1: \documentstyle[pre,aps,multicol,epsfig]{revtex}
2: %\documentstyle[preprint,pre,epsfig,aps,eqsecnum]{revtex}
3: %\pagestyle{plain}
4:
5: \begin{document}
6:
7: \draft
8:
9: \newcommand{\be}{\begin{eqnarray}}
10: \newcommand{\ee}{\end{eqnarray}}
11:
12: \title{Systematic weakly nonlinear analysis of interfacial instabilities in Hele--Shaw flows}
13:
14: \author{
15: E. Alvarez--Lacalle, J. Casademunt and J. Ort\'{\i}n. }
16: \address{
17: Departament d'Estructura i Constituents de la Mat\`eria\\
18: Universitat de Barcelona,
19: Av. Diagonal, 647, E-08028 Barcelona, Spain}
20: \date{\today}
21: \maketitle
22:
23:
24:
25: \begin{abstract}
26: We develop a systematic method to derive all orders of mode
27: couplings in a weakly nonlinear approach to the dynamics of the
28: interface between two immiscible viscous fluids in a Hele--Shaw
29: cell. The method is completely general. It includes both the
30: channel geometry driven by gravity and pressure, and the radial
31: geometry with arbitrary injection and centrifugal driving. We
32: find the finite radius of convergence of the mode--coupling
33: expansion. In the channel geometry we carry out the calculation
34: up to third--order couplings, which is necessary to account for
35: the time--dependent Saffman--Taylor finger solution and the case
36: of zero viscosity contrast. Both in the channel and the radial
37: geometries, the explicit results provide relevant analytical
38: information about the role that the viscosity contrast and the
39: surface tension play in the dynamics. We finally check the
40: quantitative validity of different orders of approximation against
41: a physically relevant, exact time--dependent solution. The
42: agreement between the low order approximations and the exact
43: solution is excellent within the radius of convergence, and
44: reasonably good even beyond that.
45: \end{abstract}
46:
47: \noindent PACS: 47.20.Ma, 47.20.Hw, 47.54.+r, 47.20.Ky
48:
49: \section{Introduction}
50:
51: The morphological instability of fluid interfaces in Hele--Shaw flows
52: \cite{Bensimon86,Saffman86} has become a paradigm of interfacial
53: pattern formation in nonequilibrium systems
54: \cite{Pelce88,LangerlesHouches,Kesslerkop88}. As opposed to the most
55: commonly studied ``bulk'' pattern forming systems \cite{Crosshoh93},
56: the inherent difficulties of the free--boundary problems associated to
57: interfacial growth makes the latter even more elusive to analytical
58: treatment. As a prototype of interfacial instabilities in
59: diffusion--limited growth problems (including for instance dendritic
60: growth, solidification of mixtures, chemical electrodeposition, flame
61: propagation, etc.) the Saffman--Taylor problem \cite{Saffman58} is a
62: relatively simple case, both theoretically and experimentally,
63: well suited to gain insight into generic dynamical features in the
64: broad context of nonlinear interface phenomena.
65:
66: The intrinsic difficulty of free--boundary problems is manifest in
67: the fact that the interface dynamics is highly non--local.
68: Furthermore, the nature of the instability (except in some cases,
69: such as in directional solidification of binary alloys) usually
70: produces non--saturated growth, which inevitably results in a
71: highly nonlinear dynamics. In bulk instabilities, when the
72: control parameter is near threshold, the traditional weakly
73: nonlinear techniques lead to a universal description of patterns
74: in terms of amplitude equations, based on center manifold
75: reduction \cite{Crosshoh93,Manneville90}. These techniques,
76: however, are not so useful for interfacial problems in which
77: nonlinearities do not saturate the growth. This is the case for
78: instance of viscous fingering. In the case of the channel
79: geometry (Saffman--Taylor problem) \cite{Bensimon86,Saffman58} the
80: interface restabilizes in a nontrivial morphology (the
81: Saffman--Taylor finger) which keeps growing at a finite rate. For
82: circular geometries \cite{Paterson81,JDChen89} the patterns do not
83: reach an equivalent steady state and the interplay of
84: tip--splitting events and screening effects may result in a
85: variety of complicated morphologies. In these problems all the
86: weakly nonlinear techniques apply only to a transient in the early
87: nonlinear regime. Nevertheless, compared to the more traditional
88: ones \cite{Crosshoh93,Manneville90}, the weakly nonlinear analysis
89: developed in this paper is not restricted to situations near the
90: instability threshold, where a separation of scales is exploited.
91: Instead, we expand on the amplitudes of the whole spectrum of
92: modes.
93:
94: In this paper we will deal both with channel and radial geometry
95: Hele--Shaw flows. In the traditional Saffman--Taylor problem
96: (channel geometry) the pressure-- and gravity--driven
97: instabilities can be formally mapped into each other in the
98: appropriate reference frames, so there is really no different
99: interface dynamics for the two physical situations. The problem,
100: then, contains two independent dimensionless parameters, namely, a
101: dimensionless surface tension $B$ and the viscosity contrast or
102: Atwood ratio $A$ \cite{Trygvason83}. In the radial geometry,
103: though, there is no such formal mapping. We will see that the
104: injection and the centrifugal forcing are not equivalent and three
105: independent parameters must be considered.
106:
107: The situation most commonly studied in the literature is the high
108: viscosity contrast limit, $A=1$, where one of the two fluids is
109: non--viscous \cite{Bensimon86} (typically air displacing a viscous
110: fluid). The singular perturbation character of the surface
111: tension $B$ has received most of the attention as responsible for
112: the subtle mechanism of {\it steady--state selection}, namely the
113: fact that surface tension ``selects'' a single finger solution out
114: of a continuum of solutions for $B=0$
115: \cite{Hong86,Shraiman86,Combescot86}. More recently, the crucial
116: role of surface tension in the \it dynamics \rm of fingering
117: patterns has been pointed out. Siegel and Tanveer
118: \cite{Tanveer93,Siegelprl96,Siegeljfm96} have shown that the
119: exact, nonsingular time--dependent solutions known for the case
120: with $B=0$ may differ significantly from the corresponding ones
121: with $B \rightarrow 0^{+}$ after a time which is of order one
122: ($B^{0}$). In practice, this implies that exact solutions of the
123: problem with $B=0$ (including those with no finite time
124: singularities) may lead to completely incorrect asymptotic
125: behaviour as compared to the regularized ones, with $B\ll 1$. A
126: careful analysis of these questions may be found in
127: Ref.\cite{Magdaleno00}. Notice however that such analysis
128: restricts to small $B$, while in many cases (for instance, for
129: fingers emerging naturally from the linear instability, with the
130: characteristic length scale of the linearly most unstable mode)
131: the effective dimensionless surface tension is necessarily $B \sim
132: 1$. Understanding the dynamics of finger competition in typical
133: experimental conditions thus requires considering relatively large
134: values of $B$, for which the perturbative techniques of
135: \cite{Tanveer93,Siegelprl96,Siegeljfm96} fail.
136:
137: On the other hand, an important role of viscosity contrast $A$
138: in the dynamics of finger competition has been observed both numerically
139: \cite{Trygvason83,Trygvason85} and experimentally
140: \cite{Maher85,Difranjima89,Difranjimb89}. A careful characterization
141: of the interface evolution has shown that for $A=0$ the finger
142: competition process is ineffective, and that the system does not
143: approach the usual single finger Saffman--Taylor attractor
144: \cite{Casademunt91,Casademunt94}. Although the nature or existence
145: of other attractors is still an open question, it seems that the
146: basin of attraction of the Saffman--Taylor solution does depend on
147: $A$, and is particularly sensitive to $A$ in the neighborhood of $A
148: \simeq 1$. In any case, it is clear that the viscosity contrast
149: plays also a crucial role in the highly nonlinear regime, and that
150: tuning $A$ in its full range is necessary to elucidate some of the
151: important open questions.
152:
153: Finally, an interesting interplay between $B$ and $A$ in
154: connection with the selection problem is apparent in that, despite
155: the fact that single finger stationary solutions of any width do
156: exist for $B=0$ regardless of viscosity contrast $A$, the only
157: single finger time dependent solution of the $A=1$ ($B=0$) problem
158: which is also a solution for any viscosity contrast $A$ is the one
159: that fills one half of the channel \cite{Francessos1,Folch00},
160: which is precisely the solution selected by surface tension in the
161: limit $B \rightarrow 0$. Whether deeper consequences concerning
162: the selection problem can be drawn from this fact is also an
163: interesting open question.
164:
165: In order to gain analytical insight into these dynamical questions we
166: propose here a systematic weakly nonlinear expansion of the problem of
167: viscous fingering in Hele--Shaw flows, including all traditional setups
168: and the most recent one of rotating flows \cite{Carrillo96}. The basic
169: motivation is to be able to extract information which is nonperturbative
170: in any of the two basic parameters, which are taken as completely
171: arbitrary. The expansion parameter will be basically the mode
172: amplitudes.
173:
174: In this paper we will focus on unstably stratified flows, for which
175: the approach is necessarily restricted to the early evolution of the
176: interface. Although some of the nontrivial dynamic effects mentioned
177: above are associated to the highly nonlinear regime, it may be useful to
178: know within a controlled approximation to what extent those or other
179: effects show up already at the early stages of nonlinear mode coupling.
180:
181: For the stably stratified case the weakly nonlinear analysis is
182: obviously valid for long times since all mode amplitudes decay with
183: time. Although this configuration may seem trivial, this is not the
184: case in some situations, for instance when some external source
185: systematically drives the interface out of its equilibrium state (planar
186: or circular). An example of this is the presence of noise sources, such
187: as the quenched noise associated to a porous medium \cite{porous}. In a
188: study of the long--wavelength, low--frequency scaling properties of the
189: interface fluctuations, the knowledge of the lowest order nonlinear
190: terms and their dependence on parameters such as viscosity contrast is
191: crucial. In this context the weakly nonlinear expansion is the starting
192: point of any Renormalization Group analysis of the relevant terms and of
193: the fixed points of the problem. This line of research is clearly
194: beyond the scope of this paper and will not be pursued here.
195:
196: The basic ideas of weakly nonlinear analysis developed here were
197: first applied to viscous fingering by Miranda and Widom
198: \cite{Mirandac98,Mirandar98} to both channel and circular geometry
199: (with fluid injection). The present work is in part an extension
200: of those previous contributions in several directions, and in part
201: a detailed study of selected particular situations to assess the
202: validity and limitations of the approach.
203:
204: First of all, we provide a fully systematic methodology which may be
205: carried out to arbitrary order. We explicitly calculate the results
206: up to third order to include important situations for which the
207: second order couplings, first discussed in
208: Refs.\cite{Mirandac98,Mirandar98}, vanish identically, such as for
209: $A=0$ or for configurations with up--down symmetry (which includes the
210: physically relevant time dependent single finger solution with width
211: one half). We give the explicit predictions both in real and Fourier
212: space.
213:
214: In our formulation we also address the case of centrifugal forcing
215: of Hele--Shaw flows \cite{Schwartz89}. The experimental study of
216: rotating Hele--Shaw flows has revealed a rich variety of new
217: phenomena \cite{Carrillo96,Carrillo99,Carrillo00}. From a
218: theoretical point of view, new classes of exact solutions with
219: $B=0$ have been found \cite{Entov96,Crowdy1}. The role of
220: rotation in the possible suppression of finite time cusp
221: singularities in the absence of surface tension has been discussed
222: in \cite{Rocco00}. From an experimental point of view, important
223: differences in pattern morphology and new dynamical effects have
224: been found for low viscosity contrasts \cite{Alvarez01}. It seems
225: thus important to have this case included in the weakly nonlinear
226: formalism.
227:
228: In addition, our study extends the earlier results of Miranda
229: and Widom \cite{Mirandac98,Mirandar98} with the discussion of the
230: convergence of the weakly nonlinear analysis. We find the explicit
231: exact criterion to assure uniform convergence of the series. Beyond
232: that condition the series is asymptotic and different resummation
233: schemes are also explored. The different orders of approximation,
234: including possible resummations, are carefully compared to exact
235: solutions for the case of a single finger configuration. We find that
236: in some cases the agreement even at relatively low orders is quite
237: remarkable.
238:
239: The layout of the rest of the paper is as follows: in Section
240: \ref{Sec:2} we introduce the formalism. Section \ref{Sec:3} deals with
241: the derivation of the weakly nonlinear equations and their application
242: to Hele--Shaw flows in channel geometry. The analysis of the radial
243: geometry is carried out in Section \ref{Sec:4}, and Section \ref{Sec:5}
244: presents a numerical analysis of exact and approximate solutions. The
245: main results and the conclusions are summarized in Section
246: \ref{conclusions}.
247:
248:
249: \section{Vortex sheet formalism}
250: \label{Sec:2}
251:
252: \subsection{Channel Geometry}
253:
254: Let us first consider the Hele--Shaw problem in the channel
255: geometry. We consider fluid 1 (viscosity $\mu_{1}$, density
256: $\rho_{1}$) below fluid 2 ($\mu_{2},\rho_{2}$) (Fig.\
257: \ref{Figcanal}). The $\hat{z}$ axis is perpendicular to the cell.
258: A velocity $V_{\infty}$ is imposed at infinity in the $\hat{y}$
259: direction. Gravity points from fluid 2 to fluid 1. The width of
260: the cell is $L$, the gap between plates is $b$, and the surface
261: tension between the fluids is $\sigma$.
262:
263: The equations of motion for the interface and the boundary
264: conditions are well known \cite{Saffman58}. Here we will use the
265: formulation of Tryggvason and Aref \cite{Trygvason83}. We
266: introduce the velocity $\vec{U}=(\vec{u}_{1} + \vec{u}_{2})/2 $ as
267: the mean of the two limiting values of the velocities from both
268: sides of the interface ($\vec{u}_{1}, \vec{u}_{2}$) at a given
269: point. This velocity $\vec{U}$ can be expressed in terms of the
270: vortex sheet distribution $\gamma$ at the interface as:
271: \be
272: \vec{U}(s,t)=\frac{1}{2\pi} P \int \frac{\hat{z} \times
273: \left( \vec{r}(s,t)-\vec{r}(s^{\prime},t) \right) }{\mid
274: \vec{r}(s,t) - \vec{r}(s^{\prime},t) \mid ^{2}} \gamma (s^{\prime} ,
275: t) ds^{\prime},
276: \label{Tr1}
277: \ee
278: where:
279: \be
280: \gamma = 2A({\bf \hat{U} \cdot \hat{s}}) + 2C{\bf \hat{y} \cdot
281: \hat{s}} + 2D \kappa _{s}, \label{2}
282: \ee
283: with $s$ the arclength, $\kappa$ the curvature, and
284: \be
285: A = \frac{\mu_{2} - \mu_{1}}{\mu_{2} + \mu_{1}}, \ \ \ C=\frac{g
286: b^{2} (\rho_{2} - \rho_{1})}{12(\mu_{2} + \mu_{1})} + AV_{\infty},
287: \ \ \ D=\frac{ \sigma b^{2}}{12(\mu_{2} + \mu_{1})}, \ \ \ \gamma
288: = (\vec{u}_{1} - \vec{u}_{2})\cdot\hat{s}. \label{cons} \ee For
289: the purposes of this work it is convenient to rewrite these
290: equations in terms of the interface height $h(x)$, following
291: \cite{Goldstein98}. The equations read:
292: \be
293: \vec{U}(x,t) = \frac{1}{2\pi} P \int_{-\infty}^{+\infty} \frac{(
294: h(x^{\prime}) - h(x) , x - x^{\prime} ) }{ (x - x^{\prime})^{2} +
295: ( h(x^{\prime})-h(x) )^{2} } \tilde{\gamma}(x^{\prime}) dx^{\prime},
296: \label{tot1}
297: \ee
298: \be
299: \tilde{\gamma} = 2D\kappa_{x} + 2C h_{x} + 2A \vec{U} \cdot ( 1 ,
300: h_{x} ), \label{gm} \ee where:
301: \be
302: \tilde{\gamma}= \gamma \sqrt{ 1 + h_{x}^2 } , \qquad \kappa = \frac{
303: h_{xx}}{( 1 + h_{x}^{2} ) ^{3/2} }.
304: \label{ka}
305: \ee
306: The dependence of $h$ and $\tilde{\gamma}$ on time is not written
307: explicitly.
308:
309: To complete the definition of the moving boundary problem the
310: continuity of the normal velocity at the interface is required.
311: This means that the velocity in the $y$ direction, $dh/dt$,
312: projected along the normal direction, is equal to the normal
313: component of the average velocity of the interface, $\vec{U} \cdot
314: \hat{n}$:
315: \be
316: \frac{dh}{dt} = U_{\hat{y}} - U_{\hat{x}} h_{x}.
317: \label{tot3}
318: \ee
319: In Section \ref{Sec:3} we will consider the interface in the
320: comoving frame ($\overline{h}=0$) and will look for the equation of
321: evolution of $h(x,t)$.
322:
323:
324: \subsection{Radial geometry}
325:
326: We proceed to write the corresponding equations for a circular
327: cell rotating with angular velocity $\Omega$. The initial
328: condition is $ R=R_{0}$ (constant) and we label the outer (inner)
329: fluid as the fluid 1 (2) which have known viscosities $\mu_1$,
330: $\mu_2$ and densities $\rho_1$, $\rho_2$ (Fig.\ \ref{Figradial}).
331:
332: If we perform a change to polar coordinates in (\ref{Tr1}), and
333: from arclength $s$ to angle $\phi$, we get for the mean velocity:
334: \be
335: U_{\hat{r}}=\vec{U}(\phi_{1},t) \cdot \hat{r}=\frac{1}{2\pi} P
336: \int_{0}^{2\pi} \frac{r_2^2 \sin(\phi_2- \phi_1)}{r_1^2
337: +r_2^2-2r_1r_2 \cos(\phi_2-\phi_1)} \tilde{\gamma} (\phi_2) d\phi_2,
338: \label{imp}
339: \ee
340: \be
341: U_{\hat{\phi}}=\vec{U}(\phi_{1},t) \cdot \hat{\phi}=\frac{1}{2\pi} P
342: \int_{0}^{2\pi} \frac{r_1r_2 - r_{2}^{2}\cos(\phi_2- \phi_1)}{r_1^2
343: +r_2^2-2r_1r_2 \cos(\phi_2-\phi_1)}\tilde{\gamma} (\phi_2) d\phi_2,
344: \label{impd}
345: \ee
346: where $\tilde{\gamma} =
347: \sqrt{1+(r_{\phi}/r)^2} (\vec{u}_1-\vec{u}_2)\cdot\hat{s}$, and we have
348: used the notation $r(\phi_{1},t) \equiv r_1$,
349: $r(\phi_{2},t) \equiv r_2$.
350:
351: In the presence of sinks or sources the velocities $\vec{u}_{1}$ and
352: $\vec{u}_{2}$ which define $\vec{U}$ include only the solenoidal
353: part of the total velocity field. For this reason, in the
354: presence of injection ($Q>0$) or withdrawal ($Q<0$) of the inner
355: fluid, Eqs. (\ref{imp}) and (\ref{impd}) must be supplemented with
356: the corresponding irrotational part of the velocity field.
357:
358: In order to obtain the expression for the vorticity as a function of
359: $\vec{u}_{i}$, we use the local equations known for this problem
360: \cite{Carrillo96,Schwartz89}:
361: \be
362: \vec{\nabla}p_i=-\frac{12\mu_i}{b^2} \left(\vec{u}_i + \frac{Q}{2\pi
363: r}\hat{r}\right) + \Omega ^2 \rho_i r \hat{r}, \qquad i=1,2
364: \ee
365: and the boundary conditions:
366: \be
367: p_2 - p_1 = \sigma \kappa, \qquad \qquad u_{\hat{n}}^{(1)} =
368: u_{\hat{n}}^{(2)}.
369: \ee
370: After some algebra, we can write an expression for the vorticity as a
371: function of $\vec{U}$:
372: \be
373: \tilde{\gamma}=\frac{b^2}{12}\frac{2\sigma}{r(\mu_1 +
374: \mu_2)}\kappa_{\phi} + 2A \vec{U}\cdot (\frac{r_{\phi}}{r} , 1 ) + 2A
375: \frac{Q}{2\pi r^2}r_{\phi} - \frac{b^2}{12} \frac{ 2 \Omega ^2(\rho_2
376: - \rho_1)}{\mu_1+\mu_2}r_{\phi}, \label{gmvip} \ee with
377: \be
378: \kappa
379: =\frac{(r^2+2r_{\phi}^2-rr_{\phi\phi})}{(r^2+r_{\phi}^2)^{\frac{3}{2}}},
380: \qquad A=\frac{\mu_2-\mu_1}{\mu_1+\mu_2}.
381: \ee
382:
383: The last step is to derive the equation of the continuity of the
384: normal velocity. Now the projection of the radial velocity along
385: the normal direction $\hat{n}$ has two contributions: the
386: solenoidal part of the average velocity, $\vec{U}$,
387: projected along $\hat{n}$, and the irrotational part of the mean interface
388: velocity, $\vec{U}_{irrot}$:
389: \be
390: \frac{dr}{dt} \hat{r} \cdot \hat{n} = \vec{U} \cdot \hat{n} +
391: \vec{U}_{irrot} \cdot \hat{n} \ \ \rightarrow \ \
392: \frac{dr}{dt}=U_{\hat{r}} - \frac{r_{\phi}}{r}U_{\hat{\phi}} +
393: \frac{Q}{2 \pi r}. \label{nvir} \ee
394:
395: This completes the system of equations for the two
396: geometries considered. For an interface with no overhangs, Eqs.
397: (\ref{tot1}), (\ref{gm}), and (\ref{tot3}) are the starting point
398: for the generalized weakly nonlinear analysis in the rectangular
399: cell. Equations (\ref{imp}), (\ref{impd}), (\ref{gmvip}), and
400: (\ref{nvir}) play the same role for the analysis in
401: the rotating circular cell.
402:
403: \section{Systematic weakly nonlinear analysis. Channel geometry}
404: \label{Sec:3}
405:
406: Our goal in this section is to introduce a systematic method to
407: derive an evolution equation of the interface in real space to a
408: given order in nonlinear couplings, in the channel geometry. The
409: different orders of mode couplings will be ordered as powers of a
410: ``book--keeping'' parameter $\varepsilon$, to be defined below. The
411: evolution of the interface will thus take the form:
412: \be
413: \frac{dh}{dt} = F[h] + \varepsilon G[h] + \varepsilon ^{2} I[h] +
414: \ldots, \label{generity} \ee where $F[h], G[h], \ldots$ are
415: nonlocal operators on the function $h(x,t)$, including
416: nonlinearities of order $n+1$ in the term of order $\varepsilon
417: ^{n}$. The small parameter $\varepsilon$ is defined as the ratio
418: of two lengths, $\varepsilon=w/L$. We take $w$ as a measure of
419: the characteristic scale of variation of the interface $h(x)$,
420: while $L$ is either the width of the cell or, alternatively, the
421: characteristic scale of variation in the $x$ direction. The
422: weakly nonlinear regime is defined by the condition $w \ll L$.
423:
424: The order $\varepsilon^0$ in Eq. (\ref{generity}) corresponds to
425: the linearized equation. The order $\varepsilon^1$, when written
426: in Fourier space, corresponds to the result of Miranda and Widom
427: \cite{Mirandac98}. Here we will perform the explicit calculation
428: to one order higher in $\varepsilon$ (up to $I[h]$) which is the
429: leading nonlinear contribution in several important cases, such as
430: the one discussed in Section \ref{subsecdit}. The systematics is
431: presented in the channel geometry for simplicity.
432:
433:
434: \subsection{Dimensionless equations, expansion and convergence}
435: \label{SectionIIIA}
436:
437: We scale the interface height with $w$, the coordinate $x$ with $L$,
438: and the time with $L/C$, where the velocity $C$ has been defined in
439: (\ref{cons}).
440:
441: The equations (\ref{tot1}), (\ref{gm}), (\ref{ka}), and
442: (\ref{tot3}) become:
443: \be
444: \vec{U}(x,t) = \frac{1}{2\pi} P \int_{-\infty}^{+\infty}
445: \frac{(\varepsilon \left[h(x^{\prime}) - h(x)\right] , x -
446: x^{\prime} ) }{ (x - x^{\prime})^{2} \left\{ 1 + \varepsilon^{2}
447: \left[\frac{ h(x^{\prime})-h(x)}{x^{\prime}-x}\right] ^2 \right\}}
448: \tilde{\gamma}(x^{\prime}) dx^{\prime}, \label{gete}
449: \ee
450: \be
451: \tilde{\gamma} = 2B\kappa_{x} + 2 \varepsilon h_{x} + 2A \vec{U}
452: \cdot ( 1 , \varepsilon h_{x} ), \label{gmvera}
453: \ee
454: where
455: \be
456: B =\frac{\sigma b^{2}}{12CL^{2}( \mu_1+\mu_2)}, \ \ \ \kappa =
457: \frac{\varepsilon h_{xx}}{( 1 + \varepsilon ^{2} h_{x}^{2} ) ^{3/2}
458: },
459: \ee
460: and
461: \be
462: \frac{dh}{dt} = \frac{1}{\varepsilon} U_{\hat{y}} - U_{\hat{x}}
463: h_{x}. \label{dhdt}
464: \ee
465:
466: The starting point of our approach is an expansion of $\vec{U}$,
467: $\tilde{\gamma}$ and $\kappa$ in powers of $\varepsilon$. Notice
468: that the term
469: $\left[ h(x^{\prime})-h(x) \right] ^2 \ \left[
470: x^{\prime}-x \right] ^2$ between curly
471: brackets in (\ref{gete}) is bounded provided that $h_x$ does not
472: diverge. Equation (\ref{dhdt}) thus takes the form:
473: \be
474: \frac{dh}{dt} = \frac{1}{\varepsilon}U_{\hat{y}}^{(0)} +
475: U_{\hat{y}}^{(1)} - h_{x} U_{\hat{x}}^{(0)} + \varepsilon
476: (U_{\hat{y}}^{(2)} - h_{x} U_{\hat{x}}^{(1)}) + \cdots,
477: \label{devel}
478: \ee
479:
480: Before pursuing the calculation in detail, let us point out that
481: the $\varepsilon$--expansion in Eq.(\ref{devel}) has a finite
482: radius of convergence. This is guaranteed by the properties of
483: uniform convergence of both the expansion of the denominator in
484: Eq.(\ref{gete}) and the curvature. These properties allow us
485: to commute the expansion with the integral in Eq.(\ref{gete}) and
486: yield a convergent series of the form (\ref{devel}). The
487: radius of convergence of the expansion of the denominator of
488: Eq.(\ref{gete}) is given by the condition
489: $\varepsilon^2
490: \left[ h(x^{\prime})-h(x) \right] ^2 \ \left[
491: x^{\prime}-x \right] ^2 < 1$,
492: while the convergence of the curvature expansion is assured by the
493: condition $\varepsilon^2 h^2_x < 1$. In the nonscaled original
494: variables, the two conditions for convergence coincide, and read
495: $MAX(|h_x|)<1$. If $|h_x|<1$ in the whole domain of integration,
496: then the $\varepsilon$--expansion converges. If the condition is
497: not fulfilled in some intervals, then Eq.(\ref{devel}) is an
498: asymptotic expansion. Even in this case, the expansion contains
499: useful information about the original problem.
500:
501: An interesting case is $A=0$ which makes the vorticity independent
502: of $\vec{U}$ in Eq.(\ref{gete}), so that Eq.(\ref{dhdt}) becomes a
503: closed equation for $h(x,t)$. Then, in Eq.(\ref{devel}) it is
504: easy to show that $U_{\hat{y}}^{(n)}=0$ when $n$ is even and
505: $U_{\hat{x}}^{(n)}=0$ when $n$ is odd, a property that makes the
506: even power terms of the expansion in Eq.(\ref{devel}) to vanish.
507:
508:
509: \subsection{First and second order expansion ($\varepsilon ^0$ and
510: $\varepsilon ^1$)}
511:
512: Following the scheme introduced in the previous section we obtain
513: from Eq.(\ref{gete}) that:
514: \be
515: U_{\hat{x}}^{(0)} = 0, \qquad U_{\hat{x}}^{(1)} = \frac{1}{2\pi} P
516: \int_{-\infty}^{+\infty} \frac{ h(x^{\prime}) - h(x) }{ (x -
517: x^{\prime})^{2}} \tilde{\gamma}^{(0)}(x^{\prime}) dx^{\prime},
518: \ee
519: \be
520: U_{\hat{y}}^{(0)} = \frac{1}{2\pi} P \int_{-\infty}^{+\infty}
521: \frac{\tilde{\gamma}^{(0)}(x^{\prime})}{x - x^{\prime}}
522: dx^{\prime}, \qquad U_{\hat{y}}^{(1)} = \frac{1}{2\pi} P
523: \int_{-\infty}^{+\infty} \frac{\tilde{\gamma}^{(1)}(x^{\prime})}{x
524: - x^{\prime}} dx^{\prime}.
525: \ee
526: Since $\tilde{\gamma}^{(0)}=2B\kappa_{x}^{(0)} + 2AU_{\hat{x}}^{(0)}=0$
527: we get $U_{\hat{x}}^{(1)}=U_{\hat{y}}^{(0)}=0$. With this result we can
528: study the first order term in the vorticity equation:
529: \be
530: \tilde{\gamma} ^{(1)}(x) =2B\kappa_{x}^{(1)} + 2h_{x} +
531: 2AU_{\hat{x}}^{(1)} + 2A h_{x}U_{\hat{y}}^{(0)}.
532: \ee
533: Taking into account the definition of the Hilbert Transform:
534: \be
535: H[f(x^{\prime})]= \frac{1}{\pi} P \int_{-\infty}^{+\infty}
536: \frac{f(x^{\prime})}{x^{\prime} - x} dx^{\prime}, \ee
537: Eq.(\ref{devel}) up to order $\varepsilon ^0$ reads:
538: \be
539: \frac{dh}{dt} =
540: - H[Bh_{x^{\prime}x^{\prime}x^{\prime}} +
541: h_{x^{\prime}}]. \label{deht}
542: \ee
543: The linear operator $F[h]$ in Eq.(\ref{generity}) thus reads explicitly:
544: \be
545: F[h] = \frac{1}{\pi} P \int_{-\infty}^{+\infty}
546: \frac{\left(h + Bh_{x^{\prime}x^{\prime}}\right)_{x^{\prime}} }{x -
547: x^{\prime}} dx^{\prime} .
548: \ee
549:
550: Writing $h(x,t)$ as a superposition of Fourier
551: modes in Eq.(\ref{deht}) we recover the linear dispersion relation:
552: \be
553: \frac{\dot{\delta} _{k}(t)}{\delta _{k}(t)} = \lambda (k) = \mid k
554: \mid (1 - Bk^{2}).
555: \ee
556: We will take $k$ as an integer but we should
557: keep in mind that, upon restoring dimensions, $k$ should
558: become $(2 \pi/L)n$, with $n$ integer.
559:
560: Let us pursue the systematics of the method by computing the
561: next order in $\varepsilon$ in Eq.(\ref{dhdt}):
562: \be
563: \frac{dh}{dt} = U_{\hat{y}}^{(1)} + \varepsilon U_{\hat{y}}^{(2)}
564: = F[h] + \varepsilon G[h].
565: \ee
566: The computation of $U_{\hat{y}}^{(2)}$ requires an expression of
567: the vorticity up to second order. This includes the
568: evaluation of $U_{\hat{x}}^{(2)}$, which must be computed from the
569: first order term of the vorticity. We get:
570: \be
571: U_{\hat{x}}^{(2)}= H\left[ \left(h(x^{\prime})
572: f(x^{\prime})\right)_{x^{\prime}}\right] -
573: h(x)H[f_{x^{\prime}}(x^{\prime})], \label{unavez}
574: \ee
575: with $f(x) \equiv (Bh_{xx} +h)_{x}=\tilde{\gamma}^{(1)}(x)/2$, and:
576: \be
577: U_{\hat{y}}^{(2)} = A \left\{ H \left[h_{x^{\prime}}
578: H[f(x^{\prime\prime})]\right] + H \left[ h(x^{\prime})
579: H[f_{x^{\prime\prime}}(x^{\prime\prime})]\right] +
580: (hf)_{x}\right\}.
581: \ee
582: Taking the Fourier transform, we obtain:
583: \be
584: \dot{\delta}_k = \lambda (k) \delta_k + \varepsilon A \mid k
585: \mid \sum_{s=-\infty}^{+\infty}\left[1 - \mbox{sgn}(ks) \right]
586: \lambda (s) \delta_s (t) \delta_{k-s} (t),
587: \label{secmod}
588: \ee
589: which coincides with the result of Miranda and Widom in Ref.
590: \cite{Mirandac98}.
591:
592:
593: \subsection{Third order expansion ($\varepsilon ^2$)}
594:
595: The expansion to order $\varepsilon^2$ is necessary to account for
596: the lowest order nonlinearities in the case of zero viscosity
597: contrast, and for other relevant situations such as the time
598: dependent Saffman--Taylor finger solutions (section
599: \ref{subsecdit}). We now have:
600: \be
601: \frac{dh}{dt} = {\cal{O}}(\varepsilon^0) +
602: {\cal{O}}(\varepsilon^1) + \varepsilon ^2 (U_{\hat{y}}^{(3)} -
603: h_xU_{\hat{x}}^{(2)} ) + \cdots \ee We have already computed
604: $U_{\hat{x}}^{(2)}$ in Eq. (\ref{unavez}). On the other hand:
605: \be
606: U_{\hat{y}}^{(3)}= - \frac{1}{2} H[
607: \tilde{\gamma}^{(3)}(x^{\prime})] + \frac{1}{2 \pi} P
608: \int_{-\infty}^{+\infty} \frac{ \left[h(x^{\prime}) -
609: h(x)\right]^2 }{ ( x^{\prime}-x)^{3}} \tilde{\gamma}
610: ^{(1)}(x^{\prime}) dx^{\prime}. \label{uyt} \ee Integrating twice
611: by parts, the last integral can also be written as a Hilbert
612: Transform. After some algebra we obtain the explicit form of the
613: operator $I[h]$ containing the cubic nonlinearities in
614: Eq.(\ref{generity}), which reads:
615: \be
616: \begin{array}{ll}
617: I[h] = & \frac{3}{2}H[ g(x^{\prime})] +
618: \frac{1}{2}H \left[ \left( f(x^{\prime}) [ h(x)-h(x^{\prime})]^2
619: \right) _{x^{\prime}x^{\prime}}\right] \\
620: & -h_x H\left[\left( f(x^{\prime}) [h(x^{\prime}) - h(x)]
621: \right) _{x^{\prime}}\right] + V[h,A]
622: \end{array}
623: \ee
624: with:
625: \be
626: V[h,A]=A^2 H \left[h(x^{\prime})H
627: [\tau_{x^{\prime\prime}}(x^{\prime\prime})] +
628: h_{x^{\prime}}(x^{\prime})H[\tau(x^{\prime\prime})]\right] + A^2
629: (h\tau)_{x},
630: \ee
631: and:
632: \be
633: g(x)=B\left(h_{xx}h_x^2\right)_x, \qquad \tau(x)=U_{\hat{x}}^{(2)} +
634: h_xU_{\hat{y}}^{(1)}=\frac {\tilde {\gamma}^{(2)}(x)} {2A}.
635: \ee
636:
637: The same result in Fourier space takes the form:
638: \be
639: \dot{\delta} _{k}(t)= {\cal{O}}(1,\varepsilon) + \varepsilon^2
640: \sum_{s,l=-\infty}^{+\infty} \delta_{l} \delta_{s-l}\delta_{k-s}
641: \left[A^2 T(k,s,l) - \frac{3}{2}B Y(k,s,l) + W(k,s,l)\right],
642: \label{triorder}
643: \ee
644: with:
645: \be
646: T(k,s,l)=\mid k \mid \mid s \mid \lambda (l) \left[1 -
647: \mbox{sgn}(ks)\right] \left[1 - \mbox{sgn}(ls)\right],
648: \ee
649: \be
650: Y(k,s,l)=\mid k \mid l^2(s-l)(k-s),
651: \ee
652: \be
653: W(k,s,l)=\left[l\left(\frac{l}{2}+k-s\right) -s k \mbox{ sgn}(ls) +
654: \frac{k^2}{2} \mbox{ sgn}(kl)\right] \lambda(l).
655: \ee
656:
657: It is clear that the viscosity contrast in Eq.(\ref{triorder})
658: appears squared because of the reflection symmetry (the
659: simultaneous change $A \rightarrow -A$ and $h \rightarrow -h$ is a
660: dynamical symmetry of the problem). Symmetry reasons alone,
661: however, do not allow to discard a three mode coupling
662: contribution when $A=0$. We see from our calculation that
663: a three mode coupling is indeed present independently of $A$.
664:
665: Following this scheme, the fourth order will carry a contribution
666: proportional to $A$, and another proportional to $A^{3}$, for
667: symmetry reasons. The fifth order will carry a contribution
668: independent of $A$ and two others, proportional to $A^2$ and
669: $A^4$, respectively. This scheme will continue for subsequent
670: orders.
671:
672:
673: \subsection{Analysis of the time dependent single finger solution}
674: \label{subsecdit}
675:
676: In this section we perform a detailed analysis of the weakly nonlinear
677: expansion in cases where exact solutions are known, namely single finger
678: configurations with $B=0$. This allows to study how exact properties of
679: solutions show up at the different orders, particularly concerning the
680: role of viscosity contrast $A$.
681:
682: At this point it is worth recalling that the case $A=1$ allows for
683: a continuum of Saffman--Taylor finger solutions corresponding to
684: different finger widths. A continuum of time dependent exact
685: solutions, leading to those stationary states, is also known for
686: $A=1$. However, for $A \neq 1$ only the one of width
687: $\lambda=\frac{1}{2}$ remains a solution, $\lambda$ being the
688: ratio of the finger width to the width of the channel. This
689: result, which has been recently addressed in Ref.\cite{Folch00}
690: although it was first discovered in Ref.\cite{Francessos1}, points
691: out an intriguing connection between the width selection problem
692: and the dynamical role of viscosity contrast. Here we will
693: analyze the interplay between $A$ and $\lambda$ in the early
694: nonlinear regime, and elucidate at what stage of the nonlinear
695: dynamics does the viscosity contrast $A \neq 1$ prevents the
696: possibility of having $\lambda\neq 1/2$.
697:
698: From now on we consider $L=2\pi$, $B=0$, $C=1$, and we take as
699: scaling velocity $(1 - \lambda) C = (1 -
700: \lambda)\equiv\frac{\eta}{2}$. Conformal mapping techniques
701: enable us to write the single finger solution of this problem in
702: the form \cite{Francessos1,Folch00}:
703: \be
704: f(w,t)=- \ln w + d(t) + \eta \ln (1 - \alpha(t) w), \label{mappi}
705: \ee where $f(w,t) = y + ix$, is an analytic function inside the
706: unit disk in the $w$--complex plane, which maps that disk into the
707: physical region occupied by the more viscous fluid. The interface
708: is obtained in a parametric form by setting $w=e^{i \theta }$.
709: The functions $\alpha (t)$, $d(t)$ verify:
710: \be
711: \dot{d}(t)=\frac{\eta}{2-\eta}\left(\frac{\eta}{2} -
712: \frac{\dot{\alpha}(t)}{\alpha (t)}\right), \label{setun}
713: \ee
714: \be
715: \frac{\dot{\alpha}(t)}{\alpha (t)} = \frac{\eta} { 2 +
716: \eta(\eta-2) + \eta(2-\eta)\frac{1+\alpha^2(t)}{1-\alpha^2(t)} }.
717: \label{setdos}
718: \ee
719: To obtain an expression for $h(x)$ in the weakly nonlinear regime of
720: the evolution, $\alpha(t)$ and $d(t)$ are expanded in powers of a
721: small parameter $\nu$:
722: \be
723: d(t) = d^{(0)}(t) + \nu d^{(1)}(t) + \nu ^2 d^{(2)}(t)+ \cdots \ \
724: , \qquad \alpha (t) = \nu \alpha^{(0)} (t) + \nu ^2 \alpha ^{(1)}
725: (t).
726: \ee
727: Introducing these expansions in Eqs.(\ref{setun}) and
728: (\ref{setdos}) we obtain:
729: \be
730: \dot{d}^{(0)}=\dot{d}^{(1)}=0 , \qquad \dot{d}^{(2)}=
731: \frac{\eta^3}{2} (\alpha^{(0)})^2(t), \qquad \dot{d}^{(3)}= \eta
732: ^3 \alpha^{(0)}\alpha^{(1)}, \ \ \cdots
733: \ee
734: \be
735: \dot{\alpha}^{(0)}=\frac{\eta}{2}\alpha^{(0)}, \qquad
736: \dot{\alpha}^{(1)}=\frac{\eta}{2}\alpha^{(1)}, \qquad
737: \dot{\alpha}^{(2)}=\frac{\eta}{2}\left(\alpha^{(2)} -\eta (2 -
738: \eta)(\alpha^{(0)})^3\right), \ \ \cdots \label{alfas} \ee which
739: will be useful later. From (\ref{mappi}) and (\ref{setun}) we
740: obtain:
741: \be
742: \begin{array}{ll}
743: y=h= & -\nu \eta \alpha^{(0)} \cos \theta -
744: \nu^2 \left( \eta \alpha^{(1)} \cos \theta +
745: \frac{\eta}{2}(\alpha^{(0)})^2 \cos 2\theta - d^{(2)} \right) \\ & -
746: \nu ^3 \left( \eta \alpha^{(2)}\cos\theta + \eta
747: \alpha^{(0)}\alpha^{(1)}\cos 2\theta + \frac{\eta}{3}
748: (\alpha^{(0)})^3 \cos 3\theta -d^{(3)} \right)+ {\rm \cal{O}}(\nu^4)
749: , \label{hexp}
750: \end{array}
751: \ee
752: \be
753: x= -\theta -\nu \eta \alpha^{(0)} \sin \theta - \nu^2 \left( \eta
754: \alpha^{(1)} \sin \theta + \frac{\eta}{2}(\alpha^{(0)})^2 \sin 2
755: \theta \right) + {\rm \cal{O}}(\nu^3).
756: \ee
757: This last equation can be inverted in a systematic way to get:
758: \be
759: \theta = -x + \nu \eta \alpha^{(0)} \sin x + \nu^2 \left( \eta
760: \alpha^{(1)} \sin x + \frac{\eta}{2} (1-\eta)(\alpha^{(0)})^2 \sin
761: 2x \right) + {\rm \cal{O}}(\nu^3).
762: \ee
763: The relation between $\theta$ and $x$ can be inverted in Eq.(\ref{hexp})
764: and the cosine functions expanded. We obtain in this way the following
765: expression for $h(x,t)$:
766: \be
767: \begin{array}{ll}
768: h(x,t)= & - \nu \eta \alpha^{(0)} \cos x - \nu^2 \eta\left( \alpha^{(1)}
769: \cos x + \frac{1-\eta}{2} (\alpha^{(0)})^2 \cos 2x\right) + \\
770: & \nu^3 \eta \left\{ (\eta-1) \alpha^{(0)} \alpha^{(1)} \cos 2x +
771: \left(\frac{3\eta}{8}(\eta-2)(\alpha^{(0)})^3 -
772: \alpha^{(2)}\right) \cos x + \right. \\
773: & \left. \left( \frac{3\eta}{8}(2-\eta) -
774: \frac{1}{3} \right) (\alpha^{(0)})^3 \cos 3x\right\} + {\rm
775: \cal{O}}(\nu^4) . \label{hexpan}
776: \end{array}
777: \ee
778:
779: In order to follow the scheme developed in the previous section,
780: we must measure $h(x,t)$ in units of its characteristic amplitude
781: $\nu$. In this way, the previous equation (\ref{hexpan}) takes
782: the form:
783: \be
784: h=h^{(0)} + \nu h^{(1)}+ \nu^2 h^{(2)} + {\rm \cal{O}}(\nu^3),
785: \label{simexp}
786: \ee
787: where $h(x,t)$ represents now $h(x,t)/ \nu$,
788: and the small parameter $\nu$ is directly comparable to the small
789: parameter $\varepsilon$ of the previous section. The expression
790: for $h(x,t)$ can be regarded as a series of modes of decreasing
791: amplitude in our expansion (\ref{generity}) (written in the units
792: of this problem), and matching the corresponding powers in either
793: $\varepsilon$ or $\nu$ we obtain:
794: \be
795: \frac{dh^{(0)}}{dt}=\frac{\eta}{2}F[h^{(0)}]
796: \ee
797: to order $\varepsilon ^{0}$.
798:
799: This identity can be verified (independently of $A$) by
800: substituting the expression of $h^{(0)}$ and using
801: Eqs.(\ref{alfas}) and (\ref{deht}) for the left and right hand
802: side respectively. At order $\varepsilon$ we have:
803: \be
804: \frac{dh^{(1)}}{dt}=\frac{\eta}{2}\left(F[h^{(1)}] +
805: G[h^{(0)},h^{(0)}]\right) .
806: \ee
807: The second term of the right hand side gives no contribution, and we
808: obtain an interesting result:
809: the equation at order $\varepsilon$ is verified independently of
810: $\eta$ and $A$. Hence, we cannot establish the difference between
811: $A=1$ (compatible with any $\eta$) and $A \neq 1$ (compatible only
812: with $\eta=1$) at this order.
813:
814: The difference between these two situations arises at order
815: $\varepsilon^2$:
816: \be
817: \frac{dh^{(2)}}{dt}= \frac{\eta}{2} \left(F[h^{(2)}] +
818: G[h^{(0)},h^{(1)}] + L[h^{(0)},h^{(0)},h^{(0)}]\right).
819: \ee
820: Since the left hand side of the equation involves only the modes $\cos
821: x$, $\cos 2x$, and $\cos 3x$, the right hand side of the equation
822: must include these same modes with the same coefficients. The
823: coefficients for $\cos 2x$ and $\cos 3x$ are easily matched. For
824: $\cos x$ the left hand side reads:
825: \be
826: \left(\frac{\eta^3(\eta-2)}{16}(\alpha^{(0)})^3 -
827: \frac{\eta^2}{2}\alpha^{(2)}\right) \cos x,
828: \ee
829: where we have used Eq. (\ref{alfas}), and the right hand side reads:
830: \be
831: \left(\frac{3\eta^3(\eta-2)}{16}(\alpha^{(0)})^3
832: -\frac{\eta^2}{2}\alpha^{(2)}\right) \cos x +
833: A\frac{\eta^3(1-\eta)}{4}(\alpha^{(0)})^3 \cos x +
834: \frac{\eta^4}{8}(\alpha^{(0)})^3 \cos x.
835: \ee
836: Hence, the requirement for matching the coefficients on both sides is:
837: \be
838: \frac{\eta^3}{4}(1-\eta)=A \frac{\eta^3}{4}(1-\eta),
839: \ee
840: which is always true if $\eta=1$ ($\lambda=1/2$), but for $\eta \neq 1$
841: ($\lambda \neq 1/2$) it requires $A=1$. This result shows that
842: the nontrivial relationship between $A$ and $\lambda$ which is
843: known from exact solutions is already manifest at
844: the early nonlinear stages of the dynamics. This clearly illustrates
845: the potential usefulness of the weakly nonlinear expansion at a purely
846: analytical level, in that a dynamical property
847: of the problem which must be satisfied at all orders (in this case the
848: incompatibility of $A \neq 1$ and $\lambda \neq 1/2$) may be
849: detected at a low order in the expansion, without having to know the
850: exact solution to all orders.
851:
852: If we pursue the expansion to higher orders,
853: the general expression for $h^{(n)}$ takes the form:
854: \be
855: h^{(n)}=\sum_{k=1}^{n+1} \beta_{k}^{(n)} (t) \cos(kx),
856: \label{eshden} \ee where each coefficient is a function of $\alpha
857: (t)$. The even modes of the solution with $\eta=1$ have zero
858: coefficients because they must remain invariant under a change of
859: sign and translation by $\pi$.
860:
861: So far in this section we have restricted ourselves to
862: $B=0$ in order to compare with exact results. However, we can carry out
863: the analysis also for $B \neq 0$. It can be shown that the structure of
864: the expression (\ref{eshden}) is preserved in this case. The
865: coefficients $\beta_{k}^{(n)} (t)$ are obtained as solutions of a set
866: of differential equations which contain the surface tension $B$. For
867: instance, to third order we have:
868: \be
869: h(x,t) = \left(\beta_1^{(0)} + \varepsilon \beta_1^{(1)} +
870: \varepsilon^2 \beta_1^{(2)}\right) \cos x + \left( \varepsilon
871: \beta_2^{(1)} + \varepsilon^2 \beta_2^{(2)}\right) \cos 2x +
872: \varepsilon^2 \beta_3^{(2)} \cos 3x,
873: \label{generalo}
874: \ee
875: which, introduced in Eq.(\ref{generity}), leads to the following
876: equations:
877: \be
878: \begin{array}{c}
879: \dot{\beta}_1^{(0)}= V_0(1-B)\beta_1^{(0)}, \qquad
880: \dot{\beta}_1^{(1)}= V_0(1-B)\beta_1^{(1)}, \qquad
881: \dot{\beta}_2^{(1)}= 2V_0(1-4B)\beta_2^{(1)}, \\ \displaystyle{
882: \dot{\beta}_2^{(2)}= 2V_0(1-4B)\beta_2^{(2)}, \qquad
883: \dot{\beta}_3^{(2)}= 3V_0(1-9B) \beta_3^{(2)} -
884: \frac{9BV_0}{8}(\beta_1^{(0)})^3, } \\
885: \displaystyle{\dot{\beta}_1^{(2)}= V_0(1-B)\beta_1^{(2)} + A
886: V_0(1-B)\beta_2^{(1)}\beta_1^{(0)} - \frac{V_0}{8}(2-5B)
887: (\beta_1^{(0)})^3,} \label{eqbet}
888: \end{array}
889: \ee once the dimensions are reintroduced. In this way we can also
890: see how surface tension perturbs the dynamics at the weakly
891: nonlinear regime. Clearly, at these early stages of the nonlinear
892: evolution there is no sign of the singular perturbation character
893: of surface tension, which, at the late stages of the evolution
894: will be responsible for the selection of the steady
895: state\cite{Bensimon86}.
896:
897:
898: \section{Application of the method to the rotating Hele--Shaw cell}
899: \label{Sec:4}
900:
901: \subsection{Dimensionless equations in the radial geometry}
902: \label{dimeq-in-rad}
903:
904: The $\varepsilon$ expansion in the channel geometry originated in
905: the different scaling of lengths in the $x$ and $y$ directions.
906: To perform an equivalent expansion in the radial geometry we first
907: split the function $r(\phi,t)$ in two parts, namely the radius of
908: the unperturbed circle and the departure from this circle:
909: \be
910: r(\phi,t) \qquad \rightarrow \qquad \sqrt{R_0^2 + \frac{Qt}{\pi}}
911: + r(\phi,t) = R(t) + r(\phi,t).
912: \ee
913: The largest length scale in the problem is given by $R(t)$, which is
914: only a constant in the
915: absence of injection. For finite injection rate, $R(t)$ defines an
916: evolving (unstable) solution. The proper scaling of
917: $r(\phi, t)$, which will define the weakly
918: nonlinear dynamics now, is not the naive extension of the scaling
919: defined in the channel geometry (namely $\varepsilon = w/R$, $w$ being
920: the typical departure from circularity). In fact, this scaling would
921: mix different orders of the weakly nonlinear analysis. The reason
922: for this was already pointed out by Miranda and
923: Widom \cite{Mirandar98}: in the radial geometry, mass
924: conservation implies that the zero mode (which in the channel geometry
925: is decoupled from the rest and drops out of the formulation)
926: has a higher order nonzero amplitude, since it must
927: satisfy a mass conservation relation which, to lowest order, reads:
928: \be
929: \delta_0 = - \frac{1}{2R} \sum_{k \neq 0} \mid \delta_k \mid ^2.
930: \label{deze}
931: \ee
932:
933: Following this observation, we split the deviation from $R$
934: in two terms, so that the radial function reads $r = R + \tilde{r} +
935: r_{0}$, where $r_0$ represents the zero mode (an expansion or
936: contraction of the circle), and $\tilde{r}$ the other modes. We
937: scale them in the form $\tilde{r} \rightarrow w\tilde{r}$,
938: $r_{0} \rightarrow (w^2/R)r_{0}$, and write the position of the
939: interface $r(\phi)$ as:
940: \be
941: r(\phi) \rightarrow R( 1 + \varepsilon r(\phi) ),
942: \label{esca}
943: \ee
944: where:
945: \be
946: r(\phi)=\tilde{r}(\phi) +\varepsilon r_{0}.
947: \label{esco}
948: \ee
949: To simplify the notation, we have dropped out the time dependence.
950:
951: The velocities will be scaled by:
952: \be
953: V_0 = \frac{1}{\mu_2 + \mu_1} \left(\frac{ Q (\mu_2 -\mu_1) }{2
954: \pi R} - \frac{b^2}{12} \Omega^2 (\rho_2 - \rho_1) R \right),
955: \ee
956: implying that time will be scaled by $R/V_0$.
957: Once equations (\ref{imp}), (\ref{impd}), (\ref{gmvip}), and
958: (\ref{nvir}) are made dimensionless, we have:
959: \be
960: \tilde{\gamma} = 2B\frac{\kappa_{\phi}}{1+\varepsilon r} + 2A\vec{U}
961: \cdot (\frac{\varepsilon r_{\phi}}{1+\varepsilon r} , 1 ) +
962: 2\varepsilon \left(\frac{C}{(1+\varepsilon r)^{2}} - D \right)
963: r_{\phi},
964: \label{qp3}
965: \ee
966: for the vorticity, where:
967: \be
968: \begin{array}{c}
969: \displaystyle{ B=\frac{b^2}{12} \frac{\sigma}{(\mu_1 + \mu_2) V_0 R^2}, \qquad
970: C=\frac{QA}{2 \pi R V_0}, \qquad D=\frac{b^2}{12} \frac{ \Omega
971: ^2 \triangle \rho R}{(\mu_1 + \mu_2) V_0},} \vspace{0.5cm} \\
972: \displaystyle{ \kappa=\frac{(1+\varepsilon r)^2 + 2 \varepsilon ^2
973: r_{\phi}^2 -\varepsilon (1+\varepsilon r)r_{\phi \phi}}{\left[
974: (1+\varepsilon r)^2 + \varepsilon ^2 r_{\phi}^2\right]
975: ^{\frac{3}{2}}} , \qquad
976: A=\frac{\mu_{2}-\mu_{1}}{\mu_{2}+\mu_{1}}, \qquad \triangle \rho =
977: \rho_2- \rho_1.} \label{kappa}
978: \end{array}
979: \ee
980: For the integrals:
981: \be
982: U_{\hat{r}}(\phi) = \frac{1}{4 \pi} P \int_{0}^{2\pi} \frac{1 +
983: 2\varepsilon r_2 + \varepsilon^2r_2^2}{\tan \left(\frac{\phi_2 -
984: \phi_1}{2}\right) \left[1 + \varepsilon f(\phi_1,\phi_2) +
985: \varepsilon^2 g(\phi_1,\phi_2)\right]}\tilde{\gamma} (\phi_2 , t)
986: d\phi_2,\label{qp1}
987: \ee
988: \be
989: U_{\hat{\phi}}(\phi) = \frac{1}{4 \pi} P \int_{0}^{2\pi} \frac{1 +
990: \varepsilon f(\phi_1,\phi_2) + \varepsilon ^2 r_1 r_2 - (1+
991: 2\varepsilon r_2 + \varepsilon ^2r_2 ^2) \cos (\phi_2 - \phi_1) }{2
992: \sin ^2 \left(\frac{\phi_2 -\phi_1}{2}\right) \left[1 + \varepsilon
993: f(\phi_1,\phi_2) + \varepsilon^2
994: g(\phi_1,\phi_2)\right]}\tilde{\gamma} (\phi_2) d\phi_2, \label{qp2}
995: \ee
996: with:
997: \be
998: f(\phi_1,\phi_2) = r_1 + r_2, \qquad g(\phi_1,\phi_2)= \frac{(r_1 -
999: r_2)^2}{4 \sin^2 \left(\frac{ \phi_2 -\phi_1 }{2}\right)} + r_1r_2,
1000: \ee
1001: and for the evolution of the deviation from $R(t)$:
1002: \be
1003: \varepsilon \frac {d\tilde{r}}{dt} + \varepsilon^2 \frac{dr_{0}}{dt}=
1004: U_{\hat{r}} - \varepsilon \frac{r_{\phi} U_{\hat{\phi}}}{1 +
1005: \varepsilon r} + \frac{Q}{2 \pi R V_0} \left[ \frac{1}{1 +
1006: \varepsilon r} - ( 1 - \varepsilon^2 r_{0})\right]. \label{popoito}
1007: \ee
1008:
1009: Eq.(\ref{popoito}) above is the only one which is sensitive to the
1010: different possible scalings of the deviation from circularity,
1011: due to the time derivatives. The choice
1012: we propose is the only one which is truly systematic. Other
1013: possibilities are not incorrect but will mix different orders of the
1014: weakly nonlinear expansion at a given order in $\varepsilon$.
1015:
1016:
1017: \subsection{The linear dispersion relation}
1018:
1019: Our goal now is to obtain the linear dispersion relation and the
1020: leading weakly nonlinear corrections, in both real and Fourier
1021: space, and to discuss the interplay between rotation and injection
1022: at the early stages of the instability.
1023:
1024: We will use the following definitions for the average of a function
1025: and for the Hilbert transform in the unit circle:
1026: \be
1027: \overline{f} = \frac{1}{2\pi} \int_{0}^{2\pi} f(\phi) d\phi, \qquad
1028: H[f(\phi^{\prime})] = \frac{1}{2\pi} P \int_{0}^{2\pi}
1029: f(\phi^{\prime}) \cot \left(\frac{\phi^{\prime} - \phi}{2}\right)
1030: d\phi^{\prime}.
1031: \ee
1032: The zero order of the vorticity and the two
1033: components of the velocity are:
1034: \be
1035: \tilde{\gamma}^{(0)} = 2AU_{\hat{\phi}}^{(0)}, \qquad
1036: U_{\hat{\phi}}^{(0)}=\frac{\overline{\tilde{\gamma}^{(0)}}}{2},
1037: \qquad U_{\hat{r}}^{(0)}= \frac{1}{2} H
1038: [\tilde{\gamma}^{(0)}(\phi^{\prime})],
1039: \ee
1040: and thus:
1041: \be
1042: \tilde{\gamma}^{(0)}= A \overline{\tilde{\gamma}^{0}}.
1043: \ee
1044: This is different from its counterpart in the channel geometry. Here
1045: the equations for the vorticity at consecutive orders require the
1046: knowledge of the average vorticity. The
1047: equations are solved by averaging on both sides. For example, at zero order,
1048: $A\neq1$ will lead to $\tilde{\gamma}^{(0)}=0$, but for $A=1$,
1049: $\tilde{\gamma}^{(0)}$ can be any constant. The presence of an
1050: arbitrary constant is not a problem, since it is eliminated at
1051: each order in the equation for the evolution of the interface.
1052:
1053: Expanding the equations up to first order, and using
1054: $\tilde{\gamma}^{(0)}=0$, the linearized equation for the evolution
1055: of the interface deviation reads:
1056: \be
1057: \frac{d\tilde{r}}{dt}=U_{\hat{r}}^{(1)} - \frac{Q}{2\pi R
1058: V_0}\tilde{r}= \frac{1}{2}H[\tilde{\gamma}^{(1)}] - \frac{Q}{2\pi R
1059: V_0}\tilde{r},
1060: \ee
1061: with:
1062: \be
1063: \tilde{\gamma}^{(1)}= -2B(\tilde{r}+\tilde{r}_{\phi\phi})_{\phi} +
1064: 2(C-D)\tilde{r}_{\phi}.
1065: \ee
1066: We can now perform a Fourier transform of the equations and, after
1067: reintroducing the adequate dimensions, we find:
1068: \be
1069: \lambda (n)= \left(\frac{b^2}{12}\frac{\Omega^2 \triangle \rho}{\mu_1
1070: + \mu_2} - \frac{Q A}{2 \pi R^2}\right)\mid n \mid - \frac{b^2}{12
1071: R^3}\frac{\sigma}{\mu_1 + \mu_2}\mid n \mid (n^2-1) - \frac{Q}{2\pi
1072: R^2} \label{lrr}.
1073: \ee
1074: This is the linear dispersion relation found in
1075: \cite{Carrillo96}.
1076:
1077: We now return to the question of the different
1078: scaling alternatives introduced at the end of Section
1079: \ref{dimeq-in-rad}. Had we used the scaling $\varepsilon=w/R^2(t)$, the
1080: term $Q/(2 \pi R^2)$ in (\ref{lrr}) would have disappeared before
1081: reintroducing the dimensions. On the contrary, for $\varepsilon=w$ this
1082: term becomes multiplied by a factor $2$. These changes in the
1083: dimensionless linear dispersion relation come from the fact that a mode
1084: $\delta_{n}$ in the scaling adopted in (\ref{esca}) and (\ref{esco})
1085: becomes $\delta_{n}/R$ in the first of the scaling alternatives above.
1086:
1087: The term $Q/(2 \pi R^2)$ at the end of Eq. (\ref{lrr}) is
1088: stabilizing for positive injection rate ($Q>0$) and destabilizing
1089: for $Q<0$. In a sense this is a purely geometric effect, which
1090: contributes to the growth or decay of modes due to the expansion
1091: or contraction of the base state. As pointed out in
1092: \cite{Rocco00} the term $Q/(2 \pi R^2)$, which can be scaled out,
1093: must be distinguished from the rest, which do describe the
1094: intrinsic instability of the problem. Disregarding that last
1095: term, it is then clear that injection and rotation ($CV_{0}$ and
1096: $DV_{0}$ respectively) play an equivalent role in the linear
1097: dispersion relation, since both appear multiplying $n$. By
1098: choosing the proper viscosity and density contrasts, they can
1099: produce exactly the same stabilizing or destabilizing effect.
1100: This is the counterpart, in the radial geometry, of the
1101: equivalence between the roles of injection and gravity in the
1102: channel geometry. In the channel geometry, however, the
1103: equivalence is exact and can thus be extended to all the nonlinear
1104: evolution (in the appropriate reference frame). This is not the
1105: case in radial geometry, as shown in the next section.
1106:
1107:
1108: \subsection{Weakly nonlinear theory with rotation}
1109:
1110: The purpose of this section is to derive the leading order
1111: nonlinear contributions for the general problem with both rotation
1112: and injection. We first recall that mass conservation related the
1113: zero mode to the other modes. In the original nonscaled variables
1114: this reads:
1115: \be
1116: \dot{\delta}_0 = - \frac{Q}{2 \pi R^2}\delta_0 - \frac{1}{R}\sum_{k
1117: \neq 0} \mid \delta_k \mid ^2 \lambda(k). \label{delzero}
1118: \ee
1119: To reproduce this result and the nonlinear couplings for the
1120: other modes we start with the equation for the interface at order
1121: $\varepsilon^2$:
1122: \be
1123: \frac{d\tilde{r}}{dt} + \varepsilon \frac{dr_{o}}{dt}= {\rm \cal{
1124: O}}(\varepsilon^{0}) + \varepsilon \left(U_{\hat{r}}^{(2)} +
1125: \frac{Q}{2\pi R V_0}\tilde{r}^2\right),
1126: \label{newman}
1127: \ee
1128: where we have used that $U_{\hat{\phi}}^{(1)}= U_{\hat{\phi}}^{(0)} =
1129: U_{\hat{r}}^{(0)} = \tilde{\gamma}^{(0)} =
1130: \overline{\tilde{\gamma}^{(0)}} = 0$. The velocity
1131: $U_{\hat{r}}^{(2)}$ can be obtained from the expansion of
1132: $U_{\hat{r}}$:
1133: \be
1134: U_{\hat{r}}^{(2)} = \frac{1}{2}H[
1135: \tilde{\gamma}^{(2)}(\phi^{\prime})] +\frac{1}{2} H[(\tilde{r}(\phi
1136: ^{\prime}) - \tilde{r}(\phi))\tilde{\gamma}^{(1)} (\phi^{\prime})],
1137: \label{vrdos}
1138: \ee
1139: where:
1140: \be
1141: \tilde{\gamma}^{(2)} = 2B\kappa ^{(2)} - 2B\tilde{r}\kappa^{(1)} +
1142: 2A \tilde{r}_{\phi}U_{\hat{r}}^{(1)} + 2 AU_{\hat{\phi}}^{(2)}
1143: -4C\tilde{r}\tilde{r}_{\phi}.
1144: \ee
1145: $\kappa ^{(2)}$ and $U_{\hat{\phi}}^{(2)}$ can be computed from
1146: (\ref{kappa}) and (\ref{qp2}) respectively. The latter involves only
1147: $\tilde{\gamma}^{(1)}$. We obtain finally:
1148: \be
1149: \begin{array}{cc}\displaystyle{ \frac{d\tilde{r}}{dt} + \varepsilon
1150: \frac{dr_{o}}{dt} = {\rm \cal{O}} (\varepsilon^0) + \varepsilon } &
1151: \left( H \left[ B \kappa_{\phi^{\prime}}^{(2)} -
1152: \frac{C+D}{2} (\tilde{r}^2)_{\phi^{\prime}} \right] +
1153: \tilde{r}H[s_{\phi^{\prime}}] + A H[H
1154: [(\tilde{r}s_{\phi^{\prime\prime}})_{ \phi^{\prime\prime}}]] \right.
1155: \\ & - \left. A H \left[\tilde{r}(\phi^{\prime}) H[s_{\phi^{\prime\prime}
1156: \phi^{\prime\prime}}] +
1157: \tilde{r}_{\phi^{\prime}}H[s_{\phi^{\prime\prime}}] \right]
1158: +\frac{Q}{2 \pi R V_0} \tilde{r}^2 \right), \label{urdos}
1159: \end{array}
1160: \ee
1161: with:
1162: \be
1163: \kappa^{(2)}= \tilde{r}^2 + \frac{\tilde{r}_{\phi}^2}{2}
1164: +2\tilde{r}\tilde{r}_{\phi\phi}, \qquad s_{\phi}=\left( B(\tilde{r}
1165: +\tilde{r}_{\phi\phi}) + (D-C)\tilde{r} \right)_{\phi}.
1166: \ee
1167:
1168: By introducing a superposition of Fourier modes in $\tilde{r}$ we
1169: directly recover Eq.(\ref{delzero}). For the other modes ($k \neq
1170: 0$) we get the result:
1171: \be
1172: \label{wnlrot}
1173: \dot{\delta}_n = \lambda(n)\delta_n + \sum_{k \neq 0,n} \delta_{k}
1174: \delta_{n-k}(F(k,n) + S(k,n) + \lambda(k) J(k,n)),
1175: \ee
1176: where:
1177: \be
1178: F(k,n) = \frac{\mid n \mid}{R} \left[ - \frac{Q A}{2 \pi R^2}
1179: \left(\frac{1}{2} - \mbox{sgn}(kn)\right) +
1180: \frac{b^2}{24}\frac{\Omega^2 \triangle \rho}{\mu_1 + \mu_2}
1181: \right],
1182: \ee
1183: \be
1184: S(k,n)= \frac{\mid n \mid}{R} \left[ \frac{b^2}{12
1185: R^3}\frac{\sigma}{\mu_1 + \mu_2}\left(1 - \frac{k}{2}(n +
1186: 3k)\right) \right],
1187: \ee
1188: \be
1189: J(k,n) =-\frac{1}{R}\left[A\mid n \mid(1 - \mbox{
1190: sgn}(kn))+1\right]. \ee For the particular case $\Omega = 0$, the
1191: expression above reproduces the result of Miranda and Widom
1192: \cite{Mirandar98}. We find that the presence of rotation does not
1193: change the formal structure of the equations, but introduces new
1194: terms.
1195:
1196: To emphasize the roles of injection and rotation, we define the
1197: quantity $H(k,n)$ as the part of the coupling matrix in the r.h.s.
1198: of Eq.(\ref{wnlrot}) which contains $Q$ and/or $\Omega$. This
1199: quantity reads:
1200: \be
1201: H(k,n)=\frac{\mid n \mid}{2R} \left( \frac{Q A}{2 \pi R^2}+
1202: \frac{b^2}{12}\frac{\Omega^2 \triangle \rho}{\mu_1 + \mu_2}\right) +
1203: J(k,n)\mid k \mid \left(\frac{b^2}{12} \frac{\Omega^2\triangle
1204: \rho}{\mu_1 + \mu_2} - \frac{QA}{2 \pi R^2}\right) + \frac{Q}{2 \pi
1205: R^3}.
1206: \ee
1207: We observe that experimental parameters occur only in two groups,
1208: namely:
1209: \be
1210: \tilde{\Omega}=\frac{b^2}{12} \frac{\Omega^2\triangle \rho}{\mu_1
1211: + \mu_2}, \qquad \tilde{Q}=\frac{QA}{2 \pi R^2}. \ee Both of them
1212: turn out to multiply the same functions of the wave numbers,
1213: except for changes in sign. However, notice that the relative
1214: sign of $\tilde{\Omega}$ and $\tilde{Q}$ is different in the
1215: linear part and in the leading nonlinear part of the dynamics.
1216: This clearly shows that the effects of injection and rotation
1217: cannot be interchanged by simple changes in parameters, and that
1218: the intrinsic new dimensionless parameter related to rotation
1219: already shows up at the early nonlinear regime.
1220:
1221: It is also remarkable that, for $Q \neq 0$, $R(t)$ is time dependent,
1222: and the relative weight of $\tilde{\Omega}$ and $\tilde{Q}$ changes with
1223: time. This may have remarkable consequences in the highly nonlinear
1224: regime. For instance, rotation can prevent the formation of finite time
1225: singularities in the case with zero surface tension \cite{Rocco00}.
1226:
1227: Finally, notice that at nonlinear order there is also a
1228: geometrical term, $Q/(2 \pi R^3)$, independent of $n$. This
1229: term cannot be scaled out by the same procedure that eliminated its
1230: counterpart at linear order.
1231:
1232:
1233: \section{Numerical study of an exact solution and its weakly nonlinear
1234: approximation} \label{Sec:5}
1235:
1236: The purpose of this section is to put the weakly nonlinear analysis to
1237: the test, as a quantitative approximation. We will check it against an
1238: exact time dependent solution of the case $B=0$ which evolves towards
1239: the Saffman--Taylor finger of width $\lambda=1/2$ in the channel
1240: geometry. This will allow us to check how fast is the convergence of
1241: the expansion and how accurate it may be even beyond its radius of
1242: convergence, when it becomes an asymptotic series.
1243:
1244: We define $F$ as the ratio between the maximum height of the interface
1245: (at the finger tip) and half the width of the cell (in our case
1246: $L=2\pi$). $F$ is a dimensionless amplitude which is proportional to
1247: $\varepsilon$ and thus measures how deep the system is into the
1248: nonlinear regime. Furhermore, since the system is described by the
1249: evolution of a curve, it may also be convenient to have a more global
1250: characterization of its configuration in order to compare the exact
1251: result and the different approximations. We propose the use of the flux
1252: $\Phi$, defined as the total amount of fluid 1 per unit length and unit
1253: time flowing across the horizontal line located at the mean interface
1254: position \cite{Magdaleno00}).
1255:
1256:
1257: \subsection{The time dependent exact solution with $B=0$
1258: and $\lambda=1/2$}
1259:
1260: An explicit solution of the problem without surface tension
1261: ($B=0$) and valid for any viscosity contrast $A$, describing the
1262: growth of a single finger which asymptotically fills half
1263: of the channel ($\lambda=1/2$), can be obtained from
1264: Eqs.(\ref{mappi}), (\ref{setun}), and (\ref{setdos}). This reads:
1265: \be
1266: f(w,t)= -\ln w + V_{o}t + a(0) + \ln \left[\sqrt{1+
1267: (b(0)^2-1)\exp(2V_{o}t)} +w\right],
1268: \label{fwt}
1269: \ee
1270: with $a(0)$ and $b(0)$ as initial constants, and $b(0)\gg 1$. Since
1271: this solution is valid for any $A$, particularly $A=0$, all terms of
1272: the weakly nonlinear equation (\ref{generity}) depending on $A$
1273: will necessarily have no contribution.
1274:
1275: We will take the following values for the parameters:
1276: \be
1277: V_{o}=\frac{1}{2}, \qquad b(0)=\frac{1}{\nu}=1000, \qquad a(0)= \ln
1278: \nu +\frac{\nu^2}{2}.
1279: \ee
1280: Fig.\ \ref{FigEx1} shows the evolution of the interface for this
1281: solution. When $F\simeq1$ the finger shape is indistinguishable
1282: from the stationary Saffman--Taylor solution
1283: (Fig.\ \ref{FigEx2}),
1284: except near the points $x=\pi/2$ and $x=3\pi/2$
1285: when the asymptotes of the finger must develop at infinite
1286: time. When $F = 1/\pi$, the slope of the interface
1287: becomes $1$ at these two points and the series (\ref{generity})
1288: loses its convergence, according to the discussion of Section
1289: \ref{SectionIIIA}.
1290:
1291: In Fig.\ \ref{FigEx3} we present the amplitude of the Fourier
1292: modes of the exact solution. It is clear that relatively few
1293: modes suffice to accurately reproduce the stationary finger shape
1294: in the tip region.
1295:
1296: \subsection{Expansion of the exact solution}
1297:
1298: In Section \ref{subsecdit} we expanded the exact solution
1299: (\ref{mappi}) in a small parameter $\nu$. This expansion revealed
1300: a hierarchy of amplitudes of the different modes
1301: for an initial condition close
1302: to the planar interface. The
1303: solution (\ref{fwt}) can be expanded similarly and yields:
1304: \be
1305: \nu = \frac{1}{b(0)}, \ \ \ \alpha (t) = - \frac{\nu}{b(t)}, \ \ \
1306: d(t)=V_{o}t + \ln b(t), \ \ \ b(t)= \sqrt{ \nu^2 + (1 -
1307: \nu^2) \exp(2V_{o}t)}.
1308: \ee
1309: In agreement with (\ref{hexpan}), and using
1310: the same parameters as in the previous section, we obtain up to third
1311: order:
1312: \be
1313: h(x,t)=\nu e^{t/2} \cos x + \nu ^3
1314: \left[\left(\frac{1}{2} e^{t/2} - \frac{1}{8}
1315: e^{3t/2}\right)\cos x - \frac{1}{24} e^{3t/2} \cos 3x \right].
1316: \ee
1317: The exact flux $\Phi$ can be easily computed from
1318: the stream function, which in its turn can be related to the form of
1319: the mapping Eq.(\ref{fwt}) (see for instance
1320: Ref.\cite{Magdaleno00}). In Fig.\ \ref{FigFp1} we compare the
1321: flux $\Phi$ of the approximate result above with the exact result
1322: coming from (\ref{fwt}). We clearly observe that the first
1323: correction to the linear regime improves the flux (to a 5 \%
1324: accuracy) from $ F \simeq 0.12 $ to $ F \simeq 0.17$. The
1325: amplitude of the mode $k=1$ (Fig.\ \ref{FigFp3}) is fairly
1326: well reproduced
1327: by the linear approximation up to $F \simeq 0.20$, and the correction
1328: of order $\nu^3$ gives a value for the mode $k=3$ reasonably good
1329: up to $F \simeq 0.20$ (Fig.\ \ref{FigFp4}) and improves the mode $k=1$
1330: almost until $F \simeq 0.25 $. Above these values of $F$ the
1331: deviation from the exact interface is exponential.
1332:
1333: Notice that the expansion up to $\nu^3$ considered here does not
1334: account for the complete three mode coupling of
1335: Eq.(\ref{triorder}) which contains part of the higher orders in
1336: $\nu$. This explains why going from order $\nu$ (linear) to order
1337: $\nu^3$ improves only slightly the shape of the finger, the flux $\Phi$,
1338: and the amplitude of the mode $k=1$.
1339:
1340:
1341: \subsection{Weakly nonlinear expansion}
1342:
1343: In this section we study the mode coupling equation
1344: (\ref{triorder}). We solve this equation in cases where only
1345: modes $k=1$ and $k=1,3$ are present, and compare the result with
1346: both the exact solution and the approximation to order $\nu^3$
1347: introduced in the two previous sections.
1348:
1349: We consider the initial condition $\beta
1350: _{1}(0)=\varepsilon=0.001$ for the first mode and zero for the
1351: rest of modes. We use Eq.\ (\ref{triorder}) recalling that, since
1352: $\beta _{i}$ represent the amplitudes of the cosine functions,
1353: $\delta_{i} = \beta_{i}/2$. The result is that the modes
1354: $k\neq 1$ remain zero and:
1355: \be
1356: \beta _{1} (t)=\frac{\varepsilon \exp(0.5t)}{\sqrt{1+ 0.25
1357: \varepsilon ^2 (\exp(t) -1)}}, \ee where we see that the mode
1358: $k=1$ saturates to a finite value as $t \to \infty$. Returning to
1359: Fig.\ \ref{FigFp1} and \ref{FigFp2}, we see that the flux $\Phi$
1360: approximates the exact value to a 5 \% uncertainty) until $F
1361: \simeq 0.23$. Fig.\ \ref{FigFp3} shows that the amplitude of the
1362: first mode starts deviating significantly from the exact solution
1363: around $F \simeq 0.40$.
1364:
1365: Next, we add the third mode to the equation (\ref{triorder}) with
1366: the initial condition $ \beta_{1}= \varepsilon + 3 \varepsilon
1367: ^3/8$ and $\beta_{3}= \varepsilon ^3/24$. The set of equations
1368: to be solved numerically is:
1369: \be
1370: && \dot{\beta}_{1}=\frac{1}{2}\beta_{1} - \frac{1}{8}\beta_{1}^3 -
1371: \frac{5}{4} \beta_{3}^2 \beta_{1} + \frac{1}{4} \beta_{1}^2
1372: \beta_{3},
1373: \\ && \dot{\beta}_{3}=\frac{3}{2} \beta_{3} - \frac{3}{4} \beta_{1}^2
1374: \beta_{3} - \frac{27}{8} \beta_{3}^3. \label{system} \ee As shown
1375: in Figs.\ \ref{FigFp1} and \ref{FigFp4} the flux does not improve
1376: significantly, while the amplitude of the mode $k=3$ remains
1377: acceptable almost until $F\simeq 0.25$, and improves the previous
1378: result by more than a 50 \% almost until $F\simeq 0.33$. Beyond
1379: this value of $F$ the third mode falls to zero due to the second
1380: term in the r.h.s. of Eq.(\ref{system}), proportional to
1381: $\beta_{1}^2 \beta_{3}$ but negative. This is a signature that
1382: higher order terms, for instance of order $\beta_{1}^4 \beta_{3}$,
1383: become important when $\beta_{1}$ is of order one.
1384:
1385: From this study we can conclude that using only two modes and the lowest
1386: order nonlinear correction for this case (involving three mode
1387: couplings), both the shape of the interface and the total flux are
1388: reasonably well described within the whole range of convergence of the
1389: series (up to $F \simeq 1/\pi$). This indicates that the weakly
1390: nonlinear description of the problem works remarkably well at the
1391: quantitative level, at least in some physically relevant cases. To what
1392: extent this conclusion holds for more complicated situations, for
1393: instance those involving competition of different fingers, remains an
1394: open question.
1395:
1396:
1397: \subsection{Partial resummation}
1398:
1399: According to Eq. (\ref{triorder}), we can write a closed equation for
1400: the mode $k=1$ which is correct up to third order couplings, of the
1401: form:
1402: \be
1403: \dot{\beta}_{1} = \frac{1}{2} \beta_{1} - \frac{1}{4} \beta_{1}^2
1404: \dot{\beta}_{1},
1405: \label{laderiva}
1406: \ee
1407: where the mode $k=3$ has been set initially to zero.
1408:
1409: The presence of $\dot{\beta}_{1}$ in the right hand side gives
1410: rise to terms of order higher than $\beta_{1}^3$. Thus, solving
1411: this equation amounts to doing a partial resummation of these
1412: higher order terms. This kind of closed differential equation was
1413: already obtained in Ref.\cite{Mirandac98}. In our
1414: approach this differential equation is obtained in a systematic way by
1415: identifying the terms $\lambda(n) \beta_{n}$ and replacing them
1416: for $\dot{\beta}_{n}$. For example, in Eq. (\ref{triorder}) the
1417: term $T(k,s,l)$ could have been partially resummated in the
1418: two mode coupling differential equation.
1419:
1420: The solution of Eq.(\ref{laderiva}) yields the transcendental
1421: form:
1422: \be
1423: t = 2 \ln \left(\frac{\beta_{1}}{\varepsilon}\right)
1424: +\frac{1}{4}(\beta_{1}^2-\varepsilon^2).
1425: \ee
1426: A numerical computation of the flux based on this equation shows a
1427: dramatic improvement up to $F \simeq 1/\pi$. It is
1428: surprising and remarkable that the equation for a single mode,
1429: including the resummation suggested by the first nonlinear
1430: correction, reproduces the exact solution to a great accuracy in
1431: the full range of convergence and defines an excellent
1432: approximation further inside the nonlinear regime. This suggests
1433: that this kind of resummation is not arbitrary, and might have a deeper
1434: physical justification.
1435:
1436: In addition, we obtain an amplitude for the mode $k=1$ within 10 \%
1437: accuracy up to $F \simeq 0.8$, and it is still quite good until
1438: $F=1$. Although the exact spectrum shows the increasing
1439: importance of the mode $k=3$ at these values of $F$, the agreement
1440: between the exact result and this approximation for a single mode
1441: $k=1$ is quite remarkable.
1442:
1443: We recall that the series converges up to $F \simeq
1444: 1/ \pi$. Beyond this value of $F$, increasing the order of
1445: mode coupling in Eq. (\ref{laderiva}) does not guarantee that the
1446: approximation improves. Accordingly, there is an optimal finite
1447: order of approximation for a given value of $\varepsilon$, as
1448: usual in asymptotic series. In this sense, our results above seem
1449: to indicate that Eq.(\ref{laderiva}) is the optimal approximation
1450: for the mode $k=1$ in the asymptotic (nonconvergent) region (i.e.,
1451: for moderately long times).
1452:
1453: In the same spirit of Eq. (\ref{laderiva}), we can also derive a
1454: closed system of differential equations which includes third order
1455: couplings for the modes $k=1,3$. Its solution shows no
1456: qualitative differences with the case of the single mode $k=1$
1457: above, as far as the flux and the amplitude of the first mode are
1458: concerned. The improvement in the amplitude of the third mode and
1459: consequently in the shape of the finger does not reach the high
1460: values of $F$ that the first mode reaches ($F \simeq 0.8$), as we
1461: can see in Fig.\ \ref{FigFp4}. In a short range of $F$ the mode
1462: $k=3$ is better than the case shown in the previous section, but
1463: deviates from the exact solution much earlier than the mode $k=1$
1464: for the same reasons than in the previous section. Specifically
1465: the mode $k=3$ is extremely good until $F \simeq 0.25$, starts
1466: deviating by more than a 10 \% at $F \simeq 0.33$ and is
1467: about 50 \% of the value at $F \simeq 0.40$.
1468:
1469: We conclude that the weakly nonlinear formalism, apart from its nominal
1470: range of validity for very small amplitudes where it converges very fast
1471: to the exact dynamics, can describe regimes beyond those
1472: expected in principle with reasonable accuracy. In the case of the
1473: growth of a single finger,
1474: for instance, we have explicitly seen that simply two modes together
1475: with an appropriate partial resummation, yields a remarkably good
1476: approximation in the full range of convergence of the expansion ($F <
1477: 1/\pi$).
1478:
1479:
1480: \section{Conclusion and perspectives}
1481: \label{conclusions}
1482:
1483: We have developed a systematic scheme to derive the successive orders of
1484: mode couplings in a weakly nonlinear regime, adapted to the study of
1485: interfacial dynamics in Hele--Shaw flows. This has been done with full
1486: generality, including both the channel geometry driven by gravity and
1487: pressure, and the radial geometry with arbitrary injection and
1488: centrifugal driving, which includes the case of a rotating Hele--Shaw
1489: cell. The method could also be applied to even more general
1490: (position dependent) drivings. The formulation in real space has
1491: enabled us to address the issue of convergence of the mode coupling
1492: expansion. We have found that the exact condition for convergence in
1493: the channel geometry is $| h_x | < 1$ in every point of the interface.
1494:
1495: We have carried out the explicit derivation of nonlinear couplings up to
1496: third order in the case of channel geometry, in order to obtain the
1497: leading nonlinear contributions in cases where the second order
1498: contribution vanishes. These include the case of zero viscosity
1499: contrast and the time dependent single finger solution of width
1500: $\lambda=1/2$.
1501:
1502: On the analytical side, we have also illustrated the usefulness of
1503: the weakly nonlinear analysis in elucidating the role of the
1504: different parameters on the dynamics, in some examples. In the
1505: case of single finger solutions growing in a channel, we have
1506: shown how an exact nontrivial property of the problem (the
1507: relationship between $\lambda$ and viscosity contrast $A$ for zero
1508: surface tension) can be extracted from an analysis order by order,
1509: without really knowing the exact solution. In the case of
1510: rotating Hele--Shaw flows, we have discussed the asymmetry between
1511: the roles of injection and rotation at the lowest nonlinear order,
1512: as opposed to what happens in the channel geometry, where the
1513: equivalence between gravity and injection driving is exact.
1514:
1515: On the numerical side we have checked the predictions of the
1516: weakly nonlinear analysis, at different orders, against exact
1517: solutions. We have found that the convergence is quite fast for
1518: small amplitudes. Furthermore, in the case of a single finger
1519: growing in a channel we have explicitly seen that the analysis to
1520: low orders (up to three mode couplings) yields a good description
1521: of the interface dynamics even close to the radius of convergence.
1522: With an appropriate resummation scheme, the prediction is
1523: relatively good even beyond that point.
1524:
1525: Some of the most interesting applications of the systematic weakly
1526: nonlinear approach have not been developed here since they deserve
1527: a separate, in--depth analysis. One of them is the study of the
1528: scaling properties of fluctuations in stably stratified Hele--Shaw
1529: flows with external noise \cite{porous}. A more direct
1530: application of our formalism which also deserves a separate
1531: analysis is the derivation of amplitude equations for the region
1532: near the instability threshold through a center manifold
1533: reduction. This would made contact with what is most commonly
1534: referred to as ``weakly nonlinear'' approach in the literature of
1535: pattern formation. A detailed study of this point with a careful
1536: discussion of the nature of the bifurcation and its implications
1537: will be presented elsewhere \cite{Alvarez02}.
1538:
1539:
1540: \section{Acknowledgements}
1541: We acknowledge financial support from the Direcci\'on General de
1542: Ense\~{n}anza Superior (Spain) under Projects PB96-1001-C02-02 and
1543: PB97-0906, E. Alvarez--Lacalle also acknowledges a grant from the
1544: Comissionat per a Universitats i Recerca of the Generalitat de
1545: Catalunya. The work was also supported by the European Commission
1546: Project ERB FMRX-CT96-0085 (Training and Mobility of Researchers).
1547:
1548: \begin{references}
1549:
1550: \bibitem{Bensimon86}
1551: D. Bensimon, L. Kadanoff, S. Liang, B.I. Shraiman, and C. Tang, Rev.
1552: Mod. Phys. {\bf 58}, 977 (1986).
1553:
1554: \bibitem{Saffman86}
1555: P. G. Saffman, J. Fluid. Mech. {\bf 173}, 73 (1986).
1556:
1557: \bibitem{Pelce88}
1558: P. Pelc\'e, {\it Dynamics of curved fronts} (Perspectives in Physics,
1559: Academic Press, San Diego, 1988).
1560:
1561: \bibitem{LangerlesHouches}
1562: J. S. Langer, {\it Lectures in the theory of pattern formation} (ed:
1563: J. Souletie, J. Vannimenus, R. Stora in Chance and Matter, Les
1564: Houches, 1986, North--Holland, Amsterdam, 1987).
1565:
1566: \bibitem{Kesslerkop88}
1567: D. Kessler, J. Koplik, and H. Levine, Adv. Phys. {\bf 37}, 255
1568: (1988).
1569:
1570: \bibitem{Crosshoh93}
1571: M. C. Cross and P. C. Hohenberg, Rev. Mod. Phys. {\bf 65}, 851
1572: (1993).
1573:
1574: \bibitem{Saffman58}
1575: P. G. Saffman and G. I. Taylor, Proc. Roy. Soc. London Ser. A \bf
1576: 245\rm, 312 (1958)
1577:
1578: \bibitem{Manneville90}
1579: P. Manneville, {\it Dissipative structures} (Springer--Verlag, Berlin,
1580: 1990).
1581:
1582: \bibitem{Paterson81}
1583: L. Paterson, J. Fluid. Mech. {\bf 113}, 513 (1981).
1584:
1585: \bibitem{JDChen89}
1586: J. D. Chen, J. Fluid. Mech. {\bf 201}, 223 (1989).
1587:
1588: \bibitem{Trygvason83}
1589: G. Tryggvason and H. Aref, J. Fluid Mech. \bf136\rm, 1 (1983).
1590:
1591: \bibitem{Hong86}
1592: D. C. Hong and J.S. Langer, Phys. Rev. Lett. \bf56\rm, 2032 (1986).
1593:
1594: \bibitem{Shraiman86}
1595: B. I. Shraiman, Phys. Rev. Lett. \bf56\rm, 2028 (1986).
1596:
1597: \bibitem{Combescot86}
1598: R. Combescot, T. Dombre, V. Hakim, Y. Pomeau, and A. Pumir, Phys.
1599: Rev. Lett. \bf56\rm, 2036 (1986).
1600:
1601: \bibitem{Tanveer93}
1602: S. Tanveer, Phil. Trans. R. Soc. Lond. A {\bf 343}, 155 (1993).
1603:
1604: \bibitem{Siegelprl96}
1605: M. Siegel and S. Tanveer, Phys. Rev. Lett. \bf76\rm, 419 (1996).
1606:
1607: \bibitem{Siegeljfm96}
1608: M. Siegel, S. Tanveer, and W. S. Dai, J. Fluid. Mech. \bf323\rm, 201
1609: (1996).
1610:
1611: \bibitem{Magdaleno00}
1612: J. Casademunt and F. X. Magdaleno , Phys. Rep. \bf337\rm, 1 (2000)
1613:
1614: \bibitem{Trygvason85}
1615: G. Tryggvason and H. Aref, J. Fluid Mech. \bf154\rm, 287 (1985).
1616:
1617: \bibitem{Maher85}
1618: J. V. Maher, Phys. Rev. Lett. {\bf54}, 1498 (1985).
1619:
1620: \bibitem{Difranjima89}
1621: M. W. Difranceso and J. V. Maher, Phys. Rev. A {\bf 39}, 4709 (1989).
1622:
1623: \bibitem{Difranjimb89}
1624: M. W. Difranceso and J. V. Maher, Phys. Rev. A {\bf 40}, 295 (1989).
1625:
1626: \bibitem{Casademunt91}
1627: J. Casademunt and D. Jasnow, Phys. Rev. Lett. {\bf 67}, 3677 (1991).
1628:
1629: \bibitem{Casademunt94}
1630: J. Casademunt and D. Jasnow, Physica D {\bf 79}, 387 (1994).
1631:
1632: \bibitem{Francessos1}
1633: P. Jacquar and P. S\'eguer, J. de M\'ecanique \bf1\rm, 367 (1962).
1634:
1635: \bibitem{Folch00}
1636: R. Folch, E. Paun\'e, and J. Casademunt (unpublished).
1637:
1638: \bibitem{Carrillo96}
1639: Ll. Carrillo, F.X. Magdaleno, J. Casademunt, and J. Ort\' {\i}n,
1640: Phys. Rev. E {\bf54}, 6260 (1996).
1641:
1642: \bibitem{porous}
1643: A controlled experiment of quenched noise in a Hele--Shaw cell with
1644: stably stratified flow has been recently proposed by A.
1645: Hern\'andez--Machado, J. Soriano, A. M. Lacasta, M. \'A.
1646: Rodr\'{\i}guez, L. Ram\'{\i}rez--Piscina, and J. Ort\'{\i}n
1647: (unpublished).
1648:
1649: \bibitem{Mirandac98}
1650: J. A. Miranda and M. Widom, Int. J. Mod. Phys. B {\bf 12}, 931
1651: (1998).
1652:
1653: \bibitem{Mirandar98}
1654: J. A. Miranda and M. Widom, Physica D {\bf 120}, 315 (1998).
1655:
1656: \bibitem{Schwartz89}
1657: L. W. Schwartz, Phys. Fluids A {\bf 1}, 167 (1989).
1658:
1659: \bibitem{Carrillo99}
1660: Ll. Carrillo, J. Soriano, and J. Ort\'{\i}n, Phys. Fluids {\bf 11},
1661: 778 (1999).
1662:
1663: \bibitem{Carrillo00}
1664: Ll. Carrillo, J. Soriano, and J. Ort\'{\i}n, Phys. Fluids {\bf 12},
1665: 1685 (2000).
1666:
1667: \bibitem{Entov96}
1668: V. M. Entov, P. I. Etingof, and D. Ya. Kleinbock, Euro. J. Appl.
1669: Math. {\bf 6}, 399 (1995).
1670:
1671: \bibitem{Crowdy1}
1672: D. G. Crowdy, Q. Appl. Maths. (2000), in press.
1673:
1674: \bibitem{Rocco00}
1675: F. X. Magdaleno, A. Rocco and J. Casademunt, Phys. Rev. E {\bf
1676: 62}, R5887 (2000).
1677:
1678: \bibitem{Alvarez01}
1679: E. Alvarez--Lacalle, J. Ort\'{\i}n, and J. Casademunt (unpublished).
1680:
1681: \bibitem{Goldstein98}
1682: R. E. Goldstein, A. I. Pesci, and M. J. Shelley, Phys. Fluids {\bf
1683: 10}, 2701 (1998).
1684:
1685: \bibitem{Alvarez02}
1686: E. Alvarez--Lacalle et al. (unpublished).
1687:
1688: \end{references}
1689:
1690: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1691: \newpage
1692:
1693: \begin{figure}
1694: \begin{center}
1695: \epsfig{file=Figure1.eps,width=6cm} \vspace{0.1cm} \caption{Sketch
1696: of the Hele-Shaw cell in channel geometry.} \label{Figcanal}
1697: \end{center}
1698: \end{figure}
1699: \begin{figure}
1700:
1701:
1702: \begin{center}
1703: \epsfig{file=Figure2.eps,width=6cm} \vspace{0.5cm} \caption{Sketch
1704: of the Hele-Shaw cell in circular geometry.} \label{Figradial}
1705: \end{center}
1706: \end{figure}
1707:
1708:
1709: \begin{figure}
1710: \begin{center}
1711: \epsfig{file=Figure3.eps,width=6cm} \vspace{0.1cm}
1712: \caption{Evolution of the interface at different values of $\pi
1713: F$.} \label{FigEx1}
1714: \end{center}
1715: \end{figure}
1716: \begin{figure}
1717:
1718:
1719: \begin{center}
1720: \epsfig{file=Figure4.eps,width=6cm} \vspace{0.5cm}
1721: \caption{Comparison of the exact solution
1722: (Eq.(\protect{\ref{fwt}})) at $F=1$ (thin line) and the
1723: Saffman-Taylor stationary finger (thick line).} \label{FigEx2}
1724: \end{center}
1725: \end{figure}
1726:
1727:
1728: \begin{figure}
1729: \begin{center}
1730: \epsfig{file=Figure5.eps,width=6cm} \vspace{0.5cm}
1731: \caption{Amplitude of the Fourier modes of the exact solution at
1732: consecutive values of $F$.} \label{FigEx3}
1733: \end{center}
1734: \end{figure}
1735:
1736:
1737:
1738:
1739: \begin{figure}
1740: \begin{center}
1741: \epsfig{file=Figure6.eps,width=6cm} \vspace{0.5cm} \caption{Flux
1742: $\Phi$ as a function of the factor F for different approximations
1743: (see text for details).} \label{FigFp1}
1744: \end{center}
1745: \end{figure}
1746:
1747: \newpage
1748: \begin{figure}
1749: \begin{center}
1750: \epsfig{file=Figure7.eps,width=6cm} \vspace{0.5cm}
1751: \caption{Relative error of the flux $\Phi$ as a function of the
1752: factor F for different approximations (see text for details).}
1753: \label{FigFp2}
1754: \end{center}
1755: \end{figure}
1756:
1757:
1758: \begin{figure}
1759: \begin{center}
1760: \epsfig{file=Figure8.eps,width=6cm} \vspace{0.5cm} \caption{The
1761: first mode amplitude as a function of the factor F for different
1762: approximations (see text for details).} \label{FigFp3}
1763: \end{center}
1764: \end{figure}
1765:
1766:
1767: \begin{figure}
1768: \begin{center}
1769: \epsfig{file=Figure9.eps,width=6cm} \vspace{0.5cm} \caption{The
1770: third mode amplitude as a function of the factor F for different
1771: approximations (see text for details).} \label{FigFp4}
1772: \end{center}
1773: \end{figure}
1774:
1775:
1776:
1777: \end{document}
1778: