $K=x+y$ and $xy$. For $K=x+y$ these solutions were first discovered by
342: Bertoin by a different argument~\cite{B_eternal}.
343:
344: \item The classical self-similar solution with $\rho=1$ attracts all
345: solutions with finite $\gamma+1$st moment.
346: In general, the domains of attraction of self-similar solutions
347: are characterized by
348: the regular variation of the $\gamma+1$st moment distribution. A positive
349: measure $\nu_0$ lies in the domain of attraction of the self-similar
350: solution $F_{\rho,\gamma}$ if and only if
351: \begin{equation}\label{1.domain}
352: \int_0^x y^{\gamma+1}\nu_0(dy) \sim x^{1-\rho}L(x), \qquad x\to \infty,
353: \end{equation}
354: for some function $L$ slowly varying at $\infty$ ($L(\lambda x)/L(\lambda)\to1$
355: as $\lambda\to\infty$ for all $x>0$). There are no other self-similar
356: solutions or domains of attraction.
357: \eit
358:
359: Our present goal is to investigate more generally the
360: scaling dynamics of Smoluchowski's equation for the solvable
361: kernels, as $t\to\infty$ for $K=2$ and $x+y$ and as $t$ approaches
362: the gelation time for $K=xy$.
363: To stress the probabilistic analogy, if our earlier
364: article was a study of stable laws, here we study infinite
365: divisibility. The basis for our work is Bertoin's characterization
366: of {\it eternal solutions\/}
367: for Smoluchowski's equation with additive kernel
368: $K=x+y$~\cite{B_eternal}. Eternal solutions for this kernel are defined
369: for all $t\in(-\infty,\infty)$
370: (i.e., they may be extended {\em backwards\/} in time globally),
371: and thus they are `infinitely divisible' under clustering dynamics.
372: Bertoin established a remarkable \LK-type representation for these
373: solutions. We generalize this result to other solvable kernels, and
374: revisit it from a dynamical systems viewpoint. In the context of the
375: framework above, we find that
376: \bit
377: \item The (proper) scaling attractor corresponds in one-to-one fashion with
378: eternal solutions of Smoluchowski's equation, and these have a
379: \LK\/ representation for each solvable kernel.
380: \item The \LK\/ representation {\em linearizes the dynamics on the
381: attractor}. Nonlinear evolution by Smoluchowski's equation on the attractor
382: is conjugate to a group of simple scaling transformations on the
383: measures that generate the representation.
384: Heuristically, the classification of the domains of
385: attraction in~\cite{MP1} shows
386: that the dynamics are extremely sensitive to the tails of
387: the initial size distribution. The \LK\/ representation
388: makes this precise: the ultimate dynamics on the
389: scaling attractor is conjugate to a continuous dilation map.
390: \item One may use the \LK\ representation to construct orbits with
391: complicated dynamics. The scaling attractor contains a dense family
392: of scaling-periodic solutions. Furthermore,
393: there are eternal solutions with
394: trajectories {\em dense} in the scaling attractor ---
395: we call these {\it Doeblin solutions}.
396: And, for any given scaling trajectory, there is a dense set of initial data
397: whose forward trajectories shadow the given one.
398: \eit
399:
400: In addition, this study of scaling limits reveals how Smoluchowski dynamics
401: ``compactifies'' in a natural way that accounts for clusters
402: of zero and infinite size (dust and gel).
403: Considering defective limits on $(0,\infty)$ that concentrate probability
404: at 0 and $\infty$ yields a well-posed dynamics of
405: ``extended solutions'' on $[0,\infty]$.
406: Proper solutions remain fundamental,
407: but considering extended solutions with dust and gel helps to understand
408: just how the tails of initial data determine long-time behavior.
409:
410: We remark that scaling-periodic solutions are analogous to the {\it
411: semi-stable} laws in probability theory \cite{semistable}.
412: Our ``Doeblin solutions'' are constructed by ``packing the tails'' of the
413: \Levy\ measure in a fashion entirely analogous to the construction of
414: Doeblin's universal laws in probability \cite[XVII.9]{Feller}.
415: In this connection, it is interesting to note
416: that the examples in~\cite[XVII.9]{Feller}, dismissed by Feller as
417: ``primarily of curiosity value,'' closely resemble
418: modern treatments of chaos, and Doeblin's construction appears
419: particularly prescient.
420:
421: The solvable cases of Smoluchowski's
422: equation correspond to sophisticated stochastic models with a rich
423: theory (see~\cite{Aldous, B_icm} for excellent reviews),
424: so perhaps it is no accident that the
425: classical probabilistic methods work so well. But let
426: us stress that our work really relies only on the analytical methods
427: for studying scaling limits that lie behind the classical limit theorems.
428: These methods are simple and powerful and should be of utility for
429: understanding scaling phenomena in other applications that have no
430: obvious probabilistic meaning.
431: Thus, no knowledge of probability is presumed and (almost) all details are
432: included (though there is no substitute for reading Feller!).
433:
434: \section{Statement of results}
435: \label{sec:results}
436: In this section, we state our results precisely in a
437: setting that unifies the treatment of dynamic scaling for all the
438: solvable kernels.
439:
440: Let $E$ denote the open interval $(0,\infty)$,
441: $\M$ the space of non-negative Radon measures on $E$, and $\PP$ the space of
442: probability measures on $E$. We will always use the weak topology on
443: $\M$ and $\PP$.
444: We also let $\Ebar$ denote the closed half line with point at
445: infinity, $\Ebar=[0,\infty]=[0,\infty)\cup\{\infty\}$,
446: and let $\bar\PP$ be the space of probability measures on $\Ebar$.
447:
448: Rigorous theories for solutions where
449: $\nu_t(dx)=n(t,x)\,dx$ is a general size-distribution measure on
450: $E=(0,\infty)$, thus accounting for both continuous and discrete
451: size distributions in one general setting, are based on the moment identity
452: \begin{equation}\label{R.moment}
453: \frac{d}{dt}\int_E f(x)\nu_t(dx) =
454: \ \frac12\int_E\int_E
455: \tilde{f}(x,y) K(x,y)\nu_t(dx)\nu_t(dy),
456: \end{equation}
457: where $f$ is a suitable test function and $\tilde
458: f(x,y)=f(x+y)-f(x)-f(y)$, see \cite{MP1,Norris,FL-wellp}.
459: Let $m_\theta(t) := \int_E x^\theta \nu_t(dx)$ denote the $\theta$-th
460: moment of $\nu_t$. By the results of \cite{MP1},
461: for a solvable kernel of homogeneity $\gamma$,
462: any initial measure $\nu_{t_0}$ with finite $\gamma$-th moment
463: $m_\gamma(t_0)$ determines a unique continuous weak solution
464: \be
465: \la
466: {R.smol}
467: \nudot = (\nu_t(dx), t \in [t_0,\Tmax) ).
468: \ee
469: For convenience we can always scale the initial data so that
470: \be
471: \label{R.normal2}
472: t_0:= \left\{ \begin{array}{rl} 1 & (K=2),\\ 0 & (K=x+y),
473: \\ -1 & (K=xy), \end{array} \right.
474: \quad\mbox{and }\ %
475: m_\gamma(t_0)= \int_E x^\gamma \nu_{t_0}(dx) =1.
476: \ee
477: Then
478: $\Tmax=\infty$ for $K=2$ and $x+y$, and $\Tmax=0$ for $K=xy$.
479: %$\Tgel = t_0 + m_2(t_0)^{-1}$ \footnote{[[=??in terms of $m_3$]]}
480: %(the gelation time) for $K=xy$~\cite[\S 2]{MP1}.
481: For each solvable kernel,
482: $m_\gamma(t)=\int_Ey^\gamma\nu_t(dy)$ is an explicitly known
483: function --- from \qref{R.moment} with $f(x)=x^\gamma$ we find
484: \begin{equation}\label{R.mgamma}
485: m_\gamma(t)= \left\{
486: \begin{array}{ll}
487: t^{-1} &(K=2),\\
488: 1 &(K=x+y),\\
489: |t|^{-1} &(K=xy).
490: \end{array} \right.
491: \end{equation}
492:
493: The solution $\nu_t$ is typically not a
494: probability measure because the total number of clusters decreases in
495: time, but there is a naturally associated
496: probability measure $F_t(dx)$ with distribution function $F_t(x)$ defined by
497: %\footnote{Is it supposed to be $(0,x)$ instead of $(0,x]$?? GM: The
498: %normalization $(0,x]$ is used by Feller, and fits better with the S-measures}
499: \begin{equation}\label{R.Fp}
500: %\mu_t((0,x))=
501: F_t(x) = \int_{(0,x]} y^\gamma \nu_t(dy) \left\slash
502: \int_E y^\gamma\nu_t(dy). \right.
503: \end{equation}
504: In this way, we regard Smoluchowski's equation
505: as defining a continuous dynamical system on the phase space $\PP$.
506:
507: \subsection{Eternal solutions}
508: For exceptional initial data $\nu_{t_0}$ we may also solve backwards
509: in time (meaning $\nu_{t_0}$ is {\em divisible\/} under clustering
510: dynamics). The maximum possible interval of existence that can be
511: obtained in this way is denoted $(\Tmin, \Tmax)$, where $\Tmin,\Tmax$
512: depend only on the kernel and $\int_E x^\gamma \nu_{t_0}(dx)$. With
513: the normalization \qref{R.normal2}, the maximum possible interval of
514: existence turns out to be
515: \be
516: \label{R.normal3}
517: (\Tmin,\Tmax) = \left\{ \begin{array}{ll} (0,\infty) & (K=2), \\
518: (-\infty, \infty) & (K=x+y), \\
519: (-\infty,0) & (K=xy). \end{array} \right.
520: \ee
521: Solutions which are defined on this maximum interval of existence
522: are the analog of infinitely divisible laws in probability.
523: \begin{defn}
524: \label{R.eternal}
525: A solution to Smoluchowski's equation that is defined for all
526: $t\in(\Tmin,\Tmax)$ is called an {\em eternal solution}.
527: \end{defn}
528:
529: \subsection{The scaling attractor}
530: A central idea in dynamical systems theory is to understand the
531: long-time behavior of solutions through the notions of
532: attractor and $\omega$-limit sets. Coagulation equations transport mass
533: from small to large scales, and all mass escapes as $t \to \Tmax$.
534: To obtain non-trivial long-time behavior we must rescale solutions. We
535: adopt the following definitions for such {\em scaling dynamical
536: systems\/}. Below, $T_n \in [t_0, \Tmax)$, $\betarsc_n >0$.
537: We will often use the same letter to denote a measure
538: and its distribution function, e.g., $F(x)=\int_{[0,x]}F(dx)$.
539:
540: \begin{defn}\label{R.omegadef}
541: The (proper) {\em scaling $\omega$-limit set} of a solution $\nu$ to
542: Smoluchowski's equation is the set of
543: probability measures $\hat F$ on $E$ for which there exist sequences
544: $T_n\to\Tmax$, $\betarsc_n \to \infty$, such that $F_{T_n}(\betarsc_nx) \to
545: \hat F(x)$ at every point of continuity of $\hat F$.
546: \end{defn}
547:
548: \begin{defn}\label{R.attractor}
549: The (proper) {\em scaling attractor}, $\attr$, is the set of probability
550: measures $\hat F$ on $E$ for which there exists
551: a sequence of solutions $\nudot\nsup$ defined for $t \in [t_0,\Tmax)$, and
552: sequences $T_n \to \Tmax$, $\betarsc_n\to\infty$, such that
553: $F\nsup_{T_n}(\betarsc_nx) \to \hat F(x)$
554: at every point of continuity of $\hat F$.
555: \end{defn}
556:
557: As a consequence of continuous dependence of solutions on
558: initial data (forward and backward in time), we will show that
559: the scaling attractor is an invariant set, and that
560: points on the proper scaling attractor and eternal solutions are
561: in one-to-one correspondence.
562:
563: \begin{thm}\label{RT.AE}
564: \begin{enumerate}
565: \item[(a)]
566: The proper scaling attractor $\attr$ is invariant: If $\nu$ is a solution
567: of Smoluchowski's equation,
568: and $F_t\in\attr$ for some $t$, then the solution is eternal and
569: $F_t\in\attr$ for all $t\in(\Tmin,\Tmax)$.
570: \item[(b)]
571: A probability measure $\hat F$ belongs to
572: $\attr$ if and only if $\hat F(dx) = x^\gamma \nu_{t_0}(dx)$
573: for some eternal solution $\nudot$, where $t_0$ is as in
574: \qref{R.normal2}.
575: \end{enumerate}
576: \end{thm}
577:
578: The perfect definition of an attractor remains elusive (see for
579: example, the discussion in~\cite[Ch. 1.6]{GH}).
580: Definition~\ref{R.attractor} is perhaps the simplest
581: for dynamical systems. It also has the virtue of
582: generalizing the probabilistic notion of domains of partial
583: attraction~\cite[XVII]{Feller}. However, some typical properties
584: that hold in finite-dimensional dynamical systems do not hold here.
585: For example, it need not be the case that every solution has a non-empty
586: scaling $\omega$-limit set. Nor is $\attr$ closed. Defective limits are
587: possible, as shown in \cite{MP1}. See \cite{JainOrey} for a
588: discussion of related issues in the probabilistic context.
589: It turns out that we can cure these defects and account for limits
590: that involve mass concentrating at zero or leaking to infinity, by
591: the simple expedient of allowing limits to be probability measures on
592: $\Ebar=[0,\infty]$.
593:
594: \begin{defn}\label{R.full}
595: The full {\em scaling $\omega$-limit set} of a solution $\nu$ to
596: Smoluchowski's equation is the set of probability measures $\hat F$ on
597: $\Ebar$ with the property in Definition~\ref{R.omegadef}.
598: The full {\em scaling attractor}, $\attrx$, is the set of
599: probability measures $\hat F$ on $\Ebar$ with the property in
600: Definition~\ref{R.attractor}.
601: \end{defn}
602:
603: The space $\bar\PP$ of probability measures on $\Ebar$,
604: equipped with the weak topology, is compact --- any
605: sequence contains a converging subsequence.
606: We will show that Smoluchowski dynamics naturally extends
607: by continuity from $\PP$ to $\bar\PP$. Such ``extended solutions''
608: have probability
609: distributions that may include atoms at $0$ and $\infty$, allowing for the
610: possibility that clusters have zero size (``dust'') or
611: infinite size (``gel'') with positive probability.
612: To interpret these physically, one
613: should recognize that of course $0$ and $\infty$ are idealizations
614: relative to a given scale of measuring cluster size.
615:
616: We defer detailed discussion of extended solutions to
617: section~\ref{S.extend}. There we extend
618: Theorem~\ref{RT.AE} to relate the full scaling attractor $\attrx$
619: (now a compact set that is the closure of $\attr$) to the set of
620: eternal extended solutions.
621: Also, the \LK\ representation and linearization theorems in the next two
622: sections have elegant extensions involving extended solutions.
623: %This representation is used to establish a basic property: $\attrx$ is the
624: %closure of $\attr$.
625: First, however, we think it appropriate to focus on standard weak
626: solutions, and develop the theory without dust in our eyes, so to speak.
627:
628:
629: \subsection{\LK\/ representations}
630:
631: In probability theory, infinitely divisible distributions
632: are parametrized by the \LK\/ representation theorem, which expresses
633: the log of the characteristic function (Fourier transform) in terms of
634: a measure that satisfies certain finiteness conditions.
635: %In particular~\cite[XIII.7]{Feller}, a probability measure
636: %$F$ supported on $\ebar$ is infinitely divisible if and only
637: %if its Laplace transform has the form
638: %$\intR \exp(-qx) F(dx)=\exp(-\Phi(q))$ where
639: %the Laplace exponent $\Phi$ admits the representation
640: In particular~\cite[XIII.7]{Feller}, a function $\omega(q)$
641: is the Laplace transform $\intR \rme^{-qx} F(dx)$
642: of an infinitely divisible probability measure
643: $F$ supported on $\ebar$ if and only if
644: $\omega(q)=\exp(-\Phi(q))$ where
645: the Laplace exponent $\Phi$ admits the representation
646: \be
647: \label{R.LK2a}
648: \Phi (q) = \int_{[0,\infty)} \frac{1-\rme^{-qx}}{x} \sm(dx)
649: \ee
650: for some measure $\sm$ on $[0,\infty)$ that satisfies
651: \be
652: \label{R.gendef}
653: \int_{[0,x]} G(dy)<\infty \quad\mbox{and}\quad
654: \int_{[x,\infty)} y^{-1}\sm(dy)<\infty \quad\mbox{for all $x>0$.}
655: \ee
656: Equivalently,
657: \begin{equation}\label{R.gendef2}
658: \int_{[0,\infty)} (1 \wedge y^{-1}) \sm(dy) <\infty.
659: \end{equation}
660: We need a name for measures with this property,
661: although none seems standard.
662: Such measures determine the jump-size distribution for
663: {\em subordinators} --- increasing continuous-time random walks with
664: stationary independent increments --- hence we call them
665: {\em \smeass}. To handle defective limits, it is
666: convenient to allow $y\inv G(dy)$ to have an atom at $\infty$.
667: %\footnote{[[Compare Feller's canonical measures? ]]}
668: \begin{defn}
669: \label{R.candef}
670: A measure $G$ on $[0,\infty)$ is an {\em \smeas}\ if
671: \qref{R.gendef2} holds.
672: A pair $(G,g_\infty)$ is called an {\em \barsmeas}\ on $\Ebar=[0,\infty]$ if
673: $G$ is an \smeas\ and $g_\infty\ge0$.
674: $g_\infty$ is called the charge of $y^{-1}G(dy)$ at $\infty$,
675: and we will abuse notation by
676: denoting the pair $(\sm,g_\infty)$ by $G$.
677: In addition, we say that an \smeas\ (or \barsmeas) $\sm$
678: is {\em divergent\/} if
679: \be
680: \label{R.candef5}
681: \sm(0)>0 \quad\mbox{or}\quad
682: \int_{E} y^{-1} \sm(dy) = \infty.
683: \ee
684: \end{defn}
685: \noindent
686:
687: Here recall we use the notation $\sm(x)=\int_{[0,x]} \sm(dy)$.
688: If $g_\infty=0$ we identify $G$ with $(G,0)$.
689: The space of \barsmeass\ has a natural weak topology which
690: will prove fundamental in our study of scaling dynamics.
691: \begin{defn}
692: \la{R.proper}
693: A sequence of \barsmeass\ $\sm\nsup$ {\em converges} to an
694: \barsmeas\ $\sm$ as $n\to\infty$, if at every point $x\in(0,\infty)$
695: of continuity of $\sm$ we have
696: \begin{equation}\label{R.candef1}
697: \int_{[0,x]} \sm\nsup(dy)\to \int_{[0,x]} \sm(dy)
698: \end{equation}
699: and
700: \begin{equation}\label{R.candef2}
701: \int_{[x,\infty]} y\inv \sm\nsup(dy)\to \int_{[x,\infty]} y\inv \sm(dy) .
702: \end{equation}
703: \end{defn}
704: The integrals in \qref{R.candef2} include the charge at $\infty$, if
705: any. We note that in view of the weak convergence implied by
706: \qref{R.candef1},
707: convergence of $\sm\nsup$ to an \smeas\ $\sm$ (having $g_\infty=0$)
708: is equivalent to \qref{R.candef1} together with the tightness condition
709: \begin{equation}\label{R.candef3}
710: \int_{[x,\infty]} y\inv \sm\nsup(dy) \to 0 \quad\mbox{as $x\to\infty$,
711: uniformly in $n$}.
712: \end{equation}
713:
714: Bertoin's main theorem in \cite{B_eternal} shows that eternal solutions for
715: $K=x+y$ are in one-to-one correspondence with divergent \smeass.
716: (More precisely, Bertoin formulates his result in terms of ``\Levy\/ pairs,''
717: separating the atom at the origin from a jump measure on $(0,\infty)$).
718: We extend this result as follows.
719: Let $\nu$ be an arbitrary solution to Smoluchowski's equation
720: for a solvable kernel of homogeneity $\gamma$.
721: Since the $\gamma$-th moment of $\nu_t$ is finite,
722: $x^{\gamma+1} \nu_t(dx)$ is an \smeas.
723: Rescaling, we associate with $\nu_t$
724: the \smeas\ $\sm_t$ defined by
725: \be
726: \label{R.cansoln1}
727: \sm_t(dx) = x^{\gamma+1} \nu_t( \lambda(t) dx),
728: %\ee
729: %\quad\mbox{where}
730: \qquad %$\lambda(t), t \in (\Tmin,\Tmax)$ is defined by
731: %\be
732: %\label{R.cansoln2}
733: \lambda(t) = \left\{ \begin{array}{rl} 1 & (K=2),\\ \rme^t &
734: (K=x+y),\\ |t|^{-1} & (K=xy). \end{array} \right.
735: \ee
736: Our choice of rescaling ensures that if the total measure $\sm_t(E)$ is
737: finite for some $t$, then it is constant:
738: $\sm_t(E) = m_{\gamma+1}(t)/\lambda(t)^{\gamma+1}=$const.
739:
740: One computes that
741: \be
742: \label{R.Sgro}
743: \intR y^{-1} \sm_t(dy) =
744: \frac{m_\gamma(t) }{ \lambda(t)^{\gamma}} =
745: \left\{ \begin{array}{rl}
746: t^{-1} & (K=2),\\ \rme^{-t} & (K=x+y), \\ |t| & (K=xy)
747: \end{array} \right.
748: \ee
749: The well-posedness theorem \cite{MP1} implies that solutions of Smoluchowski's
750: equation normalized according to \qref{R.normal2} that exist
751: on any time interval $[t,\Tmax)$
752: {\em are in one-to-one correspondence} with \smeass\ $\hat\sm$ that
753: satisfy
754: \[
755: \hat\sm(0)=0\quad\mbox{and}\quad
756: \intR y^{-1}\hat\sm(dy)= m_\gamma(t)/ \lambda(t)^{\gamma} ,
757: \]
758: via $\hat\sm=\sm_{t}$. Through studying the limit $t\dnto\Tmin$,
759: we find that eternal solutions may be characterized as follows.
760: \begin{thm}
761: \label{RT.LK}
762: \begin{enumerate}
763: \item[(a)]Let $\nu$ be an eternal solution for Smoluchowski's equation with
764: $K=2,x+y$ or $xy$. Then there is a divergent \smeas\ $\dsm$
765: such that
766: $\sm_t$ converges to $\dsm$ as $t \dnto \Tmin$.
767: \item[(b)] Conversely, for every divergent \smeas\ $\dsm$,
768: there is a unique eternal solution $\nu$
769: such that $\sm_t$ converges to $\dsm$ as $t \dnto \Tmin$.
770: \item[(c)]
771: Let $\Hmap\colon\attr\to\Mcan$
772: map the (proper) scaling attractor $\attr$
773: to the set $\Mcan$ of divergent \smeass\
774: by $\Hmap(\hat F)=\dsm$, where $\dsm$ is the divergent \smeas\
775: associated to the eternal solution $\nu$ such that
776: $\hat F(dx)=x^\gamma \nu_{t_0}(dx)$ with $t_0$ as in
777: \qref{R.normal2}. Then $\Hmap$ is a bi-continuous bijection.
778: %\footnote{GM: Is this OK? The attractor isn't
779: % closed, but the statement seems true}
780: %[[?? Let $\sm$ be divergent, $\tilde \sm$ not, and $a_n\dnto0$, then
781: %$a_n\sm+\tilde \sm$ converges to $\tilde \sm$. What gives?]]
782: % Ah, neither $\Mcan$ nor $\attr$ is closed.
783: \end{enumerate}
784: \end{thm}
785: The procedure for obtaining the eternal solution $\nu$ from the
786: divergent \smeas\ $\dsm$ is nonlinear and is different
787: for each kernel (see Theorems~\ref{CT.eternal},~\ref{AT.eternal}, and
788: \ref{MT.eternal} below). It seems natural to call Theorem~\ref{RT.LK} a
789: \LK\/ representation for the scaling attractor $\attr$ --- as we will see,
790: eternal solutions are determined through the Laplace exponents
791: associated with divergent \smeass.
792:
793: In section~\ref{S.extend}, the correspondence in Theorem~\ref{RT.LK}
794: is expanded to one between eternal extended solutions and arbitrary
795: \barsmeass.
796:
797: %The divergence condition \qref{R.candef5} seems peculiar to
798: %Smoluchowski's equation and has important probabilistic meaning in
799: %terms of the smoothness of sample paths of the associated \Levy\/ process.
800:
801: \subsection{Linearization of ultimate dynamics}
802: There are two natural group actions on the class of
803: eternal solutions that are related to scaling dynamics,
804: arising from {\em time evolution} and {\em rescaling of size}.
805: A straightforward but remarkable consequence of the scaling properties
806: of Smoluchowski's equation is that nonlinear dynamics (time evolution)
807: on the scaling attractor $\attr$ is conjugate to a simple linear
808: scaling transformation of the divergent \smeass\ that correspond by
809: Theorem~\ref{RT.LK}.
810:
811: \begin{thm}\label{RT.scale}
812: Let $\nu$ be a solution of Smoluchowski's equation with
813: $K=2$, $x+y$ or $xy$. Given scaling parameters $a>0$ and $b>0$, let
814: \begin{equation}\label{newnu}
815: \tnu_t(dx) = \left\{
816: \begin{array}{ll}
817: a \nu_{at}(b\,dx) &(K=2),\\
818: b \nu_{t+\log a}(b\,dx) &(K=x+y), \\
819: ab^2 \nu_{at}(b\,dx) &(K=xy),
820: \end{array} \right.
821: \end{equation}
822: with associated probability distribution function
823: \begin{equation}\label{newF}
824: \tilde F_t(x) = \left\{
825: \begin{array}{ll}
826: F_{at}(bx) &(K=2 \text{ or }xy),\\
827: F_{t+\log a}(bx) &(K=x+y).
828: \end{array}\right.
829: \end{equation}
830: Then $\tnu$ is again a solution.
831: % on the interval $[\tilde{t_0},
832: % \Tmax)$ where [[ugh??]] %$\tilde{t_0}$ is defined by
833: %\begin{equation}\label{newt0}
834: %\tilde{t}_0 = \left\{
835: %\begin{array}{ll}
836: %a^{-1} t_0 &(K=2 \text{ or }xy),\\
837: %t_0 - \log a &(K=x+y).
838: %\end{array}\right.
839: %\end{equation}
840: If $\nu$ is eternal and $\dsm$ its associated divergent \smeas,
841: then $\tnu$ is eternal and its associated divergent \smeas\ is given by
842: \begin{equation}\label{newdmg}
843: \tilde\dsm(x) = \left\{
844: \begin{array}{ll}
845: ab^{-1}\dsm(bx) &(K=2),\\
846: a^2b^{-1}\dsm(a^{-1}bx) &(K=x+y), \\
847: a^{-2}b^{-1}\dsm(abx) &(K=xy).
848: \end{array} \right.
849: \end{equation}
850: \end{thm}
851:
852: \begin{proof} The proof is simple, based on Theorem~\ref{RT.LK} and
853: the scaling properties of Smoluchowski's equation. First, one
854: checks that \qref{newnu} determines a solution, by
855: scaling the moment identity \qref{R.moment} in each case.
856: Next, compute that if the \smeas\ $\sm_t$ is associated with
857: $\nu_t$ as in \qref{R.cansoln1}, then the corresponding
858: \smeas\ associated with $\tnu_t$ is given by
859: \begin{equation}\label{newmg}
860: \tilde\sm_t(dx) = \left\{
861: \begin{array}{rll}
862: x\tnu_t(dx) &=\ ab^{-1}\sm_{at}(b\,dx) &(K=2),\\
863: x^2\tnu_t(\rme^t\,dx) &=\ a^2b^{-1}\sm_{t+\log a}(a^{-1}b\,dx) &(K=x+y), \\
864: x^3\tnu_t(|t|^{-1}dx) &=\ a^{-2}b^{-1}\sm_{at}(ab\,dx) &(K=xy).
865: \end{array} \right.
866: \end{equation}
867: Then take $t\downarrow\Tmin$ and apply Theorem~\ref{RT.LK} to
868: deduce \qref{newdmg}.
869: \end{proof}
870:
871: \begin{thm}\label{RT.time}
872: Let $\nu$ be an eternal solution with corresponding
873: divergent \smeas\ $\dsm$ and let $F_t$ be as in
874: \qref{R.Fp} for $K=2$, $x+y$ or $xy$. For each
875: $t\in(\Tmin,\Tmax)$, let $\dsm_t=\Hmap(F_t)$
876: be the divergent \smeas\ associated to $F_t\in\attr$.
877: Then
878: \begin{equation}\label{Ht}
879: \dsm_t(x) = \left\{
880: \begin{array}{ll}
881: t \dsm(x) &(K=2),\\
882: \rme^{2t} \dsm(\rme^{-t}x) &(K=x+y),\\
883: |t|^{-2}\dsm(|t|x) &(K=xy).
884: \end{array}\right.
885: \end{equation}
886: \end{thm}
887:
888: \begin{proof} Take $b=1$ and put $t=t_0$ in \qref{newF}, then substitute
889: $a=t$, $\rme^t$, $|t|$ for $K=2$, $x+y$, $xy$ respectively to obtain
890: $\tilde F_{t_0}=F_t$. Then the corresponding divergent
891: \smeas\ $\dsm=\Hmap(\tilde F_{t_0})$ is found from \qref{newdmg}.
892: \end{proof}
893:
894: By this theorem, we see that in terms of the divergent
895: \smeas\ that corresponds to the solution, the time evolution
896: on the scaling attractor $\attr$ is governed by the linear equations
897: \begin{alignat}{2}
898: t \D_t H_t &= H_t &\quad&(K=2),\\
899: (\D_t + x\D_x) H_t &= 2H_t &&(K=x+y), \\
900: (t \D_t - x\D_x) H_t &= -2H_t &&(K=xy).
901: \end{alignat}
902: %% Though simple, these equations are striking from a dynamical
903: %% systems viewpoint. One combines these dynamics with an
904: %% arbitrary time-dependent size rescaling in order to study
905: %% scaling limits. (Size rescaling and time evolution commute.)
906: %% This means that the study of scaling limits on the attractor
907: %% is reduced to the study of a continuous dilation map
908: %% with arbitrary choice of amplitude normalization.
909: %[[Must discuss why this indicates chaos, where the shift map is.]]
910: %(Dilation of $x$ by the factor $b$ corresponds to
911: %a shift by $\log b$ after a logarithmic
912: %change of variables $x \mapsto \log x$.)
913: \subsection{How initial tails encode scaling limits}
914:
915: The long-time scaling behavior is very sensitive to the initial
916: distribution of the largest clusters in the system, as indicated by
917: the characterization of domains of attraction via
918: \qref{1.domain}, and the linearization theorem above.
919: In fact, the long-time scaling dynamics is encoded in the tails
920: of initial data in a simple fashion related to the \LK\
921: representation.
922: % In this section we describe two manifestations of this:
923: % \begin{itemize}
924: % \item[(i)] For any point $\hat F$ in the proper scaling attractor $\attr$,
925: % its associated divergent \smeas\ $H$ is a limit of
926: % scaled initial tails of the solutions that approximate $\hat F$.
927: % \end{itemize}
928:
929: %In what follows,
930: %we consider sequences $a_n \to \Tmax, b_n \to \infty$ and rescaled
931: %solutions $\tilde{F}\nsup_t$ and
932: %$\tilde{G}\nsup_t$ defined on $[\tilde{t}\nsup_0, \Tmax)$ as in
933: %\qref{newdmg}, \qref{newt0} and \qref{newmg} with
934: %$a,b$ replaced by $a_n,b_n$. Observe that $\tilde{t}\nsup_0 \to
935: %\Tmin$. For brevity, we define
936: %\[ \tilde\sm\nsup(dx) = \tilde{\sm}_{\tilde{t}\nsup_0}(dx) =??.\]
937:
938: \begin{thm}
939: \label{RT.initial_tails}
940: Let $\hat F\in\attr$
941: %so $\hat F(dx)=x^\gamma\nu_{t_0}(dx)$ where $\nu$ is an eternal solution,
942: with associated divergent \smeas\ $H$.
943: Let $\nu\nsup$ be any sequence of solutions defined for $t\ge t_0$,
944: with associated initial \smeass\ given by
945: $\sm\nsup(dx)=x^{\gamma+1}\nu\nsup_{t_0}(dx)$.
946: Let $T_n \to \Tmax$, $\beta_n\to\infty$.
947: Then the following are equivalent:
948: \bit
949: \item[(i)] ${F}\nsup_{T_n}(\beta_n x) \to \hat F(x)$ % = x^\gamma\nu_{t_0}(dx)$
950: as $n\to\infty$, at every point of continuity.
951: %(Equivalently,
952: %$a_n\nu\nsup_{a_n}(b_n\,\cdot)$ converges weakly to $\hat F$.)
953: %% \item[(ii)]
954: %% As $n\to\infty$, $a_n \vp\nsup_1(q/b_n) \to \Vp(q)$ for all $q\ge 0$.
955: \item[(ii)] The rescaled initial \smeass\ $\tilde\sm\nsup$ defined by
956: \begin{equation}\la{R.initG}
957: \tilde\sm\nsup(x) = \left\{
958: \begin{array}{ll}
959: \beta_n^{-1}T_n \,\sm\nsup(\beta_nx)
960: %= T_n\beta_n^{-1}\int_{[0,\beta_nx]} y\nu\nsup_{t_0}(dy)
961: &(K=2),\\[6pt]
962: \beta_n\inv \rme^{2T_n}\, \sm\nsup(\rme^{-T_n}\beta_nx)
963: %= \beta_n\inv\rme^{2T_n} \int_0^{x\beta_n\exp(-T_n)}y^2\nu\nsup_0(dy)
964: &(K=x+y), \\[6pt]
965: \beta_n\inv |T_n|^{-2} \,\sm\nsup(|T_n|\beta_n x)
966: &(K=xy),
967: \end{array} \right.
968: \end{equation}
969: have the property that
970: $\tilde\sm\nsup$ converges to $\dsm$ as $n \to \infty$.
971: \eit
972: \end{thm}
973: \noindent
974: This result generalizes to the full attractor $\attrx$, with $H$
975: replaced by the corresponding \barsmeas, see section~\ref{S.extend}.
976: We remark that in the proof it is shown that for the convergence in
977: part (ii) to hold, it is necessary that $\rme^{-T_n}\beta_n\to\infty$
978: for $K=x+y$, and $|T_n|\beta_n\to\infty$ for $K=xy$.
979: %[[To see this, follow the proof of Theorem 7.1 on p 1224 of \cite{MP1}
980: %up to (7.8), considering only sequential limits. It's hard to see
981: %how one could give a direct argument without a solution formula.]]
982:
983: %% \begin{rem}
984: %% %We note that $\limsup a_n/b_n\le1$ [[why??]].
985: %% If the sequence $\nu\nsup$ is constant, then $\hat F$ belongs to the
986: %% scaling $\omega$-limit set of this solution.
987: %% [remove?: If also $a_n=b_n$ then
988: %% $\tilde\sm\nsup$ is a pure dilation of the initial mass
989: %% measure. So if the initial mass is finite then (ii) holds
990: %% with $\dsm$ an atom at the origin. Correspondingly $\nu$ is
991: %% a self-similar solution with exponentially decaying tail.
992: %% The solution $\nu\nsup$ belongs to its domain of attraction;
993: %% see \cite[Theorem 5.1]{MP1}].
994: %% \end{rem}
995:
996: \subsection{Signatures of chaos} \la{RS.chaos}
997:
998: The dilational representation of dynamics in \qref{Ht}
999: in terms of the \LK\ representation
1000: means that Smoluchowski dynamics on the scaling attractor is
1001: a continuous analog of a Bernoulli shift map, a classical paradigm for
1002: chaotic dynamics. We demonstrate the utility of this representation by
1003: constructing solutions with both chaotic and regular orbits, and
1004: by proving a shadowing theorem illustrating sensitive dependence on the tails.
1005:
1006: %\medskip
1007: %\noindent
1008: %{\em Solutions with dense limit sets.}
1009: \subsubsection{Solutions with dense limit sets}
1010: \begin{thm}\label{RT.dense}
1011: There exists an eternal solution $\nu$ whose scaling $\omega$-limit
1012: set contains every element of the full scaling attractor $\attrx$.
1013: \end{thm}
1014: \noindent
1015: We call such solutions {\em Doeblin solutions\/} by analogy with
1016: Doeblin's universal laws. The construction follows Feller closely
1017: and relies only on general principles (separability of $\PP$ and $\Mcan$, and
1018: continuity of the bijection $\Hmap$). Theorem~\ref{RT.dense}
1019: tells us that $\attrx$ cannot be decomposed into
1020: invariant subsets.
1021: \medskip
1022: %[[What about density of Doeblin solutions?]] -- done in shadowing section
1023:
1024:
1025: \subsubsection{Scaling periodic solutions}
1026: %\label{RS.periodic}
1027:
1028: %\noindent
1029: %{\em Scaling-periodic solutions.}
1030: Another classical signature of chaos is the density of periodic
1031: solutions. The notion of periodicity generalizes as follows.
1032: \begin{defn}
1033: \label{RD.periodic}
1034: Let $\nu$ be a solution and define $F_t(x)$ by
1035: \qref{R.Fp}. We say $\nu$ is {\em scaling-periodic}
1036: if for some $t_1>t_0$ and $\betarsc >1$,
1037: \begin{equation}\label{R.periodic}
1038: F_{t_1}(\betarsc x)=F_{t_0}(x) \quad\mbox{for all $x>0$.}
1039: \end{equation}
1040: \end{defn}
1041: \noindent
1042: These are analogous to semi-stable laws in probability
1043: theory~\cite{semistable}. The
1044: \LK\/ representation yields a simple characterization.
1045: \begin{thm}
1046: \label{RT.periodic}
1047: A scaling-periodic solution of Smoluchowski's equation with kernel
1048: $K=2, x+y$ or $xy$ is eternal, and its divergent \smeas\ $\dsm$ satisfies
1049: \begin{equation}
1050: \dsm(x)=a\dsm(bx)
1051: \end{equation} for some $a>0$, $b>1$ such that either
1052: \begin{enumerate}
1053: \item[(i)] $a=1$ and $\dsm$ is an atom at the origin, or
1054: \item[(ii)] $a<1$ and $ab>1$.
1055: %and $\dsm$ is determined by its restriction to $[1,b)$.
1056: \end{enumerate}
1057: Conversely, if
1058: $\dsm$ is a measure on $[0,\infty)$ with $\dsm(x)=a\dsm(bx)$,
1059: where $a>0$, $b>1$ and (i) or (ii) hold,
1060: then $\dsm$ is a divergent \smeas\ and
1061: the corresponding eternal solution is scaling-periodic.
1062: \end{thm}
1063: Case (i) is simple but important. The corresponding scaling-periodic
1064: solutions are the self-similar solutions with exponential decay. More
1065: generally, all self-similar solutions are determined by divergent
1066: \smeass\ of the power-law form $\dsm(x)= C_\rho
1067: x^{1-\rho}$, $0 < \rho \leq 1$. Scaling-periodic solutions that
1068: are not self-similar solutions are generated by (ii).
1069: Thus, there are uncountably many scaling-periodic solutions.
1070: Moreover, the \LK\/ representation allows us to prove
1071:
1072: %they are dense in the full attractor $\attrx$.
1073: \begin{thm}
1074: \label{RT.Pdense}
1075: Scaling-periodic solutions are dense in $\attrx$.
1076: \end{thm}
1077:
1078: % \noindent
1079: %
1080:
1081: \subsubsection{Shadowing and sensitive dependence on the initial tails}
1082:
1083: %\medskip
1084: %\noindent
1085: %{\em Shadowing and sensitive dependence on the initial tails.\/}
1086: We show that asymptotically similar initial tails imply {\em
1087: shadowing} of scaled solution trajectories.
1088: To study shadowing, we note that the space $\bar\PP$ of
1089: probability measures on $\Ebar$ is metrizable and compact.
1090: We let $\dist(\cdot,\cdot)$ denote any metric on $\bar\PP$ which
1091: induces the weak topology.
1092:
1093: \begin{thm} \label{RT.shad}
1094: Let $\nu$ and $\bar\nu$ denote any two solutions of Smoluchowski's
1095: equation defined on $[t_0,\Tmax)$, and let the associated initial
1096: \smeass\ be
1097: \begin{equation}\label{R.shad1}
1098: \sm(dx)=x^{\gamma+1}\nu_{t_0}(dx), \qquad
1099: \bar\sm(dx)=x^{\gamma+1}\bar\nu_{t_0}(dx),
1100: \end{equation}
1101: with Laplace exponents $\vp$ and $\bar\vp$ respectively associated
1102: as in \qref{R.LK2a}.
1103: Assume that
1104: \begin{equation}\label{R.shad2}
1105: %{\bar\sm(x)}/{\sm(x)} \sim L(x) \quad\mbox{as $x\to\infty$},
1106: \bar\vp(q)/\vp(q) \sim L(1/q) \quad\mbox{as $q\to0$,}
1107: \end{equation}
1108: where $L$ is slowly varying at $\infty$.
1109: Suppose that $b(t)\uparrow\infty$ as $t\uparrow\Tmax$, and define
1110: \begin{equation}\la{R.bartb}
1111: (\bar t,\bar b) =
1112: %\bar t=
1113: \left\{
1114: \begin{array}{ll}
1115: (\ t/L(b),\ b\ ) & (K=2),\\[5pt]
1116: (\ t - \log L(b\rme^{-t}),\ b/L(b\rme^{-t}) \ ) & (K=x+y), \\[5pt]
1117: (\ t\,L(|t|b) ,\ b L(|t|b)\ ) & (K=xy),
1118: \end{array}\right.
1119: \ee
1120: so that $\bar b/\lambda(\bar t)=b/\lambda(t)$ with
1121: $\lambda(t)$ as in \qref{R.cansoln1}.
1122: Then we have
1123: \begin{equation}\la{R.dto0}
1124: \dist( F_t(b(t)\,dx),\bar F_{\bar t}(\bar b(t)\,dx)) \to 0 \qquad
1125: \mbox{as $t\to\Tmax$.}
1126: \end{equation}
1127: \end{thm}
1128: The simplest situation, requiring no readjustment of the scaling
1129: ($\bar t=t$, $\bar b=b$) is when $L=1$.
1130: When condition \qref{R.shad2} holds, the solutions $\nu$ and $\bar\nu$
1131: have identical scaling $\omega$-limit sets, for example.
1132: If one of the solutions in this theorem, say $\bar \nu$, is
1133: self-similar, then the sufficient condition \qref{R.shad2} for
1134: shadowing in this theorem is equivalent to \qref{1.domain}
1135: (see \cite{MP1}, (5.3) and (5.7) for $K=2$, and (7.2)
1136: and (7.4) for $K=x+y$). Hence \qref{R.shad2} is also necessary,
1137: according to the classification theorem on domains of attraction. It
1138: appears that in general the sufficient condition for shadowing given
1139: in this theorem may not always be necessary. But we will not pursue
1140: this issue here.
1141:
1142: The sensitivity of solutions to initial tails in the weak topology
1143: is revealed strikingly in Theorem~\ref{RT.shad}.
1144: The topology of weak convergence is undoubtedly natural
1145: for limit theorems, for example the approach to self-similarity.
1146: On the other hand, this topology cannot distinguish the tails,
1147: as the following ``cut-and-paste'' argument shows.
1148: Let $\hat F=x^\gamma\nu_{t_0}$ and $\Fcheck=x^\gamma\nucheck_{t_0}$
1149: be given initial data for two solutions, and define
1150: \begin{equation}
1151: \hat F\nsup(x) = \left\{
1152: \begin{array}{ll}
1153: \hat F(x) \wedge \Fcheck(n) & x< n\\
1154: \check F(x) & x\ge n.
1155: \end{array}\right.
1156: \end{equation}
1157: Then $\hat F\nsup\to \hat F$ as $n\to\infty$,
1158: and $\hat F\nsup$ has Laplace exponent given by
1159: \be
1160: \frac{\vp\nsup(q)}q = \int_0^n
1161: \left( \frac{1-\rme^{-qy}}{qy}\right) y\hat F(dy)
1162: +\int_n^\infty
1163: \left( \frac{1-\rme^{-qy}}{qy}\right) y\check F(dy) ,
1164: \ee
1165: from which one sees easily that if $\check\sm(E)=\infty$, then
1166: $\vp\nsup(q)\sim\check\vp(q)$ as $q\to0$. Thus,
1167: according to Theorem \ref{RT.shad},
1168: the solution $\nu\nsup$ generated by $\hat F\nsup$ will shadow
1169: $\nucheck$. This justifies the statement made
1170: in the introduction that for any given scaling trajectory, there is a
1171: dense set of initial data whose forward trajectories shadow the given one.
1172:
1173: \subsection{Plan of the paper}
1174: In section 3 we establish some basic facts
1175: regarding convergence of the measures that generate the \LK\
1176: representation. The analysis of eternal solutions is different in
1177: detail for the constant and additive kernels, so we treat these cases
1178: in turn, establishing Theorems~\ref{RT.AE}, \ref{RT.LK}, and
1179: \ref{RT.initial_tails} for these kernels in sections 4 and 5.
1180: The multiplicative case reduces mathematically to the
1181: additive by a change of variables, and is treated briefly in section~6.
1182: We emphasize that the results in this case concern the behavior of
1183: solutions approaching the gelation time, so perhaps this is the case
1184: of most interest physically.
1185:
1186: With the \LK\ representation in hand, we then construct Doeblin
1187: solutions in section 7 and scaling-periodic solutions in section 8.
1188: Extended solutions and the full scaling attractor $\attrx$ are studied
1189: in section 9, and the shadowing theorem~\ref{RT.shad} is proved in
1190: section 10, where we also provide a streamlined treatment of
1191: the domains of attraction.
1192:
1193: \section{Laplace exponents and limits of \barsmeass}
1194: \label{S.laplace}
1195: The main analytic tool in the study of the solvable kernels is the
1196: Laplace transform. Recall that a sequence of probability measures
1197: $F\nsup$ is said to converge weakly to a
1198: probability measure $F$ if the distribution functions
1199: $F\nsup(x) \to F(x)$ at every point of continuity of the limit.
1200: It is basic~\cite[XIII.1]{Feller} that $F\nsup$ converges
1201: weakly to $F$ if and only if the Laplace transforms
1202: converge pointwise:
1203: \[\int_E \rme^{-qx} F\nsup(dx) \to \int_E \rme^{-qx} F(dx), \quad
1204: \text{for all $q > 0$.}\]
1205: We will need the following refinements of this result for \barsmeass.
1206: %The results are [[close to ones that are]]
1207: %well-known, but proofs are included for completeness.
1208: With any \barsmeas\ $\sm$ we associate ``Laplace
1209: exponents'' $\Vp$ and $\Psi$ (with $\Vp=\D_q\Psi)$ defined for
1210: $q \in \C_+=\{\lambda\in\C:\Re\lambda>0\}$ by
1211: \begin{equation}
1212: \la{L.1}
1213: \Vp(q) = \int_{\Ebar} \frac{1-\rme^{-qx}}{x} \sm(dx),
1214: \quad
1215: \Psi(q) = \int_{\Ebar} \frac{\rme^{-qx}-1 +qx}{x^2} \sm(dx) .
1216: \end{equation}
1217: If $g_0$ and $g_\infty$ denote amplitudes of the atoms of
1218: the measure $(1\wedge y\inv)\sm(dy)$ at $0$ and $\infty$, respectively,
1219: this means
1220: \begin{eqnarray}\label{L.1a}
1221: \Vp(q) &=& q g_0 + g_\infty + \int_{(0,\infty)} \frac{1-\rme^{-qx}}{x} \sm(dx),
1222: \\ \label{L.1b}
1223: \Psi(q) &=& \frac12 q^2 g_0 + q g_\infty +
1224: \int_{(0,\infty)} \frac{\rme^{-qx}-1 +qx}{x^2} \sm(dx).
1225: \end{eqnarray}
1226: We use the terminology Laplace exponent in accordance with
1227: probabilists' usage. If we need to distinguish the two types
1228: of exponents, we will refer to $\Vp$
1229: and $\Psi$ as Laplace exponents of
1230: the first and second order, respectively. Observe that
1231: \be
1232: \la{L.3}
1233: \partial_q \Vp = \partial_q^2 \Psi = \int_{[0,\infty)}\rme^{-qx}\sm(dx).
1234: \ee
1235: These functions are Laplace transforms of a positive measure,
1236: thus are completely monotone functions on $(0,\infty)$.
1237:
1238: We note that the amplitude of the atom of $(1\wedge y\inv)\sm(dy)$
1239: at $\infty$ is
1240: \begin{equation} \label{R.ginf}
1241: g_\infty = \lim_{q\to0^+}\Vp(q) = \lim_{q\to0^+} q\inv\Psi(q),
1242: \end{equation}
1243: thus the \barsmeas\ $\sm$ is an \smeas\ if and only if this vanishes.
1244: Furthermore, we claim that $\sm$ is divergent if and only if
1245: \begin{equation} \label{R.divcon}
1246: \lim_{q\to\infty} \Vp(q) = \infty, \qquad\mbox{equivalently}\quad
1247: \lim_{q\to\infty} q\inv\Psi(q)=\infty.
1248: \end{equation}
1249: To prove this,
1250: observe that
1251: \[ \Phi(q) \leq q g_0+\int_{(0,\infty)} x^{-1} \sm(dx).\]
1252: Thus, if $\lim_{q \to \infty} \Phi(q)=\infty$, then $\sm$ satisfies
1253: \qref{R.candef5}. Conversely, if $\sm$ satisfies \qref{R.candef5}, then
1254: $\lim_{q \to \infty} \Vp(q)=\infty$ by the monotone
1255: convergence theorem. The proof for $\Psi$ is similar. We integrate by
1256: parts and use Fubini's theorem to obtain
1257: \begin{eqnarray*}
1258: q^{-1}\Psi(q) &=&
1259: \frac12 q g_0 +g_\infty+ \int_{(0,\infty)}
1260: (1-\rme^{-qx}) \int_{(x,\infty)}y^{-2} \sm(dy)\, dx \\
1261: &\leq &
1262: \frac12 q g_0 +g_\infty+ \int_{(0,\infty)} y^{-1} \sm(dy).
1263: \end{eqnarray*}
1264: Thus, if $q^{-1} \Psi(q) \to \infty$ then $\sm$ satisfies
1265: \qref{R.candef5}. The converse follows from the monotone convergence
1266: theorem.
1267:
1268: \begin{thm}
1269: \label{LT.Mconv}
1270: Let $\sm\nsup$ be a sequence of \barsmeass\ with Laplace
1271: exponents $\Vp\nsup$ and $\Psi\nsup$. Then, taking $n \to \infty$, the
1272: following are equivalent:
1273: \begin{enumerate}
1274: \item[(i)] $\sm\nsup$ converges to an \barsmeas\ $\sm$
1275: with Laplace exponents $\Vp$ and $\Psi$.
1276: \item[(ii)] $\Vp(q):= \lim_{n\to\infty}\Vp\nsup(q)$ exists for each $q>0$.
1277: \item[(iii)] $\Psi(q):= \lim_{n\to\infty}\Psi\nsup(q)$ exists for each $q>0$.
1278: \end{enumerate}
1279: \end{thm}
1280:
1281: \begin{proof} %[Proof of Theorem~\ref{LT.Mconv}]
1282: (i) implies (ii): Fix $q>0$ and let $\veps >0$.
1283: \qref{R.candef2} allows us to choose
1284: $a$ such that $a$ is a point of continuity for $\sm$ and for every $n$
1285: \[
1286: \int_{[a,\infty]} \rme^{-qx} x\inv (\sm\nsup(dx)+\sm(dx))
1287: \le \rme^{-qa} C <\veps.
1288: \]%\int_a^\infty (1-\rme^{-qx}) x^{-1}\sm\nsup(dx) < \veps.\]
1289: On the other hand, \qref{R.candef1} guarantees
1290: \[\int_0^a (1-\rme^{-qx})x^{-1} \sm\nsup(dx) \to \int_0^a
1291: (1-\rme^{-qx})x^{-1} \sm(dx).\]
1292: Using again \qref{R.candef2},
1293: we conclude that for large $n$, $|\Vp\nsup(q)-\Vp(q)|<\veps$.
1294: %Therefore, $\limsup\Vp\nsup(q) -2\veps \leq \Vp(q) \leq
1295: %\liminf\Vp\nsup(q) + 2\veps$.
1296:
1297: (ii) implies (i): {\em 1. Claim:\/} $\Vp$ is analytic in $\C_+$
1298: and $\Vp\nsup \to \Vp$ uniformly on compact subsets of $\C_+$.
1299: {\em Proof:\/} Let $K \subset \C_+$ be compact. The claim follows
1300: from the estimate:
1301: \be
1302: \la{L.can1}
1303: \sup_n \sup_{q \in K} |\Vp\nsup (q)| < \infty.
1304: \ee
1305: Indeed, by Montel's theorem, (\ref{L.can1}) implies
1306: $\{\Vp\nsup\}_{n=1}^\infty$ are a normal family of analytic
1307: functions (i.e., precompact in the uniform topology). Thus, every
1308: subsequence has a further subsequence converging uniformly
1309: to an analytic function. Since every subsequence converges
1310: (pointwise) to $\Vp$, this implies $\Vp\nsup \to \Vp$ uniformly and
1311: $\Vp$ is analytic. It remains to
1312: prove \qref{L.can1}. We integrate by parts to obtain
1313: \[ q^{-1} \Vp\nsup(q) = \sm\nsup(0) +
1314: \int_{E} \rme^{-qx} \left( \int_{[x,\infty]}
1315: y^{-1}\sm\nsup(dy) \right) dx.\]
1316: Thus, for any $a >0$,
1317: $\sup_{\Re q > a} |q^{-1} \Vp\nsup(q)| \leq a^{-1} \Vp\nsup(a)$.
1318: Since $\Vp\nsup(q)$ converges for all $q > 0$, we have $\sup_n
1319: \sup_{\Re q > a} |q^{-1} \Vp\nsup(q)| < \infty$. This proves
1320: \qref{L.can1}.
1321:
1322: {\em 2. \/} Cauchy's integral formula and the claim imply
1323: $\partial^k_q \Vp\nsup \to \partial^k_q \Vp$ for every $k \in \N$.
1324: Since $\partial_q \Vp\nsup$ are completely monotone, so is the limit
1325: $\partial_q\Vp$.
1326:
1327: {\em 3.\/} Thus,
1328: $\partial_q \Vp = \int_{[0,\infty)} \rme^{-qx} \sm(dx)$
1329: is the Laplace transform of a measure $\sm$ on $[0,\infty)$.
1330: We integrate with respect to $q$ with
1331: $g_\infty:=\Vp(0^+)$ defined to be the charge of $y\inv\sm(dy)$ at $\infty$,
1332: and use Tonelli's theorem to obtain (\ref{L.1}). Note that $\int_{[1,\infty]}
1333: x^{-1} \sm(dx)<\infty$ because $\Vp(q)<\infty$ for each fixed $q$,
1334: so $\sm$ is an \barsmeas.
1335:
1336: {\em 4.\/} The convergence $\D_q \Vp\nsup \to \D_q \Vp$ is
1337: equivalent to weak convergence of $\sm\nsup$ to $\sm$ on $[0,\infty)$,
1338: meaning \qref{R.candef1} holds.
1339: This implies that for every point $x$ of continuity of $\sm$,
1340: as $n\to\infty$ we have
1341: \begin{equation}
1342: \int_{[0,x]} (1-\rme^{-qy})y\inv \sm\nsup(dy)
1343: \to \int_{[0,x]} (1-\rme^{-qy})y\inv \sm(dy),
1344: \end{equation}
1345: and together with (ii) this yields that
1346: $\int_{[x,\infty]}y\inv\sm\nsup(dy)$ is bounded and
1347: \[
1348: \int_{[x,\infty]} \rme^{-qy} y\inv \sm\nsup(dy)
1349: \to \int_{[x,\infty]} \rme^{-qy}y\inv \sm(dy) .
1350: \]
1351: From (ii) then follows \qref{R.candef2}. This proves $\sm\nsup\to\sm$.
1352:
1353: (ii) implies (iii): This
1354: is due to $\Psi\nsup(q)=\int_0^q\Vp\nsup(s)\,ds$ and monotonicity.
1355: %is step {\em 2\/} [[??]] above.
1356:
1357: (iii) implies (ii): Since $\Psi\nsup(q)=\int_0^q\Vp\nsup(s)\,ds
1358: \ge \frac12q\Vp\nsup(\frac12q)$, we find that \qref{L.can1} holds as
1359: in step {\em 1\/} above. Then for every subsequence of $\Vp\nsup$ there is a
1360: further subsequence that converges on compact sets of $\C_+$ to an analytic
1361: limit $\Vp$. This limit is unique due to (iii), and (ii) follows. (It follows
1362: also that $\Psi\nsup \to \Psi$ uniformly on compact sets.)
1363: \end{proof}
1364:
1365: \section{The constant kernel}
1366: \label{sec:const-ker}
1367: In this section we study eternal solutions and the \LK\/ representation
1368: in particular
1369: %present a detailed analysis of the scaling dynamics
1370: for the constant kernel $K=2$.
1371: This kernel is technically easiest to deal
1372: with, and the general framework is most transparent.
1373: Theorems~\ref{CT.attr}, \ref{CT.eternal} and
1374: \ref{CT.Mconv2} are the main technical results and serve to
1375: establish Theorems~\ref{RT.AE} and \ref{RT.LK} for this kernel.
1376: %With these results in hand,
1377: %we exploit the scaling properties of the \LK\/ representation to gain
1378: %a detailed understanding of the ultimate scaling dynamics in the
1379: %following [[??]] section.
1380:
1381: \subsection{Preliminaries}
1382: Smoluchowski's equation with constant
1383: kernel $K=2$ has a unique global solution in an appropriate weak sense
1384: given any initial size-distribution measure with finite zero-th
1385: moment~\cite[\S 2]{MP1}.
1386: For convenience, we adopt the normalization in \qref{R.normal2}.
1387: The moment identity \qref{R.moment} is valid for all
1388: bounded continuous functions $f$ on $\bar{E}$, and taking
1389: $f=1$ we find that the total number density of clusters is
1390: $\nu_t(E)=t\inv$.
1391: Since $t \nu_t(E)=1$,
1392: we associate to each solution a probability distribution function
1393: \begin{equation}\label{C.Fdef}
1394: F_t(x) = \int_{(0,x)} \nu_t(dx)\left\slash \int_E \nu_t(dx)\right.
1395: = t \nu_t(x).
1396: \end{equation}
1397: We also introduce the \smeass
1398: \be
1399: \label{C.cansoln1}
1400: \sm_t(dx) = x \nu_t(dx),
1401: \ee
1402: and associated Laplace exponents
1403: %[[How about $\Vp_t(q)$ or $\vp_t(q)$ instead of $\vp(t,q)$ ?]]
1404: \begin{equation}\label{C.vpdef}
1405: \vp(t,q) = \int_E (1-\rme^{-qx}) \nu_t(dx) =
1406: \int_{\Ebar} \frac{1-\rme^{-qx}}{x} \sm_t(dx).
1407: \end{equation}
1408: %and put $\vp_1(q)=\vp(1,q)$.
1409: %Note that the Laplace transform of $F$ is
1410: %\begin{equation}\label{C.Fvp}
1411: %\int_E \rme^{-qx}F_t(dx) = 1-t\vp(t,q).
1412: %\end{equation}
1413:
1414: Notice that $q\mapsto\vp(t,q)$ is strictly increasing with
1415: $\vp(t,\infty)= \nu_t(E)=t^{-1}$, and
1416: $\partial_q \vp(t,q)$ is the Laplace transform of the
1417: mass-distribution measure $x\,\nu_t(dx)$, so is completely monotone.
1418: $\vp$ solves the simple equation
1419: \begin{equation}
1420: \label{eq:vp0}
1421: \partial_t \vp = -\vp^2,
1422: \end{equation}
1423: for which the solution at any time $t>0$ is determined from data
1424: at time $t_0>0$ according to
1425: \begin{equation}
1426: \label{eq:vp1}
1427: \quad \vp(t,q) =
1428: \frac{\vp(t_0,q)}{1+(t-t_0)\vp(t_0,q)}, \quad q \geq 0,\ t >0.
1429: \end{equation}
1430: Since $0 \leq \vp(t_0,q) <t_0\inv$, we see that given $F_{t_0}=t_0\nu_{t_0}$
1431: an arbitrary probability measure,
1432: $\vp(t,q)$ is well-defined on the time-interval $(0,\infty)$.
1433: But for $0<t<t_0$, $\vp(t,q)$ may not have completely
1434: monotone derivative, and thus may not define a (positive)
1435: measure. The map $q\mapsto\D_q\vp(t,q)$ is completely monotone
1436: for all $t \in (0,\infty)$ if and only if $\nudot$ is an eternal solution.
1437:
1438: %which has the unique solution
1439: %\begin{equation}
1440: %\label{eq:vp1}
1441: %\quad \vp(t,q) =
1442: %\frac{\vp_1(q)}{1+(t-1)\vp_1(q)}, \quad q \geq 0,\ t >0.
1443: %\end{equation}
1444: %Since $0 \leq \vp_1(q) <1$, we see that for {\em any} probability
1445: %measure $\nu_1$, $\vp(t,q)$ is well-defined on the time-interval
1446: %$(0,\infty)$. But for $0<t<1$, $\vp(t,q)$ may not have completely
1447: %monotone derivative, and thus may not define a (positive)
1448: %measure. The map $q\mapsto\vp(t,q)$ is completely monotone
1449: %for all $t \in (0,\infty)$ if and only if $\nudot$ is an eternal solution.
1450:
1451: Our study of convergence properties for solution sequences is based on
1452: pointwise convergence properties of $\vp$, which are equivalent to
1453: %proper
1454: convergence properties of the \smeass\ $\sm_t$
1455: according to the results of section~\ref{S.laplace}.
1456: We begin by proving the continuous dependence of solutions on initial
1457: data, based on the evident fact that $\vp(t,q)$ is a continuous function of
1458: $\vp(t_0,q)$.
1459:
1460: \begin{thm}\label{CT.cts}
1461: (Continuous dependence on data.)
1462: For Smoluchowski's equation with constant kernel $K=2$, let $t_0>0$
1463: and let $\nu\nsup$ be a sequence of solutions defined for $t\ge t_0$.
1464: \begin{enumerate}
1465: \item[(a)] If $\nu\nsup_{t_0}$ converges weakly to a measure $\hat\nu_0$
1466: with $\hat\nu_0(E)=t_0\inv$, then for every $t\ge t_0$ we have
1467: that $\nu\nsup_t$ converges weakly to $\nu_t$, the time-$t$ solution
1468: with initial data $\nu_{t_0}=\hat\nu_0$.
1469: \item[(b)] For any $t\ge t_0$, if $\nu\nsup_{t}$ converges weakly to a
1470: measure $\hat\nu$ with $\hat\nu(E) =t\inv$, then
1471: $\nu\nsup_{t_0}$ converges weakly to a measure $\hat\nu_0$ with
1472: $\hat\nu_0(E)=t_0\inv$, and $\hat\nu=\nu_t$, the time-$t$ solution
1473: with initial data $\nu_{t_0} = \hat\nu_0$.
1474: \end{enumerate}
1475: \end{thm}
1476:
1477: \begin{proof} We prove part (b); part (a) is similar.
1478: Let $\sm\nsup_t(dx)=x\nu\nsup_t(dx)$ and $\hat\sm(dx)=x\hat\nu(dx)$,
1479: and let $\vp\nsup(t,q)$
1480: and $\hat\vp(q)$ be the associated Laplace exponents as in \qref{C.vpdef}.
1481: The hypothesis is equivalent to saying
1482: that the \smeass\ $\sm\nsup_t(dx)$ converge
1483: %properly
1484: to a non-divergent \smeas\ $\hat\sm(dx)=x\hat\nu(dx)$ with $\int_E
1485: x\inv\hat\sm(dx)=t\inv$.
1486: This is equivalent to the statement that for all $q>0$,
1487: $\vp\nsup(t,q) \to \hat\vp(q)$ as $n\to\infty$, where
1488: $\hat\vp(\infty)=t\inv$ and $\hat\vp(0^+)=0$.
1489: Then it follows that
1490: \begin{equation}\label{C.tt0}
1491: \vp\nsup(t_0,q)
1492: = \frac{\vp\nsup(t,q)}{1-(t-t_0)\vp\nsup(t,q)}
1493: \to
1494: \hat\vp_0(q):= \frac{\hat\vp(q)}{1-(t-t_0)\hat\vp(q)}.
1495: \end{equation}
1496: Since $\vp_0(0^+)=0$ and $\hat\vp_0(\infty)=t_0\inv$,
1497: we conclude that $\hat\vp_0$ is the Laplace exponent for a
1498: measure $\hat\nu_0$ on $E$ with $\hat\nu_0(E)=t_0\inv$,
1499: and that $\nu\nsup_{t_0}$ converges weakly to $\hat\nu_0$.
1500: We compare \qref{C.tt0} with the explicit solution (\ref{eq:vp1}) to
1501: see that $\hat\nu=\nu_t$,
1502: where $\nudot$ is the solution on $[t_0,\infty)$ with initial data $\hat\nu_0$.
1503: \end{proof}
1504:
1505: \subsection{The scaling attractor and eternal solutions}
1506: We are ready to prove Theorem~\ref{RT.AE} for the kernel $K=2$.
1507: First we consider part (b), the correspondence between the scaling attractor
1508: and eternal solutions.
1509:
1510: \begin{thm}\label{CT.attr}
1511: A probability measure
1512: $\hat F$ is an element of the scaling attractor for
1513: Smoluchowski's equation with constant kernel $K=2$ if and only if
1514: $\hat F = \nu_1$ for some eternal solution $\nudot$.
1515: \end{thm}
1516: \begin{proof}
1517: Let us first suppose that $\hat F = \nu_1$ for some eternal solution
1518: $\nudot$ and show that $\hat F \in \attr$.
1519: Pick arbitrary sequences $T_n, \beta_n \to \infty$, and consider
1520: the sequence of rescaled eternal solutions
1521: \be
1522: \label{C.attr-rsc}
1523: \nu_t\nsup(dx) = \frac{1}{T_n} \nu_{t/T_n} \left(\beta_n^{-1}
1524: dx\right), \quad t>0.
1525: \ee
1526: Observe that $\nu_t\nsup(E) = t^{-1}$, therefore,
1527: \[ F\nsup_{T_n}(\beta_n x) = \nu_{1}(x) = \hat F(x) \]
1528: for every $x$. Thus, $\hat F\in\attr$ by Definition~\ref{R.attractor}.
1529:
1530: Conversely, suppose $\hat F \in \attr$. We shall show that $\hat F =
1531: \nu_1$ for
1532: some eternal solution $\nudot$. Let $\hat \vp$ denote the Laplace
1533: exponent of $\hat F$, and $\nudot\nsup,T_n, \beta_n$ be as in
1534: Definition~\ref{R.attractor}. Consider the rescaled measures
1535: \[
1536: \tilde\nu\nsup_t(dx)=T_n\nu\nsup_{tT_n}(\beta_n dx).
1537: \]
1538: This rescaling yields a solution that is defined for $t\ge 1/T_n$,
1539: and by hypothesis we have that $\tilde\nu\nsup_1$ converges
1540: weakly to $\hat F$. Then by Theorem~\ref{CT.cts}, for any $t>0$
1541: we infer that $\tilde\nu\nsup_t$ converges weakly to $\nu_t$
1542: where $\nu_t(E)=t\inv$ and $\nu$ is a solution with $\nu_1=\hat F$.
1543: The solution $\nu$ is eternal since it is defined
1544: for $t\ge t_0$ for every $t_0>0$.
1545: \end{proof}
1546:
1547: Let us now prove that $\attr$ is invariant (part (a) of Theorem~\ref{RT.AE}).
1548: Suppose $\nu$ is a solution on some time interval $[t_1,\infty)$,
1549: normalized so $\nu_t(E)=t^{-1}$.
1550: Suppose $F_{T}\in\attr$ for some $T\ge t_1$.
1551: Replacing $\nu_t(dx)$ by $T\nu_{Tt}(dx)$, we may presume $T=1$ without loss
1552: of generality. By Theorem~\ref{CT.attr} above, $F_T=\tilde\nu_1$ for
1553: some eternal solution $\tilde\nu$. But then $\nu_t=\tilde\nu_t$
1554: for all $t\ge t_1$, meaning that $\nu$ is (the restriction of) an
1555: eternal solution. We obtain that $F_t\in\attr$ for all $t>0$ by
1556: a similar scaling argument.
1557:
1558:
1559: \subsection{\LK\/ representation of eternal solutions}
1560:
1561: \begin{thm}
1562: \label{CT.eternal}
1563: \begin{enumerate}
1564: \item[(a)] Let $\nu$ be an eternal solution to Smoluchowski's
1565: equation with $K=2$. Then there is divergent \smeas\ $\dsm$
1566: such that as $t\dnto 0$, the mass measure
1567: $\sm_t(dx)=x\,\nu_t(dx)$ converges to $\dsm$.
1568: %properly
1569:
1570: \item[(b)] Conversely, given any divergent \smeas\ $\dsm$ there
1571: is a unique eternal solution with the properties in part (a),
1572: defined for all $t\in(0,\infty)$ via
1573: \begin{equation}
1574: \label{C.eternal}
1575: \vp(t,q) = \frac{\Vp(q)}{1+t\Vp(q)},
1576: \qquad
1577: \Vp(q) = \int_\Ebar \frac{1-\rme^{-qx}}{x}H(dx).
1578: \end{equation}
1579: \end{enumerate}
1580: \end{thm}
1581:
1582: \begin{proof}
1583: We first show (a). Immediately from the solution formula
1584: (\ref{eq:vp1}),
1585: \be
1586: \label{eq:lim_vp1}
1587: \lim_{t\to0}\vp(t,q) =
1588: \lim_{t \to 0} \frac{\vp(1,q)}{ 1+(t-1) \vp(1,q)} =
1589: \frac{\vp(1,q)}{1-\vp(1,q)} =: \Vp(q)
1590: \ee
1591: exists for all $q>0$, with $\Vp(q) < \infty$,
1592: $\Vp(0^+)=0$, and $\Vp(\infty)=\infty$.
1593: By Theorem~\ref{LT.Mconv}, $\sm_t$ converges to an \barsmeas\/ $\dsm$
1594: with Laplace exponent $\Vp$, and $\dsm$ is a divergent \smeas\/ by
1595: the criteria in \qref{R.ginf}-\qref{R.divcon}.
1596:
1597: Let us now prove (b). Let $\dsm$ be a divergent \smeas\ with
1598: Laplace exponent $\Vp$. By \qref{eq:lim_vp1}, any eternal solution with
1599: the properties in part (a) must be determined by \qref{C.eternal}.
1600: Observe that the function $q/(1+tq)$ has completely monotone derivative
1601: for $t \in (0,\infty)$. It follows that $\partial_q \vp(t,q)$ is
1602: completely monotone when $\vp(t,q)$ is given by
1603: (\ref{C.eternal})~\cite[XIII.4]{Feller}.
1604: Moreover, with $\nu_t$ determined from \qref{C.vpdef},
1605: $\nu_t(E)= \vp(t,\infty)= t^{-1}$. Thus, $\nu_t$ is indeed an eternal solution.
1606: \end{proof}
1607:
1608: \begin{rem}
1609: Observe that $\sm_t(E)= \int_E x \nu_t(dx)$ is finite
1610: for some $t\in(0,\infty)$ if and only if it is finite for all $t$.
1611: However, it is not necessary that
1612: the mass be finite for a solution to be well-defined.
1613: \end{rem}
1614:
1615: Theorem~\ref{CT.eternal} establishes parts (a) and (b) of
1616: Theorem~\ref{RT.LK}. To establish part (c), we need to show that the map
1617: $\nu_1\mapsto\dsm $ from $\attr$ to $\Mcan$ is a bi-continuous bijection.
1618: \begin{thm}\label{CT.Mconv2}
1619: Let $\nu\nsup$ be a sequence of eternal solutions with corresponding
1620: divergent \smeass\ $\dsm\nsup$.
1621: Fix $t>0$. Then, taking $n\to\infty$, the following are equivalent:
1622: \bit
1623: \item[(i)]
1624: $\nu\nsup_t$ converges weakly to some
1625: measure $\hat\nu$ with $\hat\nu(E)=t\inv$.
1626: \item[(ii)]
1627: $\dsm\nsup$ converges to some divergent \smeas\ $\dsm$.
1628: %properly
1629: \eit
1630: If either (equivalently both) of these conditions hold, then
1631: $\hat\nu=\nu_t$ for an eternal solution with divergent \smeas\ $\dsm$.
1632: \end{thm}
1633: \begin{proof}
1634: Assume (i), so $\nu\nsup_t$ converges to $\hat\nu$ with
1635: $\hat\nu(E)=t^{-1}$. Then $\sm\nsup_t(dx)=x\nu\nsup(dx)$
1636: converges to $\hat\sm(dx)=x\hat\nu(dx)$ and
1637: the associated Laplace exponents converge: $\vp\nsup(t,q)\to\hat\vp(q)$
1638: for all $q>0$. Hence
1639: \begin{equation} \label{C.Pconv}
1640: \Vp\nsup(q) = \frac{\vp\nsup(t,q)}{1-t\vp\nsup(t,q)} \to
1641: \Vp(q):= \frac{\hat\vp(q)}{1-t\hat\vp(q)},
1642: \end{equation}
1643: as $n\to\infty$ for every $q>0$.
1644: Since $\hat\vp(0^+)=0$ and $t\hat\vp(q)\to1$ as $q\to\infty$,
1645: $\Vp(q) < \infty$ for every $q >0$, $\Vp(0^+)=0$,
1646: and $\lim_{q \to \infty}\Vp(q)=\infty$. By Theorem~\ref{LT.Mconv}
1647: and \qref{R.ginf}-\qref{R.divcon}, this proves (ii).
1648:
1649: We now show (ii) implies (i). Suppose the divergent \smeass\ $\dsm\nsup$
1650: converge to a divergent \smeas\ $\dsm$. Then
1651: Theorem~\ref{LT.Mconv} with \qref{R.ginf}-\qref{R.divcon}
1652: implies $\Vp\nsup(q)\to\Vp(q)$ for every $q>0$,
1653: $\Vp(0^+)=0$, and $\Vp(q)\to\infty$ as $q\to\infty$. Then,
1654: \[
1655: \vp\nsup(t,q)=\frac{\Vp\nsup(q)}{1+t\Vp\nsup(q)}
1656: \to \frac{\Vp(q)}{1+t\Vp(q)} =\vp(t,q)
1657: \]
1658: for every $q>0$. This yields weak convergence of $\nu\nsup_t$ to
1659: $\nu_t$, where $\nu$ is the eternal solution with Laplace exponent
1660: $\Vp$ and divergent \smeas\ $\dsm$.
1661: \end{proof}
1662:
1663: \subsection{Scaling limits and initial tails}
1664: We now prove Theorem~\ref{RT.initial_tails} for the constant
1665: kernel.
1666: \begin{proof}[Proof of Theorem~\ref{RT.initial_tails}]
1667: Introduce rescaled solutions
1668: $\tilde\nu\nsup_t(dx)=T_n\nu\nsup_{tT_n}(\beta_ndx)$,
1669: and let $\tilde F\nsup_t = t\tilde\nu\nsup_t$.
1670: Also let $\tilde\sm\nsup_t(dx)= x\tilde\nu\nsup_t(dx)$ and let
1671: $\tilde\vp\nsup(t,q)$ be the associated Laplace exponent.
1672: Let $H$ be the divergent \smeas\ corresponding to $\nu$ and
1673: $\Vp$ its Laplace exponent, and let $\vp(q)$ be the Laplace exponent
1674: of $\sm(dx)= x\nu_1(dx)$.
1675:
1676: Then statement (i) of Theorem~\ref{RT.initial_tails} is equivalent to saying $\tilde
1677: F\nsup_1\to \hat F$ weakly,
1678: meaning the \smeass\/ $\tilde\sm\nsup_1$ converge to $\sm$
1679: with $\int_E x\inv\sm(dx)=1$. This is equivalent to saying
1680: \begin{equation}\label{e.ph1}
1681: \tilde\vp\nsup(1,q)\to \vp(q), \quad q>0, \quad
1682: \mbox{where $\vp(0^+)=0$,\ $\vp(\infty)=1$.}
1683: \end{equation}
1684: On the other hand, since $T_nx\nu\nsup_1(\beta_ndx)= \tilde\sm_{1/T_n}(dx)$,
1685: statement (ii) of Theorem~\ref{RT.initial_tails} is equivalent to the assertion
1686: \begin{equation}\label{e.ph2}
1687: \tilde\vp\nsup(T_n\inv,q) \to \Vp(q), \quad q>0, \quad
1688: \mbox{where $\Vp(0^+)=0$,\ $\Vp(\infty)=\infty$.}
1689: \end{equation}
1690: But by the solution formulae \qref{eq:vp1} and \qref{C.eternal}, we have
1691: \[
1692: \tilde\vp\nsup(T_n\inv,q) = \frac{\tilde\vp\nsup(1,q)}
1693: {1+(T_n\inv-1)\tilde\vp\nsup(1,q)},
1694: \qquad
1695: \Vp(q) = \frac{\vp(q)}{1-\vp(q)}.\]
1696: Since evidently \qref{e.ph1} is equivalent to \qref{e.ph2}, (i)
1697: is equivalent to (ii).
1698: \end{proof}
1699:
1700:
1701: \subsection{The representation at $+\infty$}
1702:
1703: For the additive kernel, Bertoin showed that an eternal solution can
1704: be uniquely identified by its asymptotic behavior as $t\to\infty$ also.
1705: For the constant kernel, an analogous result follows easily
1706: from \qref{C.vpdef} and \qref{C.eternal}.
1707:
1708: \begin{thm} Let $\nu$ be an eternal solution of Smoluchowski's
1709: equation with constant kernel $K=2$, and let $\Vp$ be the Laplace
1710: exponent of the divergent \smeas\ associated with $\nu$.
1711: Then as $t\to\infty$,
1712: the measure $t^2 \nu_t$ converges weakly on $(0,\infty)$ to
1713: a measure $\Lambda_+$ with Laplace transform
1714: \[ \Vp_+(q):= \int_0^\infty \rme^{-qx}\Lambda_+(dx) = \frac{1}{\Vp(q)}
1715: = \lim_{t\to\infty} t^2 \int_0^\infty \rme^{-qx}\nu_t(dx).
1716: \]
1717: \end{thm}
1718:
1719: Clearly an eternal solution $\nu$ is uniquely determined from $\Lambda_+$
1720: through $\Vp(q)=1/\Vp_+(q)$.
1721: We see that the measure $\Lambda_+$ has a Laplace transform
1722: $\Vp_+(q)$ defined for all $q>0$, and $\Vp_+(q)\to\infty$ as
1723: $q\to0$ since $\Vp(0^+)=0$.
1724: So $\int_0^1\Lambda_+(dx)<\infty$ and $\int_E\Lambda_+(dx)=\infty$.
1725:
1726: The class of measures $\Lambda_+$ which arise in this way
1727: is characterized by the property that $\eta(q)=\D_q(1/\Vp_+(q))$
1728: is the Laplace transform of some divergent \smeas\ $\dsm$
1729: (i.e., $\eta$ is completely monotone, locally integrable on $[0,\infty)$
1730: and $\int_E \eta(q)\,dq=\infty$).
1731: There does not appear to be a simple characterization by moment
1732: conditions.
1733:
1734: \begin{rem}
1735: This representation has an interesting probabilistic interpretation;
1736: see~\cite[p.74]{B_book}.
1737: If $X_\cdot$ is a subordinator with Laplace exponent $\Vp$, then
1738: $\Vp_+ = 1/\Vp$ is the Laplace transform of the potential measure $U$,
1739: defined on Borel sets $A \subset E$ by
1740: $U(A) = \mathbb{E} \left( \int_0^\infty \mathbf{1}_{\{X_s \in A\}} ds
1741: \right)$.
1742: \end{rem}
1743:
1744:
1745: \section{The additive kernel}
1746: \label{sec:add}
1747: In this section we study the scaling dynamics for the additive kernel.
1748: Our main aims are to prove continuous dependence on initial data,
1749: establish the correspondence between points on the scaling attractor
1750: and eternal solutions, and revisit Bertoin's \LK\/ representation
1751: with convergence of \smeass\ in mind.
1752:
1753: \subsection{Solution by Laplace transform}
1754: The solution of Smoluchowski's equation with kernel $K=x+y$ by the
1755: Laplace transform is
1756: classical~\cite{Drake}, and remains the basis for rigorous
1757: work. Let $t_0\in\R$ be arbitrary.
1758: We assume $\nu_{t_0}$ is a (possibly infinite) measure with
1759: $\int_E x \nu_{t_0}(dx) < \infty$. Without loss of generality,
1760: we may assume $\int_E x\nu_{t_0}(dx)=1$.
1761: We have shown~\cite[Thm 2.8]{MP1} that (\ref{eq:smol1}) has a unique
1762: solution $\nu_t$ for $t \geq t_0$ in an appropriate weak sense, such that
1763: \be
1764: \la{A.mass1}
1765: \int_E x \nu_t(dx) = 1, \quad t \geq t_0.
1766: \ee
1767: As for the constant kernel, we use the notation
1768: \begin{equation}\label{A.vpdef}
1769: \vp(t,q) = \int_E (1-\rme^{-qx}) \nu_t(dx) , \quad q \geq 0.
1770: %\qquad
1771: %\vp_1(q) = \int_E (1-\rme^{-qx}) \nu_1(dx),
1772: \end{equation}
1773: and set $\vp_0(q)=\vp(t_0,q)$.
1774: To study scaling limits we consider the mass distribution function,
1775: which is the natural probability distribution function associated to a
1776: solution. Let
1777: \begin{equation}\label{A.Fdef}
1778: F_t(x) = \int_{(0,x]} y\, \nu_t(dy) .
1779: \end{equation}
1780: Note that the Laplace transform of $F_t$ is
1781: \begin{equation}\label{A.Fvp}
1782: \int_E \rme^{-qx}F_t(dx) = \partial_q \vp(t,q).
1783: \end{equation}
1784: Thus, $\partial_q \vp(t,q)$ is completely monotone and
1785: $\partial_q \vp(t,0)= 1$, $t \geq 0$.
1786: We know from \cite{MP1} that if we substitute
1787: $f(x)=1-\rme^{-qx}$ in (\ref{R.moment})
1788: we find that $\vp(t,q)$ solves the hyperbolic equation
1789: \begin{equation}
1790: \label{A.1}
1791: \partial_t \vp -\vp \partial_q \vp = -\vp.
1792: \end{equation}
1793:
1794: Following Bertoin, it is convenient to introduce the new variables
1795: \begin{align}
1796: \label{A.psi1}
1797: &s = \rme^t, \quad s_0= \rme^{t_0},
1798: %\\ \la{A.psi1}
1799: %&\Vp(s,q)=\vp(t,q),
1800: \quad
1801: \psi(s,q) = \frac{q}{s} - \vp\left(t, \frac{q}{s}\right).
1802: \end{align}
1803: By \qref{A.mass1} and \qref{A.vpdef}, $\psi$ is the
1804: Laplace exponent
1805: \be
1806: \la{A.psi2}
1807: \psi(s,q) = \int_{E} y^{-2} \left(\rme^{-qy} - 1 + qy\right)
1808: \sm_{\log s}(dy),
1809: \ee
1810: where $\sm_t$ denotes the \smeas\
1811: \be
1812: \la{A.psi3}
1813: \sm_t(dx) = x^2 \nu_{t} (\rme^t \,dx).
1814: \ee
1815: Observe that $\sm_t$ is not a finite measure in general, but if
1816: $\sm_{t}(E) <\infty$ for some $t$, then $\sm_t(E)$ is finite for
1817: for every $t$ for which the solution is defined, and is constant.
1818: We substitute \qref{A.psi1} in \qref{A.1} to see that $\psi$ satisfies
1819: the inviscid Burgers equation
1820: \be
1821: \la{A.psi4}
1822: \partial_s \psi + \psi \partial_q \psi =0, \quad s > s_0.
1823: \ee
1824: The values of $\psi$, $\D_q\psi$ and $\D_q^2\psi$ are positive
1825: for $s\ge s_0$, $q>0$, and $\D_q^2\psi(s,\cdot)$ is completely
1826: monotone since it is the Laplace transform of $G_t$.
1827: In addition, \qref{A.mass1}, \qref{A.psi2} and \qref{A.psi3} imply
1828: \be
1829: \la{A.psidivg}
1830: \lim_{q \to \infty} \partial_q \psi(s,q) = \int_Ex^{-1}\sm_{\log s}
1831: (dx) = s^{-1}.
1832: \ee
1833:
1834: We may describe $\psi(s,q)$ globally for $s>s_0$ by the method of
1835: characteristics. A surprising fact is that we may always solve for
1836: $\psi$ backwards in time, for all $s>0$, without developing singularities.
1837: The solution need not correspond to a positive measure $\nu_t$
1838: for $t <t_0$, however. This is analogous to the situation for $K=2$.
1839: \begin{lemma}
1840: \label{AL.charsoln}
1841: Let $t_0 \in \R$ and $\nu_{t_0} \in \M$ with $\int_E
1842: x\,\nu_{t_0}(dx)=1$, and let $\psi_0(q_0)= q_0/s_0 - \vp_0(q_0/s_0)$.
1843: There is a unique solution $\psi(s,q)$ to \qref{A.psi4} defined for every
1844: $s >0$ and $q> 0$, such that $\psi(s_0,\cdot)=\psi_0(\cdot)$.
1845: \end{lemma}
1846: \begin{proof}
1847: Applying the method of characteristics as usual,
1848: the solution $\psi=\psi(s,q)$ is determined implicitly from the equation
1849: \begin{equation}\label{A.imp}
1850: h(s,q,\psi):= \psi-\psi_0(q-(s-s_0)\psi)=0.
1851: \end{equation}
1852: We have $h(s,q,0)<0$, and $\D_\psi h>s/s_0$ since
1853: $\D_q\psi_0<s_0^{-1}$ by \qref{A.psidivg}.
1854: Since $\psi_0$ is analytic, \qref{A.imp} determines a solution
1855: of \qref{A.psi4} analytic in $(s,q)$ for all $s>0$, $q>0$.
1856: \end{proof}
1857:
1858: Equation \qref{A.imp} determines the solution at time $s$ from data at time
1859: $s_0$ and plays the same role in the analysis here as equation
1860: \qref{eq:vp1} played in the previous section. Convergence properties
1861: of solutions will be deduced from the pointwise convergence properties
1862: of the Laplace exponent $\psi$ using the theory from section \ref{S.laplace}.
1863:
1864: \begin{thm}\label{AT.cts} (Continuous dependence on data.)
1865: For Smoluchowski's equation with additive kernel $K=x+y$, let $t_0\in \R$
1866: and let $\nu\nsup$ be a sequence of solutions defined for $t\ge t_0$
1867: with $\int_E x \nu\nsup_t(dx)=1$ for all $t \geq t_0.$
1868: \begin{enumerate}
1869: \item[(a)] If $x\nu\nsup_{t_0}(dx)$ converges weakly to a measure
1870: $x\hat\nu_0(dx) $
1871: with $\int_E x \hat\nu_0(dx)=1$, then for every $t\ge t_0$ we have
1872: that $x\nu\nsup_t(dx)$ converges weakly to $x\nu_t(dx)$, the time-$t$ solution
1873: with initial data $\nu_{t_0}=\hat\nu_0$.
1874: \item[(b)] For any $t\ge t_0$, if $x\nu\nsup_{t}(dx)$ converges weakly to a
1875: measure $x\hat\nu(dx)$ with $\int_Ex\,\hat\nu(dx) =1$, then
1876: $x\nu\nsup_{t_0}(dx)$ converges weakly to a measure $x\hat\nu_0(dx)$ with
1877: $\int_E x\,\hat\nu_0(dx)=1$, and $\hat\nu=\nu_t$, the time-$t$ solution
1878: with initial data $\nu_{t_0} = \hat\nu_0$.
1879: \end{enumerate}
1880: \end{thm}
1881: \begin{proof}
1882: We prove (a); the proof of (b) is similar.
1883: Let $\sm\nsup_t(dx)=x^2\nu_t(\rme^tdx)$, and with $s=\rme^t$ let
1884: \begin{equation} \label{A.psin1}
1885: \psi\nsup(s,q) = \int_E y^{-2}(\rme^{-qy}-1+qy) \sm\nsup_t(dy).
1886: \end{equation}
1887: The family $\psi\nsup(s_0,\cdot)$ is uniformly
1888: Lipschitz, since equation \qref{A.psidivg} implies
1889: \begin{equation}\label{A.lip}
1890: \psi\nsup(s_0,0)=0,\quad
1891: 0\leq \D_q\psi\nsup(s_0,q) \le 1/s_0 \quad\mbox{ for all $q>0$.}
1892: \end{equation}
1893: The hypothesis is equivalent to saying that the \smeass\
1894: $\sm\nsup_{t_0}$ converge to a non-divergent \smeas\
1895: $\hat\sm_0(dx)=x^2\hat\nu_0(dx)$ with $\int_E x\inv\hat\sm_0(dx)=1$.
1896: By Theorem~\ref{LT.Mconv} and the criteria in \qref{R.ginf}-\qref{R.divcon},
1897: this is equivalent to the statement that for all $q>0$,
1898: $\psi\nsup(s_0,q)\to\hat\psi_0(q)$, where $\hat\psi_0$ is the
1899: (second-order) Laplace exponent for $\hat\sm_0$, with
1900: $\D_q\hat\psi_0(0^+)=0$, $\D_q\hat\psi_0(\infty)=1/s_0$.
1901: (Note $\D_q\hat\psi$ is the first-order Laplace exponent of $\hat\sm_0$.)
1902: As in \qref{A.imp} we have
1903: \begin{equation}\label{A.imp2}
1904: \psi\nsup(s,q)-\psi\nsup(s_0,q-(s-s_0)\psi\nsup(s,q))=0.
1905: \end{equation}
1906: For fixed $s,q$, the sequence $\psi\nsup(s,q)$ is bounded, and
1907: any subsequential limit $\psi_*$ must satisfy
1908: \begin{equation}\label{A.psist}
1909: \psi_* - \hat\psi_0(q-(s-s_0)\psi_*)=0,
1910: \end{equation}
1911: due to the equicontinuity of the maps
1912: $\psi\mapsto\psi\nsup(s_0,q-(s-s_0)\psi)$.
1913: But equation \qref{A.psist} has the unique solution $\psi_*=\psi(s,q)$,
1914: where $\psi$ is the solution of \qref{A.psi4} with
1915: $\psi(s_0,q)=\hat\psi_0(q)$, $q>0$.
1916: Hence the whole sequence $\psi\nsup(s,q)$ converges pointwise to
1917: $\psi(s,q)$. Moreover, differentiating \qref{A.psist} yields
1918: $\D_q\psi_*(0)=0$, $\D_q\psi_*(\infty)=1/s$, since $s_0=1$.
1919: Then the conclusion of (a) follows from
1920: Theorem~\ref{LT.Mconv}, \qref{R.ginf} and \qref{R.divcon}.
1921: \end{proof}
1922:
1923: \subsection{The scaling attractor and eternal solutions}
1924: \begin{thm}\label{AT.attractor}
1925: A probability measure $\hat F$ is an element of the scaling attractor $\attr$ for
1926: Smoluchowski's equation with additive kernel $K=x+y$ if and only if
1927: $\hat F(dx) = x\nu_0(dx)$ for some eternal solution $\nudot$.
1928: \end{thm}
1929: %
1930: \begin{proof}
1931: Suppose $\hat F(dx) = x\nu_0(dx)$ for some eternal solution
1932: $\nudot$. We show $\hat F \in \attr$.
1933: Pick arbitrary sequences $T_n, b_n \to \infty$, and consider
1934: the sequence of rescaled eternal solutions
1935: \[ \nu_t\nsup(dx) = b_n^{-1} \nu_{t-T_n} \left(b_n^{-1}
1936: dx\right), \quad t \in \R. \]
1937: The corresponding distribution functions satisfy $F\nsup_{T_n}(b_n x) =
1938: \hat F(x)$ for every $x$.
1939: Thus, $\hat F\in\attr$ by Definition~\ref{R.attractor}.
1940: %% \int_0^{b_nx} xb_n^{-1} \nu_0(b_n^{-1} dx) =
1941: %% \int_0^x y \nu_0(dy) = \hat\mu(0,x) \]
1942: %% for every $x$. Thus, Definition~\ref{A.attractor} holds.
1943:
1944: To prove the converse, suppose $\hat F \in \attr$. We show that
1945: $\hat F = F_0$ for some eternal solution $\nudot$. Let $\hat \vp$
1946: correspond to $\hat\nu$ as in \qref{A.vpdef},
1947: and $\nudot\nsup,T_n, b_n$ be as in
1948: Definition~\ref{R.attractor}. Consider the rescaled measures
1949: \[
1950: \tilde\nu\nsup_t(dx)=b_n\nu\nsup_{t+T_n}(b_n dx).
1951: \]
1952: This rescaling yields a solution that is defined for $t\ge -T_n$.
1953: By assumption,
1954: \[ \tilde F\nsup_0(x) = \int_0^x y\, \tilde\nu\nsup_0 (dy) =
1955: F\nsup_{T_n}(b_nx) \to \hat F(x),\]
1956: at all points of continuity. By
1957: Theorem~\ref{AT.cts}, this implies that for any $N\in\N$ the solutions
1958: $\nu\nsup_t$ converge weakly to $\nu_t$ for all $t \geq -N$. In
1959: particular, $\nu_t$ is a solution for $t \geq -N$ for all $N$, thus it is an
1960: eternal solution.
1961: \end{proof}
1962:
1963: Let us now prove that $\attr$ is invariant (part (a) of
1964: Theorem~\ref{RT.AE}). The proof is substantially the same as for
1965: $K=2$. Suppose $\nu$ is a solution on some time interval $[t_1,\infty)$,
1966: normalized so $ \int_E x\nu_t(dx)=1, t \geq t_1$.
1967: Suppose $F_{T}\in\attr$ for some $T\ge t_1$. We may presume $T=0$
1968: without loss (if not, we translate in time,
1969: replacing $\nu_t(dx)$ by $\nu_{t-T}(dx)$).
1970: By Theorem~\ref{AT.attractor} above, $F_T=x\tilde\nu_0$ for
1971: some eternal solution $\tilde\nu$. But then $\nu_t=\tilde\nu_t$
1972: for all $t\ge t_1$, meaning that $\nu$ is (the restriction of) an
1973: eternal solution. We obtain that $F_t\in\attr$ for every $t \in \R$ by
1974: a similar argument.
1975:
1976: \subsection{\LK\/ representation of eternal solutions}
1977: We now prove Bertoin's \LK\/ representation for eternal solutions.
1978: The proof mainly follows~\cite{B_eternal}, and
1979: is included to stress the basic framework.
1980:
1981: \begin{thm}[cf.\ Bertoin~\cite{B_eternal}]
1982: \label{AT.eternal}
1983: \begin{enumerate}
1984: \item[(a)] Let $\nu$ be an eternal solution to Smoluchowski's
1985: equation with $K=x+y$, and let $\sm_t(dx)=x^2\nu_t(\rme^tdx)$ be
1986: associated \smeass.
1987: Then there is a unique divergent \smeas\ $\dsm$
1988: such that $\sm_{t}$ converges to $\dsm$ as $t\to -\infty$.
1989: %
1990: %(dx)=\sigma^2 \delta_0(dx) +
1991: %x^2\,\Lambda(dx)$ where $\delta_0$ is a delta mass at 0 and
1992: %$(\sigma^2,\Lambda)$ is a \Levy\ pair.
1993: %Equivalently, $\psi(s,q) \to \Psi(q)$ as $s\to 0$ for all $q\ge0$.
1994: \item[(b)] Conversely, given a divergent \smeas\ $\dsm$
1995: there is a unique eternal solution with the properties in part (a),
1996: defined as follows. Let
1997: \begin{equation}\label{A.Psidef}
1998: \Psi(q) = \int_{\Ebar} \frac{\rme^{-qx}-1+qx}{x^2} H(dx)
1999: \end{equation}
2000: be the Laplace exponent of $H$,
2001: and let $\psi=\psi(s,q)$ be the solution to
2002: \begin{equation}\label{A.Imp}
2003: \psi - \Psi(q-s\psi) = 0.
2004: \end{equation}
2005: %\qref{A.psi3} with $\psi(0,q)=\Psi(q)$.
2006: Then $\nu_t$ is determined by \qref{A.psi2} and \qref{A.psi3}.
2007: \end{enumerate}
2008: \end{thm}
2009: %% \begin{rem}
2010: %% Observe that when $\sigma=0$, the procedure in (b) formally yields
2011: %% a solution to Smoluchowski's equation
2012: %% with ``initial'' size distribution $\nu_0(dx)=\Lambda(dx)$ having infinite
2013: %% total mass. [[?? At $-\infty$?]]
2014: %% \end{rem}
2015: \begin{proof}
2016: We first prove (a). By Theorem~\ref{LT.Mconv} and
2017: \qref{R.ginf}-\qref{R.divcon}, it is enough to
2018: show that $\psilevy(q):=\lim_{s \to 0} \psi(s,q)$ exists for every
2019: $q \geq 0$, with $\D_q\Psi(0)=0$ and $\D_q\Psi(\infty)=\infty$.
2020: %$\lim_{q \to \infty} q\inv \psilevy(q)=\infty$.
2021: We know $\psi\ge0$ and $\partial_q\psi\ge0$, so
2022: $\partial_s \psi(s,q) \leq 0$ for all $q$, $s$.
2023: Hence it suffices to show that for each $q>0$, $\psi(s,q)$ stays bounded
2024: as $s\dnto0$.
2025:
2026: {\em 1.\/} We first show $\psi(s,q)$ stays bounded for $q$ near $0$.
2027: Choose $q_1 >0$ such that $q_{*}:=q_1 -\psi(1,q_1) =\vp(0,q_1) >0$.
2028: Then $\psi(s,q)=\psi(1,q_1)$ along the characteristic line joining
2029: $(0,q_{*})$ and $(1,q_1)$, so $0\le\psi(s,q)\le\psi(1,q_1)$ whenever
2030: $0<s<1$ and $0<q\le q_{*}$.
2031: %Since $\psi(1,\cdot)$ is convex, the geometry of characteristics is as
2032: %in Figure~\ref{AF.char}.
2033: (See Fig.~\ref{AF.char}.)
2034: %-----------------------------------------
2035: \begin{figure}
2036: \la{AF.char}
2037: \centerline{\epsfysize=7cm{\epsffile{char.eps}}}
2038: \caption{Geometry of characteristics}
2039: \end{figure}
2040: %------------------------------------------
2041: %Thus, for any $s_0\in(0,1)$, $q_0<??$
2042: %\[ \psi(s_0,q_0) = \psi(1,q(1;s_0,q_0)) \leq \psi(1,q_*), \quad 0 < s_0
2043: %\leq 1.\]
2044: % q_0 \leq q_* -\psi(1,q_*). \]
2045:
2046: {\em 2.\/} For $q > q_*$ the complete monotonicity of $q\mapsto q^{-2}
2047: \psi(s,q)$ implies $\psi(s,q) < q^2q_*^{-2} \psi(s,q_*)$.
2048: %Thus, $ \lim_{s \to 0} q^{-2} \psi(s,q) = q^{-2}\Psi(q) \leq
2049: %q_*^{-2}\psilevy(q_*)$.
2050:
2051: {\em 3.\/} We now show $\D_q\Psi(0)=0$ and $\D_q\Psi(\infty)=\infty$.
2052: %$\lim_{q \to \infty}q^{-1}\Psi(q)=\infty$.
2053: Observe that $\Psi$ solves
2054: \[ \psilevy(q)=\psi(1,q+\psilevy(q)), \quad q> 0.\]
2055: Therefore,
2056: \begin{equation}
2057: \label{A.derivative}
2058: \partial_q \psilevy(q) = \frac{\partial_q\psi(1,q+\psilevy(q))}{1
2059: -\partial_q\psi(1,q+\psilevy(q))}.
2060: \end{equation}
2061: Since $\psi(1,q)=q-\vp(0,q)$, we have
2062: $\partial_q \psi(1,q)=1-\partial_q \vp(0,q) \to 0$
2063: as $q\to0$, $\to 1$ as $q \to \infty$. Thus,
2064: $\D_q\Psi(0)=0$ and $\D_q\Psi(\infty)=\infty$.
2065: %$\lim_{q \to \infty} \partial_q \psilevy(q) = \infty$.
2066: %This also implies $\lim_{q \to \infty} q\inv
2067: %\psilevy(q)=\infty$. Indeed, for any $n \in \N$ choose $q_n$ such that
2068: %$\partial_q \psilevy(q) \geq n$ for $q \geq q_n$. Then,
2069: %\[\psilevy(q) \geq \psilevy(q_n) + n (q -q_n), \quad q \geq q_n\]
2070: %and $\liminf_{q \to \infty} q\inv \psilevy(q) \geq n$.
2071: % Since $n$ is
2072: % arbitrary, this proves $ \lim_{q \to \infty} q\inv
2073: % \psilevy(q)=\infty.$.
2074: This proves (a).
2075:
2076:
2077: We now prove (b). Let $\dsm$ be a divergent \smeas\ and $\Psi$
2078: be defined by \qref{A.Psidef}. Note $\D_q\Psi(0)=0$ and
2079: $\D_q\Psi(\infty)=\infty$ by \qref{R.ginf}-\qref{R.divcon}.
2080: Since $\D_q\Psi(q)>0$, $\psi(s,q)$ is globally defined and
2081: analytic with $\psi(s,q)<q/s$, and \qref{A.psi4} holds for all
2082: $s>0$, $q>0$. With $\Vp$ and $\vp$ defined by \qref{A.psi1},
2083: \qref{A.1} follows.
2084:
2085: By the well-posedness theory in \cite{MP1},
2086: we obtain an eternal solution through \qref{A.Fvp},
2087: provided we show that $\D_q\vp(t,\cdot)$ is completely monotone,
2088: which implies that it is the Laplace transform of a (positive) measure
2089: that we call $x\nu_t(dx)$.
2090: From \qref{A.Imp} we obtain that $\vp=\vp(t,q)$ satisfies
2091: \be
2092: \la{A.psi11}
2093: q = \vp + \psilevy(s\vp),
2094: \ee
2095: whence
2096: \be
2097: \la{A.psi12}
2098: \D_q \vp = \frac{1}{1+s\psilevy'(s\vp)}.
2099: \ee
2100: Since $q\mapsto 1+s\psilevy'(sq)$ is positive with completely monotone
2101: derivative, the map $q\mapsto (1+s\psilevy'(sq))^{-1}$ is
2102: completely monotone \cite[XIII.4]{Feller}. We then infer that
2103: $\D_q\vp(s,\cdot)$ is completely monotone by Lemma~\ref{AL.comp} below.
2104: Since $\Psi'(0)=0$ we have the normalization \qref{A.mass1}, $\D_q\vp(t,0)=1$.
2105: This finishes the proof of existence.
2106:
2107: Note that total number of clusters $\nu_t(E)=\vp(t,\infty)=\infty$
2108: always here.
2109:
2110: Let us show that the eternal solution defined by
2111: this procedure is unique. Let $H$ be a divergent \smeas\ and suppose
2112: $\nu, \tilde{\nu}$ are two eternal solutions with \smeass\ $G_t,
2113: \tilde{G}_t$ that converge to $H$. But this is equivalent to
2114: pointwise convergence of $\psi(s,q)$ and
2115: $\tilde{\psi}(s,q)$ to $\Psi(q)$ as $s\to 0$ where $\psi$ and
2116: $\tilde{\psi}$ solve \qref{A.psi4}. But the solutions to the inviscid
2117: Burgers equation with increasing initial data are unique, thus
2118: $\psi(s,q)=\tilde{\psi}(s,q)$ and $\nu = \tilde{\nu}$.
2119: \end{proof}
2120:
2121: \begin{lemma}\la{AL.comp}
2122: Suppose $f,g\colon E\to E$, $f'=g(f)$ and $g$ is completely
2123: monotone. Then $f'$ is completely monotone.
2124: \end{lemma}
2125: \begin{proof}
2126: We prove by induction that the first $n$ derivatives of
2127: $G\circ f$ alternate in sign for every
2128: completely monotone function $G$. For $n=0$, $G(f)>0$. Suppose
2129: the statement is true for some $n\ge0$. Let $G$ be completely
2130: monotone, and note
2131: \[
2132: -(G\circ f)' = -G'(f)g(f)= \tilde G(f)
2133: \]
2134: and $\tilde G$ is completely monotone since it is the product of
2135: completely monotone functions. Using the induction hypothesis,
2136: we deduce that the first $n+1$ derivatives of $G\circ f$
2137: alternate in sign.
2138: \end{proof}
2139:
2140: To complete the proof of Theorem~\ref{RT.LK} for $K=x+y$, we need to
2141: check that the map $\nu_0 \mapsto \dsm$ from $\attr$ to $\Mcan$
2142: is a bi-continuous bijection.
2143: \begin{thm}\label{AT.Mconv2}
2144: Let $\nu\nsup$ be a sequence of eternal solutions with corresponding
2145: divergent \smeass\ $\dsm\nsup$.
2146: Fix $t \in \R$. Then, taking $n\to\infty$, the following are equivalent:
2147: \bit
2148: \item[(i)]
2149: $x\nu\nsup_t$ converges weakly to $x\hat\nu$ with $\int_E x \hat\nu(dx)=1$.
2150: \item[(ii)] $\dsm\nsup$ converges to a divergent
2151: \smeas\ $\dsm$.
2152: \eit
2153: If either (equivalently both) of these conditions hold, then
2154: $\hat\nu=\nu_t$ for an eternal solution with \smeas\ $\dsm$.
2155: \end{thm}
2156: \begin{proof}
2157: With Theorem~\ref{LT.Mconv} in hand, the proof of
2158: Theorem~\ref{AT.Mconv2} is essentially the same as that
2159: of Theorem~\ref{AT.cts}.
2160: Assume (i), so $\nu\nsup_t$ converges to $\hat\nu$ with
2161: $\int_E x\hat\nu(E)=1$. Then $\sm\nsup_t(dx)=x^2\nu_t\nsup(\rme^t dx)$
2162: converges to the \smeas\/ $\hat\sm(dx)=x^2 \hat\nu(\rme^t dx)$ and
2163: the associated Laplace exponents converge: $\psi\nsup(s,q)\to\hat\psi(q)$
2164: for all $q>0$, with $\D_q\hat\psi(0)=0$, $\D_q\hat\psi(\infty)=1/s$.
2165: Recall that $\psi\nsup (s,q)$ solves
2166: \begin{equation} \label{A.Pconv}
2167: \Psi\nsup(q-s\psi\nsup(s,q)) = \psi\nsup(s,q).
2168: \end{equation}
2169: Let $M >0$. A calculation as in \qref{A.derivative} shows that
2170: $\partial_q\Psi\nsup(q)$ is uniformly bounded in $n$ for $q \in
2171: [0,M]$. We claim that $\lim_{n \to \infty} \Psi\nsup (q - s\hat\psi(q))$ exists
2172: for every $q$. Let us restrict attention to $q \in [0,M]$. Then by \qref{A.Pconv}
2173: \[ \Psi\nsup(q-s\hat\psi(q)) = \psi\nsup(s,q) + \left(
2174: \Psi\nsup(q-s\hat\psi(q))
2175: -\Psi\nsup(q-s\psi\nsup(s,q))\right). \]
2176: The first term converges to $\hat\psi(q)$ and the second to zero by the
2177: uniform estimate on $\partial_q\Psi\nsup(q)$ on $[0,M]$. Since $M >0$
2178: was arbitrary, we may use
2179: Theorem~\ref{LT.Mconv} to deduce that $\Psi\nsup(q)$ converges to a
2180: Laplace exponent $\Psi(q)$ that satisfies
2181: \[ \Psi(q-s\hat\psi(q))=\hat\psi(q).\]
2182: As with \qref{A.derivative} and its sequel it follows that
2183: $\D_q\Psi(0)=0$ and $\D_q\Psi(\infty)=\infty$.
2184: %$\lim_{q\to \infty} \partial_q\Psi(q) = \lim_{q \to \infty} q^{-1}
2185: %\Psi(q)=\infty$. Thus [[not done??]],
2186: Thus $\Psi$ is the Laplace
2187: exponent of a divergent \smeas\/ $\dsm$, and $\dsm\nsup$ converges
2188: to $\dsm$.
2189:
2190: We now show (ii) implies (i). Suppose the divergent \smeass\ $\dsm\nsup$
2191: converge to a divergent \smeas\ $\dsm$. Then
2192: Theorem~\ref{LT.Mconv} implies $\Psi\nsup(q)\to\Psi(q)$ for every
2193: $q>0$, and $\D_q\Psi(0)=0$, $\D_q\Psi(\infty)=\infty$.
2194: %$\partial_q\Psi(q)\to\infty$ as $q\to\infty$.
2195: Then the characteristics emanating from $s=0$ converge because $q + s
2196: \Psi\nsup(q) \to q + s\Psi(q)$. Thus, $\psi\nsup(s,q) \to \psi(s,q)$,
2197: which satisfies \qref{A.Imp}.
2198: This yields weak convergence of $x\nu\nsup_t$ to
2199: $x \nu_t$, where $\nu$ is the eternal solution with divergent \smeas\ $\dsm$.
2200: \end{proof}
2201:
2202: \subsection{Scaling limits and initial tails}
2203:
2204: Let us now prove Theorem~\ref{RT.initial_tails} for the additive kernel.
2205:
2206: \begin{proof}[Proof of Theorem~\ref{RT.initial_tails}]
2207: %Let $a_n=\rme^{T_n}$, $b_n=\beta_n$.
2208: We rescale solutions via
2209: $\tilde\nu\nsup_t(dx)=\beta_n\nu\nsup_{t+T_n}(\beta_ndx)$,
2210: and let $\tilde F\nsup_t(dx) = \beta_n x \tilde\nu\nsup_t(dx)$.
2211: Also let $\tilde\sm\nsup_t(dx)= x^2\tilde\nu\nsup_t(\rme^t dx)$ and let
2212: $\tilde\psi\nsup(s,q)$ be the associated Laplace exponent as in
2213: \qref{A.psi2}. Observe
2214: $\tilde\sm\nsup$ in \qref{R.initG} is $\tilde\sm\nsup_{-T_n}$ and
2215: $\int_E x^{-1} \tilde \sm\nsup(dx) = \rme^{T_n}$.
2216: Let $H$ be the divergent \smeas\ corresponding to $\nu$ and
2217: $\Psi$ its Laplace exponent, and let $\psi(q)$ be the Laplace exponent
2218: of $\sm(dx)= x^2\nu_0(dx)$.
2219:
2220: Then (i) is equivalent to saying $\tilde F\nsup_0\to \hat F$ weakly,
2221: meaning the \smeass\/ $\tilde\sm\nsup_0$ converge to $\sm$
2222: with $\int_E x\inv\sm(dx)=1$. This is equivalent to saying
2223: \begin{equation}\label{A.ps1}
2224: \tilde\psi\nsup(1,q)\to \psi(q), \quad q>0, \quad
2225: \mbox{where $\D_q\psi(0^+)=0$, \ $\partial_q\psi(\infty)=1$.}
2226: \end{equation}
2227: On the other hand, since $\beta_nx^2\nu\nsup_0(\rme^{-T_n}\beta_ndx)=
2228: \tilde\sm_{-T_n}(dx)$,
2229: (ii) is equivalent to saying
2230: \begin{equation}\label{A.ps2}
2231: \tilde\psi\nsup(\rme^{-T_n},q) \to \Psi(q), \quad q>0, \quad
2232: \mbox{where $\D_q\Psi(0^+)=0$, \ $\partial_q \Psi(\infty)=\infty$.}
2233: \end{equation}
2234: For brevity, let $\tilde\psi\nsup (q)$ denote $\tilde\psi\nsup(1,q)$
2235: and $\tilde\Psi\nsup(q)$ denote $\tilde\psi\nsup(\rme^{-T_n},q)$. Then
2236: the implicit solution formulas to \qref{A.psi4} read
2237: \[ \psi(q) = \Psi(q-\psi(q)), \quad \psi\nsup(q) = \Psi \nsup (
2238: q-(1-\rme^{-T_n}) \psi\nsup(q)). \]
2239: As in the proof of Theorem~\ref{AT.Mconv2} we may now deduce that
2240: \qref{A.ps1} is equivalent to \qref{A.ps2}, implying (i) is equivalent
2241: to (ii). The details are omitted.
2242: \end{proof}
2243:
2244:
2245: \section{The multiplicative kernel}
2246: \label{sec:mult}
2247: In this section we study scaling dynamics approaching the gelation time
2248: for the kernel $K=xy$. The study of the multiplicative kernel
2249: can be reduced to the additive kernel by a simple change of
2250: variables. This trick is well-known (see~\cite{Drake}), and allows us
2251: to avoid separate proofs.
2252:
2253: \subsection{Solution by the Laplace transform}
2254: The self-similar solutions for $K=xy$ have infinite number and mass,
2255: but finite second moment. However, one
2256: may develop a natural well-posedness theory using only the finiteness
2257: of the second moment~\cite{MP1}. We assume $\nu_{t_0}$ is a (possibly
2258: infinite) measure with $\int_E x^2 \nu_{t_0}(dx) < \infty$. Without
2259: loss of generality, we may scale so that $\int_E x^2\nu_{t_0}(dx)=1$
2260: and $t_0=-1$ as in \qref{R.normal2}. We define the Laplace exponent
2261: (note the change from \qref{A.vpdef})
2262: \begin{equation}\label{M.vpdef}
2263: \vp(t,q) = \int_E (1-\rme^{-qx}) x\nu_t(dx) , \quad q \geq 0,
2264: %\qquad
2265: %\vp_1(q) = \int_E (1-\rme^{-qx}) \nu_1(dx).
2266: \end{equation}
2267: and write $\vp_0(q)=\vp(t_0,q)$. We may substitute \qref{M.vpdef} in the
2268: moment identity \qref{R.moment} to obtain
2269: \begin{equation}
2270: \label{M.1}
2271: \partial_t \vp - \vp \partial_q \vp = 0, \quad t \in (t_0,0).
2272: \end{equation}
2273: Equation \qref{M.1} may be transformed to \qref{A.1} by the following
2274: change of variables. Let $\vp^{\rm add}(\tau,q)$, $\tau \in [0,\infty)$,
2275: denote a solution to \qref{A.1} with initial data $\vp_0(q)$. Then the
2276: solution to \qref{M.1} is given by
2277: \be
2278: \la{M.2}
2279: \vp(t,q) = \rme^{\tau}\vp^{\rm add}(\tau,q), \quad \tau=\log(|t|^{-1}),
2280: \ee
2281: which may also be written in terms of the number density as
2282: \be
2283: \la{M.3}
2284: x \nu_{t}(dx) = e^\tau \nu^{\rm add}_\tau(dx).
2285: \ee
2286: Conservation of mass \qref{A.mass1} is now replaced by
2287: \be
2288: \la{M.mass1}
2289: \int_E x^2 \nu_t(dx) = |t|^{-1}, \quad t \in [t_0,0),
2290: \ee
2291: and the probability measure $F_t$ associated to $\nu_t$ is defined by
2292: \be
2293: F(t,x) = |t| \int_0^x y^2\nu_t(dy).
2294: \ee
2295: As in \qref{A.psi3} we define the \smeas\/
2296: \be
2297: \label{M.psi3}
2298: \sm_t(dx) = x^3 \nu_{t} (|t|^{-1} \,dx) = \sm_\tau^{\rm add}(dx),
2299: \quad t \in [t_0,0),
2300: \ee
2301: and the associated Laplace exponent
2302: \be
2303: \la{M.psi4}
2304: \psi(t,q) = \int_{\Ebar} y^{-2} \left(e^{-qy}-1+qy \right)
2305: \sm_{t}(dy) = \psi^{\rm add}\left(|t|^{-1},q \right).
2306: \ee
2307: %As before, the rescaling is chosen so that $\sm_t(E)=\sm_{t_0}(E)$
2308: %(if initially finite).
2309: \noindent
2310: The correspondences \qref{M.3}, \qref{M.psi3} and \qref{M.psi4} map
2311: normalized solutions for $K=xy$ on the
2312: time interval $t \in [-1,0)$ to normalized solutions with $K=x+y$ on the
2313: interval $\tau \in [0,\infty)$. The same change of variables may be
2314: applied to eternal solutions defined on $t \in (-\infty,0)$.
2315: By consequence, the results established so far for the additive kernel
2316: carry over in an obvious way for the multiplicative kernel.
2317: This yields continuous dependence of solutions on data
2318: (by Theorem~\ref{AT.cts}),
2319: the correspondence between the scaling attractor and eternal solutions
2320: (Theorem~\ref{RT.AE}), the \LK\/ representation (Theorem~\ref{RT.LK}),
2321: and how initial tails encode scaling limits (Theorem~\ref{RT.initial_tails}).
2322: For completeness, we make explicit the map
2323: from divergent \smeass\/ to eternal solutions implicit in
2324: Theorem~\ref{RT.LK}(b).
2325: \begin{thm}
2326: \label{MT.eternal}
2327: Given a divergent \smeas\/ $\sm$
2328: there is a unique eternal solution defined as follows. Let
2329: \be
2330: \label{M.psi5}
2331: \Psi(q) = \int_{\Ebar} \frac{e^{-qx}-1+qx}{x^2} \dsm(dx),
2332: \ee
2333: and $\psi(t,q)$, $t\in (-\infty,0)$ be the solution to
2334: \be
2335: \label{M.psi6}
2336: \psi - \Psi (q+ t^{-1}\psi)=0.
2337: \ee
2338: Then $\nu_t$ is determined by \qref{M.psi3} and \qref{M.psi4}.
2339: \end{thm}
2340:
2341: \section{Doeblin solutions}
2342: This section is inspired by Feller's treatment of Doeblin's universal
2343: laws and domains of partial attraction \cite[XVII.9]{Feller}.
2344: But apparently we must be content with using more words
2345: to prove fewer results. Our aim is to prove:
2346:
2347: %\begin{refthm}[\ref{RT.dense}]
2348: \begin{thm}\label{D.dense}
2349: There exists an eternal solution $\nu$ whose scaling $\omega$-limit
2350: set contains every element of the proper scaling attractor, $\attr$.
2351: \end{thm}
2352: %\end{refthm}
2353: We will show later that $\attrx$ is the closure of
2354: $\attr$ (see Corollary~\ref{EC.dense}). Therefore,
2355: Theorem~\ref{D.dense} establishes Theorem ~\ref{RT.dense}.
2356:
2357: The proof is based on suitably ``packing the tails''
2358: of the corresponding divergent \smeas. The following is
2359: adapted from Feller~\cite[XVII.9]{Feller}.
2360: Given an \smeas\ $\sm$ and $a,b>0$ we define
2361: a rescaled measure $\sm^{a,b}$ by
2362: %\sm^{a,b}(x)= a\sm(bx)$.
2363: \be
2364: \label{R.linear1}
2365: \sm^{a,b}(x)= a\sm(bx).
2366: \ee
2367: \begin{lemma}
2368: \label{D.packing}
2369: Let $\sm_k$ be a sequence of \smeass\ with
2370: \be
2371: \label{D.1}
2372: \int_{\Ebar} x^{-1} \sm_k(dx) \leq k.
2373: \ee
2374: Then there exist sequences $a_k,b_k$ such that $a_k \to 0$, $ a_kb_k
2375: \to \infty$,
2376: \be
2377: \label{D.2}
2378: \sm := \sum_{k=1}^\infty \sm_k^{a_k^{-1},b_k^{-1}}
2379: \ee
2380: defines an \smeas, and $\sm^{a_k,b_k}-\sm_k$ converges to zero.
2381: \end{lemma}
2382: The growth assumption \qref{D.1} is included only
2383: for concreteness and implies no real loss of generality. Our main
2384: purpose is to approximate divergent \smeass.
2385: \begin{lemma}
2386: \label{D.growth}
2387: Let $\dsm$ be a divergent \smeas.
2388: There exists a sequence of
2389: \smeass\ $\sm_k$ satisfying \qref{D.1} such
2390: that $\sm_k$ converges to $\dsm$.
2391: \end{lemma}
2392: \begin{proof}[Proof of Theorem~\ref{RT.dense}]
2393: {\em 1.\/} Let $\tilde\nudot\nsup$ be an arbitrary sequence of
2394: eternal solutions with
2395: corresponding divergent \smeass\ $\tilde{\dsm}\nsup$.
2396: Partition the integers into infinitely many subsequences,
2397: and choose $\sm_k \to \tilde{\dsm}\nsup$ for $k$ in the
2398: $n$-th subsequence as in Lemma~\ref{D.growth}.
2399:
2400: {\em 2.\/} Now define $a_k,b_k$ and $\sm$ as in Lemma~\ref{D.packing}, and
2401: put \[\dsm = \delta_0 + \sm. \]
2402: $\dsm$ has an atom at the origin, thus is the divergent \smeas\ for an
2403: eternal solution. By construction, $\sm^{a_k,b_k} \to \tilde{\dsm}\nsup$
2404: as $k \to \infty$ along the $n$-th subsequence.
2405: Moreover, since $a_k \to 0$, under rescaling
2406: $\delta_0^{a_k,b_k} = a_k\delta_0$ converges to zero.
2407: Thus, if we take limits
2408: along the $n$-th subsequence, $\dsm^{a_k,b_k}\to \tilde{\dsm}\nsup$ .
2409:
2410: {\em 3.\/} We now apply Theorem~\ref{RT.LK} together with \qref{R.linear3}.
2411: We have $\dsm^{a_k,b_k}=\Hmap(F^{a_k,b_k}_{t_0})$ and
2412: $F^{a_k,b_k}_{t_0}(x)=F_{T_k}(\beta_k x)$ where
2413: \begin{equation}
2414: (T_k,\beta_k) = \left\{
2415: \begin{array}{ll}
2416: (a_kb_k, b_k) &(K=2),\\
2417: (\log(a_kb_k), a_kb_k^2) &(K=x+y),\\
2418: (-(a_kb_k)^{-1}, a_kb_k^2) &(K=xy).
2419: \end{array}\right.
2420: \end{equation}
2421: %Set $T_k = a_kb_k,
2422: %\log(a_kb_k)$ and $(a_kb_k)^{-1}$ for $K=2,x+y$ and $xy$ respectively.
2423: %Similarly let $\beta_k = b_k$ for $K=2$ and $\beta_k=a_kb_k^2$ for
2424: %$K=x+y$ and $xy$.
2425: Observe that $T_k \to \Tmin$ and $\beta_k \to \infty$.
2426: We take limits along the $n$-th subsequence to
2427: obtain $F_{T_k}(\beta_k x) \to \tilde{F}\nsup(x)$ at every point of
2428: continuity. Hence, {\em given any sequence of eternal solutions $\nudot\nsup$
2429: there exists an eternal solution $\nudot$ whose scaling $\omega$-limit
2430: set contains each $\tilde{\nu}\nsup_1$}.
2431:
2432:
2433: {\em 4.\/} The space of divergent \smeass\ is separable. The
2434: \smeass\ which are concentrated at finitely many
2435: rational points (including 0) with rational weights form a
2436: countable set which is dense with respect to convergence of \smeass.
2437: By ordering these in a sequence $\tilde{\dsm}\nsup$ and using the
2438: construction above, we see that there exist eternal solutions $\nudot$
2439: such that for {\em every} eternal solution $\tilde\nudot$, $\tilde\nu_1$ is in
2440: the scaling $\omega$-limit set of $\nu$. This finishes the proof of
2441: the theorem.
2442: \end{proof}
2443:
2444: \subsection{The packing lemma}
2445: We will need to choose a sequence $c_k$ that grows so fast that
2446: \be
2447: \nn
2448: %\label{D.4}
2449: c_k \sum_{j=k+1}^\infty j c_j^{-1} \to 0.
2450: \ee
2451: The choice $c_k = \rme^{k^2}$ will do. For $j \geq 2$ we have the elementary
2452: estimate
2453: \[ j \rme^{-j^2} < \int_{j-1/2}^{j+1/2} y \rme^{-y^2} dy = \rme^{-j^2-1/4} \cosh j .
2454: \]
2455: Therefore for $k\geq 1$,
2456: \[ \rme^{k^2} \sum_{j=k+1}^\infty j \rme^{-j^2} < \rme^{k^2} \int_{k+1/2}^\infty y
2457: \rme^{-y^2} dy = \frac{\rme^{-k-1/4}}{2} \to 0. \]
2458: \begin{proof}[Proof of Lemma~\ref{D.packing}]
2459: {\em 1.\/} Fix $a_k b_k = c_k$. Then $\sm$
2460: defines an \smeas\ since
2461: \[ \int_{\Ebar} x^{-1} \sm(dx) = \sum_{k=1}^\infty c_k^{-1}
2462: \int_{\Ebar} x^{-1} \sm_k(dx) \leq \sum_{k=1}^\infty k c_k^{-1} < \infty.\]
2463:
2464: {\em 2.\/} Let $\Vp\ksup$ and $\Vp$ denote the Laplace exponents of $\sm_k$ and
2465: $\sm$ respectively. We use the definition \qref{R.linear1} and
2466: \qref{D.2} to obtain
2467: $\Vp(q) = \sum_{j=1}^\infty c_j^{-1} \Vp\jsup (qb_j)$. Observe
2468: that $\sm^{a_k,b_k} - \sm_k$ is a positive measure with Laplace
2469: exponent
2470: \[ c_k \Vp(qb_k^{-1}) - \Vp\ksup(q)= c_k\sum_{j \neq k} c_j^{-1}
2471: \Vp\jsup(q b_jb_k^{-1}). \]
2472:
2473: {\em 3.\/} To prove convergence to zero, it suffices to show that the
2474: right hand side converges to zero for every $q>0$. We first control
2475: the tail. Since $\int_{\Ebar} x^{-1} \sm_j(dx) \leq j$,
2476: \[ c_k\sum_{j=k+1}^\infty c_j^{-1}\Vp\jsup\left(q b_j b_k^{-1}\right)
2477: \leq c_k\sum_{j=k+1}^\infty j c_j^{-1} \to 0.\]
2478:
2479: {\em 4.\/} We now choose $b_k$ inductively to control the first $k-1$ terms in
2480: the range $0 \leq q \leq k$. Suppose $b_1,\ldots, b_{k-1}$
2481: have been chosen. Since $\Vp\jsup(q) \to 0$ as $q\to 0$, we choose
2482: $b_k$ so large that
2483: \[ a_k =c_k b_k^{-1} \leq \frac{1}{k}, \quad c_k\sum_{j=1}^{k-1}
2484: c_j^{-1}\Vp\jsup\left(k b_jb_k^{-1}\right) \leq \frac{1}{k}.\]
2485: \end{proof}
2486: \subsection{Proof of Lemma~\ref{D.growth}}
2487: First, suppose $\dsm$ has no atom at the origin.
2488: Since $\int_{x}^\infty y^{-1} \dsm(dy) \to 0$ as $x \to \infty$, we
2489: may choose a decreasing sequence
2490: $\veps_k$ such that $\int_{\veps_k}^\infty y^{-1} \dsm(dy) \leq k$. Let
2491: $\sm_k(dy) = \mathbf{1}_{y > \eps_k} \dsm(dy)$. Clearly, $\sm_k$
2492: satisfies both conditions of Definition~\ref{R.candef}.
2493:
2494: Next, let $\dsm =\delta_0$. In this
2495: case we choose $\sm_k(dx) = (x \log k )^{-1} \mathbf{1}_{x
2496: \geq k^{-1}}dx$. Then $\sm_k$ satisfies \qref{D.1} as
2497: \[ \int_{\Ebar} x^{-1} \sm_k(dx) =(\log k)^{-1} \int_{k^{-1}}^\infty x^{-2} dx
2498: = k(\log k)^{-1} \leq k, \]
2499: and
2500: \[ \Vp\ksup(q) = q (\log k)^{-1} \int_{qk^{-1}}^\infty \frac{1 -\rme^{-x}}{x^2}
2501: dx \to q = \Vp(q).\]
2502: The general case follows by superposition of these two special cases.
2503:
2504: \section{Scaling-periodic solutions}
2505: %[Scaling fixed points?]
2506: In this section we characterize scaling-periodic solutions
2507: and show that they are dense in the scaling attractor.
2508: That is, we prove Theorems~\ref{RT.periodic} and \ref{RT.Pdense}.
2509: \subsection{Characterization}
2510: \begin{proof}[Proof of Theorem~\ref{RT.periodic}]
2511: {\em 1.\/}
2512: Given a scaling-periodic solution, a solution satisfying \qref{R.periodic},
2513: we can scale it as in \qref{newnu}--\qref{newF} so that $t_0$ is as in
2514: \qref{R.normal3}. Then, under the map
2515: $F \mapsto F^{a,b}$ %from \qref{R.linear3},
2516: given by
2517: \begin{equation}
2518: \label{R.linear3}
2519: F^{a,b}_t(x) =\left\{ \begin{array}{ll}
2520: F_{abt}(bx) & (K=2),\\
2521: F_{t+\log(ab)}( ab^2x) & (K=x+y),\\
2522: F_{t/ab}(ab^2x) & (K=xy),
2523: \end{array} \right.
2524: \end{equation}
2525: for some $a,b>0$ we have $F_t = F^{a,b}_t$ for all $t\in[t_0,\Tmax)$.
2526: Explicitly,
2527: \begin{equation}
2528: (ab,b)= \left\{
2529: \begin{array}{ll}
2530: (t_1,\betarsc) &(K=2),\\
2531: (\rme^{t_1},\betarsc \rme^{-t_1}) &(K=x+y),\\
2532: (-t_1^{-1}, -\betarsc t_1) &(K=xy).
2533: \end{array}\right.
2534: \end{equation}
2535: %A scaling periodic solution must be eternal for
2536: %\qref{R.periodic} may be iterated infinitely many times.
2537: %If an eternal solution satisfies \qref{R.periodic} it must be a fixed
2538: %point of the map
2539: %$F \mapsto F^{a,b}$ from \qref{R.linear3}
2540: %Explicitly, we have $ab= t_1, \rme^{t_1}$, or
2541: %$-t_1^{-1}$; $b=\betarsc,\betarsc \rme^{-t_1}$, or $-\betarsc t_1$; for
2542: %$K=2,x+y$ and $xy$ respectively.
2543: Observe that $ab>1$ in all three cases.
2544: Iterating the map, we get that the solution must be (the restriction of)
2545: an eternal solution.
2546: By Theorem~\ref{RT.scale} and \qref{R.linear3}, $F=F^{a,b}$ is equivalent to
2547: $\dsm=\dsm^{a,b}$, that is,
2548: \be
2549: \label{P.1}
2550: \dsm(x) = a\dsm(bx), \quad x >0.
2551: \ee
2552: Without loss of generality we may suppose $b > 1$ since \qref{P.1} is
2553: equivalent to $a^{-1}\dsm(b^{-1} x) = \dsm(x)$.
2554:
2555: {\em 2.\/} Equation \qref{P.1} implies $\dsm(0_+)=a\dsm(0_+)$. If $\dsm$ has
2556: an atom at the origin, this forces $a=1$. Then $\dsm(x)=\dsm(bx)$ for every
2557: $x>0$, and since $b>1$ and $\dsm(x)$ is non-decreasing,
2558: it follows $\dsm(x)=c=\dsm(0_+)$ for all $x>0$.
2559: Therefore, if $\dsm$ has an atom at the origin, then $\dsm=c \delta_0$
2560: for some $c>0$.
2561:
2562: {\em 3.\/} Suppose $\dsm$ does not have an atom at the origin.
2563: We iterate \qref{P.1} to find that
2564: \[
2565: \int_1^{b_-}\dsm(dx)=a^j\int_{b^j}^{b_-^{j+1}}\dsm (dx), \quad
2566: \int_1^{b_-}\frac{\dsm(dx)}{x} = (ab)^j\int_{b^j}^{b_-^{j+1}}\frac{\dsm (dx)}{x}.
2567: \]
2568: In order that $\dsm$ is an \smeas\ we require
2569: \[ \int_{E} (1\wedge x^{-1})\dsm(dx) = \sum_{j<0} a^{-j}
2570: \int_1^{b_-} \dsm(dx) + \sum_{j \geq 0} (ab)^{-j}
2571: \int_1^{b_-}x^{-1}\dsm(dx) <\infty.
2572: \]
2573: Thus, $a <1$ and $ab>1$. Given $x>0$ let $k= \max\{j: b^j \leq
2574: x\}$. A similar calculation yields
2575: \[
2576: \dsm(x) = \sum_{j < k}a^{-j} \int_1^{b_-}\dsm(dx) + a^{-k} \int_1^{b^{-k}x_-} \dsm(dx). \]
2577: This shows $\dsm$ is determined by its restriction to $[1,b)$.
2578:
2579: {\em 4.\/} Conversely, suppose $\dsm=\dsm^{a,b}$ and (i) or (ii) hold.
2580: Notice that $\dsm$ is automatically divergent since it either has an
2581: atom at the origin or
2582: \[ \int_E x^{-1} \dsm(dx) = \int_1^{b_-} x^{-1}\dsm(dx) \sum_{j
2583: =-\infty}^\infty (ab)^{-j}= \infty.\]
2584: Thus, it determines an eternal solution, which by
2585: \qref{R.linear3} satisfies $F =F^{a,b}$.
2586: \end{proof}
2587:
2588: \subsection{Self-similar solutions}
2589: As remarked in Section~\ref{RS.chaos}, the case (i)
2590: is simple but important. The associated divergent \smeas\ is
2591: scale-invariant for every $b>1$ and the scaling-periodic solutions
2592: are the classical self-similar solutions with exponential tails.
2593: %\begin{equation}\label{C.sss1}
2594: %\nu_t(dx)=c^{-1}t^{-2}\rme^{-x/c t}\,dx.
2595: %\end{equation}
2596: If a scaling-periodic solution satisfies \qref{R.periodic} for
2597: every $t_1 >t_0$ (with changing $\betarsc$), it follows that
2598: for some fixed $a$ and $b$, $\dsm(x)=a^{r}\dsm(b^rx)$ for all rational and
2599: hence all real $r$. The fundamental rigidity lemma for scaling
2600: limits~\cite[VIII.8]{Feller} then implies $\dsm(x)= C_\theta x^\theta$
2601: for some $\theta \in \R$. The finiteness condition $\int_E (1 \wedge
2602: x^{-1}) \dsm (dx) < \infty$ then implies $\theta =1-\rho, \rho \in
2603: (0,1]$. If $\rho=1$, $\dsm$ is an atom at the origin corresponding to
2604: (i) above. The self-similar profiles and their domains of attraction
2605: are discussed further in Section~\ref{sec:domains}.
2606:
2607:
2608: \subsection{Density of scaling-periodic solutions}
2609: %
2610: % Let $\nu$ be an eternal solution to Smoluchowski's equation with
2611: % kernel $K=2$, $x+y$ or $xy$.
2612: % Let $a_n \to 0,b_n \to \infty $ be sequences such that $a_nb_n^{1/2}
2613: % \to 0$ and $a_nb_n\to\infty$.
2614: % Then there exist scaling-periodic solutions $\nu\nsup$ with scale parameters
2615: % $(a_n,b_n)$ such that $\nu\nsup_t$ converges weakly to $\nu_t$ for
2616: % every $t \in (\Tmin,\Tmax)$.
2617: To prove Theorem~\ref{RT.Pdense} and establish density of
2618: scaling-periodic solutions in the full scaling attractor $\attrx$,
2619: it will suffice to prove such solutions are dense in the proper
2620: scaling attractor $\attr$ (see Corollary~\ref{EC.dense}).
2621: %\begin{refthm}[\ref{RT.Pdense}]\label{CT.Pdense}
2622: %Let $\hat F\in\attrx$
2623: %be arbitrary. Let $a_n \downarrow 0,b_n \uparrow \infty $ be
2624: %sequences such that $a_nb_n^{1/2} \to 0$ and $a_nb_n\to\infty$. Then
2625: %there exist scaling-periodic solutions $\nu\nsup$ with scale
2626: %parameters $(a_n,b_n)$ such that $x^\gamma\nu\nsup_{t_0}(dx)$
2627: %converges weakly to $\hat F$ as $n\to\infty$.
2628: %\end{refthm}
2629: \begin{thm}
2630: \label{PT.Pdense}
2631: Scaling-periodic solutions are dense in $\attr$.
2632: \end{thm}
2633: \begin{proof}%[Proof of Theorem~\ref{RT.Pdense}]
2634: {\em 1.\/} Let $\hat F\in\attrx$ be arbitrary. Let $a_n \downarrow
2635: 0,b_n \uparrow \infty $ be sequences such that $a_nb_n^{1/2} \to 0$
2636: and $a_nb_n\to\infty$. We claim that there exist scaling-periodic
2637: solutions $\nu\nsup$ with scale parameters $(a_n,b_n)$ such that
2638: $F\nsup_{t_0}(dx)=x^\gamma\nu\nsup_{t_0}(dx)$ converges weakly to $\hat F$ as
2639: $n\to\infty$.
2640: Let $\dsm$ denote the divergent \smeas\ associated with $\nu$.
2641: By Theorems~\ref{RT.LK} and \ref{RT.scale} it suffices to construct
2642: divergent \smeass\ $\dsm\nsup$ such that
2643: $\dsm\nsup=a_n \dsm\nsup(b_n\,\cdot)$ and $\dsm\nsup$
2644: converges to $\dsm$.
2645:
2646: {\em 2.\/} Consider first the case where $\dsm$ has no
2647: atom at the origin. In this case we define $\dsm\nsup$ to be the
2648: scaling-invariant extension
2649: of $\dsm$ restricted to the interval $I_n:=[b_n^{-1/2},b_n^{1/2})$.
2650: Then for any $x>0$ that is a point of continuity of $\dsm$, for
2651: $n$ large we have $x\in I_n$ and
2652: \[
2653: \dsm\nsup(x)= \int_{b_n\inv}^{x} \dsm(dx) + \sum_{j<0} (a_n)^{-j}
2654: \int_{b_n\inv}^{1} \dsm(dx)
2655: \to \dsm(x)
2656: \]
2657: as $n\to\infty$. Moreover,
2658: \[
2659: \int_x^\infty \frac{\dsm\nsup(dy)}{y} =
2660: \int_x^{b_n} \frac{\dsm(dy)}{y} +
2661: \sum_{j\ge1} (a_nb_n)^{-j} \int_1^{b_n}\frac{\dsm(dy)}{y}
2662: \to \int_x^\infty \frac{\dsm(dy)}{y}.
2663: \]
2664: This establishes the desired convergence of \smeass\/.
2665:
2666: {\em 3.\/} In case $\dsm=\delta_0$, we let $\dsm\nsup$ be a sum
2667: of delta masses $\delta\nsup_j, j \in \Z$ concentrated at points
2668: $\beta_j = b_n^{j-1/2}$, so
2669: that $\dsm\nsup = \sum_{j} (a_n b_n)^j \delta\nsup_j$.
2670: Observe that there is no mass in $(b_n^{-1/2}, b_n^{1/2})$;
2671: thus for any $x>0$, for $n$ large we have
2672: \[
2673: \dsm\nsup(x) = \sum_{j\le0} (a_nb_n)^j = \frac{1}{1-a_nb_n} \to 1,
2674: \]
2675: and
2676: \[
2677: \int_x^\infty y\inv \dsm\nsup(dy) = b_n^{1/2} \sum_{j>0} a_n^j =
2678: \frac{a_n b_n^{1/2}}{1-a_n} \to 0.
2679: \]
2680: Hence the \smeass\/ $\dsm\nsup$ converge to $\delta_0$.
2681:
2682: {\em 4.\/} In the general case, we simply superpose the separate constructions.
2683: Observe that the restriction $a_n b_n^{1/2}\to0$ is only needed in the
2684: critical case when $\dsm$ has an atom at the origin.
2685: \end{proof}
2686:
2687: \section{Extended solutions, with dust and gel}
2688: \label{S.extend}
2689: \subsection{Extended solutions}
2690: A proper solution to
2691: Smoluchowski's equation satisfies $\int_E x^\gamma\nu_t(dx)=
2692: m_\gamma(t)$ with $m_\gamma(t)$ normalized as in \qref{R.mgamma}.
2693: However, a sequence of proper solutions may lose mass in the
2694: limit. We append atoms at $0$ and $\infty$ to account for these
2695: defects, considering measures on $\Ebar=[0,\infty]$ of the form
2696: \be
2697: \label{E.ext1}
2698: \mu_t = \dust(t)\delta_0 + x^\gamma \nu_t + \gel(t)
2699: \delta_\infty,
2700: \ee
2701: where $\nu_t$ is a size-distribution measure on $E$. We call the atoms $\dust$
2702: and $\gel$ the dust and gel respectively. An associated probability
2703: measure on $\Ebar$ is defined as in \qref{R.Fp}, by
2704: \be
2705: \label{E.ext2}
2706: \Fbar_t(dx)= \frac{\mu_t(dx)}{\mu_t(\Ebar)}
2707: = \frac{\dust(t)\delta_0(dx) + x^\gamma \nu_t(dx) + \gel(t) \delta_\infty(dx)}
2708: {\dust(t)+\int_E x^\gamma\nu_t(dx)+ \gel(t)} .
2709: %\int_{[0,x]} y^\gamma \nubar_t(dy) \left\slash
2710: %\int_{\Ebar} y^\gamma\nubar_t(dy). \right.
2711: \ee
2712: The \smeas\/ associated to a solution in \qref{R.cansoln1} is
2713: replaced by the \barsmeas
2714: \be
2715: \label{E.ext3}
2716: \esm_t(dx) = x^{\gamma+1} \nu_t(\lambda(t) \,dx) + \ggel(t)
2717: \delta_\infty(dx), \qquad \ggel(t) = \frac{\gel(t)}{\lambda(t)^\gamma}.
2718: \ee
2719: The measures $\mu_t$ define Laplace
2720: exponents by evident modification of equation
2721: \qref{C.vpdef} for $K=2$, namely
2722: \be
2723: \vp(t,q)=\int_{\Ebar}(1-\rme^{-qx})\mu_t(dx)
2724: = a_\infty(t)+\int_E(1-\rme^{-qx})\nu_t(dx),
2725: %\quad(K=2),
2726: \ee
2727: and of \qref{A.vpdef} and
2728: \qref{A.psi2} for $K=x+y$ and \qref{M.vpdef} for $K=xy$, both yielding
2729: \be
2730: \vp(t,q)=\int_{\Ebar}\frac{1-\rme^{-qx}}x\mu_t(dx)
2731: = a_0(t)q+\int_E(1-\rme^{-qx})\nu_t(dx).
2732: %\quad(K=x+y)
2733: \ee
2734: %\marginpar{Formulas??}
2735: The evolution
2736: equations for these exponents remain
2737: \be
2738: \label{E.ext4}
2739: \partial_t \vp = \left\{
2740: \begin{array}{rl} -\vp^2, & (K=2),\\
2741: \vp \partial_q \vp -\vp, & (K=x+y),
2742: \\ \vp \partial_q \vp, & (K=xy).
2743: \end{array} \right.
2744: \ee
2745: This motivates the following definition.
2746: \begin{defn}
2747: \label{ED.ext}
2748: %Given a probability measure $\specialF$ on $\Ebar$,
2749: A family of triples $(\nu_t,a_0(t),a_\infty(t))$, $t \in [t_0,\Tmax)$,
2750: defines an {\em extended solution} to Smoluchowski's equation
2751: for the kernels $K=2$, $x+y$ and $xy$ with initial data
2752: $(\hat\nu,\hat a_0,\hat a_\infty)$, if
2753: \begin{enumerate}
2754: \item[(a)] The measures $\mu_t$ in \qref{E.ext1} satisfy $\mu_t(\Ebar)=
2755: m_\gamma(t)$ with $m_\gamma(t)$ as in \qref{R.mgamma},
2756: for $t \in [t_0,\Tmax)$.
2757: %$x^\gamma \nubar_t (\cdot)= m_\gamma(t)\Fbar_t(\cdot)$
2758: %with $x^\gamma\nubar_t(\Ebar)= m_\gamma(t)$,
2759: %$t \in [t_0,\Tmax)$ with $m_\gamma(t)$ as in \qref{R.mgamma}.
2760: \item[(b)] \qref{E.ext4} holds for $q>0$ and $t \in (t_0,\Tmax)$.
2761: %with $x^\gamma\nubar_t=m_\gamma(t)\Fbar_t$.
2762: \item[(c)] $\mu_t \to \hat\mu
2763: =\hat a_0\delta_0+x^\gamma\hat\nu+\hat a_\infty\delta_\infty$
2764: weakly as $t \dnto t_0$.
2765: \end{enumerate}
2766: \end{defn}
2767: %A family of probability measures $\Fbar_t$,
2768: Due to the normalization in (a), we regard extended solutions
2769: as determined by the associated probability distributions
2770: $\Fbar$ in \qref{E.ext2}.
2771: We will usually denote an extended solution with values
2772: $(\nu_t,a_0(t),a_\infty(t))$ simply by $\nu$.
2773:
2774: Extended solutions provide the correct compactification in light of
2775: the following theorem. Since every proper solution also defines an
2776: extended solution, the
2777: theorem applies in particular to sequences of proper solutions.
2778: \begin{thm}
2779: \label{ET.const-compact}
2780: Let $\extsol_t\nsup$, $t\in[t_0,\Tmax)$, be probability measures
2781: associated with a sequence of extended solutions $\nu\nsup$.
2782: Then there exists a sequence $n_j\to\infty$ and
2783: probability measures $\extsol_t$ associated
2784: with an extended solution $\nu$, such that
2785: $\extsol\nsupj_t$ converges weakly to $\extsol_t$
2786: for every $t \in [t_0,\Tmax)$.
2787: \end{thm}
2788: \begin{proof}
2789: Consider the sequence of probability measures $\extsol\nsup_{t_0}$ on $\Ebar$.
2790: Then there exists a
2791: subsequence $n_j$ and a probability measure $\specialF_0$ such that
2792: $\extsol_{t_0}\nsupj$ converges weakly to $\specialF_0$.
2793: %and $\extsol_{t_0}(\Ebar)=1$.
2794: We use $\specialF_0$ to determine initial data
2795: to define an extended solution $\nu$ for $t \in
2796: [t_0,\Tmax)$. Continuous dependence on initial data as in
2797: Theorem~\ref{ET.cts} below implies the weak convergence of
2798: $\extsol\nsupj_t$ to $\extsol_t$ for every $t \in [t_0,\Tmax)$.
2799: \end{proof}
2800:
2801: We state the following result without proof, as it is an easy consequence of
2802: Definition~\ref{ED.ext}, and Theorems~\ref{CT.cts}
2803: and~\ref{AT.cts}. The notion of extended
2804: solution allows us to simplify matters, as it is no longer necessary to
2805: assume that $\hat\mu_0(\Ebar)=m_\gamma(t_0)$ , or
2806: $\hat\mu(\Ebar) =m_\gamma(t)$ as in parts (a) and (b) of
2807: Theorem~\ref{CT.cts} and ~\ref{AT.cts}.
2808: \begin{thm}\label{ET.cts}
2809: (Continuous dependence on data.)
2810: For Smoluchowski's equation with kernels $K=2$, $x+y$ or $xy$, let
2811: $t_0 \in (\Tmin,\Tmax)$ and let $\extsol\nsup$ determine a sequence of
2812: extended solutions defined for $t\in I=[t_0,\Tmax)$.
2813: \begin{enumerate}
2814: \item[(a)] If $\extsol\nsup_{t_0}$ converges weakly to a
2815: measure $\specialF_0$, then for every $t\in I$,
2816: $\extsol\nsup_t$ converges weakly to $\extsol_t$,
2817: associated with the time-$t$
2818: extended solution with initial data determined by $\extsol_{t_0}=\specialF_0$.
2819: \item[(b)] For any $t\in I$, if $\extsol\nsup_{t}$ converges
2820: weakly to a measure $\specialF$, then
2821: $\extsol\nsup_{t_0}$ converges weakly to a probability measure $\specialF_0$
2822: %$F_0$ with $\extsol_0(\Ebar)=1$,
2823: and $\specialF=\extsol_t$, associated with the
2824: time-$t$ solution with initial data determined by
2825: $\extsol_{t_0} =\specialF_0$.
2826: \end{enumerate}
2827: \end{thm}
2828:
2829: \subsection{Transformation to proper solutions}
2830: Clusters of ``zero'' or ``infinite'' size interact with other clusters
2831: in simple ways.
2832: The invariances of the evolution equations \qref{E.ext4} allow us to
2833: relate all extended solutions (except pure dust and gel)
2834: to proper solutions. Let us consider the constant and additive kernels in turn.
2835: \subsubsection{The constant kernel}
2836: The dust and gel are recovered as limits as $q \to 0$ and $\infty$
2837: respectively:
2838: \be
2839: \label{E.c_gel}
2840: \gel(t) = \vp(t,0^+), \quad \dust(t) = \mu_t(\Ebar)- \vp(t,\infty^-).
2841: \ee
2842: Since $\mu_t(\Ebar)=t\inv$,
2843: we take limits in \qref{E.ext4} to see that the dust and gel satisfy
2844: \be
2845: \label{E.c_gel2}
2846: \frac{d\gel}{dt} = -\gel^2, \quad \frac{d{\left(t^{-1}-
2847: \dust\right)}}{dt} = -
2848: (t^{-1}-\dust )^2.
2849: \ee
2850: The extended solution corresponds to purely dust and gel when
2851: $\mu_t(E)=0$, so that $\dust(t)+\gel(t)=t\inv$.
2852: We may exploit \qref{E.ext4} to show that every extended solution
2853: that is not purely dust and gel is in
2854: correspondence with a proper solution after a simple change of scale. Suppose
2855: $\vp(t,q)$ is the Laplace exponent of an extended solution.
2856: If $\gel(t_0)>0$ let
2857: %\marginpar{include calculations?}
2858: \be \la{e:ext_rsc1}
2859: \hvp(\htt,q) = \alpha(t)^{-2} \left(\vp(t,q)-\gel(t)\right),
2860: \ee
2861: where
2862: \be
2863: \htt\inv = \alpha(t)^{-2}(t\inv-\dust(t)-\gel(t)),
2864: \qquad
2865: \alpha(t) = \frac{\gel(t)}{\gel(t_0)}.
2866: \ee
2867: Then we find $\hvp(\htt,0^+)=0$, $\hvp(\htt,\infty^-)=\htt\inv$,
2868: and $\partial_{\htt}\hvp= -\hvp^2$.
2869: Thus, $\hvp(\htt,q)$ is the Laplace exponent of a proper solution
2870: defined on $[\htt_0,\infty)$.
2871:
2872: For vanishing gel ($\gel(t_0)\to0$) the transformation above
2873: simplifies, yielding $\alpha=1$, $\htt-\htt_0=t-t_0$, $\hvp=\vp$.
2874: Zero-size clusters combine trivially with other clusters,
2875: so the presence of dust only shifts time in accord with our
2876: normalization of total number.
2877: Observe that if gel is present ($\gel(t_0)>0$),
2878: the probability of being gel approaches one ($\gel(t)/\mu_t(\Ebar)\to1$)
2879: and the relative distribution of finite-size clusters approaches
2880: a state reached by the proper solution at a finite time;
2881: we have $\htt\to \htt_0+ 1/\gel(t_0)$ as $t \to \infty$.
2882:
2883: %\ba
2884: %\label{eq:ext_rsc1}
2885: %&& \hvp(\htt,q) = \alpha(t)^{-2}\left(\vp(t,q)- \gel(t) \right),
2886: %\quad
2887: %\alpha(t) = \frac{\gel(t)}{\gel(t_0)}, \\
2888: %\nn %\label{eq:ext_rsc2}
2889: %&& \htt -\htt_0 = \alpha(t) \left(t-t_0 \right), \quad \htt_0=
2890: %\frac{t_0}{1 - t_0\left(\dust(t_0)+\gel(t_0)\right)}.
2891: %\ea
2892: %Then we have $\partial_{\htt}\hvp= -\hvp^2$
2893: %and $\hvp(\infty^-,\htt)= \htt^{-1}$.
2894:
2895:
2896: \subsubsection{The additive kernel}
2897: In this case, $\D_q\vp(t,q)=\int_{\Ebar}\rme^{-qx}\mu_t(dx)$ and
2898: $\mu_t(\Ebar)=1$, so the dust and gel are given by
2899: \ba
2900: \label{eq:ext_add2}
2901: \dust (t) = \partial_q \vp(t,\infty), \quad \gel (t) = 1- \partial_q
2902: \vp(t,0^+).
2903: \ea
2904: The similarity with the constant kernel is clear if we
2905: use the time scale $s=e^t$ and the Laplace
2906: exponent $\psi(s,q)$ defined in \qref{A.psi1} and \qref{A.psi2}.
2907: Let
2908: \be
2909: \label{E.ab1}
2910: \bdust(s) = \frac{1}{s} - \partial_q\psi(s,\infty) =\frac{\dust(t)}{s}, \quad
2911: \bgel(s)=\partial_q \psi(s,0) = \frac{\gel(t)}{s}.
2912: \ee
2913: We then take limits in \qref{A.imp} to see that
2914: \be
2915: \label{E.ab2}
2916: \frac{d(s^{-1}-\bdust)}{ds} = -(s^{-1}-\bdust)^2,
2917: \quad
2918: \frac{d\bgel}{ds} = -\bgel^2,
2919: \ee
2920: which is equivalent to the following closed
2921: equations for the dust and gel:
2922: \ba
2923: \label{eq:ext_add4}
2924: \frac{d\dust}{dt} = -\dust (1-\dust), \quad \frac{d\gel}{dt} =
2925: \gel(1-\gel).
2926: \ea
2927: The extended solution is purely dust and gel when $\dust(t)+\gel(t)=1$.
2928: If it is not, we exploit the invariances of the inviscid Burgers equation
2929: \qref{A.psi4} to reduce extended solutions to proper solutions
2930: by a change of scale. Given initial data $\psi_0$ with $\partial_q
2931: \psi_0(0) = \bgel(s_0) \geq 0$ and $\partial_q \psi_0(\infty) =
2932: s_0^{-1}-\bdust(s_0) > \bgel(s_0)>0$, we define a proper solution via
2933: the change of variables
2934: \be \la{e:ext_add5}
2935: \hpsi(\hs,\hq) = \alpha(s)^{-1}\left(\psi(s,q) -\bgel(s) q\right),
2936: \ee
2937: where
2938: \[
2939: \hs\inv = \alpha(s)^{-2}\left(s\inv-\bdust(s)-\bgel(s)\right), \quad
2940: \hq=\alpha(s)q, \quad
2941: \alpha(s)= \frac{\bgel(s)}{\bgel(s_0)}.
2942: \]
2943: This ensures $\partial_{\hq} \hpsi(s,0) = 0$,
2944: $\partial_{\hq} \hpsi(s,\infty)=s^{-1}$, and
2945: $\partial_{\hs} \hpsi + \hpsi \partial_{\hq} \hpsi=0$
2946: for $\hs > \hs_0$.
2947:
2948: %\ba
2949: %\label{eq:ext_add5}
2950: %&& \hpsi(\hs,\hq) = \alpha(s)^{-1}\left(\psi(s,q) -\bgel(s) q\right), \quad
2951: %\alpha(s)= \frac{\bgel(s)}{\bgel(s_0)}\\
2952: %\nn %\label{eq:ext_add6}
2953: %&&
2954: % \hq = \alpha(s) q , \quad \hs-\hs_0 = \alpha(s) (s-s_0), \quad \hs_0 =
2955: %\frac{s_0}{1-s_0\left(\bdust(s_0)+\bgel(s_0)\right)}.
2956: %\ea
2957: %This ensures $\partial_{\hs} \hpsi +
2958: %\hpsi \partial_{\hq} \hpsi=0$, $\hs > \hs_0$, with initial data
2959: %such that $\hpsi(\hs_0,0)=0$ and
2960: %$\hpsi(\hs_0,\infty)=s_0^{-1}$. It then follows
2961: %that $\partial_{\hq} \hpsi(s,0) = 0$ and
2962: %$\partial_{\hq} \hpsi(s,\infty)=s^{-1}$ for every $\hs > \hs_0$.
2963:
2964:
2965: \subsection{\LK\/ representation}
2966: % of extended eternal solutions}
2967: %An extended eternal solution is defined as in \ref{R.eternal}.
2968: \begin{defn}
2969: \label{ED.eternal}
2970: An extended solution to Smoluchowski's equation that is defined for
2971: all $t \in (\Tmin,\Tmax)$ is called an {\em eternal extended
2972: solution}.
2973: \end{defn}
2974: The following representation theorem is the completion of
2975: Theorem~\ref{RT.LK}. We establish a bijection between the set of
2976: eternal extended solutions and the space $\Mcanx$ consisting of
2977: all \barsmeass\/ together with a point at infinity.
2978: The point at infinity corresponds to all
2979: measures such that $\int_{\Ebar}(1\wedge x^{-1})
2980: \dsm(dx)=\infty$. These measures give rise to the (unique)
2981: Laplace exponents $\Phi(q)=\Psi(q)=\infty$, $q >0$. We say a sequence
2982: of \barsmeass\/ converges to the point at infinity if $\Vp\nsup(q)
2983: \to \infty$, $q>0$ for the associated Laplace exponents.
2984: This special case corresponds to
2985: the eternal extended solution that is pure gel. It is the counterpoint
2986: to the Laplace exponents $\Phi(q)=\Psi(q)=0$, $q>0$ which generate
2987: the eternal extended solution that is pure dust.
2988: %
2989: %It is worth remarking that in
2990: % the theorem below
2991: %involving the \barsmeass\/ defined in \qref{E.ext3}. This completion
2992: %involves
2993: %We now consider the space of all \sbar
2994: % Since the proof of the \LK\/ formula relies only on the explicit
2995: % solution formulas and the topology of \barsmeass\/ we obtain the
2996: % following completion of
2997: % \
2998: \begin{thm}
2999: \label{ET.LK}
3000: \begin{enumerate}
3001: \item[(a)]Let $\nu$ be an eternal extended
3002: solution for Smoluchowski's equation with
3003: $K=2,x+y$ or $xy$. If $\nu$ is not pure gel, there is an \barsmeas\ $\dsm$
3004: such that $\esm_t$ converges to $\dsm$ as $t \dnto \Tmin$. If $\nu$
3005: is pure gel, $\esm_t$ converges to the point at infinity in $\Mcanx$.
3006: \item[(b)] Conversely, for every \barsmeas\ $\dsm$,
3007: there is a unique eternal extended solution $\nu$
3008: such that $\esm_t$ converges to $\dsm$ as $t \dnto \Tmin$. The
3009: point at infinity generates the extended solution $\nu$ that is pure gel.
3010: \item[(c)]
3011: Let $\Hmapx\colon\attrx\to\Mcanx$
3012: map the (full) scaling attractor $\attrx$
3013: to the set $\Mcanx$ of \barsmeass\
3014: by $\Hmapx(\specialF)=\dsm$, where $\dsm$ is the \barsmeas\
3015: associated to the eternal extended solution $\nu$ such that
3016: $\specialF=\Fbar_{t_0}$ with $t_0$ as in
3017: \qref{R.normal2}. Then $\Hmapx$ is a bi-continuous bijection.
3018: Moreover, $\Hmap\colon\attr \to \Mcan$ is the restriction of $\Hmapx$
3019: to $\attr$.
3020: \end{enumerate}
3021: \end{thm}
3022: The map $\Hmapx$ is defined in terms of Laplace exponents by the same
3023: formulas as for proper solutions: \qref{C.eternal} for $K=2$,
3024: \qref{A.imp} for $K=x+y$, and \qref{M.psi6} for $K=xy$.
3025: Parts (a) and (b) of the theorem are then proven just as in
3026: Theorems~\ref{CT.eternal} and \ref{AT.eternal}. The proof here is
3027: simpler, since we no longer need verify the divergence conditions on
3028: the \barsmeas\/. The proof of part (c) relies on two separate
3029: arguments. It is easy
3030: to show as in Theorems~\ref{CT.Mconv2} and \ref{AT.Mconv2} that the
3031: map $H \mapsto \Fbar_{t_0}$ is a bi-continuous bijection.
3032: However, we must also identify such $\Fbar_{t_0}$ as belonging
3033: to the attractor. Here the arguments deviate
3034: slightly from those of Section~\ref{sec:const-ker} and
3035: ~\ref{sec:add}. We use parts (a) and (b) of Theorem~\ref{ET.LK} in the
3036: proof of part (c) via the following intermediate theorem.
3037: % Extended eternal solutions are in correspondence with elements of the
3038: % (full) scaling attractor defined in \ref{R.full}.
3039: \begin{thm}
3040: \label{ET.AE}
3041: \begin{enumerate}
3042: \item[(a)]
3043: The scaling attractor $\attrx$ is invariant: If $\nu$ is an extended solution
3044: of Smoluchowski's equation,
3045: and $\Fbar_t\in\attrx$ for some $t$, then $\nu$ is eternal and
3046: $\Fbar_t\in\attrx$ for all $t\in(\Tmin,\Tmax)$.
3047: \item[(b)]
3048: A probability measure $\specialF$ on $\Ebar$ belongs to
3049: $\attrx$ if and only if $\specialF= \Fbar_{t_0}$
3050: for some extended eternal solution $\nu$.
3051: \end{enumerate}
3052: \end{thm}
3053: \begin{proof}The proof differs from earlier arguments only in the
3054: first part of Theorems~\ref{CT.attr} and~\ref{AT.attractor} (the
3055: assertion that $\Fbar_{t_0} \in
3056: \attrx$ if $\nu$ is eternal). In order to prove this, let us
3057: suppose $\nu$ is an
3058: extended eternal solution with associated \barsmeas\/ $\bar\dsm =
3059: (\dsm, g_\infty)$ where $\dsm$ is an \smeas\/ and $g_\infty$ is the
3060: charge of $y\inv H(dy)$ at $\infty$.
3061: To show that $\specialF :=\Fbar_{t_0}$ is in
3062: the scaling attractor, we must find $T_n\upto \Tmax,\beta_n \to \infty$ and
3063: a sequence of {\em proper\/} solutions such that
3064: $F\nsup_{T_n}(\beta_nx) \to \specialF(x)$ at points of continuity. We
3065: use the \LK\/ formula to find such solutions. We approximate $\bar{H}$ by
3066: the sequence of {\em divergent\/} \smeass\/ $\dsm\nsup = n^{-1}
3067: \delta_0 + \dsm + g_\infty n\delta_n$. It follows that for the corresponding
3068: (proper) eternal solutions $\nu\nsup$, the probability measures
3069: $\Fbar_t\nsup$ converge to $\Fbar_t$ for every $t> \Tmin$.
3070: Given any sequence $T_n \upto \Tmax, \beta_n \to \infty$
3071: we consider a sequence of rescaled solutions determined as in
3072: \qref{newF}, by
3073: \[ \tilde{F}\nsup_t(x) = \left\{ \begin{array}{rl}
3074: F\nsup_{t/T_n}(\beta_n^{-1} x), & (K=2), \\
3075: F\nsup_{t - T_n} (\beta_n^{-1} x), & (K=x+y), \\
3076: F\nsup_{t/|T_n|}(\beta_n^{-1}(x)), & (K=xy). \end{array} \right. \]
3077: We then have $\tilde{F}\nsup_{T_n}(\beta_nx) = F\nsup_{t_0}(x)
3078: \to \specialF(x)$ at all points of continuity.
3079:
3080: The converse implication and part (a) are proven exactly as in
3081: Theorems~\ref{CT.attr} and~\ref{AT.attractor} and the sequel.
3082: \end{proof}
3083: This also proves a property alluded to several times before.
3084: \begin{cor}
3085: \label{EC.dense}
3086: $\attrx$ is the closure of $\attr$.
3087: \end{cor}
3088: \begin{proof}
3089: If $\specialF \in \attrx$ has \barsmeas\/ $\bar\dsm$, we approximate
3090: $\bar\dsm$ by a sequence of divergent \smeass\/ as above.
3091: \end{proof}
3092:
3093: \subsection{Scaling limits and initial tails}
3094: We now state the natural extension of Theorem~\ref{RT.initial_tails}
3095: to eternal extended solutions. The proof is almost identical to that of
3096: Theorem~\ref{RT.initial_tails} except that we no longer need verify
3097: divergence of the \smeas\/.
3098: \begin{thm}
3099: \label{ET.code}
3100: Let $\specialF\in\attrx$
3101: %so $\hat F(dx)=x^\gamma\nu_{t_0}(dx)$ where $\nu$ is an eternal solution,
3102: with associated \barsmeas\ $H$.
3103: Let $\nu\nsup$ be any sequence of proper solutions defined for $t\in
3104: [t_0,\Tmax)$,
3105: with associated initial \smeass\ given by
3106: $\sm\nsup(dx)=x^{\gamma+1}\nu\nsup_{t_0}(dx)$.
3107: Let $T_n \to \Tmax$, $\beta_n\to\infty$.
3108: Then the following are equivalent:
3109: \bit
3110: \item[(i)] ${F}\nsup_{T_n}(\beta_n x) \to \specialF(x)$
3111: as $n\to\infty$, at every point of continuity.
3112: \item[(ii)] The rescaled initial \smeass\ $\tilde\sm\nsup$ defined by
3113: \qref{R.initG} converge to the \barsmeas\/ $\dsm$ as $n \to \infty$.
3114: % \begin{equation}\la{R.initG}
3115: % \tilde\sm\nsup(x) = \left\{
3116: % \begin{array}{ll}
3117: % \beta_n^{-1}T_n \,\sm\nsup(\beta_nx)
3118: % %= T_n\beta_n^{-1}\int_{[0,\beta_nx]} y\nu\nsup_{t_0}(dy)
3119: % &(K=2),\\[6pt]
3120: % \beta_n\inv \rme^{2T_n}\, \sm\nsup(\rme^{-T_n}\beta_nx)
3121: % %= \beta_n\inv\rme^{2T_n} \int_0^{x\beta_n\exp(-T_n)}y^2\nu\nsup_0(dy)
3122: % &(K=x+y), \\[6pt]
3123: % \beta_n\inv |T_n|^{-2} \,\sm\nsup(|T_n|\beta_n x)
3124: % &(K=xy),
3125: % \end{array} \right.
3126: % \end{equation}
3127: % have the property that
3128: % $\tilde\sm\nsup$ converges to $\dsm$ as $n \to \infty$.
3129: \eit
3130: \end{thm}
3131:
3132: \section{Initial tails and ultimate scaling dynamics}
3133: In this section, we present two applications of the principle that
3134: ultimate scaling dynamics are encoded in the initial tails (as
3135: formalized in theorems~\ref{RT.initial_tails} and~\ref{ET.code}).
3136: The first is a proof of the shadowing theorem~\ref{RT.shad}. The second is a
3137: streamlined proof of the classification of domains of
3138: attraction in~\cite{MP1} that avoids the use of Karamata's Tauberian theorem.
3139:
3140: \subsection{Initial tails and shadowing}
3141: \begin{proof}[Proof of Theorem~\ref{RT.shad}]
3142: {\em 1.\/} As in section~2, we let $\dist(\cdot,\cdot)$ denote any metric on
3143: $\bar\PP$ which induces the weak topology.
3144: Suppose that for Smoluchowski's equation with kernel $K=2$, $x+y$ or
3145: $xy$, $\nu$ and $\bar\nu$ are two solutions defined on $[t_0,\Tmax)$,
3146: and make the assumptions stated in the theorem.
3147: Suppose that \qref{R.dto0} fails, i.e., that
3148: \begin{equation}\la{d.notdto0}
3149: \dist( F_t(b(t)\,dx),\bar F_{\bar t}(\bar b(t)\,dx)) \not\to 0 \qquad
3150: \mbox{as $t\to\Tmax$.}
3151: \end{equation}
3152: Then since $\bar\PP$ is compact, by passing to subsequences we can find
3153: sequences $T_n\uparrow\Tmax$ and $\beta_n=b(T_n)$ and {\em different}
3154: probability measures $\hat F$, $\check F\in \bar\PP$, such that as
3155: $n\to\infty$
3156: we have
3157: \begin{equation}\la{d.seq1}
3158: F_{T_n}(\beta_n x)\to \hat F(x), \qquad
3159: \bar F_{\bar T_n}(\bar \beta_n x)\to \check F(x),
3160: \end{equation}
3161: at every point of continuity of the limit. Here the values $\bar
3162: T_n$, $\bar\beta_n$ are those that correspond via the map
3163: $(t,b)\mapsto(\bar t,\bar b)$ stated in the theorem.
3164: Relabeling if necessary, we may assume $0<\hat F(x)$ for some finite
3165: $x$, i.e., $\hat F$ does not represent pure gel.
3166: Therefore, according to the extended \LK\ representation theorem~\ref{ET.LK},
3167: there exists an \barsmeas\ $\dsm$ that corresponds to $\hat F$.
3168:
3169: {\em 2.\/} Let
3170: \begin{equation}\la{d.alp}
3171: \alpha_n = \begin{cases}
3172: \beta_n & (K=2),\cr
3173: \beta_n\rme^{-T_n} & (K=x+y),\cr
3174: \beta_n|T_n| & (K=xy),
3175: \end{cases}
3176: \quad
3177: \lambda_n = \begin{cases}
3178: T_n & (K=2).\cr
3179: \rme^{T_n} & (K=x+y),\cr
3180: |T_n|\inv & (K=xy).
3181: \end{cases}
3182: \end{equation}
3183: and similarly define $\bar\alpha_n$, $\bar\lambda_n$ in terms of
3184: $\bar \beta_n$, $\bar T_n$. Note that $\bar\alpha_n=\alpha_n$.
3185: We claim that $\alpha_n\to\infty$. This is evident for $K=2$, and
3186: once we prove it for $K=x+y$ it follows for $K=xy$
3187: by the transformation formula \qref{M.2}.
3188: For $K=x+y$, one can prove $\alpha_n\to\infty$ by following the
3189: beginning of the proof of Theorem 7.1 in \cite{MP1} up to (7.8)
3190: using only subsequential convergence. From (7.8) one deduces
3191: $\lambda\rme^{-t}\to \infty$, which corresponds here to
3192: $\alpha_n\to\infty$.
3193: %[[Ok - desirable to give clear, self-contained proof.]]
3194:
3195: {\em 3.\/} Define rescaled initial \smeass\ (see \qref{R.initG}) by
3196: \begin{equation}\la{d.scG}
3197: \sm\nsup (x) = \lambda_n \alpha_n\inv \sm(\alpha_n x),
3198: \qquad
3199: \bar\sm\nsup (x) = \bar\lambda_n \alpha_n\inv \sm(\alpha_n x),
3200: \end{equation}
3201: According to the extended encoding theorem~\ref{ET.code}
3202: %\marginpar{[[needs statement and proof!]]; added 8/16/06}
3203: the \smeass\ $\sm\nsup$ converge to $\dsm$.
3204: Let $\vp\nsup$, $\bar\vp\nsup$ and $\Phi$ be the
3205: first-order Laplace exponents associated
3206: to $\sm\nsup$, $\bar\sm\nsup$ and $\dsm$
3207: respectively as in \qref{L.1}.
3208: We have $\bar\lambda_n = \lambda_n/L(\alpha_n)$, and
3209: with $\hat L(1/q)=\bar\vp(q)/\vp(q)$, the hypothesis
3210: \qref{R.shad2} ensures $\hat L$ is slowly varying and
3211: $\hat L\sim L$.
3212: Hence, for any $q\in(0,\infty)$ we have
3213: \[
3214: \bar\vp\nsup(q) = \bar\lambda_n\bar\vp(q/\alpha_n)
3215: = \vp\nsup(q)
3216: \frac{\hat L(\alpha_n /q)}{\hat L(\alpha_n)}
3217: \frac{\hat L(\alpha_n)}{L(\alpha_n)}
3218: \to \Phi(q)
3219: \]
3220: as $n\to\infty$. By Theorem~\ref{LT.Mconv}, it follows
3221: $\bar\sm\nsup$ converges to $\dsm$, and
3222: by the extended \LK\ representation theorem,
3223: this yields $\hat F=\check F$ in \qref{d.seq1}.
3224: This contradicts \qref{d.notdto0} and finishes the proof.
3225: \end{proof}
3226:
3227: \subsection{Self similar solutions and domains of attraction}
3228: \label{sec:domains}
3229: % %Here we characterize the domains of attraction
3230: % [[The idea is to reprove the characterization theorem as an
3231: % application of shadowing, showing how the condition is necessary in
3232: % this case. ]]
3233: The self-similar solutions are the simplest examples of eternal solutions.
3234: All self-similar solutions are generated by the \smeass\/
3235: $\dsm(x)=C x^{1-\rho}$ with $\rho\in(0,1]$, $C>0$,
3236: with corresponding Laplace exponents
3237: \be \la{Do.Hlap}
3238: \Phi(q) =
3239: Cq^\rho
3240: \frac{ \Gamma(2-\rho)}{\rho},
3241: \qquad
3242: \Psi(q) = C q^{1+\rho}
3243: \frac{\Gamma(2-\rho)}{\rho(1+\rho)}.
3244: \ee
3245: Thus, there is a one-parameter family
3246: up to trivial scalings. By \qref{C.eternal} and \qref{A.psi11},
3247: for appropriate $C$, the Laplace exponent $\vp=\vp(t,q)$ of the
3248: solution satisfies
3249: \be
3250: \vp = \frac{\Phi}{1+t\Phi} = \frac{q^\rho}{1+tq^\rho}
3251: \qquad(K=2),
3252: \ee
3253: \be
3254: q = \vp+\Psi(\rme^t\vp) = \vp + (\rme^t \vp)^{1+\rho}
3255: \qquad (K=x+y).
3256: \ee
3257: The self-similar solutions were described in ~\cite{MP1},
3258: and can all be captured by expressing the associated probability distribution
3259: in the form
3260: \be
3261: F(t,x) = F_{\rho,\gamma}(x/\lgr_{\rho,\gamma}(t)),
3262: \ee
3263: for $\gamma=0,1,2$, where the scale factors are
3264: \be
3265: \lgr_{\rho,0}(t)=t^{1/\rho},\quad \lgr_{\rho,1}(t)=\rme^{t/\beta}, \quad
3266: \lgr_{\rho,2}(t)=|t|^{-1/\beta},
3267: \ee
3268: with $\beta=\rho/(1+\rho)$, and the probability distributions
3269: $F_{\rho,\gamma}$ are explicitly
3270: \begin{eqnarray}
3271: \label{eq:f_rho_const}
3272: F_{\rho,0}(x) &=&
3273: \sum_{k=1}^\infty \frac{ (-1)^{k+1}x^{\rho k}}{\Gamma(1+\rho k)},
3274: \\
3275: \label{eq:f_rho_add}
3276: F_{\rho,1}(x)= F_{\rho,2}(x) &=&
3277: \frac1\pi \sum_{k=1}^\infty
3278: \frac{(-1)^{k-1}x^{k\beta}}{k!}\Gamma(1+k-k\beta)
3279: \frac{\sin k\pi\beta}{k\pi\beta}.\quad
3280: \end{eqnarray}
3281:
3282: We now restate the characterization of the domains of attraction of these
3283: self-similar solutions obtained in~\cite{MP1}. We say a probability measure
3284: on $E$ is nontrivial if it is not concentrated at the origin.
3285:
3286: \begin{thm}
3287: \label{DT.domains}
3288: Let $F_t$ denote the probability measure associated to a solution to
3289: Smoluchowski's coagulation equations with $K=2, x+y,$ or
3290: $xy$.
3291: \bit
3292: \item[(a)] Assume there is a rescaling $b(t)\to \infty$ and a
3293: nontrivial probability measure $\hat F$ on $E$
3294: such that $F_t(b(t)x) \to \hat F(x)$ at all points of
3295: continuity. Then there is $\rho \in (0,1]$, and a function $L$
3296: slowly varying at infinity such that
3297: \be
3298: \la{DE.dom}
3299: \sm_{t_0}(x) =\int_{(0,x]}y^{\gamma+1}\nu_{t_0}(dy)
3300: \sim x^{1-\rho}L(x), \quad x \to \infty.
3301: \ee
3302: \item[(b)] Conversely, assume \qref{DE.dom} holds. Then there is a
3303: rescaling $b(t) \to \infty$ such that
3304: \be
3305: \la{DE.dom2}
3306: \lim_{t \to \infty} \dist\left(F_t(b(t)\,dx), F_{\rho,\gamma}(dx)
3307: \right)=0.
3308: %F_{\rho,\gamma}(\lambda_{\rho,\gamma}(t)\,dx)\right) = 0.
3309: \ee
3310: \eit
3311: \end{thm}
3312: Theorem~\ref{DT.domains} illustrates the {\em rigidity of scaling
3313: limits\/}. If we insist on the existence of a proper limit as $t \to
3314: \infty$ (as opposed to subsequential limits), the only possibility is
3315: that $\hat F(x) = F_{\rho,\gamma} (ax)$ for some $\rho \in (0,1]$ and
3316: $a \in (0,\infty)$.
3317: (For degenerate limits, see the remark below.)
3318: Theorems~\ref{RT.initial_tails} and \ref{RT.shad}
3319: shed more light on this result as they clarify the main hypothesis
3320: (\qref{DE.dom} above) and allow us to avoid the use of
3321: Karamata's Tauberian theorem in the proof.
3322:
3323: \begin{proof}
3324: Let us first prove (a). Suppose there is a (possibly discontinuous)
3325: rescaling $b(t) \to \infty$ such that $\lim_{t \to \infty} F_t(b(t)x) =
3326: \hat{F}(x)$ at all points of continuity of $\hat{F}$. Then $\hat{F}
3327: \in \attr$, so it is associated to a divergent \smeas\/ $\dsm$.
3328: % We use Theorem~\ref{RT.initial_tails} with $\beta_n=b(T_n)$ where
3329: % $T_n=n$ ($K=2,x+y$) and $T_n=n^{-1}$ ($K=xy$).
3330: % % \begin{equation}\label{R.domains1}
3331: % % T_n = \left\{
3332: % % \begin{array}{ll}
3333: % % n/t_0 &(K=2 \text{ or }xy),\\
3334: % % e^n &(K=x+y).
3335: % % \end{array}\right.
3336: % % \end{equation}
3337: % % and $\beta_n=b(T_n)$.
3338: % the measures
3339: % $\tilde{F}^{(t)}(x) = F_t(b(t)x)$, that is
3340: % \begin{equation}\label{R.domains1}
3341: % a(t) = \left\{
3342: % \begin{array}{ll}
3343: % t/t_0 &(K=2 \text{ or }xy),\\
3344: % e^t &(K=x+y).
3345: % \end{array}\right.
3346: % \end{equation}
3347: Theorem~\ref{RT.initial_tails} (ii) now implies the convergence
3348: of the \smeass\/ $\tilde\sm\tsup \to \dsm$ where
3349: \begin{equation}\label{R.dom2}
3350: \tilde\sm\tsup (x) = \tlam\alpha\inv \sm_{t_0}(\alpha x),
3351: \end{equation}
3352: \be \la{Do.alp}
3353: \alpha(t) = \begin{cases}
3354: b(t) & (K=2),\cr
3355: b(t)\rme^{-t} & (K=x+y),\cr
3356: b(t)|t| & (K=xy),
3357: \end{cases}
3358: \quad
3359: \tlam(t) = \begin{cases}
3360: t & (K=2).\cr
3361: \rme^{t} & (K=x+y),\cr
3362: |t|\inv & (K=xy).
3363: \end{cases}
3364: \ee
3365: As we have seen in the proof of Theorem~\ref{RT.shad},
3366: $\alpha(t)$ diverges as $t \to \Tmax$ in
3367: each case. Then by \qref{R.dom2}, the Laplace exponent $\vp_0$
3368: for $\sm_{t_0}$ satisfies
3369: \be \la{Do.vps}
3370: \tlam \vp_0(q/\alpha) \to \Phi(q)
3371: \ee
3372: as $t\to\Tmax$, where $\Phi$ is the Laplace exponent of $\dsm$.
3373: Taking $t\to\Tmax$ along a sequence $t_n$ such that
3374: $\tlam(t_{n+1})/\tlam(t_n)\to1$, by
3375: a fundamental rigidity lemma~\cite[VIII.8.3]{Feller}, we infer
3376: that the only possible limits are power-laws, meaning
3377: $\Phi(q)=cq^\rho$ for some $\rho\ge0$. Since $H$ is a nontrivial
3378: \smeas, we must have $0<\rho\le1$ and $c>0$.
3379: Moreover we infer $\vp_0$ is regularly varying at 0, meaning
3380: $\vp_0(q)=q^\rho \hat L(q)$, where
3381: $\hat L(aq)/\hat L(q)\to1$ as $q\to0$ for every $a>0$.
3382: Note that by \qref{Do.vps},
3383: \be\la{Do.tlam}
3384: \tlam \sim c\alpha^\rho/\hat L(1/\alpha),
3385: \quad
3386: c_n=\tlam(t_n)\vp_0(1/\alpha(t_n))\to c.
3387: \ee
3388:
3389: With $t_n$ as described and $\alpha_n=\alpha(t_n)$,
3390: we claim $\alpha_{n+1}/\alpha_n\to1$ as $n\to\infty$.
3391: Let $a>1$ and suppose $\alpha_{n+1}/\alpha_n>a$ for infinitely many $n$.
3392: Then since $\vp_0$ is strictly increasing,
3393: along this subsequence we have
3394: \[
3395: \frac{c_{n+1}}{\tlam(t_{n+1})}
3396: \frac{\tlam(t_n)}{c_n}
3397: =
3398: \frac{\vp_0(1/\alpha_{n+1})}{\vp_0(1/\alpha_n)}
3399: \le
3400: \frac{\vp_0(a\inv/\alpha_n)}{\vp_0(1/\alpha_n)}
3401: \to a^{-\rho} <1.
3402: \]
3403: But the left-hand side converges to 1. Hence
3404: $\limsup \alpha_{n+1}/\alpha_n \le1$. Similarly we deduce
3405: $\liminf \alpha_{n+1}/\alpha_n \ge1$, establishing the claim.
3406:
3407: We may now apply the rigidity lemma~\cite[VIII.8.3]{Feller} to \qref{R.dom2}
3408: to infer that $G_{t_0}$ is regularly varying at $\infty$, meaning
3409: \qref{DE.dom} holds. (The value of $\rho$ must be the same here, due to
3410: \qref{Do.tlam} and \qref{Do.Hlap}.) This proves part (a).
3411:
3412: To prove the converse, we assume that \qref{DE.dom}
3413: holds. Since \qref{DE.dom} holds we may choose
3414: increasing rescalings $\alpha(t) \to \infty$ and
3415: $\tlam(t)$ such that the \smeass\/
3416: $G\tsup(x)=\tlam\alpha^{-1} G_{t_0}(\alpha x)$ converge to $\dsm
3417: =x^{1-\rho}$. Let $b(t)$ be defined by \qref{Do.alp} for the
3418: various kernels. It then follows that $F_t(b(t)x) \to \Fgr(x)$ for every
3419: $x >0$. Since the metric is equivalent to weak convergence we also
3420: have \qref{DE.dom2}.
3421: \end{proof}
3422:
3423: \begin{rem}
3424: A remaining nontrivial possibility discussed in \cite{MP1}
3425: is that of a defective limit on $E$, which we may now take to mean
3426: that $F_t(b(t)x)\to\hat F(x)$ where $\hat F$ is a probability measure
3427: on $\Ebar=[0,\infty]$, with $0<\hat F(\infty^-)<1$, meaning that gel
3428: appears in the limit.
3429: If this is the case, then $\hat F$ is an element of the full scaling
3430: attractor $\attrx$, and by Theorem~\ref{ET.code}, the rescaled
3431: \smeass\/ $\tilde\sm\tsup\to H$, the \barsmeas\/ associated to
3432: $\hat F$. Moreover,
3433: $y\inv H(dy)$ must have nonzero charge $h_\infty$ at $\infty$,
3434: and hence $\Phi(0+)>0$. This means that the in the proof above,
3435: the rigidity lemma must yield $\rho=0$, i.e., we must have
3436: $\Phi(q)=c>0$, corresponding to an eternal extended solution
3437: consisting of a pure dust/gel mixture.
3438: By the discussion in \cite{MP1} (see Remarks 5.4 and 7.4) a necessary and
3439: sufficient condition for this to occur is that
3440: \be
3441: \int_{[x,\infty)} y\inv \sm_{t_0}(dy) \sim L(x), \quad x\to\infty,
3442: \ee
3443: where $L$ is slowly varying at $\infty$.
3444: \end{rem}
3445:
3446:
3447:
3448: \section*{Acknowledgements}
3449: This material is based upon work supported by the National Science
3450: Foundation under grant nos.\ DMS 00-72609, DMS 03-05985, DMS 06-04420
3451: and DMS 06-05006. GM thanks
3452: the IMA for partial support during the preparation of the manuscript.
3453: RLP thanks the DFG for partial support through a Mercator
3454: professorship at Humboldt University.
3455:
3456: %\vspace{-6pt}
3457: \pagebreak
3458:
3459: \bibliography{mp3}
3460:
3461: \end{document}
3462:
3463:
3464: