1: \documentclass[aps,prl,twocolumn]{revtex4}
2: %\documentclass[letterpaper,fullpage,aps,prl,twocolumn]{revtex4}
3: \usepackage{graphicx,amsmath}
4: %\topmargin +0.0in
5:
6: \newcommand{\Pe}{\mathrm{P\hspace{-1pt}e}}
7: \newcommand{\Peo}{\mathrm{P\hspace{-1pt}e}\hspace{-0.5pt}_o}
8: \newcommand{\re}{\mbox{Re\hspace{1pt}}}
9:
10:
11: \begin{document}
12: \title{ Interfacial dynamics in transport-limited dissolution }
13:
14: \author{Martin Z. Bazant}
15: \affiliation{
16: Department of Mathematics,
17: Massachusetts Institute of Technology,
18: Cambridge, MA 02139}
19:
20: \date{\today}
21:
22: \begin{abstract}
23: Various model problems of ``transport-limited dissolution'' in two
24: dimensions are analyzed using time-dependent conformal maps. For
25: diffusion-limited dissolution (reverse Laplacian growth), several
26: exact solutions are discussed for the smoothing of corrugated
27: surfaces, including the continuous analogs of ``internal
28: diffusion-limited aggregation'' and ``diffusion-limited erosion''. A
29: class of non-Laplacian, transport-limited dissolution processes are
30: also considered, which raise the general question of when and where
31: a finite solid will disappear. In a case of dissolution by
32: advection-diffusion, a tilted ellipse maintains its shape during
33: collapse, as its center of mass drifts obliquely away from the
34: background fluid flow, but other initial shapes have more
35: complicated dynamics.
36: \end{abstract}
37:
38: \maketitle
39:
40: The analysis of interfacial dynamics using conformal maps (``Loewner
41: chains'') is a classical subject, which is finding unexpected
42: applications in physics~\cite{gruzberg04,bauer04,handbook05}. For
43: dynamics controlled by Laplacian fields, there is a vast literature on
44: continuous models of viscous fingering~\cite{bensimon86,howison92},
45: and stochastic models of diffusion-limited aggregation
46: (DLA)~\cite{hastings98,handbook05} and fractal curves in critical
47: phenomena~\cite{gruzberg04,bauer04}. Conformal-map dynamics has also
48: been formulated for a class of non-Laplacian growth phenomena of both
49: types~\cite{bazant03}, driven by conformally invariant transport
50: processes~\cite{bazant04}. For growth limited by advection-diffusion
51: in a potential flow, the connection between continuous and stochastic
52: growth patterns has been elucidated~\cite{david05}, and the continuous
53: dynamics has also been studied in cases of freezing in flowing
54: liquids~\cite{maksimov76,kornev88,kornev94,alimov98}.
55:
56: In all of these examples, the moving interface separates a ``solid''
57: region, where singularities in the map reside, a ``fluid'' region,
58: where the driving transport processes occur and the map is
59: univalent. (In viscous fingering, these are the inviscid and viscous
60: fluid regions, respectively.) Most attention has been paid to problems
61: of ``transport-limited growth'', where the solid region grows into the
62: fluid region, since the dynamics is unstable and typically leads to
63: cusp singularities in finite time~\cite{meyer82,shraiman84,howison86}
64: (without surface tension~\cite{bensimon86}). Here, we consider various
65: time-reversed problems of ``transport-limited dissolution'' (TLD),
66: where the solid recedes from the fluid region, e.g. driven by
67: advection-diffusion in a potential flow. These are stable processes,
68: so we focus on continuous dynamics, without surface tension.
69:
70: Stochastic diffusion-limited dissolution (DLD), sometimes called
71: ``diffusion-limited erosion'' (DLE) or ``anti-DLA'', has been
72: simulated by allowing random walkers in the fluid to annihilate
73: particles of the solid upon contact
74: ~\cite{paternson84,tang85,meakin86,krug91}, rather than aggregating as
75: in DLA~\cite{witten81}. Outward radial DLE on a lattice, or
76: ``internal DLA'' (IDLA), where the random walkers start at the origin
77: and cause a fluid cavity to grow in an infinite solid, has also been
78: studied by mathematicians, who proved that the asymptotic shape is a
79: sphere in any dimension~\cite{lawlor92}.
80:
81: We begin our analysis by summarizing some exact solutions for
82: continuous DLD, which could help to understand fluctuations and the
83: long-time limits of DLE and IDLA. Although these solutions are known
84: in somewhat different forms for Laplacian
85: growth~\cite{meyer82,shraiman84,howison86}, it is instructive to
86: summarize them prior to considering problems of TLD by
87: advection-diffusion, to highlight the effects of fluid flow.
88:
89: {\it Dissolution of surface corrugation. --} Let $G(w,t)$ be a
90: conformal map from the left half plane to the fluid region, where
91: steady (Laplacian) diffusion with uniform flux at $-\infty$ drives
92: dissolution of the solid. The interfacial dynamics is given by the
93: (dimensionless) Polubarinova-Galin (PG)
94: equation~\cite{handbook05,howison92},
95: \begin{equation}
96: \re\left(\overline{G^\prime}{G_t}\right) = 1, \ \ \mbox{ on } \re w = 0.
97: \end{equation}
98: An exact solution has the form,
99: \begin{equation}
100: G(w,t) = w + h(t) + \log(1+a(t)e^w) \label{eq:fingers}
101: \end{equation}
102: with $|a(t)|<1/2$ (real) where
103: \begin{eqnarray}
104: a &=& \frac{Ce^{-t}}{(1-a^2)^{3/2}} \sim Ce^{-t} + \frac{3}{2}C^3e^{-3t} + \ldots\\
105: \frac{dh}{dt} &=& \frac{1+a-a^2}{1-a^2}, \ h \sim t + C(e^{-t}-1) +
106: \ldots \label{eq:ah}
107: \end{eqnarray}
108: and $C = a(0)(1-a(0)^2)^{3/2}$. This solution, shown in
109: Fig.~\ref{fig:fingers}, describes the decay of surface corrugations in
110: electropolishing~\cite{meakin86,krug91,wagner54,edwards53} or the
111: displacement of an inviscid fluid by an immiscible viscous fluid in a
112: Hele-Shaw cell~\cite{paterson84,tang85,bensimon86}.
113: % Upon time-reversal, the interface forms $2/3$-power cusps as $a\to 1/2$
114: %~\cite{bensimon86,meyer82,shraiman84,howison86}.)
115: and closely ressembles simulations of ``mean-field DLE'' (Fig. 2 of
116: ~\cite{tang85}).
117:
118: The linear stability of a flat interface in DLD is contained in the
119: long-time limit of Eqs.~(\ref{eq:fingers})-(\ref{eq:ah}). With the
120: dimensions restored ($w \mapsto wk$, $G \mapsto Gk$, and $t \mapsto
121: kvt$), the interface shape becomes sinusoidal,
122: \begin{equation}
123: x(y) \sim vt + C e^{-kvt} \cos(ky),
124: \end{equation}
125: for a Fourier mode of wavenumber $k$ and mean interfacial velocity
126: $v$. We thus recover the classical
127: result~\cite{tang85,krug91,wagner54} that $kv$ is the exponential
128: decay rate of mode $k$, as observed in experiments~\cite{edwards53}.
129:
130: \begin{figure}
131: \includegraphics[width=.9\linewidth]{fingers}
132: \caption{ \label{fig:fingers} Diffusion-limited dissolution of a
133: corrugated surface from left to right, for $h(0)=0$, $a(0)=.3$,
134: $t=0,.25,.50,\ldots,2.50$ in Eq.~(\ref{eq:fingers}). }
135: \end{figure}
136:
137:
138: {\it Outward dissolution of clover-like shapes. -- } Next we consider
139: the continuous analog of IDLA: DLD in a radial geometry driven by a
140: constant diffusive flux from the origin. This could model quasi-steady
141: melting of an infinite solid around a point source of heat, or the
142: injection of a viscous fluid into a Hele-Shaw cell, displacing an
143: inviscid fluid. The radial PG equation is~\cite{handbook05,howison92},
144: \begin{equation}
145: \re\left(\overline{w g^\prime}
146: g_t \right) = 1, \ \ \mbox{ on }
147: |w|=1, \label{eq:pgr}
148: \end{equation}
149: where $ z=g(w,t)$ is a univalent mapping of the {\it interior} of
150: the unit disk.
151: % Substituting the Taylor expansion,
152: % \begin{equation}
153: % g(w,t) = \sum_{n=0}^\infty a_n(t) w^n \ \ \mbox{for } |w| \leq 1.
154: %\label{eq:taylor}
155: %\end{equation}
156: %yields a system of nonlinear ordinary differential equations for the
157: %coefficients.
158: A tractable case is the clover-like $(N-1)$-fold perturbation of a
159: circle,
160: \begin{equation}
161: g(w,t) = a_1(t) w + a_N(t) w^N \label{eq:gout}
162: \end{equation}
163: first analyzed by Meyer for the (time-reversed) Hele-Shaw
164: problem~\cite{meyer82}. For $a_1(0)=1$ and $0 < c = a_N(0) < 1/N$
165: (real), the solution has the implicit form,
166: \begin{equation}
167: a_1^2 + N a_N^2 = 1 + Nc^2+2t \ \mbox{ and } \ a_1^Na_N=c, \label{eq:aNout}
168: \end{equation}
169: Since $a_1(t)$ increases to $\infty$ from $a_1(0)=1$ and $a_N(t)$
170: decreases to $0$ from $a_N(0)=c<1/N < 1$, the following recursion
171: converges very quickly,
172: \begin{equation}
173: a_1 = \sqrt{2t + 1 + Nc^2(1-a_1^{-2N})}, \label{eq:rec}
174: \end{equation}
175: and yields asymptotic approximation upon recursive substitution.
176: %
177: % \begin{equation}
178: % a_1 \sim \sqrt{2t + 1+Nc^2(1-(2t+1)^{-N})} \label{eq:a1outsim}
179: % \end{equation}
180: An example shown in Figure \ref{fig:out4}, illustrates the rapid
181: smoothing of a four-leafed clover shape.
182:
183: From Eqs.~(\ref{eq:aNout})-(\ref{eq:rec}) we see that the $(N-1)$-fold
184: radial perturbation decays as a power-law, $a_N \propto t^{-N/2}$, in
185: contrast to the exponential decay of perturbations of a flat
186: interface. We conjecture that the stochastic interface in IDLA is
187: asymptotic to the continuous DLD solution above, up to small
188: logarithmic fluctuations~\cite{lawlor95}, where $w^N$ is the smallest
189: perturbation in the initial cluster shape. In general, the continuous
190: dynamics of DLD may also accurately approximate the ensemble-averaged
191: stochastic dynamics of IDLA, which is not the case for DLA and other
192: unstable aggregation processes~\cite{david05}.
193:
194: \begin{figure}
195: \includegraphics[width=1.2\linewidth]{out4}
196: \caption{ \label{fig:out4} Outward radial diffusion-limited dissolution
197: driven by a point sink at the origin for an initial four-lobed
198: perturbation of a circle. This exact solution is given by
199: Eqs.~(\ref{eq:out})-(\ref{eq:outeqns}) with $N=5$, $c=0.15$, $t=0,
200: 0.2, 0.4,\ldots,1.0$. }
201: \end{figure}
202:
203: An (apparently new) explicit solution is possible for $N=2$. In that
204: case, Equation (\ref{eq:aNout}) reduces to a depressed
205: cubic~\cite{nick93}, $a_2^3 + 3p a_2 = 2 q$, solved by the
206: formula of Cardano and dal Ferro,
207: \begin{equation}
208: a_1 = \sqrt[3]{q + \sqrt{q^2-p^3}} + \sqrt[3]{q-\sqrt{q^2-p^3}}.
209: \end{equation}
210: where $p=-(t+c^2+1/2)/4$ and $q=-c/4$.
211:
212: {\it Inward dissolution of star-like shapes. -- } Next we briefly
213: consider inward DLD with a constant diffusive flux at infinity, which
214: simply is the time-reverse of Laplacian growth in a radial
215: geometry~\cite{howison92}. Now the map in Eq.~(\ref{eq:pgr}) must be
216: univalent {\it outside} the unit disk, with a Laurent expansion,
217: \begin{equation}
218: g(w,t) = \sum_{n=-\infty}^1 a_n(t) w^n, \ \ |w|\geq 1 \label{eq:ls}
219: \end{equation}
220: There are well-known solutions for $(N+1)$-fold
221: perturbations~\cite{shraiman84,howison86},
222: \begin{equation}
223: g(w,t) = a_1(t) w + a_{-N}(t) w^{-N} \ \ \mbox{ for } \ |w|\geq 1, \label{eq:in}
224: \end{equation}
225: where, without loss of generality, $a_1(0)=1$ and $a_{-N}(0) = c<1/N$
226: is real. Again, the Laurent coeficients satisfy a pair of nonlinear
227: equations, which is most easily solved as a fixed-point iteration,
228: \begin{equation}
229: a_1 = \sqrt{1-2t+Nc^2(a_1^{2N}-1)} \ \mbox{ and } \ a_{-N} = c a_1^N. \label{eq:ineqs}
230: \end{equation}
231: The only qualitative difference with outward DLD is that the solid
232: collapses to a point in a finite time, $t_c = (1-Nc^2)/2$. For $N=1$,
233: an ellipse ($0<c<1$) or circle ($c=0$) maintains its shape during
234: collapse,
235: \begin{equation}
236: g(w,t) = \sqrt{1 - \frac{t}{t_c}} \left(w + \frac{c}{w}\right)
237: \end{equation}
238: For $N\geq 1$, the shape approaches a circle prior to collapse,
239: according to the asymptotic formula,
240: \begin{equation}
241: a_1(t) \sim \sqrt{2(t_c-t) + Nc^2(2(t_c-t))^N}
242: \end{equation}
243: %a_{-N}(t) &\sim& c\left(2(t_c-t) + Nc^2(2(t_c-t))^N\right)^{N/2}
244: %\end{eqnarray}
245: with $a_{-N}(t)=c a_1(t)^N$. The collapse of a four-pointed
246: shape ($N=3$) is shown in Figure~\ref{fig:in4}.
247:
248: \begin{figure}
249: \includegraphics[width=1.2\linewidth]{in4}
250: \caption{ \label{fig:in4} Inward radial diffusion-limited dissolution
251: driven by a sink at $\infty$ for a four-pointed shape. This exact
252: solution is given by Eqs.~(\ref{eq:in})-(\ref{eq:ineqs}) with $N=3$,
253: $c=0.3$, $t=0, 0.04, 0.08,\ldots,0.36, 0.3649$; the collapse occurs
254: at $t_c = 0.365$. }
255: \end{figure}
256:
257:
258: {\it Advection-diffusion-limited dissolution. --} The dynamics of
259: dissolution become more interesting when driven by non-Laplacian (but
260: conformally invariant~\cite{bazant04}) transport
261: processes~\cite{bazant03}, where right-hand side of the PG equation
262: (\ref{eq:pgr}) is replaced by the nonuniform, time-dependent flux to
263: the interface, $\sigma(w,t)$. An important example is dissolution by
264: advection-diffusion in a potential flow, e.g. the erosion of rock by
265: flowing water, the evaporation of a fiber coating in a flowing gas, or
266: the melting of a solid column in a flowing liquid. The time-reversed
267: growth problem has been studied extensively in the contexts of
268: freezing~\cite{maksimov76,kornev88,kornev94,alimov98} and
269: advection-diffusion-limited aggregation~\cite{bazant03,david05}, but
270: it seems that dissolution -- which leads to collapse in finite time --
271: has not been analyzed. For a given initial shape and background flow,
272: {\it when and where will collapse occur?}
273:
274: Consider a finite solid of constant concentration and arbitrary shape
275: in a two dimensional potential flow of zero concentration and uniform
276: velocity far away. The relative importance of advection to diffusion
277: is measured by the P\'eclet number, $Pe_0 = UL/D$, for a background
278: fluid velocity $U$, diffusivity $D$ and length $L$. The time-dependent
279: P\'eclet number, $Pe(t) = Pe_0 a_1(t)$, is defined by the conformal
280: radius, $a_1(t)$, in Eq.~(\ref{eq:ls}). When solid dissolves
281: ($a_1(t)\to 0$), diffusion eventually dominates, and it is natural to
282: focus on the low-$Pe$ limit.
283:
284: The flux profile $\sigma(\theta,\Pe)$ on the absorber has been studied
285: extensively, and very accurate asymptotic approximations are
286: available~\cite{choi05}. (A numerical code in matlab is also at
287: http://advection-diffusion.net.) From the low-$\Pe$ approximation,
288: \begin{equation}
289: \sigma \sim \frac{I_0(Pe)}{K_0(Pe/2)}e^{Pe\,\cos\theta} -
290: \Pe\left(\cos\theta + \int_0^{Pe} dt e^{t\cos\theta} \frac{I_1(t)}{t}\right) \label{eq:sig}
291: \end{equation}
292: which is uniformly accurate in angle $\theta$ up to $Pe=10^{-1}$, let
293: us keep only the leading terms,
294: \begin{equation}
295: \sigma \sim
296: \frac{1+Pe\cos\theta}{-\log(Pe/4)-\gamma}-Pe\cos\theta \label{eq:sigsim}
297: \end{equation}
298: where $\gamma = 0.577215\ldots$ is Euler's constant.
299:
300: In the final stage of collapse where $-\log Pe \gg \gamma$, the interface
301: is asymptotically circular, $g(w,t) \sim a_1(t) w$. From
302: (\ref{eq:sigsim}), the radius in this regime satisfies, $a_1 da_1/dt
303: \sim -1/\log a_1$, and thus has the asymptotic form,
304: \begin{equation}
305: a_1 \sim \sqrt{\frac{4(t-t_c)}{\log\frac{1}{t-t_c}-\log\log
306: \frac{1}{t-t_c} + \ldots }}
307: \end{equation}
308:
309: To describe the dynamics starting at small $Pe$ and ending just prior
310: to collapse, it is reasonable to further set $\log Pe\approx$
311: constant. In this intermediate regime, the interfacial dynamics is
312: given by
313: \begin{equation}
314: \re(\overline{w g^\prime} g_t ) = -1 + b\, a_1(t) \cos\theta \ \
315: \mbox{ for } \ w=e^{i\theta} \label{eq:pgb}
316: \end{equation}
317: (with a suitable choice of units). The $\cos\theta$ term has the
318: effect of exciting new modes in the shape of the dissolving
319: solid, which are not
320: present in the initial condition. We will see that a circle, $g(w,0) =
321: w$, translates away from the flow, $g(w,t) = w + a_0(t)$, as its
322: upstream side dissolves more quickly prior to collapse. It can also be
323: shown that an $(N+1)-$fold perturbation of a circle (\ref{eq:in}) will
324: translate and develop an $N-$fold perturbation, $g(w,t) =
325: a_1(t)w+a_0(t) + a_{-N+1}/w^{N-1} + a_N(t)/w^N$. Higher Fourier modes
326: in $\sigma(w,t)$ excite additional terms in the Laurent
327: expansion of the shape, $g(w,t)$.
328:
329: \begin{figure}
330: \includegraphics[width=1.2\linewidth]{inflow}
331: \caption{ \label{fig:inflow} Inward advection-diffusion-limited
332: dissolution of a tilted elliptical solid in a uniform background
333: flow from left to right ($c=0.5i$, $b = 0.8$, $t=0, 0.04,
334: 0.08,\ldots,0.36, 0.3749$). The major axis of the ellipse remains
335: oriented at $\pi/4$, while its center of mass moves at an angle of
336: $\pi/6$ relative to the background flow. At time $t_c = 0.375$, the
337: solid disappears at the point $a_0(t_c) = 0.4+0.2i$. }
338: \end{figure}
339:
340: {\it Collapse of a tilted ellipse in a uniform flow. --} To illustrate
341: these principles, let us consider a solid ellipse, $g(w,0) = w + c/w$
342: ($c<1$), which is tilted at an angle $\phi=(\mbox{arg} c)/2$ with
343: respect to a background flow in the positive $x$ direction. During
344: dissolution, the shape remains an ellipse, but its center of mass,
345: $a_0(t)$, move. An example is shown in Fig.~\ref{fig:inflow}.
346:
347: The Laurent expansion is given by
348: %\begin{equation}
349: $g(w,t) = a_1(t)w + a_0(t) + a_{-1}(t)/w$,
350: %\label{eq:inflow}
351: %\end{equation}
352: where $a_1(t)$ is real, but $a_0(t)$ and $a_{-1}(t)$, are
353: complex. Substituting into (\ref{eq:pgb}) and integrating shows that
354: the conformal has a square-root singularity,
355: \begin{equation}
356: a_1(t) = \sqrt{1 - \frac{t}{t_c}}, \ \ t_c = \frac{1-|c|^2}{2},
357: \end{equation}
358: since the area decreases linearly to zero (in the approximation of
359: constant total flux): $A(t) = A(0)-2\pi t$, where $A(0) =
360: \pi(1-|c|^2)$. The collapse time depends on the initial shape through
361: $|c|$. The slowest collapse, $t_c=1/2$, occurs for a circle, $c=0$,
362: while the collapse time tends to zero for a very elongated ellipse,
363: $|c| \to 1$, regardless of orientation.
364:
365: For a constant total flux, the collapse time $t_c$ does not depend on
366: the bias introduced by the flow velocity (through $b$), although the
367: flow affects the time scale through the initial P\'eclet number,
368: $Pe_0$. Mathematically, this is a general consequence of the conformal
369: invariance of the transport process~\cite{bazant03}, which causes the
370: total flux (Nusselt number) to depend only on the conformal radius and
371: not the asymmetric shape of the particle~\cite{choi05}. Physically,
372: the enhancement of dissolution on the upstream side of the solid is
373: cancelled by the reduction in dissolution on the downstream side.
374:
375: During dissolution, the ellipse keeps its shape and its orientation
376: with respect to the flow direction since $a_{-1}(t) = c
377: a_1(t)$. However, the center of mass moves away from the flow, but
378: also away from the end of the ellipse which protrudes upstream. The
379: velocity of the center of mass is constant and in the $1+c$ direction,
380: \begin{equation}
381: a_0(t) = \frac{b(1+c)t}{1-|c|^2}.
382: \end{equation}
383: Note that the center of mass does not move along the major axis of the
384: ellipse at angle $\phi$, but instead at an oblique angle $\theta$
385: given by $\sin \theta = |c| \sin(2\phi-\theta)$. The final collapse
386: occurs at the point, $a_0(t_c) = b(1+c)/2$. In the simplest case,
387: $c=0$, a circle maintains its shape while its center of mass
388: translates away from the flow at (dimensionless) velocity $b$ until
389: collapse occurs at $x=b/2$ at time $t_c=b/2$.
390:
391: For any initial shape, the center of mass will drift away from the
392: flow, as well as away from any protrusions in the direction of the
393: flow. For non-elliptical shapes, it is nontrivial to predict the exact
394: time and place of collapse, even for the simplified dynamics of
395: Eq.~(\ref{eq:pgb}). For more general transport-limited
396: dynamics~\cite{bazant03}, predicting the collapse seems like an
397: interesting open problem.
398:
399:
400:
401:
402: The author thanks J. Propp for an introduction to IDLA.
403:
404: \begin{thebibliography}{99}
405:
406: \bibitem{gruzberg04} I. Gruzberg and L. P. Kadanoff,
407: J. Stat. Phys. {\bf 114}, 1183 (2004).
408:
409: \bibitem{bauer04} M. Bauer and D. Bernard, cond-mat/0412372.
410:
411: \bibitem{handbook05} M. Z. Bazant and D. Crowdy, in {\it Handbook of
412: Materials Modeling}, Vol. I, ed. by S. Yip, Art. 4.10 (Springer,
413: 2005).
414:
415: \bibitem{bensimon86} D. Bensimon, L. P. Kadanoff, S. Liang,
416: B. I. Shraimain, and C. Tang, Rev. Mod. Phys. {\bf 58}, 977 (1986).
417:
418: \bibitem{howison92} S. D. Howsion, Euro. J. Appl. Math. {\bf 3}, 209 (1992).
419:
420: \bibitem{hastings98} M. Hastings and L. Levitov, Physica D {\bf 116},
421: 244 (1998).
422:
423: \bibitem{bazant03} M. Z. Bazant, J. Choi, and B. Davidovitch,
424: Phys. Rev. Lett. {\bf 91}, 04503 (2003).
425:
426: \bibitem{bazant04} M. Z. Bazant, Proc. Roy. Soc. Lond. A {\bf 460},
427: 1433 (2004).
428:
429: \bibitem{david05} B. Davidovitch, J. Choi, and M. Z. Bazant,
430: Phys. Rev. Lett. {\bf 95}, 075504 (2005).
431:
432: \bibitem{maksimov76} V. A. Maksimov, Prikl. Mat. Mekh. {\bf 40}, 264 (1976).
433:
434: \bibitem{kornev88} K. G. Kornev and V. A. Chugunov,
435: Prikl. Mat. Mekh. {\bf 52}, 773 (1988).
436:
437: \bibitem{kornev94} K. Kornev and G. Mukhamadullina,
438: Proc. Roy. Soc. Lond. A {\bf 447}, 281 (1994).
439:
440: \bibitem{alimov98} M. Alimov, K. Kornev, and G. Mukhamadullina, SIAM
441: J. Appl. Math. {\bf 59}, 387 (1998).
442:
443: \bibitem{meyer82} G. H. Meyer, in {\it Numerical Treatment of Free
444: Boundary Value Problems}, p. 202 (Birkh\"auser, 1982).
445:
446: \bibitem{shraiman84} B. I. Shraiman and D. Bensimon, Phys. Rev. A {\bf 30},
447: 2840 (1984).
448:
449: \bibitem{howison86} S. D. Howison, SIAM J. Appl. Math. {\bf 46}, 20
450: (1986).
451:
452: \bibitem{paternson84} L. Paterson, Phys. Rev. Lett. {\bf 52}, 1621
453: (1984).
454:
455: \bibitem{witten81} T. Witten and L. M. Sander, Phys. Rev. Lett. {\bf
456: 47}, 1400 (1981).
457:
458: \bibitem{tang85} C. Tang, Phys. Rev. A {\bf 31}, 1977 (1985).
459:
460: \bibitem{meakin86} P. Meakin and J. M. Deutch, J. Chem Phys. {\bf 85}, 2320
461: (1986).
462:
463: \bibitem{krug91} J. Krug and P. Meakin, Phys. Rev. Lett. {\bf 66}, 703 (1991).
464:
465: \bibitem{lawlor92} G. Lawlor, M. Bramson, and D. Griffeath,
466: Ann. Prob. {\bf 20}, 2117 (1992).
467:
468: \bibitem{wagner54} C. Wagner, J. Electrochem. Soc. {\bf 101}, 225
469: (1954).
470:
471: \bibitem{edwards53} J. Edwards, J. Electrochem. Soc. {\bf 100}, 189C,
472: 223C (1953).
473:
474: \bibitem{lawlor95} G. F. Lawlor, Ann. Probab. {\bf 23}, 71 (1995).
475:
476: \bibitem{nick93} R. W. D. Nickalls, Math. Gazette {\bf 77}, 354 (1993).
477:
478: \bibitem{choi05} J. Choi, D. Margetis, T. M. Squires, and
479: M. Z. Bazant, J. Fluid Mech. {\bf 536}, 155 (2005).
480:
481: \end{thebibliography}
482:
483: \end{document}
484:
485: %
486:
487:
488: {\it Conclusions. --}
489:
490: reverse laplacian growth: exact solutions to compare to DLE, IDLA
491: analysis and deterministic models; by comparing exact continuous
492: solution to stochastic growth simulations, can check mean-field
493: approximation, as in ~\cite{david05} [and Sander for DLA in a
494: channel]. due to the stable growth, it may be that the naive mean
495: field approx works fine...
496:
497: advection-diffusion-limited dissolution. new set of basic questions:
498: for given tranport process, when and where does collapse of a finite
499: solid occur?
500:
501:
502: ------
503:
504:
505:
506:
507: Following a suggestion of Jim Propp, in these notes we derive some
508: exact solutions for general problems of ``reverse Laplacian growth'',
509: where an interface advances {\it away} from a region of quasi-steady
510: heat or mass transport which drives the dynamics, rather than toward
511: it as in Laplacian growth, DLA, viscous fingering, etc. As a result,
512: the well-known finite-time singularities of Laplacian growth are
513: replaced by stable dynamics leading to simple, smooth shapes.
514:
515: These continuous models might describe a variety of physical
516: situations, such as the electro-dissolution of an electrode, corrosion
517: of a solid limited by the diffusion of a catalyst, the slow melting of
518: a solid limited by heat diffusion, and the displacement of a less
519: viscous fluid by a more viscous fluid in a Hele-Shaw cell or porous
520: medium. Surface tension can be neglected since the dynamics is stable
521: and tends to minimize interfacial curvature, as we shall see below.
522:
523: In two dimensions, conformal-map dynamics (a reverse ``Loewner
524: chain'', which expands the region of interest) can be formulated for
525: interfacial dynamics limited by any conformally invariant transport
526: process, such as advection-diffusion in a potential flow or
527: electrochemical conduction. We suggest the term ``transport-limited
528: dissolution'' to describe all such cases, when the dynamics is
529: continuous, although here we will only consider the simplest case of
530: steady linear diffusion (described by a harmonic concentration field).
531:
532: In each problem, the interfacial dynamics can be either discrete and
533: stochastic, or continuous and deterministic. The former case, on a
534: lattice, corresponds to what probabilists call ``internal DLA'' for
535: outward radial dynamics or ``diffusion-limited erosion'' for inward
536: radial dynamics. Here, we will present some simple exact solutions to
537: the continuous analogs of these two discrete models, although these
538: naive mean-field approximations may not precisely describe the average
539: shape of the interface in the discrete, stochastic models, as in DLA
540: and ADLA.
541: -----
542:
543: Paterson~\cite{paterson84} introduced the model of ``anti-DLA''
544: where random walkers in the fluid region annihiliate particles of the
545: solid region upon contact, rather than aggregating as in
546: DLA~\cite{witten81}. Tang~\cite{tang85} simulated a ``mean-field''
547: generalization of anti-DLA (where many walkers must strike the same
548: solid site before annihilation) and compared to numerical solutions of
549: Laplacian dissolution. The smoothing and residual roughness of
550: anti-DLA surfaces was studied by Meakin and Deutch~\cite{meakin86}, as
551: model of electrochemical polishing. Krug and Meakin~\cite{krug91}
552: introduced the term ``diffusion-limited erosion'' (DLE) for anti-DLA
553: and analyzed the roughness of a flat front. Bram
554:
555: