physics0512043/VM.tex
1: \documentclass{elsart}
2: \usepackage{graphicx}
3: \usepackage{dcolumn}
4: \usepackage{amsmath}
5: \usepackage{amssymb}
6: 
7: \begin{document}
8: \begin{frontmatter}
9: \title{How a Long Bubble Shrinks: a Numerical Method for an Unforced
10: Hele-Shaw Flow}
11: \author{Arkady Vilenkin and  Baruch Meerson}
12: \address{The Racah Institute  of  Physics, The Hebrew University
13: of Jerusalem, Jerusalem 91904, Israel}
14: \begin{abstract}
15: We develop a numerical method for solving a free boundary problem
16: which describes shape relaxation, by surface tension, of a long and
17: thin bubble of an inviscid fluid trapped inside a viscous fluid in a
18: Hele-Shaw cell. The method of solution of the exterior Dirichlet
19: problem employs a classical boundary integral formulation. Our
20: version of the numerical method is especially advantageous for
21: following the dynamics of a very long and thin bubble, for which an
22: asymptotic scaling theory has been recently developed. Because of
23: the very large aspect ratio of the bubble, a direct implementation
24: of the boundary integral algorithm would be impractical. We modify
25: the algorithm by introducing a new approximation of the integrals
26: which appear in the Fredholm integral equation and in the integral
27: expression for the normal derivative of the pressure at the bubble
28: interface. The new approximation allows one to considerably reduce
29: the number of nodes at the almost flat part of the bubble interface,
30: while keeping a good accuracy. An additional benefit from the new
31: approximation is in that it eliminates numerical divergence of the
32: integral for the tangential derivative of the harmonic conjugate.
33: The interface's position is advanced in time by using explicit node
34: tracking, whereas the larger node spacing enables one to use larger
35: time steps. The algorithm is tested on two model problems, for which
36: approximate analytical solutions are available.
37: \end{abstract}
38: 
39: \begin{keyword}
40: Laplace's equation; Dirichlet problem; Fredholm integral equation of
41: the second kind; free boundary problem, Hele Shaw flow, surface
42: tension
43: \end{keyword}
44: %\PACS 02.70.Pt 47.11.+j 47.15.Gf 47.15.Hg
45: \end{frontmatter}
46: \maketitle
47: 
48: \section{Introduction}
49: Let a bubble of low-viscosity fluid (say, air) get trapped inside a
50: high-viscosity fluid (say, oil) in a quasi-two-dimensional Hele-Shaw
51: cell: two parallel plates with a narrow gap between them. What will
52: happen to the shape of the bubble, if the (horizontal) plates of the
53: Hele-Shaw cell are perfectly smooth, and the two fluids are
54: immiscible? The answer depends on the initial bubble shape. A
55: perfectly circular bubble (or an infinite straight strip) will not
56: change, while a bubble of any other shape will undergo
57: surface-tension-driven relaxation until it either becomes a perfect
58: circle, or breaks into two or more bubbles, which then become
59: perfect circles. The bubble shape relaxation process is non-local,
60: as it is mediated by a flow in the external viscous fluid. The
61: two-dimensional surface-tension-driven flow can be called an
62: unforced Hele-Shaw (UHS) flow. This is in contrast to forced
63: Hele-Shaw flows that have been in the focus of hydrodynamics and
64: nonlinear and computational physics for the last two decades
65: \cite{Langer1,BKD,Kessler,Casademunt1,Casademunt2}. In rescaled
66: variables, the UHS flow is described by the solution of the
67: following free boundary problem, see \textit{e.g.} Refs.
68: \cite{CLM,VMS}:
69: \begin{equation} \label{LPLC}
70: \nabla^{2}p(q)=0\qquad \mbox{for}\qquad q\in E,
71: \end{equation}
72: \begin{equation} \label{GT}
73: p(q)=\mathcal{K}\qquad \mbox{for}\qquad q\in \gamma,
74: \end{equation}
75: %{\cal K}
76: \begin{equation} \label{VN}
77: v_{n}(q)=-\nabla_{n}p(q)\qquad \mbox{for}\qquad q\in \gamma,
78: \end{equation}
79: where $E$ is an unbounded region of the plane, external to the
80: bubble interface $\gamma$, $v_{n}$ is the normal velocity of the
81: interface, the index $\textsc{n}$ denotes the component of vectors
82: normal
83: to the interface, and $K$ is the local curvature of the %{\cal K}
84: interface. The pressure $p$ is bounded at infinity. The free
85: boundary problem (\ref{LPLC})-(\ref{VN}) splits into two
86: sub-problems:
87: \begin{enumerate}
88: \item{Solving the exterior Dirichlet problem (\ref{LPLC}) and
89: (\ref{GT}) and  calculating $v_{n}(q)$ from Eq. (\ref{VN}).}
90: \item{Advancing the interface $\gamma$ in time with the known
91: $v_{n}(q)$.}
92: \end{enumerate}
93: 
94: The free boundary problem (\ref{LPLC})-(\ref{VN}) represents an
95: important example of area-preserving curve-shortening motion
96: \cite{ConstantinPugh}, but it is not integrable. Moreover, the only
97: analytical solution to this problem, available until recently, was
98: the approximate solution following from a linear stability analysis
99: of a slightly deformed circular or flat interface \cite{linear}.
100: Recently an asymptotic scaling theory has been developed for a
101: non-trivial case when the inviscid fluid occupies, at $t=0$, a
102: half-infinite (or, physically, very long) strip \cite{VMS}. It
103: turned out that this somewhat unusual initial condition provides a
104: useful characterization of the UHS flow, as the evolving strip,
105: which develops a dumbbell shape, exhibits approximate
106: self-similarity with non-trivial dynamic exponents \cite{VMS}.
107: Predictions of the scaling analysis have been verified numerically
108: in Ref. \cite{VMS} by using a boundary integral algorithm, tailored
109: to the very large aspect ratio of the bubble. The present paper
110: describes this algorithm in detail.
111: 
112: A multitude of numerical methods have been suggested in the recent
113: years for simulating different variants of Hele-Shaw flows. Boundary
114: integral methods, which deal directly with the interface between the
115: two fluids, are advantageous compared to methods of finite elements
116: and finite differences. Methods based on conformal mapping
117: techniques have long been used in this class of problems (see,
118: \textit{e.g.} Refs. \cite{BKD,DKZ,Tanv}). However, they apply most
119: naturally to the case of \textit{zero} surface tension and are less
120: convenient when surface tension is non-zero \cite{HLSH}.  Still
121: another numerical strategy is phase field methods. Folch \textit{et
122: al.} \cite{Folch1,Folch2} developed a phase field method for an
123: arbitrary ratio of the viscosities of the two fluids. Unfortunately,
124: their method becomes inefficient when the viscosity contrast is too
125: high \cite{Folch2}. To remind the reader, the viscosity contrast is
126: infinite in the case under consideration in the present work.
127: Glasner \cite{Glasner} developed a phase field method for a
128: description of a bubble of a high-viscosity fluid trapped in a
129: low-viscosity fluid. We are unaware of any phase-field approach
130: which would deal with the opposite case, which is under
131: investigation in the present work: a low-viscosity bubble in a
132: high-viscosity fluid.
133: 
134: The present work suggests a numerical algorithm for solving the free
135: boundary problem (\ref{LPLC})-(\ref{VN}) in the special case of a
136: very long bubble. It is well-known (but still remarkable) that the
137: exterior Dirichlet problem (\ref{LPLC}) and (\ref{GT}) can be
138: formulated in terms of a Fredholm integral equation of the second
139: kind for an effective density of the dipole moment \cite{POTTH}. A
140: na\"{\i}ve formulation, however, would lead to non-existence of
141: solution by the Fredholm alternative \cite{GGM}. To overcome this
142: difficulty, Greenbaum \textit{et al.} \cite{GGM} implemented in
143: their algorithm a modification of the Fredholm equation due to
144: Mikhlin \cite{Mikhlin}. The modified Fredholm equation has a unique
145: solution for any smooth $\gamma$ and integrable $\mathcal{K}$
146: \cite{Mikhlin}. Greenbaum \textit{et al.} developed an efficient
147: numerical algorithm (which is also valid for multiple bubbles) by
148: discretization. However, the geometry of a very long and thin
149: bubble, that we are mostly interested in, defines widely different
150: length scales in the problem. Rapid variations of the dipole moment
151: density at the highly curved ends of the bubble naturally
152: necessitate a small spacing between the interface nodes. It is less
153: natural, however, that, in a straightforward approach, one must keep
154: the node spacing much smaller than the bubble thickness \textit{over
155: the whole bubble interface}. Indeed, as we show below, the typical
156: length scale of the variation of the kernel of the integral equation
157: is comparable to the bubble \textit{thickness} which, during the
158: most interesting part of the long bubble dynamics, remains almost
159: unchanged.  Apart from being computationally inefficient, the
160: straightforward approach would cause a problem for explicit tracking
161: of nodes, as the stability criterion, intrinsic in the explicit
162: method, demands a time step less then a constant $\mathcal{O}(1)$
163: multiplied by the node spacing cubed \cite{Beale}. In this work we
164: turned this obstacle into advantage, by employing the fact that the
165: length scale of variation of the solution, over the most of the
166: bubble interface, is much greater than the bubble thickness. We
167: constructed a new approximation of the integral entering the
168: Fredholm equation, by representing the sought dipole moment density
169: as a piecewise constant function, and the bubble interface shape
170: function as a piecewise linear function. As a result, the integral
171: is approximated by a sum, each term of which is equal to a local
172: value of the dipole moment density multiplied by an integral of the
173: kernel between two neighboring nodes. Fortunately, the latter
174: integral can be calculated analytically. The new approximation
175: allowed us to considerably increase the node spacing over the most
176: of the bubble interface, while keeping a good accuracy.
177: 
178: Having found an approximated solution $p$ in the form of a double
179: layer potential, one needs to compute the normal derivative of the
180: solution $\nabla_{n}p(q\in \gamma)$. In a straightforward
181: realization of the boundary integral formulation this would result
182: in a hypersingular integral, see Ref. \cite{GGM}. To overcome this
183: difficulty, one resorts to theory of analytic functions and computes
184: the harmonic conjugate $V(q)$. By virtue of the Cauchy-Riemann
185: equations, the tangential derivative of $V(q)$ is equal to the
186: desired normal derivative of $p$.  The harmonic conjugate $V(q)$ has
187: the form of a principal value integral, over the interface, of the
188: dipole moment density multiplied by a kernel, which is a function of
189: coordinates of two points, $q$ and $g$, belonging to the interface.
190: This kernel diverges when the integration variable $g$ coincides
191: with $q$. Here we again employ the large scale difference at the
192: flat part of the bubble and use the same approximation as in the
193: Fredholm equation. As an additional benefit, the numerical
194: divergence of the integrand of the harmonic conjugate $V$ is
195: avoided. As a result, we do not need to use even nodes to compute
196: $V$ at odd nodes and vice-versa, as suggested in Ref. \cite{GGM}.
197: 
198: Here is a layout of the rest of the paper. Section 2 deals with the
199: numerical solution of the exterior Dirichlet problem, and with the
200: computation of the normal derivative of the solution at the
201: interface. We briefly review the boundary integral method for an
202: exterior Dirichlet problem and motivate the need for its
203: modification for very long bubbles. Then we formulate our discrete
204: approximation. In Section 3 we briefly describe a simple explicit
205: integration which we used to track the bubble interface. Section 4
206: presents the results of code testing, while Section 5 presents the
207: Conclusions.
208: 
209: \section{Exterior Dirichlet Problem}
210: \subsection{Boundary integral formulation}
211: Following Mikhlin \cite {Mikhlin}, we seek the solution $p(q)$ of
212: the problem(\ref{LPLC}) and (\ref{GT}) for a simply connected bubble
213: in a double layer potential representation:
214: \begin{equation} \label{PofQ}
215: p(q)=\frac{1}{2\pi}\oint_{\gamma}\left[1+K(q,g)\right] \mu(g)\,
216: dg\;\;\; \mbox{for}\;\;q\in E,
217: \end{equation}
218: where $\mu(g)$ is an unknown dipole density at the point $g$ of the
219: interface, and $dg$ is the element of arclength including the point
220: $g$. The kernel $K(q,g)=\cos \alpha/|\vec{r}(q,g)|$ follows from
221: classical potential theory \cite{POTTH}. Here $\alpha$ is the angle
222: between the outward normal to the interface at the point $g$ and the
223: vector $\vec{r}(q,g)$, see Fig.~\ref{fig1}.
224: 
225: \begin{figure}%[ptb]
226: \centerline{\includegraphics[width=9.5cm,clip=] {FIG1.eps}}
227: \caption{Geometry of the kernel $K(q,g)$. $\vec{\tau}$ and $\vec{n}$
228: are the tangential and outward normal directions, respectively.}
229: \label{fig1}
230: \end{figure}
231: 
232: 
233: The boundary condition (\ref{GT}) can be rewritten as an integral
234: equation for $\mu(q)$, see, \textit{e.g.}, \cite{POTTH}:
235: \begin{equation} \label{INTEQ1}
236: -\mu(q) + \frac{1}{\pi}\oint_{\gamma}\left[1+K(q,g)\right] \mu(g)\,
237: dg = 2\mathcal{K}(q)\,.
238: \end{equation}
239: %{\cal K}
240: %The solution, described by Eq. (\ref{PofQ}), is discontinuous at the
241: %interface $\gamma$, and Eq. (\ref{INTEQ1}) satisfies the jump
242: %relation \cite{GGM,POTTH}.
243: That is, to compute $p(q)$  in Eq. (\ref{PofQ}), one needs to solve
244: the integral equation (\ref{INTEQ1}).  Mikhlin \cite{Mikhlin} showed
245: that Eq. (\ref{INTEQ1}) has a unique solution for any smooth
246: $\gamma$ and integrable $\mathcal{K}$, while $p(q)$ from Eq.
247: (\ref{PofQ}) is a harmonic function in the exterior, satisfying the
248: boundary condition Eq. (\ref{GT}). This representation was employed
249: by Greenbaum \textit{et al.} \cite{GGM} for numerical analysis.
250: %\cal K
251: 
252: For the purposes of the free boundary problem
253: (\ref{LPLC})-(\ref{VN}) one only needs the value of
254: $\nabla_{n}p(q)$, $p\in \gamma$. A straightforward calculation of
255: $\nabla_{n}$ from the double layer potential would yield a
256: hypersingular integral, see below. One circumvents this difficulty
257: by resorting to theory of analytic functions, see Ref. \cite{GGM}
258: and references therein. Suppose $\mu(q)$ is known and introduce the
259: quantity
260: $$\tilde{p}(q)=\frac{1}{2\pi}\oint_{\gamma}\mu(g)K(q,g)\,dg\,,$$
261: which differs from $p(q)$ only by a constant, as
262: %=p(q)-\mbox{const},$$ since
263: $\oint_{\gamma}\mu(g)\,dg=\mbox{const}.$ Obviously,
264: $\nabla_{n}\tilde{p}=\nabla_{n}p$.  It is known \cite{GGM} that
265: $\tilde{p}(q)$ is the real part of the Cauchy integral
266: $$\frac{1}{2\pi i}\oint_{\gamma}\frac{\mu(\zeta)}{\zeta-z}d\zeta=\tilde{p}(z)+iV(z),$$
267: where we have identified the points $q$ and $g$ on the plane with
268: respective complex numbers $z$ and $\zeta$. Then $\tilde{p}$ and its
269: harmonic conjugate $V$ satisfy the Cauchy-Riemann equations, so that
270: \begin{equation}\label{CR}
271: \tilde{p}_{n}=V_{\tau}\,,
272: \end{equation}
273: where the indices $n$ and $\tau$ stand for the normal and tangential
274: derivatives, respectively. The kernel $K(q,g)$ can be written as
275: follows:
276: $$K(q,g)\,dg=\frac{-(y_{g}-y_{q})\,dx_{g}+(x_{g}-x_{q})\,dy_{g}}{r^{2}(q,g)}\,,$$
277: where $x$ and $y$ are the Cartesian coordinates of the respective
278: points. After a simple algebra we obtain
279: \begin{equation}\label{Vofz}
280: V(q)=-\frac{1}{2\pi}\oint_{\gamma}\frac{\mu(x_{g},y_{g})(x_{g}-x_{q})}{r^{2}(q,g)}\,dx_{g}+
281: \frac{\mu(x_{g},y_{g})(y_{g}-y_{q})}{r^{2}(q,g)}\,dy_{g}\,.
282: \end{equation}
283: 
284: \subsection{Discrete approximation}
285: 
286: Let us parameterize the closed interface $\gamma$ of the bubble:
287: $x=x(\sigma)$, $y=y(\sigma)$, $0 \leq \sigma \leq M$, $x(0)=x(M)$,
288: $y(0)=y(M)$, where $x$ and $y$ are the Cartesian coordinates of a
289: point belonging to the interface. In the parametric form Eq.
290: (\ref{INTEQ1}) becomes
291: \begin{equation} \label{INTEQPR}
292: - \mu(\sigma) +
293: \frac{1}{\pi}\int_{0}^{M}\mu(\xi)\left[1+\kappa(\sigma,\xi)
294: \right]\sqrt{\dot{x}^{2}+\dot{y}^{2}} \,d\xi  =
295: 2\mathcal{K}(\sigma)\,,
296: \end{equation}
297: %{\cal K}
298: where
299: \begin{equation}\label{Kksi}
300: \kappa(\sigma,\xi)=\frac{\dot{y}[x(\xi)-x(\sigma)]-\dot{x}[y(\xi)-y(\sigma)]}
301: {\{[x(\sigma)-x(\xi)]^{2}+[y(\sigma)-y(\xi)]^{2}\}\sqrt{\dot{x}^{2}+\dot{y}^{2}}}\,,
302: \end{equation}
303: while $\dot{x}=dx/d\xi$ and $\dot{y}=dy/d\xi$. The harmonic
304: conjugate takes the form
305: \begin{equation}\label{Vofksi}
306: V(\sigma)=-\frac{1}{2\pi}\int_{0}^{M}\mu(\xi)
307: \frac{\dot{x}[(x(\xi)-x(\sigma)]+\dot{y}[(y(\xi)-y(\sigma)]}{[(x(\xi)-x(\sigma)]^{2}+
308: [(y(\xi)-y(\sigma)]^{2}}\,d\xi.
309: \end{equation}
310: Note that the kernel $\kappa$ is continuous as $\xi \to \sigma$. On
311: the contrary, the integrand in  the last expression diverges as $\xi
312: \to \sigma$, and the integral exists only as a principal value.
313: \begin{figure}%[ptb]
314: \centerline{\includegraphics[width=8.5cm,clip=] {FIG2.eps}}
315: \caption{The discrete approximation scheme. Here $\xi_{j}=j.$}
316: \label{fig2}
317: \end{figure}
318: In the main case of our interest the bubble length is much greater
319: than its thickness $\Delta$. In the almost flat parts of the
320: interface $\dot{y} \approx 0$. Now, $y(\sigma)-y(\xi)\sim \Delta$
321: when the points $\sigma$ and $\xi$ belong to the different (upper
322: and lower) parts of the interface, while $y(\sigma)-y(\xi)\approx 0$
323: when they belong to the same part of the interface. Then, using the
324: relation $\dot{x}=dx/d\xi$, we can estimate the kernel $\kappa$ as
325: \begin{equation}\label{kappadelta}
326: \kappa(\sigma,\xi)\,d\xi\approx\frac{-\Delta}{[x(\sigma)-x(\xi)]^{2}+\Delta^{2}}\,dx\,,
327: \end{equation}
328: when $\sigma$ and $\xi$ belongs to the different parts of the
329: interface, while $\kappa \approx 0$ when they belong to the same
330: part. Equation (\ref{kappadelta}) shows that the typical scale of
331: variation of the kernel (\ref{Kksi}) over the almost flat part of
332: the interface is of order of the bubble thickness $\Delta$. A
333: similar estimate applies to the fraction entering the integrand of
334: Eq. (\ref{Vofksi}). A straightforward discretization would then
335: require a node spacing much less than $\Delta$. Instead, we rewrite
336: Eq. (\ref{INTEQPR}) as
337: %{\cal K}
338: \begin{equation}\label{inteqsm}
339: -\mu(\sigma) +
340: \frac{1}{\pi}\sum_{j=0}^{m-1}\int_{\xi_{j}}^{\xi_{j+1}}\mu(\xi)[1+\kappa(\sigma,\xi)
341: ]\sqrt{\dot{x}^{2}+\dot{y}^{2}}\,d\xi  = 2\mathcal{K}(\sigma)\,,
342: \end{equation} where
343: $\xi_{0}=0$, $\xi_{m}=M$, and $\xi_{j+1}>\xi_{j}$, $j=0, 1, 2,
344: \dots, m-1$. Introduce a piecewise linear approximation for $x(\xi)$
345: and $y(\xi)$ (see Fig. \ref{fig2}):
346: \begin{equation}\label{xappr}
347: x(\xi)=k_{j}^{x}\xi+b_{j}^{x}, \quad y(\xi)=k_{j}^{y}\xi+b_{j}^{y}, %\quad \xi_{j}\leq\xi\leq\xi_{j+1},\quad j=0, 1, 2, \dots, m-1,
348: \end{equation}
349: where $\xi_{j}\leq\xi\leq\xi_{j+1},\quad j=0, 1, 2, \dots, m-1,$
350:  $k_{j}^{x}=(x_{j+1}-x_{j})/(\xi_{j+1}-\xi_{j})$,
351: $b_{j}^{x}=(\xi_{j+1}x_{j}-\xi_{j}x_{j+1})/(\xi_{j+1}-\xi_{j})$,
352: $k_{j}^{y}=(y_{j+1}-y_{j})/(\xi_{j+1}-\xi_{j})$,
353: $b_{j}^{y}=(\xi_{j+1}y_{j}-\xi_{j}y_{j+1})/(\xi_{j+1}-\xi_{j})$,
354: $x_{j}=x(\xi_{j})$, $y_{j}=y(\xi_{j}),$ and a piecewise constant
355: approximation for $\mu$:
356: \begin{equation}\label{muappr}
357: \mu(\xi)=\mu_{j}=\mbox{const}, \qquad \xi_{j}\leq\xi\leq\xi_{j+1},
358: \end{equation}
359: Note that $\dot{x}(\xi_{j}\leq\xi\leq\xi_{j+1})=k^{x}_{j},$
360: $\dot{y}(\xi_{j}\leq\xi\leq\xi_{j+1})=k^{y}_{j}$. The kernel
361: (\ref{Kksi}) is therefore approximated as
362: $$\kappa(\sigma,\xi)=\frac{k^{y}_{j}[k^{x}_{j}\xi+b^{x}_{j}-x(\sigma)]-
363: k^{x}_{j}[k^{y}_{j}\xi+b^{y}_{j}-y(\sigma)]}
364: {\{[k^{x}_{j}\xi+b^{x}_{j}-x(\sigma)]^{2}+[k^{y}_{j}\xi+b^{y}_{j}-y(\sigma)]^{2}\}S_{j}}
365: %=$$$$
366: =\frac{Q_{j}(\sigma)/S_{j}}
367: {S_{j}^{2}\xi^{2}+B_{j}(\sigma)\xi+C_{j}(\sigma)},$$
368: where
369: $$
370: S_{j}=\sqrt{(k^{x}_{j})^{2}+(k^{y}_{j})^{2}}, \quad
371: Q_{j}(\sigma)=k^{y}_{j}[b^{x}_{j}-x(\sigma)]-k^{x}_{j}[b^{y}_{j}-y(\sigma)]\,,
372: $$
373: $$
374: B_{j}(\sigma)=2\{k^{x}_{j}[b^{x}_{j}-x(\sigma)]+k^{y}_{j}[b^{y}_{j}-y(\sigma)]\},\;\;\mbox{and}\;\;
375: C_{j}(\sigma)=[b^{x}_{j}-x(\sigma)]^{2}+[b^{y}_{j}-y(\sigma)]^{2}\,.
376: $$
377: The integrals in (\ref{inteqsm}) can be calculated analytically:
378: $$\int_{\xi_{j}}^{\xi_{j+1}}\mu(\xi)[1+\kappa(\sigma,\xi)
379: ]\sqrt{\dot{x}^{2}+\dot{y}^{2}}\,d\xi
380: %=$$$$
381: =\mu_{j}\int_{\xi_{j}}^{\xi_{j+1}}\left[S_{j}+\frac{Q_{j}(\sigma)}
382: {S_{j}^{2}\xi^{2}+B_{j}(\sigma)\xi+C_{j}(\sigma)}\right]\, d\xi = $$
383: $$=\mu_{j}\left\{S_{j}(\xi_{j+1}-\xi_{j})
384: %+$$$$
385: +\frac{1}{Q_{j}(\sigma)}
386: \left[\arctan\frac{2S_{j}^{2}\xi_{j+1}+B_{j}(\sigma)}{2Q_{j}(\sigma)}-
387: \arctan\frac{2S_{j}^{2}\xi_{j}+B_{j}(\sigma)}{2Q_{j}(\sigma)}\right]\right\}.$$
388: It is convenient to define $\xi_{j}=j$, then $\xi_{j+1}-\xi_{j}=1$.
389: In our discretization scheme
390: $x(\sigma)=x_{i+1/2}=k_{i}^{x}(i+1/2)+b_{i}^{x}$ and
391: $y(\sigma)=y_{i+1/2}=k_{i}^{y}(i+1/2)+b_{i}^{y}$. Let us denote
392: $Q_{j}(\sigma)=Q_{ij}$, $B_{j}(\sigma)=B_{ij}$, and
393: $C_{j}(\sigma)=C_{ij}.$ The integrals in Eq. (\ref{inteqsm}) are
394: $$\int_{\xi_{j}}^{\xi_{j+1}}\mu(\xi)[1+\kappa(\sigma,\xi)
395: ]\sqrt{\dot{x}^{2}+\dot{y}^{2}}\,d\xi=\pi \mu_{j}\textsc{A}_{ij},$$
396: where
397: %{\cal A}{\cal A}
398: $$\textsc{A}_{ij}=\frac{1}{\pi}\left\{S_{j}
399: %+$$$$
400: +\frac{1}{Q_{ij}}\arctan\frac{4S_{j}^{2}Q_{ij}}
401: {4Q_{ij}+(2S_{j}^{2}j+B_{ij})[2S_{j}^{2}(1+j)+B_{ij}]}\right\}.$$ We
402: have arrived at a set of linear algebraic equations with respect to
403: $\mu_{j}$, which is our approximation of the integral equation
404: (\ref{INTEQ1}):
405: \begin{equation} \label{INTEQdscr}
406: \sum_{j=0}^{m-1}(\textsc{A}_{ij}-\delta_{ij})\mu_{j}=2\mathcal{K}_{i},\quad
407: i=0, 1, 2, 3, \dots, m-1\,.
408: \end{equation}
409: %{\cal A}{\cal K}
410: %{\cal K}
411: We approximate the interface curvature $\mathcal{K}(\sigma)$ by
412: finite differences:
413: %{\cal K}{\cal K}
414: $$\mathcal{K}(\sigma=i+1/2)=\mathcal{K}_{i}=\frac{\ddot{y}_{i+1/2}\dot{x}_{i+1/2}-\ddot{x}_{i+1/2}\dot{y}_{i+1/2}}
415: {[(\dot{y}_{i+1/2})^{2}+(\dot{x}_{i+1/2})^{2}]^{3/2}},$$ where
416: $\dot{x}_{i+1/2}=k_{i}^{x}$, $\dot{y}_{i+1/2}=k_{i}^{y}$,
417: $\ddot{x}_{i+1/2}=x_{i+2}-x_{i+1}-x_{i}+x_{i-1}$, and
418: $\ddot{y}_{i+1/2}=y_{i+2}-y_{i+1}-y_{i}+y_{i-1}$.
419: %{\cal A}
420: Importantly, our approximation scheme yields the principal value of
421: the integral (\ref{Vofksi}) automatically. Furthermore, we can
422: directly compute the coefficients $\textsc{A}_{ij}$, using the same
423: expression for $i\neq j$ and $i=j$, where the kernel (\ref{Kksi})
424: has a removable discontinuity. The method suggested in \cite{GGM}
425: prescribes instead to use an analytic evaluation of the kernel at
426: the point of removable discontinuity.
427: 
428: We solved the algebraic equations  (\ref{INTEQdscr})  by  an
429: iterative refinement method after a LU factorization of  the matrix.
430: As the maximum number of equations in the examples that we
431: considered (see below) did not exceed $1100$, there was no need to
432: use more sophisticated methods.
433: 
434: 
435: \subsection{Grid}
436: Most of our results were obtained with the version of the code which
437: assumed a four-fold symmetry of the bubble. This allowed us to work
438: with a one quarter of the interface and reduce the number of nodes
439: by 4. In the beginning of the bubble relaxation, the solution varies
440: rapidly in the region of the lobes, and very slowly in the flat
441: region of the bubble. Therefore one should employ here a non-uniform
442: grid. At later times, when the aspect ratio of the bubble becomes
443: comparable to unity, the code switches to a uniform grid. For the
444: non-uniform grid we used an exponential spacing. Here the node
445: spacing grows exponentially from the lobe's end to the middle of the
446: flat part of the bubble. To  generate the node distribution we use
447: the following procedure. Let the quarter of the interface perimeter
448: be $\Pi$, the specified number of nodes be $m$, and the specified
449: \textit{smallest} spacing in the lobe region be $h_{0}$.
450: 
451: If $\Pi>h_{0}(m-1)$, the exponential grid is used. Here we introduce
452: the quantity $\eta$ which satisfies the condition
453: \begin{equation}\label{eta}
454: \Pi=h_{0}+\eta h_{0}+\eta^{2} h_{0} + \ldots + \eta^{m-2} h_{0}=
455: \frac{h_{0}(1-\eta^{m-1})}{1-\eta}\,,
456: \end{equation}
457: solve Eq. (\ref{eta}) numerically  for $\eta$, use a discrete
458: arclength parametrization: $\xi_{1}=0,\quad
459: \xi_{2}=h_{0},\dots,\xi_{k}=\eta^{k-2}
460: h_{0},\dots,\xi_{m}=\eta^{m-2} h_{0}$, and calculate the arrays
461: $x=x(\xi_{i})$ and $y=y(\xi_{i})$, where $i=1, 2, \dots, m$.
462: 
463: In the process of the interface evolution $\Pi$ decreases with time,
464: so one can reduce the number of the grid nodes. Furthermore, as the
465: nodes in our code move like lagrangian particles (see Section 3),
466: the node spacing in the lobe region decreases with time even faster.
467: If left unattended, this would cause instability of the node
468: tracking (see Section 3), as the maximum allowed time step is
469: proportional to the node spacing cubed \cite{Beale}.   Therefore,
470: when the minimum node spacing decreases below $\xi_{2}=0.8 h_{0}$,
471: we redistribute the nodes: we look for the new value of $\eta$,
472: corresponding to the updated value of $\Pi$, calculate the new array
473: of $\xi$, and determine the new arrays $x$ and $y$ by linear
474: interpolation.
475: 
476: When the perimeter goes down so that $\Pi\leq h_{0}(m-1)$, we switch
477: to a uniform grid. Here we calculate a new $m$: $m=[\Pi/h_{0}]+1$,
478: where $[a]$ is an integer number such that $0\leq a-[a]<1$, and fine
479: tune $h_0$ so that $h_{0}=\Pi/(m-1)$.
480: 
481: Finally, the choice of $h_0$ is dictated by a compromise between the
482: desired accuracy and the value of $m$ which determines the size of
483: the matrix $A_{ij}$.
484: 
485: \subsection{Calculation of the normal velocity}
486: After the set of linear equations (\ref{INTEQdscr}) is solved, and
487: the quantities $\mu_{i}$ are found, we compute the harmonic
488: conjugate $V$. The same approximation, applied to Eq.
489: (\ref{Vofksi}), yields:
490: $$V_{i}=-\frac{1}{2\pi}\sum_{j=0}^{m-1}\mu_{j}F_{ij}\,,$$
491: where
492: $$F_{ij}=\int_{j}^{j+1}
493: \frac{k_{j}^{x}(k_{j}^{x}\xi+b_{j}^{x}-x_{i+1/2})+k_{j}^{y}(k_{j}^{y}\xi+b_{j}^{y}-y_{i+1/2})}
494: {(k_{j}^{x}\xi+b_{j}^{x}-x_{i+1/2})^{2}+(k_{j}^{y}\xi+b_{j}^{y}-y_{i+1/2})^{2}}\,d\xi
495: $$
496: $$=\frac{1}{2}\int_{j}^{j+1}\frac{2S_{j}^{2}\xi+B_{ij}}
497: {S_{j}^{2}\xi^{2}+B_{ij}\xi+C_{ij}}\,d\xi
498: %=$$$$
499: =\frac{1}{2}\ln\frac{S_{j}^{2}(j+1)^{2}+B_{ij}(j+1)+C_{ij}}{S_{j}^{2}j^{2}+B_{ij}j+C_{ij}}\,,$$
500: where the quantities $S_{j}$, $B_{ij}$ and $C_{ij}$ were defined
501: earlier. Again, the integral is calculated analytically. The
502: resulting formula for $V_{i}$ is the following:
503: \begin{equation} \label{Vfinal}
504: V_{i}=-\frac{1}{4\pi}\sum_{j=0}^{m-1}\mu_{j}
505: \ln\frac{S_{j}^{2}(j+1)^{2}+B_{ij}(j+1)+C_{ij}}{S_{j}^{2}j^{2}+B_{ij}j+C_{ij}},
506: \end{equation}
507: where $V_{i}=V(\sigma=i+1/2),$ $i=0,1,2,3,\dots, m-1$, see Fig.
508: \ref{fig2}. Note that for $\xi=i+1/2$ the denominator of the
509: integrand in $F_{ij}$ vanishes, and the integrand diverges. To
510: overcome this problem, Ref. \cite{GGM} suggested to divide the mesh
511: into odd and even nodes and compute $V$ at the odd points by summing
512: over the even nodes, and vice-versa. Our analytical integration
513: yields the correct principal value of the integral, so there is no
514: need  to use the recipe of Ref. \cite{GGM}.
515: 
516: To compute the normal velocity of the interface we use the
517: Cauchy-Riemann equation (\ref{CR}) and approximate the derivative of
518: $V$ with respect to the arclength:
519: $$v_{n}(\sigma=i)=\frac{(V_{i+1}-V_{i})}
520: {\tilde{s}_{i}}\,,$$
521: where
522: $$\tilde{s}_{i}=\frac{1}{2}\left[\sqrt{(x_{i+1}-x_{i})^{2}+(y_{i+1}-y_{i})^{2}}+
523: %$$$$+
524: \sqrt{(x_{i}-x_{i-1})^{2}+(y_{i}-y_{i-1})^{2}}\right].$$
525: \section{Interface Tracking}
526: To track the interface, we use an explicit first-order integration:
527: $$x_{i}(t+\Delta t)=x_{i}(t)+\Delta t\, v_{n}(\sigma=i,t)\cos n_{i},$$
528: $$y_{i}(t+\Delta t)=y_{i}(t)+\Delta t\, v_{n}(\sigma=i,t)\sin n_{i},$$
529: where
530: $$\cos n_{i}=\frac{\dot{y}_{i}}{\sqrt{\dot{x}^{2}_{i}+\dot{y}^{2}_{i}}}\,, \quad
531: \sin
532: n_{i}=-\frac{\dot{x}_{i}}{\sqrt{\dot{x}^{2}_{i}+\dot{y}^{2}_{i}}}\,,$$
533: $\dot{x}_{i}=(\dot{x}_{i+1/2}+\dot{x}_{i-1/2})/2=(k^{x}_{i}+k^{x}_{i-1})/2$,
534: and $\dot{y}_{i}=(k^{y}_{i}+k^{y}_{i-1})/2$. We have assumed the
535: counter-clockwise direction of the interface parametrization, see
536: Fig. \ref{fig1}.
537: 
538: It is important to prescribe the time step $\Delta t$ properly. We
539: employ an ad-hoc criterion which demands that the node displacement
540: at each grid point be considerably less then the curvature radius
541: $R_i$ at that point: $\min |(\Delta t\, v_{n}(i))/R_{i}|\leq
542: \varepsilon,\quad 0\leq i \leq m-1.$ That is, we consider the
543: curvature radius $R_i$ as a natural local length scale of the
544: problem. A more convenient form of this criterion is
545: \begin{equation} \label{epsilon}
546: \Delta t=\varepsilon \min\{|R_{i}/v_{n}(i)|\}\,,
547: \end{equation}
548: where $\varepsilon$ is an input parameter which has to be
549: sufficiently small to satisfy the requirements of stability and
550: desired accuracy. In the exact formulation (\ref{LPLC})-(\ref{VN})
551: the bubble area must be constant in the process of relaxation. The
552: area conservation can be conveniently used for accuracy control of
553: the code.
554: 
555: \section{Numerical Results}
556: We present here some simulation results produced with our code for
557: two different sets of initial conditions. One of them describes the
558: decay of a small sinusoidal perturbation of a perfectly circular
559: bubble of inviscid fluid. An approximate analytical solution to this
560: problem is given by the linear stability analysis \cite{linear}, and
561: we used this solution to test the code.
562: 
563: The second initial condition describes a very long and thin strip of
564: inviscid fluid. In the process of its shrinking the bubble develops
565: a dumbbell shape, while the characteristic dimensions of the
566: dumbbell exhibit asymptotic scaling laws found in Ref. \cite{VMS}.
567: 
568: \subsection{Relaxation of a slightly perturbed circle}
569: Let the initial shape of the interface be a circle with a small
570: sinusoidal perturbation:
571: $$
572: \rho(\varphi,0)=R_{0}[1+\delta(0)\sin(n\varphi)]\,,$$ where $\rho$
573: and $\varphi$ are the polar radius and angle, respectively, $R_{0}$
574: is the radius of the unperturbed interface, while $\delta(0)$ and
575: $n$ are the initial amplitude and azimuthal number of the
576: perturbation. The analytical solution provided by the linear theory
577: \cite{Paterson} is
578: $$
579: \rho(\varphi,t)=R_{0}[1+\delta(t)\sin(n\varphi)]\,,
580: $$
581: where the amplitude of the perturbation is
582: $$\delta(t)=\delta(0)\exp\left[-\frac{n(n^{2}-1)}{R_{0}^{3}}t\right],$$
583: A typical numerical result is presented in Fig. \ref{fig3}. The
584: parameters are $R_{0}=100,$ $\delta(0)=0.01$ and $n=4$. In the case
585: of $n=4$ the interface has a four-fold symmetry which allows a
586: direct application of our code. In this simulation the quarter of
587: the interface was described by 100 nodes. The initial spacing was
588: uniform. The code did not have to use the mesh interpolation in this
589: example. The parameter regulating the time step was
590: $\varepsilon=5\cdot10^{-5}$. As one can see, a very good agreement
591: with the analytical result is obtained.
592: \begin{figure}%[ptb]
593: \centerline{\includegraphics[width=7.5cm,clip=] {FIG3.eps}}
594: \caption{Shown in the logarithmic scale is the perturbation
595: amplitude $\delta$ as a function of time. The squares are the
596: simulation results, the solid line is the analytical prediction. }
597: \label{fig3}
598: \end{figure}
599: 
600: \subsection{Relaxation of a long and thin bubble}
601: In the second setting the initial interface shape is a very long
602: rectangular strip. In the example we report here the initial strip
603: thickness was 1, and the initial length 2000. Here we could compare
604: the numerical results with the predictions of a recent asymptotic
605: scaling analysis \cite{VMS}. The interface shapes at different times
606: are presented in Fig. \ref{fig4}. It can be seen that the shrinking
607: strip acquires the shape of a dumbbell (or petal). At much later
608: times it approaches circular shape. By the end of the simulation (at
609: $t=48000$) the relative deviation of the observed shape from the
610: perfect circle, $[\rho_{max} (\varphi) - \rho_{min}
611: (\varphi)]/\rho_{min} (\varphi) \approx  0.013$.
612: \begin{figure}%[ptb]
613: \includegraphics[width=13.5 cm,clip=] {FIG4.eps} \caption{Figure a shows  a snapshot
614: of one half of the simulated system at $t=0$, $3670$, $7 020$, and
615: $24 840$. Notice the large difference between the horizontal and
616: vertical scales. Figure b shows the lobe of the dumbbell to scale at
617: $t=7 020$, while Figure c shows the computed bubble shape at late
618: times: $t=30 900$, $34 200$ and $48 000$.} \label{fig4}
619: \end{figure}
620: \begin{figure}%[ptb]
621: \centerline{\includegraphics[width=11.0cm,clip=] {FIG5.eps}}
622: \caption{Figure a shows, in a log-log scale, the retreat distance
623: $L$ versus time and its power-law fit $2.75 \,t^{0.60}$. Figure b
624: shows, in a log-log scale, the maximum dumbbell height, $h^{max}$
625: (the empty circles), and the position of the maximum, $x_1^{max}$
626: (the filled circles), versus time, as well as their power-law fits
627: $0.66\,t^{0.21}$ and $0.94\,t^{0.20}$, respectively.}
628: \label{lengths}
629: \end{figure}
630: 
631: The asymptotic scaling analysis \cite{VMS} deals with the
632: intermediate stage of the relaxation. Introduce the retreat distance
633: $L(t)=1000-x_{tip}(t)$, where $x_{tip}(t)$ is the maximum abscissa
634: of all points belonging to the interface. One prediction of Ref.
635: \cite{VMS} is that, at intermediate times, $L(t)\propto t^{3/5}$.
636: Figure \ref{lengths}a shows a very good agreement of this prediction
637: with the simulation result. Additional predictions of asymptotic
638: scaling analysis deal with the time dependence of the maximum
639: dumbbell elevation $h^{max}(t)$, and of the abscissa of the
640: corresponding point of the interface $x^{max}(t)$. Let us introduce
641: a new variable: $x_{1}(x,t)=x_{tip}(t)-x$, the distance along the
642: $x$-axis between the tip of the dumbbell and a point $x$. In
643: particular, $x_{1}^{max}(t)=x_{tip}(t)-x^{max}(t)$. A comparison of
644: the simulation results with the predicted intermediate-time scaling
645: laws $h^{max}(t)\propto x_{1}^{max}(t) \propto t^{1/5}$ is shown in
646: Figure \ref{lengths}b, and again a very good agreement is observed.
647: 
648: To verify the self-similarity of the dumbbell shape in the lobe
649: region, predicted in Ref. \cite{VMS}, we introduce a new function
650: $h(x_1,t)$ so that $h[x_{1}(x,t),t]=y(x,t)$. Figure \ref{SS} shows
651: the spatial profiles of $h$ rescaled to the values of $h^{max}$
652: versus $x_1/x_1^{max}$ at three different times. The observed
653: collapse in the lobe region confirms the expected self-similarity.
654: 
655: \begin{figure}%[ptb]
656: \centerline{\includegraphics[width=8.0cm,clip=] {FIG6.eps}}
657: \caption{Self-similarity of the lobe. Shown is the shape function
658: $h(x_1,t)$, rescaled to the maximum dumbbell elevation, versus the
659: coordinate $x_1$, rescaled to the abscissa of the maximum, at times
660: $160.3$ (the filled circles), $1000$ (the squares), and $3010$ (the
661: empty circles).} \label{SS}
662: \end{figure}
663: The initial number of nodes in this simulation was 1100, and the
664: smallest spacing in the lobe region was 0.4.  With the grid
665: interpolation employed, the time-step parameter $\varepsilon=0.005$
666: proved sufficiently small to guarantee stability and good accuracy.
667: As the curvature of the interface goes down during the evolution,
668: the required time step increases significantly. It was  $1.7 \times
669: 10^{-3}$ at $t=0$, $0.22$ at $t=3670$ and increased up to about 10
670: by the end of the simulation, at $t=48000$. We used the small
671: observed area loss of the bubble for accuracy control. The observed
672: area loss was less then $0.5\%$ for $t<10000$. By the end of the
673: simulation, at $t=48000$, the area loss reached only $2.8\%$.
674: 
675: \section{Conclusion}
676: We have developed and tested a new numerical version of the boundary
677: integral method for an exterior Dirichlet problem, which is
678: especially suitable for long and thin domains. The method allows one
679: to significantly reduce the number of the interfacial nodes. The new
680: method was successfully tested in a numerical investigation of the
681: shape relaxation, by surface tension, of a long and thin bubble,
682: filled with an inviscid fluid and immersed in a viscous fluid in a
683: Hele-Shaw cell. Here we confirmed the recent theoretical predictions
684: on the self-similarity and dynamic scaling behavior during an
685: intermediate stage of the bubble dynamics.
686: 
687: \section*{Acknowledgment}
688: This work was supported by the Israel Science Foundation, Grant No.
689: 180/02.
690: 
691: \begin{thebibliography}{111}
692: \bibitem{Langer1} J.S. Langer, in \textit{Chance and Matter},
693: edited by J. Souletie, J. Vannimenus, and R. Stora, Elsevier,
694: Amsterdam, 1987.
695: \bibitem{BKD}D. Bensimon, L.P. Kadanoff, S.D. Liang, B.I. Shraiman, C. Tang, Viscous flow in two dimensions, Rev. Mod.
696: Phys. \textbf{58} (1986) 977-999.
697: \bibitem{Kessler} D.A. Kessler, J. Koplik, H. Levine, Pattern selection in fingered growth phenomena, Adv.
698: Physics \textbf{37} (1988) 255-339.
699: \bibitem{Casademunt1} J. Casademunt, F.X. Magdaleno, Dynamics and selection of fingering
700: patterns. Recent developments in the Saffman-Taylor problem, Phys.
701: Rep. \textbf{337} (2000) 1-35.
702: \bibitem{Casademunt2} J. Casademunt, Viscous fingering as a paradigm of
703: interfacial pattern formation: Recent results and new challenges,
704: Chaos 14 (2004) 809-824.
705: \bibitem{CLM} M. Conti, A. Lipshtat, B. Meerson, Scaling anomalies in the coarsening
706: dynamics of fractal viscous fingering patterns, Phys. Rev. E
707: \textbf{69} (2004) 031406 (1-4).
708: \bibitem {VMS}A. Vilenkin, B. Meerson, P.V. Sasorov, Scaling and self-similarity in an
709: unforced flow of inviscid fluid trapped inside a viscous fluid in a
710: Hele-Shaw cell, Phys. Rev. Lett. (submitted).
711: \bibitem{ConstantinPugh} P. Constantin, M. Pugh, Global solutions for small data
712: to the Hele-Shaw Problem, Nonlinearity \textbf{6} (1993) 393-415.
713: \bibitem{linear} The damping rates of small sinusoidal
714: perturbations of circular and flat interfaces are given by the
715: zero-flow-rate limit of Eq.~(11) of Ref. \cite{Paterson} (for the
716: circular interface) and Eq.~(10) of Ref. \cite{ST} (for the flat
717: interface).
718: \bibitem{HLSH} T. Y. Hou, J.S. Lowengrub, M.J. Shelley, Boundary integral methods
719: for multicomponent fluids and multiphase materials,  J. Comput.
720: Phys. \textbf{169} (2001) 302-362.
721: \bibitem {DKZ} W.-S. Dai, L.P. Kadanoff, S.M. Zhou, Interface dynamics and the motion
722: of complex singularities, Phys. Rev. A \textbf{43} (1991) 6672-6682.
723: \bibitem{Tanv} S. Tanveer, Surprises in Viscous Fingering, J. Fluid Mech. \textbf{409} (2000) 273-308.
724: \bibitem{Folch1} R. Folch, J. Casademunt, A. Hernandez-Machado, L.
725: Ramirez-Piscina, Phase-field model for Hele-Shaw flows with
726: arbitrary viscosity contrast. I. Theoretical approach, Phys. Rev. E
727: \textbf{60}  (1999) 1724-1733.
728: \bibitem{Folch2} R. Folch, J. Casademunt, A. Hernandez-Machado, L.
729: Ramirez-Piscina, Phase-field model for Hele-Shaw flows with
730: arbitrary viscosity contrast. II. Numerical study, Phys. Rev. E
731: \textbf{60}  (1999) 1734-1740.
732: \bibitem {Glasner} K. Glasner, A diffuse inerface approach to Hele-Shaw flow,
733: Nonlinearity \textbf{16} (2003) 49-66.
734: \bibitem {POTTH} A. Tikhonov, A. Samarskii, Equations of Mathematical Physics, Pergamon Press, Oxford, 1963.
735: \bibitem {GGM} A. Greenbaum, L. Greengard, G.B. McFadden, Laplace's equation
736: and the Dirichlet-Neumann map in multiply connected domains, J.
737: Comput. Phys. \textbf{105} (1993) 267-278.
738: \bibitem{Mikhlin} S.G. Mikhlin,  Integral Equations,
739: London, Pergamon, 1957.
740: \bibitem {Beale} J.T. Beale, T.Y. Hou, J.S. Lowengrub, On the well-posedness of
741: two fluid interfacial flows with surface tension, in Singularities
742: in Fluids, Plasmas and Optics, edited by R. Caflish and G.
743: Papanicolaou, NATO Adv. Sci. Inst. Ser. A, Kluwer Academic,
744: Amsterdam, 1993, p. 11.
745: \bibitem {Paterson} L. Paterson, Radial fingering
746: in a Hele Shaw cell, J. Fluid Mech. \textbf{113} (1981) 513-529.
747: \bibitem{ST} P.G. Saffman, G.I. Taylor, The penetration of a fluid into a porous medium or
748: Hele-Shaw cell containing a more viscous liquid, Proc. R. Soc.
749: London, Ser. A  \textbf{245} (1958) 312-329.
750: 
751: \end{thebibliography}
752: 
753: \end{document}
754: