1:
2: \documentclass[prl,twocolumn]{revtex4}
3: \usepackage{graphicx}
4: \usepackage{psfrag}
5: \usepackage{psfig}
6:
7: \begin{document}
8: \title{Asymptotic behaviour of the Rayleigh--Taylor instability}
9: \author{Laurent Duchemin$^1$, Christophe Josserand$^2$ and Paul Clavin$^3$\\
10: \small $^1$ Department of Applied Mathematics and Theoretical Physics, University of Cambridge,
11: Cambridge CB3 0WA, United Kingdom\\
12: \small $^2$ Laboratoire de Mod\'elisation en M\'ecanique\\
13: \small UPMC-CNRS UMR 7607, 4 place Jussieu, 75252 Paris C\'edex 05 France\\
14: \small $^3$ IRPHE, Universit\'es d'Aix-Marseille I \& II-CNRS, 49 rue Joliot-Curie, BP 146, 13384 Marseille Cedex France\\
15: }
16:
17: \begin{abstract}
18: We investigate long time numerical simulations of the inviscid Rayleigh-Taylor instability at Atwood
19: number one using a boundary integral method. We are able to attain the asymptotic behavior for the
20: spikes predicted by Clavin \& Williams\cite{clavin} for which we give a simplified demonstration. In
21: particular we observe that the spike's curvature evolves like $t^3$ while the overshoot in acceleration
22: shows a good agreement with the suggested $1/t^5$ law. Moreover, we obtain consistent results
23: for the prefactor coefficients of the asymptotic laws. Eventually we exhibit the self-similar behavior of
24: the interface profile near the spike.
25: \end{abstract}
26: \maketitle
27:
28: \section{Introduction}
29: The Rayleigh-Taylor (RT) instability appears when, under gravity, an heavy liquid is placed over a lighter one\cite{rayleigh}.
30: This instability is crucial for our understanding of different phenomena in fluid mechanics: mixing, thermal
31: convection (\cite{kada} and cited ref. herein) and also finger number selection in splashes\cite{gueyffier}.
32: It is also important in inertial confinment fusion where the mass ablation provides a stabilizing effect to
33: the interface instability\cite{sanz}. Without ablation, after the exponential growth of the perturbations due
34: to the linear RT instability, nonlinear profiles develop through the formation of bubbles of lighter fluid
35: rising into the heavier one and falling
36: spikes of the heavier liquid penetrating the lighter one. In the
37: general situations of viscous fluids which are immiscible and/or have Atwood number not equal to unity ($A_T=(\rho_h
38: -\rho_l)/(\rho_h+\rho_l)$ with $\rho_h$ and $\rho_l$ being the density of the heavier and lighter fluids
39: respectively), famous mushrooms-like structures grow for
40: larger times\cite{kada,sohn1,sohn2}. The limit of an inviscid fluid above a vacuum ($A_T=1$) without surface tension plays
41: a specific role since no stabilizing effects are present in the linear dynamics. Numerous theoretical
42: and numerical work have focused on this idealized limit in order to track insights into the instability
43: itself\cite{layzer,zhang,hazac,abar,mika,inog}. It has been
44: shown using a conformal mapping that a finite time singularity might appear in the conformal
45: plane\cite{tanveer} and
46: it is also suspected that for some sufficiently irregular initial conditions finite time singularities should also
47: be observed in the physical plane. However, starting with sufficiently smooth initial conditions, the
48: asymptotic dynamics\cite{mika,zhang,inog} presents a constant velocity rising bubble separated by free falling tiny spikes as
49: displayed on figure \ref{profils}. Although the
50: rising bubble motion has been described using local properties of the flow\cite{gonch}, the asymptotic dynamics
51: of the spikes is far from being well understood. The single mode approach gives a fair
52: description of the constant velocity of the rising bubble ($v_b=\sqrt{g/(3k)}$ where $g$ is the
53: acceleration of the gravity and $k$ the wavenumber of the perturbation) but gives only partial results
54: for the spike\cite{zhang}. The fluid there obeys free fall dynamics to a good approximation and the
55: pressure field of the flow leads to an overshoot in the acceleration. The accelerated
56: motion of the liquid stretches the spike geometry and one expects self-similar behaviour of the
57: tip of the spikes.
58: \begin{figure}[h]
59: \begin{center}
60: \psfrag{vertical velocity}{\Large Vertical velocity}
61: \psfrag{time}{\Large time}
62: \psfrag{-pi}{$-\pi$}
63: \psfrag{0}{$0$}
64: \psfrag{-1}{$-1$}
65: \psfrag{-2}{$-2$}
66: \psfrag{1}{$1$}
67: \psfrag{2}{$2$}
68: \psfrag{3}{$3$}
69: \psfrag{4}{$4$}
70: \psfrag{5}{$5$}
71: \psfrag{6}{$6$}
72: \psfrag{7}{$7$}
73: \psfrag{8}{$8$}
74: \psfrag{5}{$5$}
75: \psfrag{10}{$10$}
76: \psfrag{pi}{$\pi$}
77: \psfrag{2p}{$2\pi$}
78: \psfrag{3p}{$3\pi$}
79: \psfrag{4p}{$4\pi$}
80: \psfrag{5p}{$5\pi$}
81: \psfrag{6p}{$6\pi$}
82: \psfrag{6.5}{$t=6.5$}
83: \psfrag{7.5}{$t=7.5$}
84: \psfrag{9.5}{$t=9.5$}
85: \centerline {\includegraphics[width=6cm]{all.eps}}
86: \end{center}
87: \caption{Snapshots of the interface subject to the Rayleigh-Taylor instability for time
88: ranging from $t=0$ to $10$, starting with a small amplitude sine mode (left). On the
89: right is shown the velocity of several points along the interface, non-dimensionalized with the stationnary
90: bubble rising velocity $\sqrt{g/3k}$, as a function of time.\label{all}}
91: \label{profils}
92: \end{figure}
93: Recently, an asymptotic theory using a parallel flow description of the velocity field near the
94: spikes has been constructed \cite{clavin}. The interface dynamics is nonlinear for large time and can
95: be described using the theory of characteristics which gives rise to finite time singularity
96: solutions. In the case of regular dynamics a self-similar
97: description of the peak is obtained for large time: the maximal curvature of the interface at the
98: peak tip is found to behave like the cubic power of time $t^3$. Moreover, the spike position,
99: following the free fall $ \frac12 g t^2$ at leading order, is shown to converge to the constant
100: acceleration $g$ with an overshoot in acceleration decreasing like $t^{-5}$. In this letter, we
101: present a numerical study of the Rayleigh-Taylor instability which focuses on the large time dynamics
102: of the spikes in order to investigate the self similar dynamics predicted in \cite{clavin}.
103: We consider the dynamics for an inviscid liquid (heavy) with an exterior fluid of zero density
104: ($A_t=1$) and no surface tension. The numerics use a boundary integral method (BIM later on).
105: Due to strong numerical instabilities, a careful treatment of the
106: interface using conformal mapping is needed as explained below. The results are then
107: shown and compared with the theory.
108:
109: \section{Asymptotic analysis and numerical method}
110:
111: We consider the two-dimensional motion of an inviscid fluid
112: above a vacuum, subject to a negative acceleration $-g$. A periodic
113: sine perturbation of the interface of wave number $k$ is implemented as initial conditions.
114: Neglecting surface tension, the equations of motion have no
115: control parameter after rescaling the time, the position and the velocity potential $\varphi$ by
116: factors $\sqrt{gk}$, $k$ and $\sqrt{k^3/g}$ respectively.
117: The interface is described by $y=\alpha(x,t)$, where $y$ is the direction along the
118: gravity and $x$ orthogonal to it (see figure \ref{map}).
119: The velocity field ${\bf U}=(u,v)$ satisfies the dimensionless Euler equation
120: $$ \frac{d{\bf U}}{dt}=-{\bf \nabla}P +{\bf e_y} $$
121: where $P(x,y,t)$ is the pressure, ${\bf e_y}$ the non-dimensional acceleration due to gravity and
122: the fluid density $\rho=1$.
123: The kinetic equation for the interface reads~:
124: $$ \frac{\partial \alpha(x,t)}{\partial t}+u\frac{\partial \alpha(x,t)}{\partial x}=v$$
125: with the velocity field $(u,v)$ evaluated at the interface $(x,\alpha(x,t))$.
126: Starting at time $t=0$ with a small sine amplitude interface, we observe for large time
127: that the fluid particles located in the vicinity of the tiny spikes come from an almost free fall from the
128: initial interface region. Therefore, following \cite{clavin},
129: we assume quasi-parallel steady flow for the velocity field which gives then in the tip region $ |u|\ll |v|$ and~:
130: $$ v \sim \sqrt{2y}$$
131: with $y \sim \frac12 t^2$ for large time.
132: Writing a perturbation expansion of the velocity field in the tip region $|x| \ll y$, we in fact consider:
133: $$ v=\sqrt{2(y+f(x,y,t))} $$
134: with $ f(x,y,t) \ll y$. Taking a Taylor expansion in $x$ of the perturbation $f$, we obtain by symmetry:
135: $$ v = \sqrt{2y}+\frac{f_0(y,t)}{\sqrt{2y}}+\frac{x^2}{2} \frac{f_2(y,t)}{\sqrt{2y}} + O(x^4) $$
136: We limit our expansion to the second order in $x$ for the velocity field later on.
137: Incompressibility gives~:
138: $$ u=-\left(\sqrt{\frac{1}{2y}}+\frac{\partial (f_0(y,t)/\sqrt{2y})}{\partial y}\right)x + O(x^3).$$
139: At the leading order (where we neglect even the perturbation $f(x,y,t)$) we obtain the following
140: expression for the interface location:
141: $$ \frac{\partial \alpha(x,t)}{\partial t}-\frac{x}{\sqrt{2\alpha(x,t)}} \frac{\partial \alpha(x,t)}{\partial x}=\sqrt{2\alpha(x,t)} $$
142: which can be solved using the methods of charasteristics (see \cite{clavin}).
143: Writing $\alpha(x,t)=t(\frac{t}{2} - \gamma(x,t))$ and noting that $\gamma(x,t)\ll t/2$ in the spike region,
144: we obtain, after linearisation~:
145: $$ \frac{\partial \gamma(x,t)}{\partial t}-\frac{x}{t}\frac{\partial \gamma(x,t)}{\partial x}=0 $$
146: which has self-similar solution of the form $ \gamma(x,t)=\theta(xt)$. A first conclusion can be drawn
147: about the curvature of the interface at the tip, $\kappa=-\partial^2\alpha/\partial x^2|_{x=0}$, which is
148: thus found to increase as the cubic power of time~:
149: \begin{equation}
150: \kappa=t^3 \theta''(0)
151: \label{cubic}
152: \end{equation}
153:
154: The next order terms of the expansion allow the determination of the function $f_0(y,t)$ near the tip.
155: Using the constant value of the pressure at the interface we use the projection of the Euler
156: equation at the interface on its local tangent~:
157: $$ \frac{du}{dt} + \frac{\partial \alpha(x,t)}{\partial x} \frac{dv}{dt} = \frac{\partial \alpha(x,t)}{\partial x}. $$
158: Since on the interface $dP(x,\alpha(x,t),t)/dx=0$.
159: We develop this equation at first non-zero order (which will end up to be the first order in $x$)
160: with the expansion $\theta(xt)=\theta(0)+x^2t^2 \theta''(0)/2+O(x^4)$.
161: Remembering that $|f| \ll y$, we can neglect also the larsge scale terms $\partial^2 (f_0(y,t)/\sqrt{2y})/\partial t \partial y$ and $\sqrt{2y}\partial^2 (f_0(y,t)/\sqrt{2y})/\partial y^2$ with respect to the others.
162: We obtain finally for the tip position $y=y_s$~:
163: $$ \frac{\partial f_0(y_s,t)}{\partial t} +\sqrt{2y_s}\frac{\partial f_0(y_s,t)}{\partial y}=\frac{df_0(y_s,t)}{dt}=
164: \sqrt{\frac{2}{y_s}} \frac{1}{\kappa} $$
165:
166: Recalling that:
167: $ \frac{dy_s}{dt}=\sqrt{2y_s}+\frac{f_0(y_s,t)}{\sqrt{2y_s}} $
168: we obtain for the tip acceleration at leading order:
169: \begin{equation}
170: \frac{d^2y_s}{dt^2}=1+\frac{1}{\sqrt{2 y_s}} \frac{df_0(y_s,t)}{dt}=1+\frac{2}{t^5 \theta''(0)}
171: \label{quint}
172: \end{equation}
173: which corresponds to an overshoot in the spike acceleration decreasing as the fifth
174: power of time.
175:
176: The numerical method is elaborated using the incompressible and potential properties
177: of the flow. The velocity field can thus be evaluated everywhere when the velocity
178: potential is known on the interface thanks to Cauchy's theorem, in the spirit of pionnering works\cite{vinje,cokelet,BMO,MZ}.
179: The non-dimensional Bernoulli equation on the free surface reads~:
180: \begin{equation}
181: \frac{\partial \varphi}{\partial t} = -\frac{1}{2}(\nabla \varphi)^2 + y,
182: \end{equation}
183: where the velocity potential $\varphi$ is a harmonic function in the
184: fluid domain $\Omega$~:
185: \begin{equation}
186: \Delta \varphi = 0
187: \label{lap}
188: \end{equation}
189:
190: The kinematic condition on the free surface expresses the fact that
191: fluid particles move with the same normal velocity than the free surface
192: itself~:
193: \begin{equation}
194: \frac{d \bf x}{dt} \cdot {\bf n} = \nabla \varphi \cdot {\bf n}
195: \label{kin}
196: \end{equation}
197:
198: Knowing $\varphi$ on the free surface at a given time-step,
199: we search for the solution of equation (\ref{lap}) that satisfies
200: this boundary condition (\ref{kin}). We use the complex potential $\beta(z) = \varphi + i \psi$
201: and the conformal map $f(z) = exp(-i z)$ (Cf. Figure \ref{map}),
202: where $z=x+iy$ and $\psi$ is the stream function. The conformal map transforms the periodic
203: domain $\Omega$ into the closed domain $M$.
204: Since $\psi$ is harmonic inside $\Omega$, $\beta(z)$ is analytic
205: inside $\Omega$ and therefore $\gamma(\zeta)=\beta(f(z))$
206: is analytic inside $M$. Using Cauchy's theorem, we obtain a
207: Fredholm equation of the second kind for the stream function $\psi$ which is solved using
208: discretization of the free surface ($\partial \Omega$ and thus $\partial M$).
209: This linear system of equations is solved using a $LU$ decomposition.
210: Once we know $\psi$ on each point on $\partial M$, the complex velocity of each marker in the physical plane is given by~:
211: \begin{equation}
212: \frac{d \beta}{dz} = u - i \, v
213: \end{equation}
214: where $u$ and $v$ are the horizontal and vertical velocities
215: respectively.
216: This complex velocity is computed with a finite difference scheme
217: using the values of the complex potential on the collocation points
218: on $\partial \Omega$.
219: \begin{figure}[h]
220: \psfrag{omega}{$\Omega$}
221: \psfrag{zeta}{$\zeta=f(z)$}
222: \psfrag{x}{$y$}
223: \psfrag{y}{$x$}
224: \psfrag{function}{$f(z) = e^{-i z}$}
225: \psfrag{M}{$M$}
226: \centerline {\includegraphics[width=8cm]{map_new.eps}}
227: \caption{\label{map} Conformal map used to transform the physical periodic plane $\Omega$ into a closed domain $M$.}
228: \end{figure}
229: The position of the surface markers (kinematic condition)
230: and the value of the velocity potential on each of these markers
231: (Bernoulli equation) are then updated in time using a fourth
232: order Runge-Kutta method.
233:
234: \section{Results and discussions}
235: We have performed numerical simulations of the Rayleigh-Taylor instability using the numerical
236: method described above. We start with a small amplitude sine-mode. The unavoidable
237: numerical noise cannot be damped by the numerics and the calculations always end up subject to
238: numerical instabilities.
239: Nevertheless, we emphasize that the numerical scheme used here is remarkably robust and
240: can be accurately evolved to reach the large time where the scalings predicted by the theory \cite{clavin} are valid.
241: Comparing our simulations with recent numerical works\cite{sohn1,sohn2,hazac}, we have been able to run the
242: dynamics at least twice as far which corresponds roughly to an increase of a factor of $8$ in the tip's curvature.
243:
244: The position of the spike is shown on figure \ref{spike} as function of time. We observe that the
245: asymptotics dynamics are very well approximated by the relation $ y_s =\frac12 g(t-t_0)^2$ as shown in
246: the inset to the figure with $t_0=3.74$. This remarkable behavior, in good agreement with the free fall
247: hypothesis,
248: suggests that $t_0$ is the time delay accounting for the initial exponential development of the instability.
249: We will therefore present further data on the curvature dependance and the acceleration of the tip as functions
250: of this delayed time $t-t_0$ instead of $t$.
251: \begin{figure}[h]
252: \centerline {
253: \includegraphics[width=7cm]{spike_zoom2.eps}}
254: \caption{Position of the spike $y_s(t)$ as a function of time. The inset shows in a log-log plot of the spike
255: position (black curve) as function of time $t-t_0$ with $t_0=3.74$ obtained by a second order
256: polynomial fit of $y_s$. the dashed line shows the expected behavior $\frac12 t^2$. }
257: \label{spike}
258: \end{figure}
259: The curvature $\kappa_s$ at the tip is then shown on figure \ref{courbure}. The large time asymptotic
260: behavior is similarly found to follow the cubic law (see equation \ref{cubic}) with $\theta''(0)=1.5$.
261: \begin{figure}[h]
262: \centerline {
263: \includegraphics[width=7cm]{courspike.eps}}
264: \caption{Spike curvature $\kappa_s$ calculated at the tip $y=y_s$ as function of the delayed time $t-t_0$
265: in a log-log plot. The dashed line displays the cubic law (\ref{cubic}) with $\theta''(0)=1.5$.}
266: \label{courbure}
267: \end{figure}
268: In addition, the acceleration of the tip is computed by finite differences on the tip velocity and
269: the overshoot in the acceleration is presented on figure \ref{accel}. We observe that the results
270: look noisier than the two previous ones. Two factors can explain such noise: firstly, we are
271: looking to a finite difference which decreases to zero so that the numerical errors are relatively more
272: important. However,
273: we note that the overshoot in acceleration shows a good agreement with the $1/t^5$ law, noting
274: that no adjustable parameter is used in this comparison.
275: \begin{figure}[h]
276: \centerline {
277: \includegraphics[width=7cm]{overshoot.eps}}
278: \caption{Overshoot in acceleration, defined as the difference between the tip acceleration and the
279: gravity. The plot is in log-log scale and with the delayed time $t-t_0$. The dashed line shows the
280: theoretical prediction (\ref{quint}) using the value of $\theta''(0)$ obtained from figure \ref{courbure}}
281: \label{accel}
282: \end{figure}
283: Moreover, the self similar structure of the interface near the tip has been exhibited on figure \ref{similar}.
284: We observe after the proper rescaling on the left part of the figure that the interface
285: profiles collapse onto a single curve near the spike.
286: \begin{figure}[h]
287: \centerline {
288: \includegraphics[width=6cm]{selfsim.eps}}
289: \caption{Self-similar structure of the tip: the interface profile around the spike have been superimposed
290: on the right side of the figure for different time $t$ ranging from $4$ to $12$. The left side of the figure
291: shows the same curves rescaled by factor $1/(t-t_0)$ and $(t-t_0)$ for the $x$ and $y$ coordinates
292: respectively, following the scaling behavior predicted by the theory.}
293: \label{similar}
294: \end{figure}
295:
296:
297: We have thus exhibited large times numerical simulations of the Rayleigh-Taylor instability which
298: present asymptotic scaling behavior in agreement with theoretical predictions using Taylor
299: expansions of the free fall velocity field at the spike\cite{clavin}. Although our numerics always stops
300: due to numerical instability, we have been able to reach large time enough to exhibit the cubic power in time dependance for the spike
301: curvature and the inverse of the quintinc power of time decreasing of the overshoot in acceleration.
302:
303: It is our pleasure to thank J. Ashmore for useful comments.
304: We acknowledge also the support of CEA through the contract CEA/DIF N° 4600051147/P6H29.
305:
306: \begin{thebibliography}{99}
307: \expandafter\ifx\csname natexlab\endcsname\relax\def\natexlab#1{#1}\fi
308: \bibitem{rayleigh} Lord Rayleigh, {\em Scientific Papers II (Cambridge University Press, Cambridge,
309: United Kingdom, 1900)}, p 200.
310:
311: \bibitem{kada} B. Castaing, G. Gunaratne, F. Heslot, L. Kadanoff, A. Libchaber, S. Thomae, X. Wu, S. Zaleski and G. Zanetti, {\em J. Fluid Mech.} {\bf 204}, 1-30 (1989).
312:
313: \bibitem{gueyffier} D. Gueyffier and S. Zaleski, {\em C. R. Acad. Sci. Paris IIb} {\bf 326}, 839-844 (1998).
314:
315: \bibitem{sanz} J. Sanz, J. Ramirez, R. Ramis, R. Betti and R.P.J. Town, {\em Phys. Rev. Lett.} {\bf 89},
316: 195002 (2002).
317:
318: \bibitem{sohn1} S.-I. Sohn, {\em Phys. Rev. E} {\bf 67}, 026301 (2003).
319:
320: \bibitem{sohn2} S.-I. Sohn, {\em Phys. Rev. E} {\bf 69}, 036703 (2004).
321:
322: \bibitem{layzer} D. Layzer, {\em Astrophys. J.} {\bf 122}, 1 (1955).
323:
324: \bibitem{zhang} Q. Zhang, {\em Phys. Rev. Lett.} {\bf 81}, 3391 (1998).
325:
326: \bibitem{hazac} G. Hazak,{\em Phys. Rev. Lett.} {\bf 76}, 4167 (1996).
327:
328: \bibitem{abar} S.I. Abarzhi,{\em Phys. Rev. Lett.} {\bf 81}, 337 (1998).
329:
330: \bibitem{mika} K.O. Mikaelian,{\em Phys. Rev. Lett.} {\bf 80}, 508 (1998).
331:
332: \bibitem{inog} N. Inogamov,{\em Astrophys. Space Phys.} {\bf 10}, 1-335 (1999).
333:
334: \bibitem{tanveer} S. Tanveer, {\em Proc. R. Soc. Lond. A} {\bf 441}, 501-525 (1993).
335:
336: \bibitem{gonch} V.N. Goncharov, {\em Phys. Rev. Lett.} {\bf 88}, 134502 (2002).
337:
338: \bibitem{clavin} P. Clavin \& F. Williams accepted for publication in J. Fluid Mech. (2005).
339:
340: \bibitem{vinje} T. Vinje \& P. Brevig, {\em Adv. Water Resources} {\bf 4}, 77 (1981).
341:
342: \bibitem{cokelet} M.S. Longuet-Higgins \& E.D. Cokelet {\em Proc. R. Soc. Lond. A} {\bf 350}, 1--26 (1976).
343:
344: \bibitem{BMO} G.R. Baker, D.I. Meiron and S.A. Orszag, {\em Phys. Fluids} {\bf 23}, 1485 (1980).
345: \bibitem{MZ} R. Menikoff and C. Zemach, {\em J. Comput. Phys.} {\bf 51}, 28 (1983).
346:
347: %\bibitem{hecht}
348:
349: \end{thebibliography}
350:
351: \end{document}
352:
353: We can write Cauchy's theorem on $\partial M$~:
354: \begin{equation}
355: \int_\mathcal{\partial M} \frac{\gamma(\zeta)}{\zeta-\zeta_e}d\zeta = 0,
356: \label{cauchy}
357: \end{equation}
358: where $\zeta_e$ is a point outside $\partial M$.
359: It can be shown that the real part of equation \ref{cauchy} --
360: when $\zeta_e$ approaches $\partial M$ --
361: is a Fredholm equation of the second kind for the stream function $\psi$,
362: which is known to have a unique solution.
363: If we discretize the free surface in the physical plane with collocation
364: points, the images of these points are on $\partial M$ and we can
365: write a discrete version of equation \ref{cauchy}~:
366: \begin{equation}
367: \sum_{j=1}^N \left( Re(\Gamma_{k,j})\psi_j - Im(\Gamma_{k,j})\varphi_j \right) = 0
368: \label{discrete}
369: \end{equation}
370: where the coefficients $\Gamma_{k,j}$ are known if we consider $\gamma$
371: to be linear between two collocation points on $\partial M$.
372: Having $N$ collocation points $\zeta_k$ on $\partial M$,
373: we can write $N$ equations like \ref{discrete}
374: corresponding $\zeta_e \to \zeta_k$.
375:
376: :
377:
378: $$ -t^3\theta"(0) (\frac{\partial F_0(x_s,t)}{\partial t} +\frac{\partial f_0(x_s,t)}{\partial x})
379: -\frac{\partial^2 F_0(x,t)}{\partial t \partial x}+f_2(x_s,t)+\frac{g}{2x_s}=0 $$
380:
381: which simplifies into
382: