1: %
2: % This is file JFM2esam.tex
3: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
4: % first release v1.0, 20th October 1996
5: % release v1.01, 29th October 1996
6: % release v1.1, 25th June 1997
7: % (based on JFMsampl.tex v1.3 for LaTeX2.09)
8: % Copyright (C) 1996, 1997 Cambridge University Press
9:
10:
11: \NeedsTeXFormat{LaTeX2e}
12:
13:
14: %\documentclass[referee]{jfm}
15: \documentclass{jfm}
16: \usepackage{graphicx,natbib}
17:
18:
19: % See if the author has AMS Euler fonts installed: If they have, attempt
20: % to use the 'upmath' package to provide upright math.
21:
22:
23: \ifCUPmtlplainloaded \else
24: \checkfont{eurm10}
25: \iffontfound
26: \IfFileExists{upmath.sty}
27: {\typeout{^^JFound AMS Euler Roman fonts on the system,
28: using the 'upmath' package.^^J}%
29: \usepackage{upmath}}
30: {\typeout{^^JFound AMS Euler Roman fonts on the system, but you
31: dont seem to have the}%
32: \typeout{'upmath' package installed. JFM.cls can take advantage
33: of these fonts,^^Jif you use 'upmath' package.^^J}%
34: \providecommand\upi{\pi}%
35: }
36: \else
37: \providecommand\upi{\pi}%
38: \fi
39: \fi
40:
41:
42: % See if the author has AMS symbol fonts installed: If they have, attempt
43: % to use the 'amssymb' package to provide the AMS symbol characters.
44:
45:
46: \ifCUPmtlplainloaded \else
47: \checkfont{msam10}
48: \iffontfound
49: \IfFileExists{amssymb.sty}
50: {\typeout{^^JFound AMS Symbol fonts on the system, using the
51: 'amssymb' package.^^J}%
52: \usepackage{amssymb}%
53: \let\le=\leqslant \let\leq=\leqslant
54: \let\ge=\geqslant \let\geq=\geqslant
55: }{}
56: \fi
57: \fi
58:
59:
60: % See if the author has the AMS 'amsbsy' package installed: If they have,
61: % use it to provide better bold math support (with \boldsymbol).
62:
63:
64: \ifCUPmtlplainloaded \else
65: \IfFileExists{amsbsy.sty}
66: {\typeout{^^JFound the 'amsbsy' package on the system, using it.^^J}%
67: \usepackage{amsbsy}}
68: {\providecommand\boldsymbol[1]{\mbox{\boldmath $##1$}}}
69: \fi
70:
71:
72:
73: %%% Example macros (some are not used in this sample file) %%%
74:
75:
76: % For units of measure
77: \newcommand\dynpercm{\nobreak\mbox{$\;$dynes\,cm$^{-1}$}}
78: \newcommand\cmpermin{\nobreak\mbox{$\;$cm\,min$^{-1}$}}
79:
80:
81: % Various bold symbols
82: \providecommand\bnabla{\boldsymbol{\nabla}}
83: \providecommand\bcdot{\boldsymbol{\cdot}}
84: \newcommand\biS{\boldsymbol{S}}
85: \newcommand\etb{\boldsymbol{\eta}}
86:
87:
88: % For multiletter symbols
89: \newcommand\Real{\mbox{Re}} % cf plain TeX's \Re and Reynolds number
90: \newcommand\Imag{\mbox{Im}} % cf plain TeX's \Im
91: \newcommand\Rey{\mbox{\textit{Re}}} % Reynolds number
92: \newcommand\Pran{\mbox{\textit{Pr}}} % Prandtl number, cf TeX's \Pr product
93: \newcommand\Pen{\mbox{\textit{Pe}}} % Peclet number
94: \newcommand\Ai{\mbox{Ai}} % Airy function
95: \newcommand\Bi{\mbox{Bi}} % Airy function
96:
97:
98: % For sans serif characters:
99: % The following macros are setup in JFM.cls for sans-serif fonts in text
100: % and math. If you use these macros in your article, the required fonts
101: % will be substitued when you article is typeset by the typesetter.
102: %
103: % \textsfi, \mathsfi : sans-serif slanted
104: % \textsfb, \mathsfb : sans-serif bold
105: % \textsfbi, \mathsfbi : sans-serif bold slanted (doesnt exist in CM fonts)
106: %
107: % For san-serif roman use \textsf and \mathsf as normal.
108: %
109: \newcommand\ssC{\mathsf{C}} % for sans serif C
110: \newcommand\sfsP{\mathsfi{P}} % for sans serif sloping P
111: \newcommand\slsQ{\mathsfbi{Q}} % for sans serif bold-sloping Q
112:
113:
114: % Hat position
115: \newcommand\hatp{\skew3\hat{p}} % p with hat
116: \newcommand\hatR{\skew3\hat{R}} % R with hat
117: \newcommand\hatRR{\skew3\hat{\hatR}} % R with 2 hats
118: \newcommand\doubletildesigma{\skew2\tilde{\skew2\tilde{\Sigma}}}
119: % italic Sigma with double tilde
120:
121:
122: % array strut to make delimiters come out right size both ends
123: \newsavebox{\astrutbox}
124: \sbox{\astrutbox}{\rule[-5pt]{0pt}{20pt}}
125: \newcommand{\astrut}{\usebox{\astrutbox}}
126:
127:
128: \newcommand\GaPQ{\ensuremath{G_a(P,Q)}}
129: \newcommand\GsPQ{\ensuremath{G_s(P,Q)}}
130: \newcommand\p{\ensuremath{\partial}}
131: \newcommand\tti{\ensuremath{\rightarrow\infty}}
132: \newcommand\kgd{\ensuremath{k\gamma d}}
133: \newcommand\shalf{\ensuremath{{\scriptstyle\frac{1}{2}}}}
134: \newcommand\sh{\ensuremath{^{\shalf}}}
135: \newcommand\smh{\ensuremath{^{-\shalf}}}
136: \newcommand\squart{\ensuremath{{\textstyle\frac{1}{4}}}}
137: \newcommand\thalf{\ensuremath{{\textstyle\frac{1}{2}}}}
138: \newcommand\Gat{\ensuremath{\widetilde{G_a}}}
139: \newcommand\ttz{\ensuremath{\rightarrow 0}}
140: \newcommand\ndq{\ensuremath{\frac{\mbox{$\partial$}}{\mbox{$\partial$} n_q}}}
141: \newcommand\sumjm{\ensuremath{\sum_{j=1}^{M}}}
142: \newcommand\pvi{\ensuremath{\int_0^{\infty}%
143: \mskip \ifCUPmtlplainloaded -30mu\else -33mu\fi -\quad}}
144:
145:
146: \newcommand\etal{\mbox{\textit{et al.}}}
147: \newcommand\etc{etc.\ }
148: \newcommand\eg{e.g.\ }
149:
150:
151:
152: \newtheorem{lemma}{Lemma}
153: \newtheorem{corollary}{Corollary}
154:
155:
156:
157: %-----title and author----------------------
158:
159:
160: \title[On the aggregation of inertial particles in random flows]{On the
161: aggregation of inertial particles in random flows}
162:
163:
164: \author[B. Mehlig, M. Wilkinson, K. Duncan, T. Weber and M. Ljunggren]
165: %{B. Mehlig$^1$, M. Wilkinson$^2$, K. Duncan$^2$, T. Weber$^1$ and M. Ljunggren$^1$}
166: {B.\ns M\ls E\ls H\ls L\ls I\ls G$^1$,
167: M.\ns W\ls I\ls L\ls K\ls I\ls N\ls S\ls O\ls N$^2$,
168: K.\ns D\ls U\ls N\ls C\ls A\ls N$^2$,
169: T.\ns W\ls E\ls B\ls E\ls R$^1$,
170: \and
171: M.\ns L\ls J\ls U\ls N\ls G\ls G\ls R\ls E\ls N$^1$}
172:
173:
174:
175: \affiliation{$^1$Theoretical Physics,
176: School of Physics and Engineering Physics,
177: G\"oteborg University/Chalmers, 41296 G\"{o}teborg, Sweden\\[\affilskip]
178: $^2$Department of Applied Mathematics,
179: The Open University, Walton Hall, Milton Keynes, MK7 6AA, England}
180:
181:
182: \pubyear{}
183: \volume{}
184: \pagerange{}
185: \date{??}
186: \setcounter{page}{1}
187:
188:
189: % -----------------------------
190:
191:
192: \begin{document}
193:
194:
195:
196: \maketitle
197:
198:
199: \begin{abstract}
200: We describe a criterion for particles suspended in a
201: randomly moving fluid to aggregate. Aggregation
202: occurs when the expectation value of a random variable
203: is negative. This variable evolves
204: under a stochastic differential equation.
205: We analyse this equation in detail in the limit where the
206: correlation time of the velocity field of the fluid is very short, such that
207: the stochastic differential equation is a Langevin equation.
208: \end{abstract}
209:
210:
211: \section{Introduction}
212: \label{sec: 1}
213: \par
214: \par
215: \noindent{\sl 1.1 Illustration and context}
216: \par
217: Figure~\ref{fig: 1}
218: illustrates a simulation of the
219: distribution of small particles suspended in a three-dimensional
220: random flow: the particles are modelled as points, but they
221: are shown as small spheres to make them visible in the figure.
222: The suspended particles do not interact (that is, the motion of each
223: particle is independent of the coordinates of the other particles,
224: the equations of motion are given below).
225: We show the initial configuration (Figure~\ref{fig: 1}{\bf a}),
226: and two snapshots of the particle positions after a long time,
227: with differing values of the fluid viscosity,
228: (Figures~\ref{fig: 1}{\bf b} and {\bf c}). In one case
229: the particles aggregate, in the sense that the trajectories of
230: different particles coalesce. In the other their distribution
231: shows some degree of clustering, but their trajectories never
232: coalesce.
233: In this paper we present an analysis of the transition between
234: aggregating and non-aggregating phases,
235: which we term the \lq path-coalescence transition'.
236:
237:
238: \begin{figure}
239: \vspace*{4mm}
240: \centerline{\includegraphics[width=4.15cm,clip]{fig1a.bitmap.eps}
241: \hspace*{4mm}\includegraphics[width=4.15cm,clip]{fig1b.bitmap.eps}
242: \hspace*{4mm}\includegraphics[width=4.15cm,clip]{fig1c.bitmap.eps}}
243: \caption{\label{fig: 1}
244: Illustrating the aggregation of non-interacting particles
245: in a random three-dimensional flow: the motion
246: is defined by equations (\ref{eq: 1.1}) to (\ref{eq: 1.3}),
247: (using a simplified form explained in section \ref{sec: 3}:
248: see equation (\ref{eq: 3.23})). The left
249: panel shows the initial configuration at $t=0$, the center and right panels
250: show final configurations (at $t=180\,\tau$) for two different
251: values of $\gamma$.
252: In the case shown in the center panel,
253: trajectories coalesce, until eventually all particles
254: follow the same trajectory. The particles in the right
255: panel exhibit density
256: fluctuations, but the trajectories do not coalesce.
257: }
258: \end{figure}
259:
260:
261: There are numerous experimental observations that
262: when small particles are suspended in a complex and apparently
263: random flow, their density becomes non-uniform.
264: Clustering of particles into regions of high density
265: has been observed in experiments on particles floating on the surface
266: of liquids with a complex or turbulent flow \cite[]{Som93,Cre04}, and
267: also turbulent flow in channels \cite[]{Fes94,vHa98}.
268: The conditions
269: under which this occurs are not yet fully understood.
270: The aggregation effect illustrated in figure 1 is
271: an extreme form of clustering.
272: There appears to be less experimental work on
273: aggregation, but coalescence of suspended water droplets
274: is clearly very important in the formation of rain drops
275: in clouds \cite[]{Sha03}.
276: %This process is still not very well understood,
277: Even less is known about aggregation. In particular
278: it is not clear when a model of particles suspended in a random flow
279: can exhibit aggregation, and when additional physical
280: phenomena (such as differential drift velocities under gravitational
281: forces, or Brownian diffusion) must be invoked.
282:
283:
284: Earlier theoretical work on clustering in random flows
285: has used Fokker-Planck equations determining
286: moments of the particle density of advected
287: particles \cite[]{Kly99}. This \lq passive scalar'
288: approach does not allow for the effects of
289: inertia of the particles. It has been supplemented by a
290: perturbative analysis of the deviations of particles from
291: advected trajectories, as proposed by \cite{Max87}.
292: These two approaches are combined in papers by
293: \cite{Elp96} and \cite{Fal03}.
294: Numerical work indicates that clustering
295: in solenoidal flows occurs most readily when the correlation
296: time of the velocity field is comparable with the
297: time constant associated with the viscous drag \cite[]{Sig02}.
298: The aggregation (as opposed
299: to clustering) of
300: particles by random flows has received relatively little
301: attention. \cite{Deu85} appears to
302: have been the first to propose that particles
303: subjected to a smooth random flow can coalesce, and showed
304: numerical evidence that this can happen in one dimension.
305: He argued that there is a transition between coalescing
306: and non-coalescing phases, and identified the dimensionless
307: parameter which determines the phase transition in one dimension.
308:
309:
310: This paper describes results obtained from a new approach
311: to characterising particle aggregation in random flows.
312: It is based upon calculating a Lyapunov exponent describing
313: the rate of separation of nearby particles from a solution
314: of a system of stochastic differential equations: the Lyapunov
315: exponent is the expectation value of one of the variables.
316: In the limit as the correlation time of the flow approaches
317: zero, the stochastic differential equations can be reduced
318: to a pair of Langevin equations.
319: Our results
320: for three-dimensional flows which are discussed here build upon earlier work
321: by two of the present authors in one dimension
322: \cite[]{Wil03} (where we solved Deutsch's model
323: exactly) and two dimensions \cite[]{Meh04}.
324: The three-dimensional case is most important
325: for physical applications, but involves substantial additional
326: technical complications.
327:
328:
329: We remark that \cite{Pit01} also considered the two-dimensional
330: case, and quotes an analytic expression for the maximal
331: Lyapunov exponent. His expression is
332: incorrect in two dimensions, because the generating function
333: that he uses is divergent for the equilibrium distribution.
334: His calculation can be adapted to give the correct expression
335: in the one dimensional case, as quoted in \cite{Meh04}.
336:
337:
338: The definition of Lyapunov exponents is explained by
339: \cite{Eck79}, who also discuss the method we use for
340: extracting the Lyapunov exponents from direct numerical simulations.
341: We note that in the case where inertia of the particles can be neglected,
342: our results reduce to calculating the Lyapunov exponent for a spatially
343: correlated Brownian motion, which was discussed
344: from a mathematical point of view by \cite{LeJ85}
345: and \cite{Bax86}.
346: \par
347: \
348: \par
349: \
350: \par
351: \noindent{\sl 1.2 The model}
352: \par
353: The most natural model for theoretical investigation
354: is the motion of spherical particles (radius $a$, mass $m$)
355: moving in a random velocity field ${\bf u}({\bf r},t)$ with
356: specified isotropic and homogeneous statistical properties.
357: The particles are assumed not to affect either the flow,
358: or each other's motion, and to experience a drag force
359: given by Stokes's law: the force ${\bf f}_{\rm dr}$ on a
360: particle moving with velocity ${\bf v}$ relative to the
361: fluid is ${\bf f}_{\rm dr}=-6\pi \eta a{\bf v}$, where $\eta$
362: is the viscosity of the fluid.
363: We will simplify the problem by assuming that the
364: particles are made of a material which is much
365: denser than the fluid in which they are suspended:
366: this enables us to neglect the inertia of the displaced
367: fluid. Accordingly, we consider a large number of suspended
368: particles, initially with random positions and zero
369: velocity, having equations of motion
370: %
371: %
372: \begin{eqnarray}
373: \label{eq: 1.1}
374: \dot{\bf r}&=&{1\over{m}}{\bf p}\ ,\qquad
375: \dot {\bf p}=-\gamma [{\bf p}-m{\bf u}({\bf r},t)]\,.
376: \end{eqnarray}
377: %
378: %
379: The random velocity field ${\bf u}({\bf r},t)$ could be either
380: externally imposed (for example, if a gas is driven by
381: an ultrasonic noise source), or self-generated (as in the
382: case of turbulence). The equations of motion with displaced
383: mass effects included are discussed by \cite{Lan58}.
384: Our neglect of displaced mass effects is justified for
385: aerosol systems.
386:
387:
388: In order to fully specify the problem we must define
389: the statistical properties of the random velocity field
390: ${\bf u}({\bf r},t)$.
391: The random force ${\bf f}=m\gamma{\bf u}$ is generated
392: from a vector potential ${\bf A}=(A_1,A_2,A_3)$ and a scalar
393: potential $\phi=A_0$:
394: %
395: %
396: \begin{equation}
397: \label{eq: 1.2}
398: {\bf f}=\nabla \phi+\nabla \wedge {\bf A}
399: \ .
400: \end{equation}
401: %
402: %
403: The scalar fields $A_i({\bf r},t)$ have isotropic, homogeneous and
404: stationary statistics. We assume that these fields are
405: statistically independent, and that they
406: all have the same correlation function,
407: except that the intensity of $\phi=A_0$ exceeds that of the
408: other fields by a factor $1/\alpha^2$:
409: %
410: %
411: \begin{equation}
412: \label{eq: 1.3}
413: \langle A_i({\bf r},t)A_j({\bf r}',t')\rangle=
414: \delta_{ij}[(1-\alpha^2)\delta_{i0}+\alpha^2]
415: C(\vert{\bf r}-{\bf r}'\vert,t-t')
416: \ .
417: \end{equation}
418: %
419: %
420: The random force ${\bf f}({\bf r},t)$ is characterised by its typical
421: magnitude $\sigma$, and by the correlation length $\xi$ and correlation
422: time $\tau$ of the correlation function $C(R,t)$. In the case
423: of a well-developed turbulent flow, the velocity field
424: has a power-law spectrum with upper and lower cutoffs
425: \cite[]{Fri97}.
426: If a random velocity field modeling fully-developed
427: turbulence is used in our model, it is most appropriate
428: to take the correlation length and time to be those of
429: the \lq dissipation scale', that is, the cutoff with the
430: smaller length scale.
431:
432:
433: The system of equations is
434: characterised by three independent dimensionless parameters,
435: which we take as
436: %
437: %
438: \begin{equation}
439: \label{eq: 1.4}
440: \nu=\gamma \tau\ ,\ \ \chi={\sigma \tau^2\over{m\xi}}\ ,\ \ \ \alpha
441: \ .
442: \end{equation}
443: %
444: %
445: The parameter $\nu $ is a dimensionless measure of the degree of
446: damping,
447: $\chi $ is a dimensionless measure of the strength of the forcing
448: term and $\alpha $ measures the relative magnitudes of potential
449: and solenoidal components of the velocity field (which is purely
450: potential when $\alpha=0$, and purely solenoidal in the limit as
451: $\alpha \to \infty$).
452: \par
453: \
454: \par
455: \
456: \par
457: \
458: \par
459: \noindent{\sl 1.3 Description of our results}
460: \par
461: We show that the phase transition is determined by a
462: Lyapunov exponent, $\lambda_1$, describing the separation
463: of nearby particles: their trajectories coalesce if $\lambda_1<0$.
464: Here we describe a new and general approach
465: to this problem, reducing the determination of the
466: Lyapunov exponent to the analysis of a simple
467: dynamical system, described by a system of
468: ordinary differential equations containing stochastic
469: forcing terms: the Lyapunov exponent is found to be proportional
470: to the expectation value of one coordinate.
471: These stochastic differential
472: equations are derived in section \ref{sec: 2}. They
473: introduce an apparent paradox: the structure of the equations
474: appears to be identical in two-dimensional
475: and three-dimensional flows, suggesting that
476: the path-coalescence transition is fundamentally
477: the same in two and three dimensions. That would be
478: a very surprising conclusion.
479:
480:
481: In order to demonstrate and illuminate our method, in sections
482: \ref{sec: 3} and \ref{sec: 4}
483: we pursue the solution of this problem in considerable
484: depth in one limiting case, namely the limit where the correlation
485: time $\tau $ of the random velocity field is small and the random
486: force is sufficiently weak
487: (strictly, we consider the case where $\chi \ll \nu \ll 1$).
488: In this limit, the system of ordinary
489: differential equations described in section \ref{sec: 2} becomes a
490: system of two coupled Langevin equations.
491: In section \ref{sec: 3} we show that there is in fact a
492: difference between the three-dimensional problem and the
493: two-dimensional case studied in \cite{Meh04}, which is a rather subtle
494: example of the difficulties in applying Langevin approaches
495: to nonlinear equations \cite[]{vKa92}.
496:
497:
498: In section \ref{sec: 4} we discuss a perturbation theory for the
499: Lyapunov exponent describing the phase transition, expanded
500: in powers of a parameter
501: $\epsilon=\chi\nu^{-3/2}$ which is a dimensionless
502: measure of the inertia of the particles. The perturbation
503: theory is constructed by transforming the Langevin equation
504: first into a Fokker-Planck equation, and then into a
505: non-Hermite an perturbation of a three-dimensional isotropic
506: quantum harmonic oscillator. We are then able to use the
507: harmonic-oscillator creation and annihilation operators
508: \cite[]{Dir30} to
509: express the perturbation theory in a purely algebraic
510: form, enabling us to compute the coefficients to any desired
511: order. We investigate the phase diagram (the line in parameter space
512: separating coalescing and non-coalescing phases), using
513: both Monte Carlo averaging of the Langevin equation
514: and the results of our
515: perturbation theory. We find that aggregation can only
516: occur if the random flow has a certain degree of compressibility,
517: which increases as the effects of inertia increase, until
518: there is no coalescing phase even for a purely potential flow field.
519:
520:
521: The analysis in sections \ref{sec: 2} to \ref{sec: 4}
522: considers the case where
523: all of the suspended particles have the same mass $m$ and damping
524: rate $\gamma$.
525: This ideal can be approached quite accurately in model experiments,
526: but in most applications in the natural world and in technology,
527: the suspended particles will have different masses, sizes and shapes.
528: In section \ref{sec: 5} we discuss the effect of dispersion in the
529: distribution of masses in the one-dimensional case: these
530: arguments can be adapted to treat higher dimensions and
531: dispersion of the damping constant $\gamma$.
532: We argue that
533: path coalescence is not destroyed by mass dispersion
534: (although of course it is no longer a sharp transition).
535: In this paper we give a quite comprehensive discussion of the
536: the case where the correlation time of the random flow
537: is very short, and the stochastic differential equations
538: derived in section 2 can be approximated by Langevin equations.
539: Section 6 discusses how our approach can be extended to
540: other cases.
541:
542:
543:
544: \section{Equations determining the Lyapunov exponent}
545: \label{sec: 2}
546:
547:
548: To determine whether particles cluster together,
549: we consider two nearby trajectories with spatial
550: separation $\delta {\bf r}$ and momenta differing by $\delta {\bf p}$. The
551: linearised equations of motion derived from (\ref{eq: 1.1}) are
552: %
553: %
554: \begin{eqnarray}
555: \label{eq: 2.1}
556: \delta \dot {\bf r}&=&{1\over m}\delta {\bf p}\,,\qquad
557: \delta \dot {\bf p}=-\gamma \delta {\bf p}+\tilde F(t)\delta {\bf r}
558: \ .
559: \end{eqnarray}
560: %
561: %
562: Here $\tilde F(t)$ is proportional to the strain-rate matrix
563: of the velocity field, with elements
564: %
565: %
566: \begin{equation}
567: \label{eq: 2.1a}
568: F_{ij}(t)={\partial f_i\over{\partial r_j}}({\bf r}(t),t)=m\gamma
569: {\partial u_i\over{\partial r_j}}({\bf r}(t),t)
570: \ .
571: \end{equation}
572: %
573: %
574: It is convenient to
575: parameterise $\delta {\bf r}$ and $\delta {\bf p}$ as follows:
576: %
577: %
578: \begin{eqnarray}
579: \label{eq: 2.2}
580: \delta {\bf r}&=&X{\bf n}_1\,,\qquad
581: \delta {\bf p}=X(Y_1{\bf n}_1+Y_2{\bf n}_2)\,,
582: \end{eqnarray}
583: %
584: %
585: where ${\bf n}_1$ and ${\bf n}_2$ are orthogonal unit vectors, which depend
586: upon time.
587: The parameter $X$ is a scale factor: trajectories coalesce if
588: $X$ decreases with probability unity in the long-time limit.
589: In the three-dimensional case, we find it convenient
590: to introduce the third
591: element
592: ${\bf n}_3={\bf n}_1\wedge {\bf n}_2$ of a time-dependent orthonormal basis
593: so that
594: %
595: %
596: \begin{eqnarray}
597: \label{eq: 2.4}
598: {\bf n}_i.{\bf n}_j&=&\delta_{ij}
599: \qquad \mbox{and}\qquad
600: {\bf n}_i=\varepsilon_{ijk}{\bf n}_j\wedge{\bf n}_k
601: \ .
602: \end{eqnarray}
603: %
604: %
605: Differentiating (\ref{eq: 2.2}), and substituting the resulting
606: expressions into (\ref{eq: 2.1}) gives
607: %
608: %
609: \begin{eqnarray}
610: \label{eq: 2.5}
611: \delta \dot {\bf r}&=&\dot X{\bf n}_1+X \dot {\bf n}_1
612: \nonumber \\
613: %
614: %
615: &=&{1\over m}X(Y_1{\bf n}_1+Y_2{\bf n}_2)
616: \\
617: %
618: %
619: \delta \dot {\bf p}&=&\dot X(Y_1{\bf n}_1+Y_2{\bf n}_2)
620: +X(\dot Y_1{\bf n}_1+\dot Y_2{\bf n}_2)+X(Y_1\dot{\bf n}_1+Y_2\dot {\bf n}_2)
621: \nonumber \\
622: %
623: %
624: &=&-\gamma X(Y_1{\bf n_1}+Y_2{\bf n_2})+\tilde F(t){\bf n}_1
625: \ .
626: \nonumber
627: \end{eqnarray}
628: %
629: %
630: Projecting $\delta \dot {\bf r}$ onto the unit vectors ${\bf n}_i$
631: gives the following three scalar equations of motion
632: %
633: %
634: \begin{eqnarray}
635: {\bf n}_1.\delta \dot {\bf r}&=&\dot X={1\over m}Y_1 X
636: \nonumber
637: \\
638: %
639: %
640: \label{eq: 2.6}
641: {\bf n}_2.\delta \dot {\bf r}&=&X({\bf n}_2.\dot {\bf n}_1)={1\over m}Y_2X
642: \\
643: %
644: %
645: {\bf n}_3.\delta \dot {\bf r}&=&X \dot {\bf n}_1.{\bf n}_3=0
646: \ .
647: \nonumber
648: \end{eqnarray}
649: %
650: %
651: The last of these equations implies that $\dot {\bf n}_1\wedge{\bf n}_2=0$,
652: so that $\dot {\bf n}_1$ is proportional to ${\bf n}_2$: we write
653: %
654: %
655: \begin{equation}
656: \label{eq: 2.7}
657: \dot {\bf n}_1=\dot \theta {\bf n}_2
658: \end{equation}
659: %
660: %
661: so that the equation for ${\bf n}_2.\delta \dot {\bf r}$ gives
662: %
663: %
664: \begin{equation}
665: \label{eq: 2.8}
666: \dot \theta={1\over m}Y_2
667: \ .
668: \end{equation}
669: %
670: %
671: The first equation of (\ref{eq: 2.6}) indicates that $X$ is a
672: product of random variables, and therefore has a log-normal distribution,
673: that is, the logarithm of $X$ has a Gaussian probability density.
674: In the limit as $t\to \infty$, the mean and variance of $\log_{\rm e}X$
675: are both linear functions of time:
676: %
677: %
678: \begin{equation}
679: \label{eq: 2.9a}
680: \langle \log_{\rm e}X\rangle\sim \lambda_1 t+c_1\ ,\ \ \
681: {\rm var}(X)=\mu t+c_2
682: \end{equation}
683: %
684: %
685: where $\lambda_1$, $\mu$, $c_1$ and $c_2$ are constants.
686: If $\lambda_1<0$, the probability of $\log_{\rm e}X$ exceeding any
687: specified value approaches zero as $t\to \infty$, implying that
688: trajectories of nearby particles almost always coalesce.
689:
690:
691: The Lyapunov exponent $\lambda_1$ is the mean value of the derivative
692: ${\rm d}\log_{\rm e}X/{\rm d}t$, so that the first equation
693: of (\ref{eq: 2.6}) gives
694: %
695: %
696: \begin{equation}
697: \label{eq: 2.9}
698: \lambda_1={1\over m}\langle Y_1\rangle
699: \ .
700: \end{equation}
701: %
702: %
703: Now consider the three projections of $\delta \dot {\bf p}$, as given
704: by equation (\ref{eq: 2.5}):
705: %
706: %
707: \begin{eqnarray}
708: {\bf n}_1.\delta \dot {\bf p}&=&\dot XY_1+X\dot Y_1
709: +XY_2({\bf n}_1.\dot {\bf n}_2)
710: \nonumber \\
711: %
712: %
713: &=&-\gamma XY_1+{\bf n}_1.\tilde F(t){\bf n}_1X
714: \nonumber
715: \\
716: %
717: %
718: \label{eq: 2.10}
719: {\bf n}_2.\delta \dot {\bf p}&=&\dot XY_2+X\dot Y_2
720: +XY_1(\dot {\bf n}_1.{\bf n}_2)
721: \nonumber \\
722: %
723: %
724: &=&-\gamma XY_2+{\bf n}_2.\tilde F(t){\bf n}_1X
725: \\
726: %
727: %
728: {\bf n}_3.\delta \dot {\bf p}&=&
729: XY_1({\bf n}_3.\dot {\bf n}_1)+XY_2({\bf n}_3.\dot {\bf n}_2)
730: \nonumber \\
731: %
732: %
733: &=&{\bf n}_3.\tilde F(t){\bf n}_1 X
734: \nonumber
735: \ .
736: \end{eqnarray}
737: %
738: %
739: We introduce the notation
740: %
741: %
742: \begin{equation}
743: \label{eq: 2.11}
744: {\bf n}_i(t).\tilde F(t){\bf n}_j(t)=F'_{ij}(t)
745: \end{equation}
746: %
747: %
748: and note that the statistics of the transformed matrix elements
749: $F'_{ij}(t)$ are the same as those of the original elements
750: $F_{ij}(t)$, because the statistics of the velocity field
751: are isotropic.
752: Using eqs.~(\ref{eq: 2.6}) to (\ref{eq: 2.8}) and
753: $({\bf n}_i.\dot {\bf n}_j)+(\dot {\bf n}_i.{\bf n}_j)=0$ to simplify, we
754: find the following equations of motion for the variables $Y_i$
755: %
756: %
757: \begin{eqnarray}
758: \label{eq: 2.12}
759: \dot Y_1&=&-\gamma Y_1+{1\over m}(Y_2^2-Y_1^2)+F'_{11}(t)
760: \\
761: %
762: %
763: \dot Y_2&=&-\gamma Y_2-{2\over m}Y_1Y_2+F'_{21}(t)
764: \nonumber
765: \ .
766: \end{eqnarray}
767: %
768: %
769: Finally, the equation for ${\bf n}_3.\delta \dot {\bf p}$ gives
770: %
771: %
772: \begin{equation}
773: \label{eq: 2.13}
774: {\bf n}_2.\dot {\bf n}_3=-{1\over Y_2}F'_{31}(t)
775: \ .
776: \end{equation}
777: %
778: %
779: Eqs.~(\ref{eq: 2.9}) and (\ref{eq: 2.12}) are the principal results
780: of this paper. Eq.~(\ref{eq: 2.9}) shows that the Lyapunov exponent
781: (the sign of which determines whether or not path coalescence occurs)
782: is given
783: by the expectation value of a random variable $Y_1$ of a simple,
784: finite dimensional stochastic dynamical system, described
785: by eqs.~(\ref{eq: 2.12}). This dynamical system is
786: almost completely de-coupled from the other variables:
787: the equations for $Y_1$ and $Y_2$ do not depend upon $X$, and the
788: vectors ${\bf n}_i(t)$ only enter these equations through the evaluation
789: of the random matrix elements $F'_{ij}$. We pointed out that
790: statistics of these elements are independent of the orientation
791: of the orthogonal triplet $({\bf n}_1,{\bf n}_2,{\bf n}_3)$.
792:
793:
794: In the two-dimensional case the analysis leading to (\ref{eq: 2.12})
795: proceeds along similar lines, and leads to the same
796: pair of equations \cite[]{Meh04}.
797: The only difference
798: is that the equation (\ref{eq: 2.13}) is absent
799: in the two-dimensional case.
800: This suggests that the expression
801: for the Lyapunov exponent $\lambda_1=\langle Y_1\rangle/m$
802: should be the same in two and three dimensions.
803: This would be a surprising conclusion, but it is not obvious how
804: it can be averted. However, it does prove to be false,
805: as will be demonstrated in the next section for the limiting case
806: where
807: %$\nu\ll 1$ and $\chi/\nu \ll 1$.
808: $\chi \ll \nu \ll 1$.
809:
810:
811: \section{The Langevin approximation}
812: \label{sec: 3}
813:
814:
815: Let us now consider how to treat
816: eqs.~(\ref{eq: 2.12}) in the limit where the correlation
817: time $\tau$ of $\tilde F(t)$ is very short.
818: %(that is, $\nu \ll 1$).
819: Because the random field ${\bf f}({\bf r},t)$ is fluctuating very rapidly,
820: the position ${\bf r}(t)$ of a particle at time $t$ is independent
821: of the instantaneous value of the force ${\bf f}({\bf r},t)$,
822: so that the value
823: if $F_{ij}(t)=\partial f_i/\partial x_j({\bf r}(t),t)$ at the position
824: of the particle is statistically indistinguishable from a random
825: sample of the field $\partial f_i/\partial x_j$.
826: We also assume that the gradients of the fluctuating forces
827: (the quantities $F'_{11}$ and $F'_{22}$ in equations (\ref{eq: 2.12}))
828: are sufficiently small that the typical magnitude of the
829: displacement of the variables $Y_1$, $Y_2$ occurring during
830: the correlation time $\tau$ is small compared to the typical
831: magnitude of these variables. This condition is expressed
832: in terms of the dimensionless variables $\chi$ and $\nu$
833: later in this section. Under these conditions of short
834: correlation time and small amplitude,
835: equations (\ref{eq: 2.12}) may be replaced by
836: Langevin equations. At first sight, this would appear to lead
837: to
838: %
839: %
840: \begin{eqnarray}
841: \label{eq: 3.1}
842: {\rm d}Y_1&=&\biggl[-\gamma Y_1+{1\over m}(Y_2^2-Y_1^2)\biggr]{\rm d}t
843: +{\rm d}f_1
844: \\
845: %
846: %
847: {\rm d}Y_2&=&\biggl[-\gamma Y_2-{2\over m}Y_1Y_2\biggr]{\rm d}t+{\rm d}f_2
848: \nonumber
849: \end{eqnarray}
850: %
851: %
852: where the ${\rm d}f_i$ are increments of a Brownian process, satisfying
853: $\langle {\rm d}f_i\rangle=0$ and
854: $\langle {\rm d}f_i{\rm d}f_j\rangle=2D_{ij}{\rm d}t$, for
855: some constant diffusion coefficients $D_{ij}$.
856: In the two-dimensional case, this expectation is correct
857: \cite[]{Meh04}, and
858: eqs.~(\ref{eq: 3.1}) are the appropriate Langevin equations.
859: In the three-dimensional case, we will see that an additional
860: drift term must be added to the second of these equations.
861: This is a consequence of the fact that, in the three-dimensional
862: case, $Y_2$ is a non-linear function of the components of the vector
863: $\delta {\bf p}-Y_1\delta {\bf r}$ because it is the magnitude
864: of this vector. This is an example of the difficulties that
865: arise when treating Langevin equations involving non-linear
866: functions of noise terms \cite[]{vKa92}.
867:
868:
869: In order to determine the correct Langevin equations to
870: model (\ref{eq: 2.12}),
871: let us consider the integral of the stochastic forcing terms
872: ${\rm d}f_i$ over a time
873: interval $\delta t$ which is long compared to $\tau$, but
874: short enough that the change in the variables $Y_i$ occurring
875: in time $\delta t$ can be neglected. We define
876: %
877: %
878: \begin{equation}
879: \label{eq: 3.2}
880: \delta f_i(t)=\int_t^{t+\delta t}\!\!\!\!{\rm d}t'\ F'_{i1}(t')
881: \end{equation}
882: %
883: %
884: and find
885: %
886: %
887: \begin{equation}
888: \label{eq: 3.3}
889: \langle \delta f_i\delta f_j\rangle=2D_{ij}\delta t+O(\tau)
890: \end{equation}
891: %
892: %
893: where
894: %
895: %
896: \begin{equation}
897: \label{eq: 3.4}
898: D_{ij}={\textstyle{1\over 2}}\int_{-\infty}^\infty {\rm d}t\
899: \langle F'_{i1}(t)F'_{j1}(0)\rangle
900: \ .
901: \end{equation}
902: %
903: %
904: The calculation of $\langle \delta f_i(t)\rangle$ is more subtle.
905: We have
906: %
907: %
908: \begin{equation}
909: \label{eq: 3.5}
910: \langle \delta f_i\rangle=\int_0^{\delta t}{\rm d}t\
911: \langle {\bf n}_i(t).\tilde F(t) {\bf n}_1(t)\rangle
912: \ .
913: \end{equation}
914: %
915: %
916: We must take account of the fact that the unit vectors ${\bf n}_i(t)$
917: are rotating: we write
918: %
919: %
920: \begin{equation}
921: \label{eq: 3.6}
922: {\bf n}_i(t)=\sum_{k=1}^3 R_{ik}(t){\bf n}_k(0)
923: \end{equation}
924: %
925: %
926: where $R_{ik}(t)$ are elements of a rotation matrix.
927: We now write (\ref{eq: 3.5}) in the form
928: %
929: %
930: \begin{eqnarray}
931: \label{eq: 3.7}
932: \langle \delta f_i\rangle
933: &=&\int_0^{\delta t}{\rm d}t\ \sum_{k=1}^3\sum_{l=1}^3
934: \langle R_{ik}(t)R_{1l}(t)F'_{kl}(t)\rangle
935: \ .
936: \end{eqnarray}
937: %
938: %
939: Now consider the rotation of the unit vectors:
940: using (\ref{eq: 2.9}) and (\ref{eq: 2.13}) we have
941: %
942: %
943: \begin{eqnarray}
944: \nonumber
945: {\bf n}_1(t)&=&{\bf n}_1(0)+\dot \theta {\bf n}_2(0)t+O(t^2)\,,
946: \\
947: %
948: %
949: \label{eq: 3.8}
950: {\bf n}_2(t)&=&{\bf n}_2(0)-\dot \theta {\bf n}_1(0)t+
951: {1\over{Y_2}}\int_0^t{\rm d}t'\ F'_{31}(t'){\bf n}_3(0)+O(t^2)\,,
952: \\
953: %
954: %
955: {\bf n}_3(t)&=&{\bf n}_3(0)
956: -{1\over{Y_2}}\int_0^t{\rm d}t'\ F'_{31}(t'){\bf n}_2(0)\,,
957: +O(t^2)
958: \nonumber
959: \end{eqnarray}
960: %
961: %
962: so that
963: %
964: %
965: \begin{equation}
966: \label{eq: 3.9}
967: \tilde R(t)=\left( \begin{array}{ccc}
968: 1 & \dot \theta t& 0 \\
969: -\dot \theta t& 1 & Y_2^{-1}\int_0^t {\rm}dt'\, F'_{31} \\
970: 0 & -Y_2^{-1}\int_0^t{\rm d}t'\,F'_{31}&1
971: \end{array} \right)
972: +O(t^2)
973: \ .
974: \end{equation}
975: %
976: %
977: We obtain
978: %
979: %
980: \begin{eqnarray}
981: \label{eq: 3.10}
982: \langle \delta f_i\rangle&=&\int_0^{\delta t}{\rm d}t'\ \sum_k
983: \langle R_{ik}(t')F'_{k1}(t')+\dot \theta t F'_{k2}(t')\rangle
984: \nonumber \\
985: %
986: %
987: &=&\int_0^{\delta t}{\rm d}t\sum_k\langle R_{ik}(t)F'_{k1}(t)\rangle
988: +O(\delta t^2)
989: \ .
990: \end{eqnarray}
991: %
992: %
993: This yields
994: %
995: %
996: \begin{eqnarray}
997: \label{eq: 3.11}
998: \langle \delta f_1\rangle&=&\int_0^{\delta t}{\rm d}t\
999: \langle F'_{11}(t)+\dot \theta t F'_{21}(t)\rangle=0
1000: \nonumber \\
1001: %
1002: %
1003: \langle \delta f_2\rangle&=&\int_0^{\delta t}{\rm d}t\
1004: \langle -\dot \theta F'_{11}(t)+F'_{21}(t)+{1\over{Y_2}}\int_0^t{\rm d}t'\
1005: F'_{31}(t)F'_{31}(t')\rangle
1006: \nonumber \\
1007: %
1008: %
1009: &=&{1\over{Y_2}}\int_0^{\delta t}{\rm d}t\int_0^t{\rm d}t'
1010: \langle F'_{31}(t)F'_{31}(t')\rangle
1011: \nonumber \\
1012: %
1013: %
1014: &=&{1\over{Y_2}}D_{31}\delta t
1015: \ .
1016: \end{eqnarray}
1017: %
1018: %
1019: The Langevin equations therefore contain an additional drift term
1020: due to the fact that $\langle \delta f_2\rangle$ is non-zero:
1021: the correct Langevin equations in three dimensions are
1022: %
1023: %
1024: \begin{eqnarray}
1025: \label{eq: 3.13}
1026: {\rm d}Y_1&=&\biggl[-\gamma Y_1+{1\over m}(Y_2^2-Y_1^2)\biggr]{\rm d}t
1027: +{\rm d}\zeta_1\,,
1028: \nonumber
1029: \\
1030: %
1031: %
1032: {\rm d}Y_2&=&\biggl[-\gamma Y_2+{D_{31}\over{Y_2}}-{2\over{m}}Y_1Y_2\biggr]
1033: {\rm d}t+{\rm d}\zeta_2\ ,
1034: \end{eqnarray}
1035: %
1036: %
1037: with
1038: %
1039: %
1040: \begin{eqnarray}
1041: \label{eq: 3.14}
1042: \langle {\rm d}\zeta_i\rangle&=&0\ ,\qquad
1043: \langle {\rm d}\zeta_i{\rm d}\zeta_j\rangle=2D_{ij}{\rm d}t\ .
1044: \end{eqnarray}
1045: %
1046: %
1047: The diffusion constants $D_{ij}$ were defined in equation
1048: (\ref{eq: 3.4}). Note that in two dimensions, however, (\ref{eq: 3.1})
1049: remains valid because the term arising from the rotation
1050: of ${\bf n}_3$ is absent.
1051:
1052:
1053: Now consider the evaluation of the diffusion constants
1054: in terms of the statistics of the force ${\bf f}({\bf r},t)$.
1055: The elements of the force-gradient matrix $\tilde F$ are
1056: %
1057: %
1058: \begin{equation}
1059: \label{eq: 3.15}
1060: F_{ij}={\partial^2\phi\over{\partial x_i\partial x_j}}+
1061: \epsilon_{ilk}{\partial^2 A_k\over{\partial x_j\partial x_l}}
1062: \end{equation}
1063: %
1064: %
1065: where $\epsilon_{ilk}$ is the ``Kronecker $\epsilon$-symbol'' describing
1066: the parity of the permutation of the indices $ilk$. Define
1067: %
1068: %
1069: \begin{equation}
1070: \label{eq: 3.16}
1071: D_0={\textstyle{1\over 2}}\int_{-\infty}^\infty {\rm d}t\
1072: \biggl\langle {\partial^2\phi\over{\partial x^2}}(t)
1073: {\partial^2 \phi\over{\partial x^2}}(0)\biggr\rangle
1074: \ .
1075: \end{equation}
1076: %
1077: %
1078: Then define $D_1=D_{11}$, $D_2=D_{21}=D_{31}$, so that
1079: %
1080: %
1081: \begin{eqnarray}
1082: \label{eq: 3.17}
1083: D_1&=&D_0\biggl(1+{2\alpha^2\over 3}\biggr)
1084: \qquad\mbox{and}\qquad
1085: D_2=D_0\biggl({1\over 3}+{4\alpha^2\over 3}\biggr)
1086: \ .
1087: \end{eqnarray}
1088: %
1089: %
1090: We introduce a more convenient dimensionless measure of the relative
1091: importance of solenoidal and potential fields
1092: %
1093: %
1094: \begin{equation}
1095: \label{eq: 3.18}
1096: \Gamma\equiv {D_2\over {D_1}}={1+4\alpha^2\over{3+2\alpha^2}}
1097: \end{equation}
1098: %
1099: %
1100: and find ${1\over 3}\le \Gamma \le 2$ in the three-dimensional
1101: case because $0\le \alpha\le \infty$.
1102: It is convenient to re-scale the Langevin equations into dimensionless
1103: form: write
1104: %
1105: %
1106: \begin{equation}
1107: \label{eq: 3.19}
1108: {\rm d}t' =\gamma {\rm d}t
1109: \ ,\ \ \
1110: x_i=\sqrt{\gamma\over{D_i}}Y_i
1111: \ ,\ \ \
1112: {\rm d}w_i=\sqrt{\gamma\over{D_i}}{\rm d}\zeta_i
1113: \end{equation}
1114: %
1115: %
1116: and define
1117: %
1118: %
1119: \begin{equation}
1120: \label{eq: 3.19a}
1121: \epsilon={D_1^{1/2}\over{m\gamma^{3/2}}}
1122: \ .
1123: \end{equation}
1124: %
1125: %
1126: With these changes of variables, the Langevin equations become
1127: %
1128: %
1129: \begin{eqnarray}
1130: \label{eq: 3.20}
1131: {\rm d}x_1&=&[-x_1+\epsilon(\Gamma x_2^2-x_1^2)]{\rm d}t'+{\rm d}w_1
1132: \\
1133: %
1134: %
1135: {\rm d}x_2&=&[-x_2+x_2^{-1}-2\epsilon x_1x_2]{\rm d}t'+{\rm d}w_2
1136: \nonumber
1137: \end{eqnarray}
1138: %
1139: %
1140: with
1141: %
1142: %
1143: \begin{eqnarray}
1144: \label{eq: 3.21}
1145: \langle {\rm d}w_i\rangle&=&0\,,
1146: \qquad
1147: \langle {\rm d}w_i{\rm d}w_j\rangle=2\delta_{ij}{\rm d}t'
1148: \ .
1149: \end{eqnarray}
1150: %
1151: %
1152: Eqs.~(\ref{eq: 3.20},\ref{eq: 3.21}) must be solved to determine the expectation
1153: value of $x_1$ in the steady state. The Lyapunov exponent is then given by
1154: %
1155: %
1156: \begin{equation}
1157: \label{eq: 3.22}
1158: \lambda_1=\gamma \epsilon \langle x_1 \rangle
1159: \ .
1160: \end{equation}
1161: %
1162: %
1163: \par
1164: Figure~\ref{fig: 2}{\bf a} compares the Lyapunov exponent
1165: obtained from a Monte Carlo simulation of equations
1166: (\ref{eq: 3.20}--\ref{eq: 3.22}) with
1167: a direct numerical simulation of a random flow
1168: described by equation (\ref{eq: 1.1}).
1169: The Lyapunov exponents
1170: determined from eqs.~(\ref{eq: 3.20}) and (\ref{eq: 3.21})
1171: for $\Gamma = 1/3,1$ and $2$ are plotted as red lines.
1172: The results
1173: are compared to numerical simulations of (\ref{eq: 1.1}),
1174: using a method described
1175: in \cite{Eck79} to determine the Lyapunov exponent.
1176: Because we are concerned with the limit where the
1177: correlation time $\tau$ is taken to zero, the random
1178: flow was generated using a discrete series
1179: of uncorrelated random impulses, acting over
1180: a small time step $\delta t \gg \tau$: the impulse
1181: %
1182: %
1183: \begin{equation}
1184: \label{eq: 3.23}
1185: {\bf f}_n({\bf r}) =\int_{n\delta t}^{(n\!+\!1)\delta t}\!\!
1186: {\rm d}t'\, {\bf f}({\bf r}_{t'},t')
1187: \end{equation}
1188: %
1189: %
1190: at time $n\,\delta t$ is taken to
1191: be of the form (\ref{eq: 1.2}) in terms of scalar fields
1192: $\phi_n({\bf r})$ and ${\bf A}_n({\bf r})$ satisfying
1193: %
1194: %
1195: \begin{equation}
1196: \label{eq: 3.24}
1197: \langle \phi_n({\bf r})\phi_{n'}({\bf r}')\rangle
1198: =\sigma^2\, \xi^2\, \delta
1199: t\,\exp(|{\bf r}-{\bf r}'|^2/2\xi^2)\delta_{nn'}
1200: \end{equation}
1201: %
1202: %
1203: and similarly for ${\bf A}_n({\bf r})$. This implies
1204: $D_0 = 3\sigma^2/(2m^2\gamma^3\xi^2)$.
1205:
1206:
1207: Now we discuss the conditions under which the Langevin
1208: equations (\ref{eq: 3.20}) and (\ref{eq: 3.21}) are a valid
1209: approximation of (\ref{eq: 2.12}) and (\ref{eq: 2.13}).
1210: For this purpose it is sufficient to consider the one-dimensional
1211: version of equations (\ref{eq: 3.13}), namely
1212: %
1213: %
1214: \begin{equation}
1215: \label{eq: 3.25}
1216: \dot Y=-\gamma Y-{1\over m}Y^2+F(t)
1217: \end{equation}
1218: %
1219: %
1220: (this equation appears with a different notation in \cite{Wil03}).
1221: The Langevin equations are valid provided the changes
1222: in the value of $Y$ over the correlation
1223: time $\tau$ is small compared to the typical values of
1224: this quantity. This criterion can obviously only be
1225: satisfied if the correlation time is sufficiently short
1226: that $\nu =\gamma\tau\ll 1$. The criterion also requires
1227: the stochastic force $F(t)$ to be sufficiently weak.
1228: The deterministic part of the velocity, $-\gamma Y-Y^2/m$, is
1229: positive in the interval from $Y=-\gamma m$ to $Y=0$.
1230: The criterion on the strength of $F\sim \sigma/\xi$ is
1231: that the displacement over time $\tau$ should be small compared
1232: to the width of that interval, that is $\vert F\vert \tau\ll \gamma m$.
1233: Using the fact that $\vert F\vert\sim \sigma/\xi$, we obtain
1234: the following criteria for the validity of the Langevin
1235: approximation:
1236: %
1237: %
1238: \begin{equation}
1239: \label{eq: 3.26}
1240: {\chi\over {\nu }}\ll 1 \ ,\ \ \ \nu\ll 1.
1241: \end{equation}
1242: %
1243: %
1244: \par
1245: For completeness, we end this section by mentioning how
1246: equations (\ref{eq: 3.20}) and (\ref{eq: 3.21}) differ in one
1247: and -two dimensions. The one-dimensional case was considered
1248: in \cite[]{Wil03}: the Lyapunov exponent is given by
1249: $\lambda=\langle Y\rangle/m$, with $Y$ satisfying (\ref{eq: 3.25}).
1250: In two dimensions, as we have already remarked,
1251: the term $x_2^{-1}$ is absent
1252: from the second equation of (\ref{eq: 3.20}), and
1253: ${1\over 3}\le \Gamma\le 3$ \cite[]{Meh04}).
1254:
1255:
1256: \section{Perturbation theory}
1257: \label{sec: 4}
1258: We now show how to obtain an asymptotic approximation
1259: for the Lyapunov exponent using eqs.~(\ref{eq: 3.20}) and
1260: (\ref{eq: 3.21}). These equations are equivalent to a
1261: two-dimensional Fokker-Planck
1262: equation (\cite{vKa92}) for a probability density $P(x_1,x_2;t')$, of the form
1263: %
1264: %
1265: \begin{equation}
1266: \label{eq: 4.1}
1267: \partial_{t'} P=D\nabla^2P-\nabla.({\bf v}P)=\hat {\cal F} P
1268: \ .
1269: \end{equation}
1270: %
1271: %
1272: Here the diffusion constant
1273: $D=1$ and the drift velocity is ${\bf v}=(v_1,v_2)$ with components
1274: $v_1=-x_1+\epsilon (\Gamma x_2^2-x_1^2)$
1275: and $v_2=-x_2+x_2^{-1}-2\epsilon x_1x_2$.
1276: We write $\hat {\cal F}=\hat {\cal F}_0+\epsilon \hat {\cal F}_1$, and
1277: seek a steady-state solution satisfying $\hat {\cal F}P=0$ by
1278: perturbation theory in $\epsilon$.
1279: In order to simplify the application of perturbation theory,
1280: it is convenient to make a transformation so that the unperturbed
1281: Fokker-Planck operator $\hat {\cal F}_0$ is transformed into a
1282: Hermitian operator. Rather than proceeding to the Hermitian
1283: form directly, we first map the two-dimensional Fokker-Planck
1284: equation to a three-dimensional equation with a rotational symmetry
1285: (we seek a solution which is invariant under rotation).
1286: After making this transformation, we find that the corresponding
1287: Hermitian operator in three-dimensional space is the Schr\" odinger
1288: operator of an isotropic three-dimensional harmonic oscillator.
1289: The perturbation analysis can then be performed very easily, using
1290: the algebra of harmonic-oscillator raising and lowering operators,
1291: described in \cite[]{Dir30}. To shorten equations,
1292: we will use a variant of the Dirac
1293: notation scheme: in summary, functions $a$, $b$ are symbolised
1294: by vectors $\vert a)$, $\vert b)$, linear operators are
1295: denoted by a \lq hat', e.g. $\hat{\cal A}$, and the integral
1296: over all space of the product of two functions is denoted by
1297: the inner product $(a\vert b)$.
1298:
1299:
1300: In the original form, the action of the unperturbed part of the
1301: Fokker-Planck operator on a function $P$ is
1302: %
1303: %
1304: \begin{equation}
1305: \label{eq: 4.2}
1306: \hat {\cal F}_0 P=(\partial_1^2+\partial_2^2)P
1307: +\partial_1(x_1P)+\partial_2[(x_2-x_2^{-1})P]
1308: \ .
1309: \end{equation}
1310: %
1311: %
1312: We transform this by defining the action of $\hat{\cal F}_0'$ on
1313: a function $P'=P/x_2$ as follows
1314: %
1315: %
1316: \begin{eqnarray}
1317: \label{eq: 4.3}
1318: \hat{\cal F}_0'P'&=&{1\over{x_2}}\hat {\cal F}_0 P
1319: \nonumber \\
1320: %
1321: %
1322: &=&{1\over{x_2}}(\partial_1^2+\partial_2^2)(x_2P')
1323: +{1\over{x_2}}\partial_1(x_1x_2P')+{1\over{x_2}}\partial_2[(x_2^2-1)P']
1324: \nonumber \\
1325: %
1326: %
1327: &=&\partial_1[(\partial_1+x_1)P']
1328: +{1\over{x_2}}\partial_2[x_2(\partial_2+x_2)P']
1329: \ .
1330: \end{eqnarray}
1331: %
1332: %
1333: We now consider $\hat{\cal F}'_0$ to be an operator acting
1334: in three-dimensional space, with cylindrical polar coordinates
1335: $(r,\varphi,z)$. We identify $r=x_2$, and $z=x_1$, and take $P'$
1336: to be a function which is restricted so that it
1337: has cylindrical symmetry, being independent of $\varphi$.
1338: With this interpretation, we can add differentials with respect to
1339: $\phi$ to the definition of $\hat {\cal F}_0'$, and write
1340: %
1341: %
1342: \begin{eqnarray}
1343: \label{eq: 4.4}
1344: \hat {\cal F}_0'&=&{1\over r}\partial_r\bigl[r(\partial_r+r)\bigr]+{1\over{r^2}}
1345: \partial_\varphi^2+\partial_z(\partial_z+z)
1346: =\nabla.({\bf x}+\nabla)
1347: \end{eqnarray}
1348: %
1349: %
1350: which is the Fokker-Planck operator for isotropic diffusion
1351: in three-dimensional space (with $D=1$), with a drift velocity
1352: ${\bf v}=-{\bf x}$. Thus we have transformed the two-dimensional
1353: Fokker-Planck equation to a three-dimensional one with a very
1354: simple unperturbed velocity field. It is convenient
1355: to work with Cartesian coordinates ${\bf x} = (x,y,z)$ in the three-dimensional
1356: space, having the usual relation to the cylindrical polar coordinates $(r,\varphi,z)$.
1357: The Fokker-Planck equation is then
1358: %
1359: %
1360: \begin{eqnarray}
1361: \label{eq: 4.5}
1362: \partial_{t'} P'&=&\nabla^2 P'+\nabla.[({\bf r}-\epsilon{\bf v}_1')P']
1363: =\hat {\cal F}'P'=[\hat {\cal F}_0+\epsilon \hat {\cal F}_1]P'
1364: \end{eqnarray}
1365: %
1366: %
1367: where, in Cartesian coordinates,
1368: ${\bf v}_1'$ has components
1369: %
1370: %
1371: \begin{eqnarray}
1372: \nonumber
1373: {v}_{11}'& =&-2xz\,,\\
1374: \label{eq: 4.6}
1375: v_{12}' &=&-2yz\,,\\
1376: \nonumber
1377: v_{13}' &=&-z^2+\Gamma (x^2+y^2)\,.
1378: \end{eqnarray}
1379: %
1380: %
1381: We now transform the Fokker-Planck $\hat {\cal F}'$ operator so that
1382: $\hat {\cal F}_0'$ is transformed into a very simple Hermitian operator,
1383: by writing
1384: %
1385: %
1386: \begin{equation}
1387: \label{eq: 4.7}
1388: \hat {\cal H}=\exp(\Phi_0/2)\hat {\cal F}'\exp(-\Phi_0/2)
1389: \qquad\mbox{with}\qquad
1390: \Phi_0={\textstyle{1\over 2}}(x^2+y^2+z^2)
1391: \ .
1392: \end{equation}
1393: %
1394: %
1395: We find (on writing $(x,y,z)=(z_1,z_2,z_3)$)
1396: %
1397: %
1398: \begin{equation}
1399: \label{eq: 4.8}
1400: \hat {\cal H}_0=\exp(\Phi_0/2)\hat {\cal F}_0'\exp(-\Phi_0/2)
1401: =\sum_{j=1}^3 [\partial^2_{z_j}-{\textstyle{1\over 4}}z_j^2
1402: +{\textstyle{1\over 2}}]
1403: \end{equation}
1404: %
1405: %
1406: so that $\hat {\cal H}_0$ is (apart from a negative
1407: multiplicative factor) the Hamiltonian operator for a
1408: three-dimensional harmonic oscillator.
1409: The spectrum of $\hat {\cal H}_0$ is the set of non-positive
1410: integers $(0,-1,-2,-3,..)$. The eigenfunctions of $\hat {\cal H}_0$
1411: are generated by raising and lowering operators \cite[]{Dir30}:
1412: %
1413: %
1414: \begin{equation}
1415: \label{eq: 4.9}
1416: \hat a_i={\textstyle{1\over 2}}z_i+\partial_{z_i}
1417: \ ,\ \ \
1418: \hat a_i^+={\textstyle{1\over 2}}z_i-\partial_{z_i}
1419: \ .
1420: \end{equation}
1421: %
1422: %
1423: The transformed perturbation operator is
1424: %
1425: %
1426: \begin{eqnarray}
1427: \label{eq: 4.10}
1428: \hat {\cal H}_1&=&-\sum_{j=1}^3 \hat a_j^+ \hat v'_{1j}
1429: \nonumber \\
1430: %
1431: %
1432: &=&2\hat a_1^+\hat z_1\hat z_3+2\hat a_2^+\hat z_2\hat z_3
1433: -\hat a_3^+[\hat z_3^2-\Gamma(\hat z_1^2+\hat z_2^2)]
1434: \ .
1435: \end{eqnarray}
1436: %
1437: %
1438: Instead of solving the Fokker-Planck equation $\hat {\cal F}'P'=0$
1439: we attempt to solve $\hat {\cal H}\,Q=0$, where $Q=\exp(\Phi_0/2)P'$.
1440:
1441: Now consider how to obtain the Lyapunov exponent from
1442: the function $Q$. We have $\lambda_1=\gamma \epsilon \langle z_3\rangle$.
1443: We calculate the average of $z_3$ as follows
1444: %
1445: %
1446: \begin{eqnarray}
1447: \label{eq: 4.11}
1448: \langle z_3\rangle&=&\int_0^\infty {\rm d}r\int_{-\infty}^\infty {\rm d}z\
1449: P(r,z)
1450: \nonumber \\
1451: %
1452: %
1453: &=&\int_0^\infty r\,{\rm d}r\int_0^{2\pi}{\rm d}\varphi\int_{-\infty}^\infty
1454: {\rm d}z\ z\, P'(r,z)
1455: \nonumber \\
1456: %
1457: %
1458: &=&\int_{-\infty}^\infty \int_{-\infty}^\infty \int_{-\infty}^\infty
1459: {\rm d}x\ {\rm d}y\ {\rm d}z\ z\, \exp(-\Phi_0/2)Q(x,y,z)
1460: \ .
1461: \end{eqnarray}
1462: %
1463: %
1464: Now we change the notation, using a variant of the Dirac notation
1465: to represent the function $Q$ by a \lq ket vector' $\vert Q)$.
1466: Allowing for the possibility that $\vert Q)$ is not normalised,
1467: we write
1468: %
1469: %
1470: \begin{equation}
1471: \label{eq: 4.12}
1472: \langle z_3\rangle
1473: ={(\phi_{000}\vert \hat z_3\vert Q)\over{(\phi_{000}\vert Q)}}
1474: ={(\phi_{000}\vert \hat a_3\vert Q)\over{(\phi_{000}\vert Q)}}
1475: \end{equation}
1476: %
1477: %
1478: where $\vert \phi_{000})$ is the ground-state eigenfunction
1479: of ${\cal H}_0$, given by the function
1480: $\exp(-\Phi_0/2)=\exp[-(z_1^2+z_2^2+z_3^2)/4]$.
1481:
1482:
1483: We calculate $\vert Q)$ by perturbation theory: writing
1484: %
1485: %
1486: \begin{equation}
1487: \label{eq: 4.13}
1488: \vert Q)=\vert Q_0)+\epsilon \vert Q_1)+\epsilon^2\vert Q_2)+...
1489: \end{equation}
1490: %
1491: %
1492: we find that the functions $\vert Q_k)$ satisfy the recursion relation
1493: %
1494: %
1495: \begin{equation}
1496: \label{eq: 4.14}
1497: \vert Q_{k+1})=-\hat {\cal H}_0^{-1}\hat {\cal H}_1 \vert Q_k)
1498: \ .
1499: \end{equation}
1500: %
1501: %
1502: At first sight this appears to be ill-defined because one of the
1503: eigenvalues of ${\cal H}_0$ is zero, so that the inverse of
1504: ${\cal H}_0$ is only defined for the subspace of functions which
1505: are orthogonal to the ground state, $\vert \phi_{000})$.
1506: However, because all of the components of ${\cal H}_1$ have
1507: a creation operator as a left factor, the function $\hat{\cal H}_1\vert \psi)$
1508: is orthogonal to $\vert \psi)$ for any function $\vert \psi )$, so
1509: that (\ref{eq: 4.14}) is in fact well-defined.
1510: The iteration starts with $\vert Q_0)=\vert \phi_{000})$.
1511: The functions $\vert Q_k)$ should all have rotational symmetry
1512: about the $z$-axis. The angular-momentum operator
1513: $\hat {\cal J}_3=\hat p_1 \hat z_2-\hat p_2 \hat z_1$
1514: commutes with both $\hat {\cal H}_0$ and $\hat {\cal H}_1$
1515: (that is, $[\hat {\cal H}_0,\hat {\cal J}_3]=0$ and
1516: $[\hat {\cal H}_1,\hat {\cal J}_3]=0$
1517: where we use square brackets for the commutator,
1518: $[\hat {\cal A},\hat {\cal B}]=
1519: \hat {\cal A}\hat {\cal B}-\hat {\cal B}\hat {\cal A}$).
1520: The operators $\hat {\cal H}_0$ and $\hat {\cal H}_1$ can therefore
1521: be simultaneously reduced to block diagonal form, with blocks
1522: labelled by eigenvalues of $\hat {\cal J}_3$. We can restrict
1523: ourselves to the subspace where the eigenvalue of $\hat {\cal J}_3$
1524: is zero. The functions of this subspace
1525: are generated from the ground state using transformed raising
1526: and lowering operators, defined as follows:
1527: %
1528: %
1529: \begin{eqnarray}
1530: \label{eq: 4.15}
1531: \hat{\alpha}_+ &=& \frac{1}{\sqrt{2}} (\hat{a}_1 + {\rm i}
1532: \hat{a}_2)\,,\qquad
1533: \hat{\alpha}_- = \frac{1}{\sqrt{2}} (\hat{a}_1 - {\rm i} \hat{a}_2)
1534: \ .
1535: \end{eqnarray}
1536: %
1537: %
1538: The transformed operators $\hat a_+$ and $\hat a_-$ satisfy
1539: $[\hat a_\pm,\hat a^\dagger_\pm]=\hat I$ (where $\hat I$ is the
1540: identity operator), which is the fundamental
1541: relation describing harmonic-oscillator raising and lowering
1542: operators.
1543: Expressing $\hat {\cal H}_0$ and $\hat {\cal J}_3$ in terms
1544: of these operators, we find
1545: %
1546: %
1547: \begin{equation}
1548: \label{eq: 4.16}
1549: \hat{\cal J}_3 = \hat{\alpha}_-^\dagger\hat{\alpha}_-
1550: - \hat{\alpha}_+^\dagger \hat{\alpha}_+
1551: \ ,\ \ \
1552: \hat{\cal H}_0 = -(\hat{\alpha}_-^\dagger\hat{\alpha}_-
1553: + \hat{\alpha}_+^\dagger \hat{\alpha}_+ +
1554: a_3^\dagger a_3)\,.
1555: \end{equation}
1556: %
1557: %
1558: Using results from \cite{Dir30}, we see that
1559: both $\hat {\cal H}_0$ and $\hat {\cal J}_3$ are linear combinations
1560: of harmonic-oscillator Hamiltonians, $\hat a_-^\dagger\hat a_-$,
1561: $\hat a^\dagger_+\hat a_+$ and $\hat a^\dagger_3\hat a_3$.
1562: The $n$th eigenfunction $\vert\phi_n)$ of a harmonic-oscillator Hamiltonian
1563: $\hat a^\dagger\hat a$ is obtained from its ground state $\vert \phi_0)$
1564: by repeated application of the raising operator $\hat a^\dagger$:
1565: %
1566: %
1567: \begin{equation}
1568: \label{eq: 4.16a}
1569: \vert \phi_n)={1\over{\sqrt{n!}}}(\hat a^\dagger)^n\vert \phi_0)
1570: \end{equation}
1571: %
1572: %
1573: and this eigenfunction has eigenvalue $n$.
1574: Thus eigenfunctions of $\hat{\cal H}_0$ and $\hat{\cal J}_3$
1575: with zero angular momentum are constructed as follows
1576: %
1577: %
1578: \begin{equation}
1579: \label{eq: 4.17}
1580: |\psi_{nm}) = \frac{1}{n!} \frac{1}{\sqrt{m!}}
1581: (\hat \alpha_-^\dagger)^n
1582: (\hat \alpha_+^\dagger)^n
1583: (\hat a_z^\dagger)^m |\phi_{000})
1584: \end{equation}
1585: %
1586: %
1587: for $n=0,1,\ldots$ and $m=0,1,\ldots$. The corresponding
1588: eigenvalues of $\hat {\cal H}_0$ are $-2n-m$.
1589: The functions $\vert Q_k)$
1590: are expanded in terms of the $\vert \psi_{nm})$,
1591: with coefficients $a_{nm}^{(k)}$:
1592: %
1593: %
1594: \begin{equation}
1595: \label{eq: 4.18}
1596: \vert Q_k)=\sum_{n=0}^\infty \sum_{m=0}^\infty a^{(k)}_{nm}\vert
1597: \psi_{nm})
1598: \ .
1599: \end{equation}
1600: %
1601: %
1602: By projecting equation (\ref{eq: 4.14}) onto the vector
1603: $\vert\psi_{nm})$ and using the fact that the eigenvectors
1604: $\vert \psi_{n'm'})$ of $\hat {\cal H}_0$ form a complete
1605: basis, the iteration
1606: can be expressed as follows [for $(n,m)\ne (0,0)$]:
1607: %
1608: %
1609: \begin{equation}
1610: \label{eq: 4.19}
1611: a^{(k+1)}_{nm}
1612: =\sum_{n'=0}^\infty\sum_{m'=0}^\infty
1613: \frac{(\psi_{nm}\vert \hat {\cal H}_1\vert \psi_{n'm'})}{2n+m}
1614: a^{(k)}_{n',m'}
1615: \ .
1616: \end{equation}
1617: %
1618: %
1619: The matrix elements $(\psi_{nm}\vert\hat{\cal H}_1\vert\psi_{n'm'})$
1620: are readily computed using the algebraic properties of the
1621: raising and lowering operators, as discussed in \cite{Dir30}.
1622: The coefficients $a^{(k)}_{nm}$ are then calculated recursively.
1623: This allows us to obtain the functions $|Q_k)$.
1624: The lowest order is
1625: $|Q_0) = |\phi_{000})$. Its contribution to $\lambda_1$
1626: vanishes in view of (\ref{eq: 4.10}). The leading order is
1627: %
1628: %
1629: \begin{eqnarray}
1630: \label{eq: 4.20}
1631: |Q_1)
1632: &=& -\frac{4}{3} |\psi_{11}) - |\psi_{01}) - \frac{\sqrt{6}}{3} |\psi_{03})
1633: +2\Gamma |\psi_{01}) + \frac{2\Gamma}{3} |\psi_{11})\,.
1634: \end{eqnarray}
1635: %
1636: %
1637: The next order, $|Q_2)$, does not contribute to $\lambda_1$ since
1638: $\hat{\cal H}_1|Q_1)$ does not
1639: contain $|\psi_{01})$. In fact,
1640: only odd orders contain $|\psi_{01})$ and thus
1641: give non-zero contributions
1642: to $\lambda_1$. We also find that the denominator
1643: in (\ref{eq: 4.13}) is unity at all orders.
1644: The final result is:
1645: %
1646: %
1647: \begin{equation}
1648: \label{eq: 4.21}
1649: \lambda_1 =\gamma\epsilon \sum_{l=1}^\infty c_l(\Gamma)\, \epsilon^{2l-1}
1650: \end{equation}
1651: %
1652: %
1653: where the first five coefficients $c_l(\Gamma)$ are
1654: %
1655: %
1656: \begin{eqnarray}
1657: \label{eq: 4.22}
1658: c_1(\Gamma) &=& -1+2\Gamma \\
1659: c_2(\Gamma) &=& -5+20\, \Gamma-16\,\Gamma^2 \nonumber\\
1660: c_3(\Gamma) &=& -60+360\,\Gamma-568\,\Gamma^2+272\,\Gamma^3 \nonumber\\
1661: c_4(\Gamma) &=& -1105 + 8840\,\Gamma
1662: -61936/3\,\Gamma^2+58432/3\,\Gamma^3 -19648/3\,\Gamma^4
1663: \nonumber\\
1664: c_5(\Gamma)
1665: &=&-27120+271200\,\Gamma-7507040/9\,\Gamma^2+3492160/3\,\Gamma^3\nonumber\\
1666: &&\hspace*{5mm}-2316032/3\,\Gamma^4+ 1785856/9 \,\Gamma^5
1667: \nonumber\\
1668: &\vdots&\nonumber
1669: \end{eqnarray}
1670: %
1671: %
1672: The coefficients in (\ref{eq: 4.22}) have a growth which is
1673: typical of asymptotic series, as discussed by \cite{Din74}.
1674: Figure~\ref{fig: 2}{\bf b} shows
1675: approximations to the Lyapunov exponent $\lambda_1$ for
1676: $\Gamma = 0.45$. Shown are six different
1677: partial sums of the series (\ref{eq: 4.22}), including terms
1678: up to $l_{\rm max}$, with $l_{\rm max} = 1,\ldots,6$.
1679: For a given value of $\epsilon$, there is an optimal
1680: value of $l_{\rm max}$, which we term $l_{\rm max}^\ast$,
1681: defined by the criterion that the term in (\ref{eq: 4.22})
1682: with index $l_{\rm max}^\ast$
1683: is smallest in magnitude. The function $l_{\rm max}^\ast(\epsilon)$
1684: can be inverted, its inverse $\epsilon^\ast(l_{\rm max})$
1685: is the value of $\epsilon$ for which the $l_{\rm max}$ term is optimal.
1686: For values of $\epsilon$
1687: less than the $\epsilon^\ast(l_{\rm max})$ the results
1688: are shown as solid lines. Beyond this optimal value
1689: of $\epsilon$, the results are shown as dashed lines.
1690: The results show that the series agrees well with the numerical
1691: simulation up $\epsilon^\ast$, as would be expected for an
1692: asymptotic series.
1693:
1694:
1695: %One notable difference between the three-dimensional calculation
1696: %presented here and the two-dimensional case considered in
1697: %\cite[]{Meh04} is that in the two-dimensional case the
1698: %phase boundary has a power series in $\epsilon$ which vanishes
1699: %identically (and the phase line is therefore non-analytic).
1700:
1701:
1702: \begin{figure}
1703: \vspace*{4mm}
1704: \centerline{\includegraphics[width=\textwidth,clip]{fig2.eps}}
1705: \caption{\label{fig: 2}
1706: {\bf a} Lyapunov exponent as a function
1707: of $\epsilon$: results
1708: from eqs.~(\ref{eq: 3.20}), (\ref{eq: 3.21}) and (\ref{eq: 3.22})
1709: are shown as solid lines, those
1710: from direct simulations as symbols, $\Gamma=1/3$ ($\circ$), $\Gamma=1$
1711: ($\Box$), and $\Gamma=2$ ($\diamond$).
1712: {\bf b} Lyapunov exponent
1713: $\lambda_1/(\epsilon\gamma)$ versus $\epsilon$ for $\Gamma=0.45$.
1714: Shown are results from simulations ($\circ$) as well as from
1715: the asymptotic series (\ref{eq: 4.22}) summed to orders
1716: $l_{\rm max}=1,\ldots,6$: up to $\epsilon^\ast(l_{\rm max})$ full lines
1717: and from $\epsilon^\ast(l_{\rm max})$ dashed lines.
1718: {\bf c} Phase diagram in the $\epsilon$-$\Gamma$-plane.
1719: Shown are results from numerical simulations of (\ref{eq: 2.1}),
1720: $\circ$, summation of the asymptotic series (\ref{eq: 4.22}) summed to
1721: $k^\ast(\epsilon,\Gamma)$, blue line, as well as results
1722: from Langevin simulations (red line).
1723: }
1724: \end{figure}
1725:
1726:
1727: The phase boundary in the $\epsilon$-$\Gamma$-plane is determined
1728: by the condition $\lambda_1= 0$. Figure~\ref{fig: 2}{\bf c} shows
1729: this phase boundary as determined from truncating
1730: the series (\ref{eq: 4.22}) at the optimal order
1731: $l_{\rm max}^\ast(\epsilon)$ (blue line). This asymptotic
1732: result is shown for values of $\epsilon$ up to $\approx 0.2$.
1733: Beyond this range, the asymptotic approximation becomes
1734: increasingly inaccurate. Also shown, in the same plot,
1735: are results obtained from the Langevin equations
1736: [eqs.~(\ref{eq: 3.20}), (\ref{eq: 3.21}) and (\ref{eq: 3.22})],
1737: and from direct simulations ($\circ$). The results show that the
1738: coalescing phase disappears as the effect of inertia is
1739: increased: the coalescing disappears altogether
1740: for the case of pure potential flow ($\Gamma={1\over 3}$) at
1741: $\epsilon\approx 0.33$.
1742:
1743:
1744: One notable difference between the three-dimensional calculation
1745: presented here and the two-dimensional case considered in
1746: \cite[]{Meh04} is that in the two-dimensional case the
1747: phase boundary has a power series in $\epsilon$ which vanishes
1748: identically (and the phase line is therefore non-analytic).
1749:
1750:
1751: \section{Effect of dispersion of particle masses}
1752: \label{sec: 5}
1753:
1754:
1755: In most naturally occurring aerosols the suspended particles
1756: have different mass $m$, and particles of differing sizes
1757: will also have different values of the damping coefficient
1758: $\gamma$. It is important to consider whether particles
1759: still have a tendency to coalesce even when the particles have
1760: differing values of $m$ and $\gamma$: we argue that
1761: the path coalescence effect is not destroyed by small
1762: mass dispersion. The argument can be adapted to
1763: dispersion of the damping coefficient, reaching the same
1764: conclusion.
1765:
1766:
1767: Assume that the path-coalescence effect occurs for particles of
1768: mass $m$. Compare the motion of this reference particle with that
1769: of an initially nearby particle with mass $m+\delta m$.
1770: The reference particle has equation of motion
1771: %
1772: %
1773: \begin{equation}
1774: \label{eq: 5.1}
1775: m\ddot x=-\gamma m\dot x+f(x,t)
1776: \ .
1777: \end{equation}
1778: %
1779: %
1780: Writing $f(x+\delta x,t)=f(x,t)+F(t)\delta x+O(\delta x^2)$, the
1781: equation of motion for the other particle is
1782: %
1783: %
1784: \begin{equation}
1785: \label{eq: 5.2}
1786: (m+\delta m)(\ddot x+\delta \ddot x)
1787: =-\gamma m(\dot x +\delta \dot x)+f(x,t)+F(t)\delta x +O(\delta x^2)
1788: \ .
1789: \end{equation}
1790: %
1791: %
1792: Collecting the terms which are first order in $\delta x$, we
1793: obtain a linearised equation of motion for $\delta x$:
1794: %
1795: %
1796: \begin{equation}
1797: \label{eq: 5.3}
1798: -m\delta \ddot x-\gamma m\delta \dot x+F(t)\delta x
1799: =\delta m\, \ddot x
1800: \ .
1801: \end{equation}
1802: %
1803: %
1804: This is an inhomogeneous differential equation for the separation
1805: $\delta x$ between two particles, with a driving term proportional
1806: to their mass difference $\delta m$.
1807: The solution of this equation can be constructed from a Green's
1808: function satisfying $G(t,t_0)$
1809: %
1810: %
1811: \begin{equation}
1812: \label{eq: 5.4}
1813: -m{{\rm d}^2G\over{{\rm d}t^2}}
1814: -\gamma m{{\rm d}G\over{{\rm d} t}}+F(t)G=\delta(t-t_0)
1815: \end{equation}
1816: %
1817: %
1818: with $G(t,t_0)=0$ for $t<t_0$. The solution of equation (\ref{eq: 5.3}) is
1819: %
1820: %
1821: \begin{equation}
1822: \label{eq: 5.5}
1823: \delta x(t)=\delta m\int_0^t{\rm d}t'\ G(t,t')
1824: {{\rm d}^2 x(t')\over{{\rm d}t'^2}}
1825: \ .
1826: \end{equation}
1827: %
1828: %
1829: For $t>t_0$, equation (\ref{eq: 5.4}) is the equation for small displacements
1830: of trajectories of particles with the same mass. We know that in
1831: the path-coalescing phase the solutions have a negative value of the
1832: Lyapunov exponent $\lambda_1$, and that they therefore decay
1833: exponentially at large time.
1834: In the case where $G(t,t')$ is bounded by an exponentially
1835: decreasing function, such that
1836: $\vert G(t,t')\vert <A\exp(-\lambda_1\vert t-t'\vert)$,
1837: equation (\ref{eq: 5.5}) remains finite as $t\to \infty$.
1838: For sufficiently large $A$, the probability of this inequality
1839: being violated is extremely small.
1840: This indicates that in the path-coalescing
1841: phase the solution (\ref{eq: 5.5}) remains finite as $t\to \infty$,
1842: except for very rare events. The conclusion is that, when
1843: $\lambda_1 <0$, two initially close particles with nearly
1844: equal mass will remain in close proximity for a very
1845: long time.
1846:
1847:
1848:
1849: \section{Discussion}
1850: \label{sec: 6}
1851:
1852:
1853: In this paper we described the path-coalescence transition,
1854: and showed that the transition point is determined by the change
1855: of sign of a Lyapunov exponent. We showed that in general the Lyapunov
1856: exponent is determined from an expectation value of a
1857: variable of a simple dynamical system, equations (\ref{eq: 2.12}),
1858: which is driven by stochastic forcing functions.
1859: We considered the solution of these equations in a particular
1860: limiting case, where the dimensionless parameters
1861: satisfy $\chi \ll \nu \ll 1$, by mapping the
1862: continuous differential equations into a pair of coupled Langevin
1863: equations. We used these Langevin equations to produce a rather
1864: complete description of the phase transition in that limit.
1865: The remainder of these concluding remarks
1866: discuss how equations (\ref{eq: 2.12}) can be used in the
1867: case where these inequalities are not satisfied.
1868:
1869:
1870: In order to solve these differential equations it is necessary
1871: to characterise the statistics of the stochastic driving terms
1872: $F'_{ij}(t)$. These terms contain information about the strain-rate
1873: of the field evaluated at a point along the reference particle
1874: trajectory. There are two possibilities:
1875:
1876:
1877: \begin{description}
1878:
1879:
1880: \item Case A: the statistics of the
1881: strain-rate tensor along a trajectory may be indistinguishable
1882: from those sampled along a randomly chosen trajectory.
1883:
1884:
1885: \item Case B: the
1886: trajectory may select regions where the strain-rate tensor
1887: has atypical properties, for example by tracking points where
1888: the velocity vector ${\bf u}$ vanishes.
1889:
1890:
1891: \end{description}
1892:
1893:
1894: If case A is realised there are two further possibilities:
1895:
1896:
1897: \begin{description}
1898:
1899:
1900: \item Case A1: the trajectory ${\bf r}(t)$ is sufficiently
1901: slowly moving that the displacement over time $\tau$ is small
1902: compared to $\xi$. In this case the statistics of $F'_{ij}$
1903: are those of a randomly chosen static point, and the correlation
1904: time of $F'_{ij}(t)$ will be $\tau$.
1905:
1906:
1907: \item Case A2: if the trajectory ${\bf r}(t)$ is moving sufficiently
1908: rapidly that its displacement in time $\tau $ is large compared
1909: to $\xi$, then the correlation time of $F'_{ij}(t)$ will be smaller
1910: than $\tau$ because the loss of correlations results primarily
1911: from changing the position at which $\partial u_i/\partial x_j({\bf r},t)$
1912: is sampled.
1913:
1914:
1915: \end{description}
1916:
1917:
1918: The limit which was investigated in detail in this paper
1919: ($\chi\ll\nu\ll 1$) is an example of case A1.
1920: In cases where $\chi$ and $\nu$ approach different limits however,
1921: all three possibilities can occur in the system described by
1922: equations (\ref{eq: 1.1}) to (\ref{eq: 1.3}).
1923: %, making this a very rich model.
1924:
1925:
1926:
1927: \bibliographystyle{jfm}
1928: \begin{thebibliography}{12}
1929: \bibitem[Falkovitch, {\sl et. al} (2001)]{Fal03}
1930: {\sc Balkovsky, E., Falkovitch, G. \& Fouxon, A.}, 2003,
1931: {Intermittent distribution of inertial particles in turbulent flows},
1932: {\it Phys. Rev. Lett.}, {\bf 86}, 2790-3.
1933:
1934:
1935: \bibitem[Baxendale \& Harris (1986)]{Bax86}
1936: {\sc Baxendale, P. \& Harris, T.}, 1986,
1937: {Isotropic stochastic flows},
1938: {\it Ann. Probab.}, {\bf 14}, 1155-79.
1939:
1940:
1941: \bibitem[Cressman {\sl et al.} (2004)]{Cre04}
1942: {\sc Cressman, J. R., Goldburg, W. I. \& Schuhmacher, J.}, 2004,
1943: {Dispersion of tracer particles in a compressible flow},
1944: {\it Europhys. Lett. } {\bf 66}, {219}.
1945:
1946:
1947: \bibitem[Deutsch (1985)]{Deu85}
1948: {\sc Deutsch, J. M.}, 1985,
1949: {Aggregation-disorder transitio induced by fluctuating random forces},
1950: {\it J. Phys. A}, {\bf 18}, 1457.
1951:
1952:
1953: \bibitem[Dingle (1973)]{Din74}
1954: {\sc Dingle, R. B.}, 1974,
1955: {Asymptotic expansions: their derivation and interpretation},
1956: Academic Press, New York.
1957:
1958:
1959: \bibitem[Dirac (1930)]{Dir30}
1960: {\sc Dirac, P. A. M.}, 1930,
1961: {The principles of quantum mechanics},
1962: Oxford University Press.
1963:
1964:
1965: \bibitem[Eckmann \& Ruelle (1979)]{Eck79}
1966: {\sc Eckmann, J-P. \& Ruelle, D.}, 1985,
1967: {\it Rev. Mod. Phys.}, {\bf 57}, 617-56.
1968:
1969:
1970: \bibitem[Elperin {\sl et al}(1996)]{Elp96}
1971: {\sc Elperin, T., Kleeorin, N. \& Rogachevskii, I.}, 1996,
1972: {Self-excitation of fluctuation of inertial particle concentration
1973: in turbulent fluid flow},
1974: {\it Phys. Rev. Lett.}, {\bf 77}, 5373.
1975:
1976:
1977: \bibitem[Fessler {\sl et al}(1994)]{Fes94}
1978: {\sc Fessler, J. R., Kulick, J. D. \& Eaton, J. K.}, 1994,
1979: {Preferential concentration of heavy particles in a
1980: turbulent channel flow},
1981: {\it Phys. Fluid}, {\bf 6}, {3742}.
1982:
1983:
1984: \bibitem[Frisch (1997)]{Fri97}
1985: {\sc Frisch, U.}, 1997,
1986: {Turbulence},
1987: Cambridge Univeristy Press.
1988:
1989:
1990: \bibitem[Klyatskin \& Gurarie (1999)]{Kly99}
1991: {\sc Klyatskin, V. I. \& Gurarie, D.}, 1999,
1992: {Coherent phenomena in stochastic dynamical systems},
1993: {\it Physics Uspekhi}, {\bf 42}, 165.
1994:
1995:
1996: \bibitem[Landau \& Lifshitz (1958)]{Lan58}
1997: {\sc Landau, L. D. \& Lifshitz, E. M.}, 1958,
1998: {Fluid Mechanics}, Pergamon, Oxford.
1999:
2000:
2001: \bibitem[Le Jan (1985)]{LeJ85}
2002: {\sc Le Jan, Y.}, 1985,
2003: {On isotropic Brownian motions},
2004: {\it Z. Wahsch. verw. Gebiete}, {\bf 70}, 609-20.
2005:
2006:
2007: \bibitem[Maxey (1987)]{Max87}
2008: {\sc Maxey, M. R.}, 1987,
2009: {The gravitational settling of aerosol particles in homogenous
2010: turbulence and random flow fields}
2011: {\it J. Fluid Mech.} {\bf 174}, 441-465.
2012:
2013:
2014: \bibitem[Mehlig \& Wilkinson (2004)]{Meh04}
2015: {\sc Mehlig, B. \& Wilkinson M.}, 2004,
2016: {Coagulation by random velocity fields as a Kramers problem},
2017: {\it Phys. Rev. Lett.}, {\bf 92}, 250602.
2018:
2019:
2020: \bibitem[Piterbarg (2001)]{Pit01}
2021: {\sc Piterbarg, L.}, 2001,
2022: {The top Lyapunov exponent for a stochastic flow modelling
2023: the upper ocean turbulence},
2024: {\it SIAM J. Appl. Math.}, {\bf 62}, 777-800.
2025:
2026:
2027: \bibitem[Shaw (2003)]{Sha03}
2028: {\sc Shaw, R. A.}, 2003,
2029: {Particle-turbulence interactions in atmospheric clouds},
2030: {\it Annu. Rev. Fluid Mech.}, {\bf 35}, 183-227.
2031:
2032:
2033: \bibitem[Sigurgeirsson \& Stuart (2002)]{Sig02}
2034: {\sc Sigurgeirsson, H. and Stuart, A. M.}, 2002,
2035: {A model for preferential concentration},
2036: {\it Phys. Fluids}, {\bf 14}, 4352-61.
2037:
2038:
2039: \bibitem[Sommerer \& Ott (1993)]{Som93}
2040: {\sc Sommerrer, J \& Ott, E.}, 1993,
2041: {Particles floating on a random flow: a dynamically comprehensible
2042: physical fractal},
2043: {\it Science}, {\bf 359}, {334}.
2044:
2045:
2046: \bibitem[Wilkinson \& Mehlig (2003)]{Wil03}
2047: {\sc Wilkinson, M. \& Mehlig, B.}, 2003,
2048: {Path-coalescence transition and its applications},
2049: {\it Phys. Rev. E}, {\bf 68}, 040101(R).
2050:
2051:
2052: \bibitem[van Haarlem {\sl et al.} (1998)]{vHa98}
2053: {\sc van Haarlem, B., Boersma, B. J. \& Nieuwstadt, F. T. M.}
2054: {Direct numerical simulations of particle deposition onto
2055: a free-slip and no-slip surface}
2056: {\it Phys. Fluids}, {\bf 10}, 2608-2620.
2057:
2058:
2059: \bibitem[van Kampen (1992)]{vKa92}
2060: {\sc van Kampen, N. G.}, 1992,
2061: {Stochastic processes in physics and chemistry},
2062: North-Holland.
2063:
2064:
2065: \end{thebibliography}
2066:
2067: \end{document}