0805.1242/analysis_thin_films/analysis_thin_films.tex
1: 
2: \chapter{Nonlinear dynamics of phase separation in ultra-thin films: Analysis}
3: \label{ch:analysis_thin_films}
4: 
5: \section{Overview}
6: \label{sec:analysis_thin_films:overview}
7: %
8: %
9: In this chapter we recall the Navier--Stokes Cahn--Hilliard (NSCH) equations
10: derived
11: in Sec.~\ref{sec:background:NSCH}, shifting focus from the passive tracer,
12: to the active one.  We study these equations in the incompressible
13: limit.  We shall specialize to thin films, in which the binary liquid forms
14: a thin layer on a substrate.  Then, if the vertical gradients are small compared
15: to horizontal ones, a long-wavelength version of the NSCH equations can be
16: obtained, representing a much simplified problem.  We present the
17: derivation of these long-wavelength or thin-film Stokes Cahn--Hilliard equations.
18:  We analyze these equations and obtain results concerning the existence,
19:  regularity, and uniqueness of solutions.
20: %
21: %
22: %
23: \section{The thin-film Stokes Cahn--Hilliard equations}
24: \label{sec:model}
25: %
26: \noindent 
27: %
28: In this section we recall the incompressible Navier--Stokes Cahn--Hilliard
29: equations from Sec.~\ref{sec:background:NSCH} and discuss the passage from
30: these equations to the long-wavelength approximation in two dimensions.
31: %
32: We enumerate the scaling rules necessary to obtain the simplified
33: equations.  
34: %
35: Finally, we arrive at a set of equations that describe phase separation in
36: a thin film acted on by arbitrary body forces.
37: 
38: The incompressible NSCH equations derived in Sec.~\ref{sec:background:NSCH}
39: describe the coupled effects of phase separation
40: and flow in a density-matched binary fluid.  If $\bm{v}$ is the fluid velocity
41: and $c$ is the concentration of the mixture, where $c=\pm1$ indicates total
42: segregation, then these fields evolve in the following way,
43: %
44: %
45: \begin{subequations}
46: \begin{gather}
47:         \frac{\partial \bm{v}}{\partial t}+\bm{v}\cdot\nabla\bm{v}=\nabla\cdot\bm{T}-\frac{1}{\rho_0}\nabla\phi,\\
48:         \frac{\partial c}{\partial t}+\bm{v}\cdot\nabla c=D \Delta\left(c^3-c-\gamma\Delta{c}\right),\\
49:         \nabla\cdot\bm{v}=0,
50: %
51: \end{gather}%
52: \label{eq:NSCH}%
53: \end{subequations}%
54: %
55: %
56: where
57: \begin{equation}
58:         {T}_{ij} =-\frac{p}{\rho_0}\delta_{ij}+\nu\left(\frac{\partial
59:         v_i}{\partial
60:         x_j}+\frac{\partial
61:         v_j}{\partial x_i}\right)-\beta\gamma\frac{\partial c}{\partial x_i}\frac{\partial
62:         c}{\partial x_j}
63: \label{eq:NSCH_tensor}%
64: \end{equation}%
65: %
66: %
67: is the stress tensor, $p$ is the fluid pressure, $\phi$ is the body potential
68: and $\rho_0$ is the constant mixture density.  The constant $\nu$ is the kinematic
69: viscosity, $\nu=\eta/\rho_0$, where $\eta$ is the viscosity.  Additionally,
70: $\beta=\varepsilon_0/\rho_0$ is a constant with units of $[\mathrm{Energy}][\mathrm{Mass}]^{-1}$,
71: $\sqrt{\gamma}$ is a constant that gives the typical width of interdomain
72: transitions, and $D$ a diffusion coefficient with dimensions $[\mathrm{Length}]^2[\mathrm{Time}]^{-1}$.
73: 
74: 
75: If the system has a free surface in the vertical or $z$-direction and has
76: infinite or periodic boundary conditions (BCs) in the lateral or $x$-direction,
77: then the vertical BCs we impose are 
78: %
79: %
80: \begin{subequations}
81: \begin{equation}
82: u = w = c_z = c_{zzz}\text{ on }z=0,
83: \end{equation}
84: %
85: %
86: while on the free surface $z=h\left(x,t\right)$ they are
87: %
88: %
89: %
90: %
91: \begin{equation}
92: \hat{n}_i \hat{n}_j{T}_{ij} = -\Gamma\kappa,\qquad \hat{n}_i \hat{t}_j {T}_{ij} =
93: -\frac{\partial\Gamma}{\partial s},
94: \label{eq:BC_stress}
95: \end{equation}
96: %
97: \begin{equation}
98: w=\frac{\partial h}{\partial t}+u\frac{\partial h}{\partial x},
99: \label{eq:BC_height}
100: \end{equation}
101: %
102: \begin{equation}
103: \hat{n}_i \partial_i c = 0, \qquad \hat{n}_i \partial_i \Delta c = 0,
104: \label{eq:BC_c}
105: \end{equation}%
106: \label{eq:BCs}%
107: \end{subequations}%
108: %
109: %
110: %
111: where $\hat{\bm{n}}$ is the unit normal to the surface, $\hat{\bm{t}}$,
112: is the unit vector tangent to the surface, $s$ is the
113: surface coordinate, $\Gamma$ is the surface tension and $\kappa$ is the
114: mean curvature,
115: %
116: %
117: %
118: \begin{multline*}
119: \phantom{aaaaa}
120: \kappa=\nabla\cdot\hat{\bm{n}}
121: \\
122: =\left(\partial_x,\partial_z\right)\cdot\left(\frac{-\partial_x{h}}{\sqrt{1+\left(\partial_x{h}\right)^2}},\frac{1}{\sqrt{1+\left(\partial_x{h}\right)^2}}\right)=\frac{\partial_{xx}h}{\left[1+\left(\partial_x{h}\right)^2\right]^{\frac{3}{2}}}.
123: \phantom{aaaaa}
124: \end{multline*}
125: %
126: %
127: This choice of BCs guarantees the conservation of the total mass and volume,
128: respectively 
129: %
130: %
131: %
132: \begin{equation}
133: M =\int_{\text{Dom}\left(t\right)} c\left(\bm{x},t\right)d^2x,\qquad V=\int_{\text{Dom}\left(t\right)}d^2x.
134: \label{eq:mass_consv}
135: \end{equation}
136: %
137: %
138: %
139: Here 
140: $\text{Dom}\left(t\right)%=\{\left(x,y\right)|x\in\left[0,L\right],z\in\left[0,h\left(x,t\right)\right]\}
141: $
142: represents the time-varying domain of integration, whose time dependence
143: is due to the variable nature of the free surface height.  Note that in view
144: of the concentration BC~\eqref{eq:BC_c}, the stress BC~\eqref{eq:BC_stress}
145: does not contain $c\left(\bm{x},t\right)$ or its derivatives.
146: 
147: These equations admit a great simplification if the fluid forms a thin layer
148: of mean thickness $h_0$, for then the scale of lateral variations
149: $\lambda$ is large compared with the scale of vertical variations $h_0$.
150:  This is shown in Fig.~\ref{fig:defn_sketch}.
151:  Specifically, the parameter $\Smalll = h_0/\lambda$ is small, and after nondimensionalization
152:  of Eq.~\eqref{eq:NSCH}, we expand its solution in terms of this parameter,
153:  keeping only the lowest-order terms.  For a review of this method and its
154:  applications, see~\cite{Oron1997}.  For simplicity, we work in two
155:  dimensions, but the generalization to three dimensions is easily effected.
156: %
157: %
158: \begin{figure}
159: \centering
160: %
161:     \includegraphics[width=.8\textwidth]{defn_sketch}
162: \caption{Definition sketch for the long-wavelength approximation in two dimensions.
163:  The approximation relies on the smallness of the parameter $\Smalll=h_0/\lambda$.}
164: \label{fig:defn_sketch}
165: \end{figure}
166: %
167: %
168: %
169: 
170: %
171: % 
172: In terms of the small parameter $\Smalll$, the equations nondimensionalize
173: as follows.   The diffusion timescale is $t_0 = \lambda^2 / D = h_0^2 /
174: \left(\Smalll^2 D\right)$ and we choose this to be the unit of time.  Then
175: the unit of horizontal velocity is $u_0 = \lambda/ t_0 = \Smalll D / h_0$
176: so that $u = \left(\Smalll D / h_0\right)U$, where variables in upper
177:  case denote dimensionless quantities.  Similarly the vertical velocity
178:  is $w = \left(\Smalll^2 D / h_0\right)W$.  For the equations of
179:  motion to be half-Poiseuille at $O\left(1\right)$ (in the absence of the
180:  backreaction) we choose $p = \left(\eta D / h_0^2\right)P$ and $\phi = \left(\eta
181:  D / h_0^2\right)\Phi$.  Using these scaling rules, the dimensionless momentum
182:  equations are
183: %
184: %
185: %
186: %
187: %
188: \begin{multline}
189: %
190: %
191: \Smalll Re \left(\frac{\partial U}{\partial T} + U\frac{\partial
192: U}{\partial X}+ W\frac{\partial U}{\partial Z}\right) 
193: %
194: %
195: = -\frac{\partial}{\partial X}\left(P+\Phi\right)+\Smalll^2\frac{\partial^2
196: U}{\partial X^2}+\frac{\partial^2 U}{\partial Z^2} \\
197: %
198: %
199: -\tfrac{1}{2}\frac{\beta\gamma}{\nu D}\frac{\partial}{\partial
200: X}\bigg[\Smalll^2\left(\frac{\partial c}{\partial X}\right)^2
201: %
202: %
203: +\left(\frac{\partial c}{\partial Z}\right)^2\bigg]-\frac{\beta\gamma}{\nu
204: D}\frac{\partial c}{\partial X}\bigg[\Smalll^2\frac{\partial^2 c}{\partial X^2}+\frac{\partial^2 c}{\partial Z^2}\bigg],
205: %
206: %
207: %
208: %
209: \label{eq:thin_film1}
210: \end{multline}
211: %
212: %
213: %
214: %
215: %
216: %
217: %
218: \begin{multline}
219: %
220: \Smalll^3 Re \left(\frac{\partial W}{\partial T} + U\frac{\partial
221: W}{\partial X} + W\frac{\partial W}{\partial Z}\right) 
222: %
223: %
224: = -\frac{\partial}{\partial Z}\left(P+\Phi\right)+\Smalll^4\frac{\partial^2
225: W}{\partial X^2}+\Smalll^2\frac{\partial^2 W}{\partial Z^2}  \\
226: %
227: %
228: -\tfrac{1}{2}\frac{\beta\gamma}{\nu D}\frac{\partial}{\partial
229: Z}\bigg[\Smalll^2\left(\frac{\partial c}{\partial X}\right)^2+\left(\frac{\partial
230: c}{\partial Z}\right)^2\bigg]
231: %
232: %
233: -\frac{\beta\gamma}{\nu D}\frac{\partial c}{\partial Z}\bigg[\Smalll^2\frac{\partial^2
234: c}{\partial X^2}+\frac{\partial^2 c}{\partial Z^2}\bigg],
235: %
236: %
237: %
238: \label{eq:thin_film2}
239: \end{multline}
240: %
241: %
242: where
243: %
244: %
245: \begin{equation}
246: Re = \frac{\Smalll D}{\nu} = \frac{\left(\Smalll D / h_0\right)h_0}{\nu} =
247: O\left(1\right).
248: \end{equation}
249: %
250: The choice of ordering for the Reynolds number $Re$ allows us to recover
251: half-Poiseuille flow at $O\left(1\right)$.  We delay choosing the ordering
252: of the dimensionless group $\beta\gamma/D\nu$ until we have examined the
253: concentration equation, which in nondimensional form is
254: %
255: %
256: %
257: %
258: \begin{multline}
259: %
260: %
261: \Smalll^2 \left(\frac{\partial c}{\partial T} + U\frac{\partial c}{\partial
262: X} + W\frac{\partial c}{\partial Z}\right)\\
263: %
264: %
265: = \Smalll^2 \frac{\partial^2}{\partial X^2}\left(c^3-c\right)+\frac{\partial^2}{\partial Z^2}\left(c^3-c\right)
266: %
267: %
268: -\Smalll^4 C_{\mathrm{n}}^2\frac{\partial^4c}{\partial X^4}-C_{\mathrm{n}}^2\frac{\partial^4
269: c}{\partial Z^4}
270: %
271: %
272: -2\Smalll^2 C_{\mathrm{n}}^2 \frac{\partial^2}{\partial X^2}\frac{\partial
273: c}{\partial Z^2},
274: %
275: %
276: \label{eq:conc0}
277: \end{multline}
278: %
279: %
280: %
281: %
282: where $C_{\mathrm{n}}^2=\gamma/h_0^2$.  By switching off the backreaction in
283: the momentum equations (corresponding
284: to $\beta\gamma/D\nu\rightarrow0$), we find the trivial solution to the
285: momentum equations, $U = W = \partial_X\left(P+\Phi\right) = \partial_Z\left(P+\Phi\right)
286: = 0$, $H = 1$.  The concentration boundary conditions are then $c_Z = c_{ZZZ}
287: = 0$ on $Z = 0,1$ which forces $c_Z\equiv0$ so that the Cahn--Hilliard equation
288: is simply
289: %
290: %
291: \[
292: \frac{\partial c}{\partial T} = \frac{\partial^2}{\partial X^2}\left(c^3-c\right)-\Smalll^2 C_{\mathrm{n}}^2\frac{\partial^4c}{\partial
293: X^4}.
294: \]
295: %
296: To make the lubrication approximation consistent, we take 
297: %
298: \begin{equation}
299: \Smalll C_{\mathrm{n}}= \tilde{C}_{\mathrm{n}}= \Smalll\sqrt{\gamma} /
300: h_0 = O\left(1\right).
301: \label{eq:gamma_h}
302: \end{equation}
303: %
304: %
305: %
306: %
307: We now carry out a long-wavelength approximation to Eq.~\eqref{eq:conc0},
308: writing
309: $U=U_0+O\left(\Smalll\right)$, $W=W_0+O\left(\Smalll\right)$, $c=c_0+\Smalll
310: c_1+\Smalll^2 c_2+...$.  We examine the boundary conditions on $c\left(\bm{x},t\right)$
311: first.  They are $\hat{\bm{n}}\cdot\nabla c = \hat{\bm{n}}\cdot\nabla\Delta
312: c=0$
313: on $Z=0,H$; on $Z=0$ these conditions are simply $\partial_Z c = \partial_{ZZZ}
314: c=0$, while on $Z=H$ the surface derivatives are determined by the relations
315: %
316: %
317: \[
318: \hat{\bm{n}}\cdot\nabla\propto-\Smalll^2 H_X\partial_X+\partial_Z,
319: \]
320: %
321: %
322: \[
323: \hat{\bm{n}}\cdot\nabla\Delta\propto-\Smalll^4
324: H_X\partial_{XXX}-\Smalll^2H_X\partial_X\partial_{ZZ}+\Smalll^2\partial_{XX}\partial_Z+\partial_{ZZZ}.
325: \]
326: %
327: %
328: Thus, the BCs on $c_0$ are simply $\partial_Z c_0 = \partial_{ZZZ} c_0=0$
329: on $Z=0,H$, which forces $c_0=c_0\left(X,T\right)$.  Similarly, we
330: find  $c_1=c_1\left(X,T\right)$, and
331: % %
332: % %
333: % \[
334: % c_2\left(X,Z,T\right)=\tilde{c}_2\left(X,T\right)+\tfrac{1}{2}Z^2\frac{H_X}{H}\frac{\partial
335: % c_0}{\partial X}.
336: % \]
337: % %
338: % %
339: % Hence,
340: %
341: %
342: \[
343: \frac{\partial c_2}{\partial Z}=Z\frac{H_X}{H}\frac{\partial
344: c_0}{\partial X},\qquad
345: %
346: %
347: \frac{\partial^2 c_2}{\partial Z^2}=\frac{H_X}{H}\frac{\partial
348: c_0}{\partial X},\qquad \text{for any }Z\in \left[0,H\right].
349: \]
350: %
351: %
352: %
353: In the same manner, we derive the results $\partial_{ZZZZ}c_2=\partial_{ZZZZ}c_3=0$.
354: Using these facts, equation~\eqref{eq:conc0} becomes
355: %
356: %
357: %
358: %
359: %
360: % \begin{multline*}
361: % %
362: % %
363: % \varepsilon^2\left(\frac{\partial c_0}{\partial T} + U_0\frac{\partial
364: % c_0}{\partial
365: % X} + W_0\underbrace{\frac{\partial c_0}{\partial Z}}_{=0}\right)
366: % %
367: % %
368: % = \varepsilon^2 \frac{\partial^2}{\partial X^2}\left(c_0^3-c_0\right)
369: % %
370: % %
371: % +\left(3c_0^2-1\right)\frac{\partial^2 c_2}{\partial Z^2}
372: % %
373: % %
374: % \\-\epsilon^2 \tilde{C}_{\mathrm{n}}^2\frac{\partial^4c}{\partial X^4}
375: % %
376: % %
377: % -\frac{\tilde{C}_{\mathrm{n}}^2}{\varepsilon^2}\frac{\partial^4}{\partial
378: % Z^4}\left(c_0+\varepsilon c_1+\varepsilon^2 c_2+\varepsilon^3 c_3+\varepsilon^4
379: % c_4+...\right)
380: % %
381: % %
382: % \\-2\tilde{C}_{\mathrm{n}}^2\frac{\partial^2}{\partial X^2}\frac{\partial^2}{\partial
383: % Z^2}\left(c_0+\varepsilon c_1+\varepsilon^2 c_2+\varepsilon^3 c_3+\varepsilon^4
384: % c_4+...\right),
385: % \end{multline*}
386: %
387: %
388: %
389: %
390: %
391: %
392: %
393: %
394: % \begin{multline*}
395: % %
396: % %
397: % \Smalll^2\left(\frac{\partial c_0}{\partial T} + U_0\frac{\partial c_0}{\partial
398: % X}\right)
399: % %
400: % %
401: % = \Smalll^2 \frac{\partial^2}{\partial X^2}\left(c_0^3-c_0\right)
402: % %
403: % %
404: % \\+\left(3c_0^2-1\right)\frac{H_X}{H}\frac{\partial c_0}{\partial X}
405: % %
406: % %
407: % -\Smalll^2 \tilde{C}_{\mathrm{n}}^2\frac{\partial^4c}{\partial X^4}
408: % %
409: % %
410: % -{\tilde{C}_{\mathrm{n}}^2}\Smalll^2\frac{\partial^4 c_4}{\partial
411: % Z^4}
412: % %
413: % %
414: % -2\Smalll^2\tilde{C}_{\mathrm{n}}^2\frac{\partial^2}{\partial X^2}\bigg[\frac{H_X}{H}\frac{\partial
415: % c_0}{\partial X}\bigg].
416: % \end{multline*}
417: %
418: %
419: %
420: %
421: %We divide out by $\Smalll^2$,
422: %
423: %
424: %
425: %
426: \begin{multline*}
427: %
428: %
429: \frac{\partial c_0}{\partial T} + U_0\frac{\partial c_0}{\partial
430: X} =
431: %
432: %
433: \\
434: \frac{\partial^2}{\partial X^2}\left(c_0^3-c_0\right)
435: %
436: %
437: -\tilde{C}_{\mathrm{n}}^2\frac{\partial^4c}{\partial X^4}
438: %
439: %
440: +\left(3c_0^2-1\right)\frac{H_X}{H}\frac{\partial c_0}{\partial X}
441: %
442: %
443: -2\tilde{C}_{\mathrm{n}}^2\frac{\partial^2}{\partial X^2}\frac{H_X}{H}\frac{\partial
444: c_0}{\partial X}
445: %
446: %
447: -{\tilde{C}_{\mathrm{n}}^2}\frac{\partial^4 c_4}{\partial Z^4}.
448: %
449: %
450: %
451: %
452: %
453: \end{multline*}
454: %
455: %
456: %
457: %
458: We now integrate this equation from $Z=0$ to $Z=H$ and use the boundary conditions
459: %
460: %
461: %
462: \[
463: \frac{\partial^3 c_4}{\partial Z^3}=0\qquad\text{on }Z=0,
464: \]
465: %
466: %
467: %
468: \begin{eqnarray*}
469: %
470: %
471: \frac{\partial^3 c_4}{\partial Z^3}&=&
472: %
473: %
474: %H_X\frac{\partial^3 c_0}{\partial X^3}+
475: %
476: %
477: %H_X\left(\frac{\partial}{\partial
478: %X}\frac{\partial^2 c_2}{\partial Z^2}\right)_{Z=H}
479: %
480: %
481: %-\left(\frac{\partial^2}{\partial
482: %X^2}\frac{\partial c_2}{\partial Z}\right)_{Z=H},\\
483: %
484: %
485: %&=&
486: H_X\frac{\partial^3 c_0}{\partial X^3}+
487: %
488: %
489: H_X\frac{\partial}{\partial X}\left(\frac{H_X}{H}\frac{\partial c_0}{\partial
490: X}\right)
491: %
492: %
493: -H\frac{\partial^2}{\partial X^2}\left(\frac{H_X}{H}\frac{\partial c_0}{\partial
494: X}\right)\qquad\text{on }Z=H.
495: \end{eqnarray*}
496: %
497: %
498: %
499: %
500: After rearrangement, the concentration equation becomes
501: %
502: %
503: %
504: %
505: \begin{multline*}
506: %
507: %
508: H\frac{\partial c_0}{\partial T}+H\overline{U_0}\frac{\partial c_0}{\partial
509: X} = 
510: %
511: %
512: \\
513: H\frac{\partial^2}{\partial X^2}\left[c_0^3-c_0-\tilde{C}_{\mathrm{n}}^2\frac{\partial^2
514: c_0}{\partial X^2}-\tilde{C}_{\mathrm{n}}^2\frac{H_X}{H}\frac{\partial c_0}{\partial
515: X}\right]
516: %
517: %
518: +\frac{\partial H}{\partial X}\frac{\partial}{\partial X}\left[c_0^3-c_0-\tilde{C}_{\mathrm{n}}^2\frac{\partial^2
519: c_0}{\partial X^2}-\tilde{C}_{\mathrm{n}}^2\frac{H_X}{H}\frac{\partial c_0}{\partial
520: X}\right],
521: \end{multline*}
522: %
523: %
524: %
525: where
526: %
527: %
528: \[
529: \overline{U_0} = \frac{1}{H}\int_0^H U_0\left(X,Z,T\right)dZ,
530: \]
531: %
532: %
533: is the vertically-averaged velocity.  Introducing
534: %
535: \[
536: \mu 
537: %= \left[c_0^3-c_0-\tilde{C}_{\mathrm{n}}^2\frac{\partial^2
538: %c_0}{\partial X^2}-\tilde{C}_{\mathrm{n}}^2\frac{H_X}{H}\frac{\partial c_0}{\partial
539: %X}\right]
540: %
541: %
542: =c_0^3-c_0-\frac{\tilde{C}_{\mathrm{n}}^2}{H}\frac{\partial}{\partial X}\left(H\frac{\partial
543: c_0}{\partial X}\right),
544: \]
545: %
546: %
547: %
548: the thin-film Cahn--Hilliard equation becomes
549: %
550: %
551: %
552: \begin{equation}
553: \frac{\partial c_0}{\partial T}+\overline{U_0}\frac{\partial c_0}{\partial
554: X} = \frac{1}{H}\frac{\partial}{\partial X}\left(H\frac{\partial\mu}{\partial
555: X}\right).
556: \end{equation}
557: 
558: We are now able to perform the long-wavelength approximation to Eqs.~\eqref{eq:thin_film1}
559: and~\eqref{eq:thin_film2}.  At lowest order, Eq.~\eqref{eq:thin_film2} is
560: %
561: %
562: \[
563: \frac{\partial}{\partial Z}\left(P+\Phi\right)+r\frac{\partial
564: c_0}{\partial Z}\left[\frac{\partial^2 c_0}{\partial X^2}+\frac{\partial^2
565: c_2}{\partial Z^2}\right]+\tfrac{1}{2}r\frac{\partial}{\partial
566: Z}\left(\frac{\partial c_0}{\partial X}\right)^2=0,\qquad
567: r=\frac{\Smalll^2\beta\gamma}{D\nu},
568: \]
569: %
570: %
571: which is simply $\partial_Z\left(P+\Phi\right)=0$, and hence
572: %
573: %
574: \[
575: P+\Phi = P_{\mathrm{surf}}+\Phi_{\mathrm{surf}}\equiv P\left(X,H\left(X,T\right),T\right)+\Phi\left(X,H\left(X,T\right),T\right).
576: \]
577: %
578: %
579: %
580: Here
581: %
582: %
583: \begin{equation}
584: r=\frac{\Smalll^2\beta\gamma}{D\nu}=O\left(1\right)
585: \end{equation}
586: %
587: %
588: is the dimensionless measure of the backreaction strength.  Thus, Eq.~\eqref{eq:thin_film1}
589: becomes
590: %
591: %
592: %
593: \[
594: \frac{\partial^2 U_0}{\partial Z^2}=\frac{\partial}{\partial X}\left(P_{\mathrm{surf}}+\Phi_{\mathrm{surf}}\right)+r\frac{\partial}{\partial
595: X}\left(\frac{\partial c_0}{\partial X}\right)^2+r\frac{\partial c_0}{\partial
596: X}\frac{\partial^2 c_2}{\partial Z^2},
597: \]
598: %
599: %
600: %
601: and using $\partial_{ZZ}c_2=\left(H_X/H\right)\left(\partial c_0/\partial
602: X\right)$ this is
603: %
604: %
605: %
606: \begin{equation}
607: \frac{\partial^2 U_0}{\partial Z^2}=\frac{\partial}{\partial X}\left(P_{\mathrm{surf}}+\Phi_{\mathrm{surf}}\right)+\frac{r}{H}\frac{\partial}{\partial
608: X}\left[H\left(\frac{\partial c_0}{\partial X}\right)^2\right].
609: \label{eq:d2U}
610: \end{equation}
611: %
612: %
613: %
614: At lowest order, the BC~\eqref{eq:BC_stress} becomes
615: %
616: %
617: \begin{equation}
618: \frac{\partial U_0}{\partial Z}=\frac{\partial\tilde{\Gamma}}{\partial{X}}\qquad\text{on
619: }Z=H,
620: \label{eq:gamma_bc}
621: \end{equation}
622: %
623: %
624: %
625: where $\tilde{\Gamma}$ is the dimensionless, spatially-varying surface tension.
626:  Combining Eqs.~\eqref{eq:d2U} and~\eqref{eq:gamma_bc} yields the relation
627: %
628: %
629: %
630: \[
631: \frac{\partial U_0}{\partial Z}=\frac{\partial\tilde{\Gamma}}{\partial X}+\left(Z-H\right)\bigg\{
632: %
633: %
634: \frac{\partial}{\partial X}\left(P_{\mathrm{surf}}+\Phi_{\mathrm{surf}}\right)+\frac{r}{H}\frac{\partial}{\partial
635: X}\left[H\left(\frac{\partial c_0}{\partial X}\right)^2\right]
636: \bigg\}.
637: %
638: %
639: \]
640: %
641: %
642: %
643: Making
644: use of the BC $U_0=0$ on $Z=0$ and integrating again, we obtain the result
645: %
646: %
647: %
648: \begin{equation}
649: U_0\left(X,Z,T\right)=Z\frac{\partial\tilde{\Gamma}}{\partial X}+\left(\tfrac{1}{2}Z^2-HZ\right)\bigg\{
650: %
651: %
652: \frac{\partial}{\partial X}\left(P_{\mathrm{surf}}+\Phi_{\mathrm{surf}}\right)+\frac{r}{H}\frac{\partial}{\partial
653: X}\left[H\left(\frac{\partial c_0}{\partial X}\right)^2\right]
654: \bigg\}.
655: %
656: %
657: \label{eq:U_0}
658: \end{equation}
659: %
660: %
661: %
662: The vertically-averaged velocity is therefore
663: %
664: %
665: %
666: \[
667: \overline{U_0} = \tfrac{1}{2}H\frac{\partial\tilde{\Gamma}}{\partial X}-\tfrac{1}{3}H^2\bigg\{
668: %
669: %
670: \frac{\partial}{\partial X}\left(P_{\mathrm{surf}}+\Phi_{\mathrm{surf}}\right)+\frac{r}{H}\frac{\partial}{\partial
671: X}\left[H\left(\frac{\partial c_0}{\partial X}\right)^2\right]
672: \bigg\}.
673: %
674: %
675: \]
676: %
677: %
678: %
679: %
680: %
681: Using the standard Laplace--Young free-surface boundary condition, this becomes
682: %
683: %
684: %
685: \[
686: \overline{U_0} = \tfrac{1}{2}H\frac{\partial\tilde{\Gamma}}{\partial X}-\tfrac{1}{3}H^2\bigg\{
687: %
688: %
689: \frac{\partial}{\partial X}\left(-\frac{1}{C}\frac{\partial^2 H}{\partial X^2}+\Phi_{\mathrm{surf}}\right)+\frac{r}{H}\frac{\partial}{\partial
690: X}\left[H\left(\frac{\partial c_0}{\partial X}\right)^2\right]
691: \bigg\},
692: \]
693: %
694: %
695: where
696: %
697: %
698: \begin{equation}
699: C = \frac{\nu\rho D}{h_0\sigma_0\Smalll^2}=O\left(1\right).
700: \label{eq:C}
701: \end{equation}
702: %
703: %
704: Finally, by integrating the continuity equation in the $Z$-direction, we
705: obtain, in a standard way, an equation for free-surface variations,
706: %
707: %
708: \[
709: \frac{\partial H}{\partial X}+\frac{\partial}{\partial X}\left(H\overline{U_0}\right)=0.
710: \]
711: %
712: %
713: %
714: %
715: 
716: 
717: Let us assemble our results, restoring the lower-case fonts and
718: omitting ornamentation over the nondimensional quantities,
719: %
720: %
721: \begin{subequations}
722: \begin{equation}
723: \frac{\partial h}{\partial t}+\frac{\partial J}{\partial x}=0,\\
724: \end{equation}
725: %
726: %
727: \begin{equation}
728: \frac{\partial}{\partial t}\left(c h\right)+\frac{\partial}{\partial x}\left(Jc\right)=\frac{\partial}{\partial{x}}\left(h\frac{\partial\mu}{\partial{x}}\right),
729: \end{equation}
730: %
731: %
732: where
733: \begin{equation}
734: J=\tfrac{1}{2}h^2\frac{\partial\Gamma}{\partial{x}}-\tfrac{1}{3}h^3\bigg\{\frac{\partial}{\partial{x}}\left(-\frac{1}{C}\frac{\partial^2{h}}{\partial{x}^2}
735: +\phi\right)+\frac{r}{h}\frac{\partial}{\partial{x}}\left[h\left(\frac{\partial{c}}{\partial{x}}\right)^2\right]\bigg\},
736: \end{equation}
737: %
738: %
739: %
740: %
741: \begin{equation}
742: \mu=c^3-c-\frac{C_{\mathrm{n}}^2}{h}\frac{\partial}{\partial{x}}\left(h\frac{\partial{c}}{\partial{x}}\right),
743: \end{equation}%
744: \label{eq:analysis_thin_films:model}%
745: \end{subequations}%
746: %
747: %
748: %
749: and where we have the following nondimensional constants,
750: %
751: %
752: %
753: \[
754: r=\frac{\Smalll^2\beta\gamma}{D\nu},\qquad C_{\mathrm{n}}=\frac{\Smalll\sqrt{\gamma}}{h_0},\qquad
755: C=\frac{\nu\rho D}{h_0\sigma_0\Smalll^2}.
756: \]
757: %
758: %
759: %
760: %
761: These are the thin-film Stokes Cahn--Hilliard (SCH) equations.  The integral
762: quantities $M$ and $V$ defined in Eq.~\eqref{eq:mass_consv} are manifestly
763: conserved, while the free surface and concentration are coupled.
764: 
765: We note that the relation $C_{\mathrm{n}}=\Smalll\sqrt{\gamma}/h_0=O\left(1\right)$
766: is the condition that the mean thickness of the film be much smaller than
767: the transition layer thickness.  In experiments involving the smallest film
768: thicknesses attainable ($10^{-8}$ m)~\cite{Sung1996}, this condition is automatically
769: satisfied.
770: %
771: The condition is also realized in ordinary thin films when external effects
772: such as the air-fluid and fluid-substrate interactions do not prefer one
773: binary fluid component or another.  In this case, the vertical extent of the domains
774: becomes comparable to the film thickness at late times, the thin film
775: behaves in a quasi two-dimensional way, and the model equations
776: are applicable.
777: %
778: % However, the model also applies to ordinary thin films whose interactions
779: % with the substrate and surrounding air are non-preferential.  
780: %
781: % For during
782: % the late stages of coarsening in such films, the vertical extent of the
783: % domains is comparable to the film thickness, and the system becomes quasi
784: % two-dimensional.
785: 
786: The choice of potential $\phi$ determines the behaviour of solutions.
787: If interactions between the fluid and the substrate and air interfaces are
788: important, the potential should take account of the Van der Waals forces
789: present.  A simple model potential is thus
790: \[
791: \phi = Ah^{-p},
792: \]
793: where $A$ is the dimensionless Hamakar constant and typically $p=3$~\cite{Oron1997}.
794:  Here $A$ can be positive or negative, with positivity indicating a net attraction
795:  between the fluid and the substrate and negativity indicating a net repulsion.
796:   This choice of potential can also have a regularizing effect, preventing
797:   a singularity or rupture from occurring in Eq.~\eqref{eq:model}.  
798:   
799: For $\phi=-|A|/h^3$ (repulsive Van der Waals interaction), the system of
800: equations~\eqref{eq:model} has a Lyapunov functional $F=F_1+F_2$, where
801: %
802: %
803: \[
804: F_1=\int_{\Omega}{dx}\, \left[\frac{1}{2C}\left(\frac{\partial{h}}{\partial{x}}\right)^2+\tfrac{1}{2}\frac{|A|}{h^2}\right],\qquad
805: %
806: F_2 = \frac{r}{C_{\mathrm{n}}^2}\int_{\Omega}{dx}\,  h\left[\tfrac{1}{4}\left(c^2-1\right)^2+\tfrac{1}{2}{C_{\mathrm{n}}^2}\left(\frac{\partial{c}}{\partial{x}}\right)^2\right],
807: \]  
808: %
809: %
810: and where $\Omega=\left[0,L\right]$ defines the lateral extent of the system.
811:  In this chapter we take $\Omega$ to be a periodic domain, which fixes the
812:  boundary conditions, although other choices of boundary conditions are possible.
813: By differentiating these expressions with respect to time, we obtain the
814: relation
815: % %
816: % %
817: % %
818: % %
819: % \begin{eqnarray*}
820: % %
821: % \dot{F}_2&=&
822: % %
823: % %
824: % %
825: % %\int\frac{\partial h}{\partial t}\left[\tfrac{1}{4}\left(c^2-1\right)^2+\frac{C_{\mathrm{n}}^2}{2}\left(\frac{\partial
826: % %c}{\partial x}\right)^2\right]dx
827: % %
828: % %+\int h\left(c^3-c+C_{\mathrm{n}}^2\frac{\partial c}{\partial x}\frac{\partial}{\partial
829: % %x}\right)\frac{\partial c}{\partial t} dx,\\
830: % %
831: % %
832: % %
833: % %&=&
834: % %-\int\frac{\partial}{\partial x}\left(uh\right)\left[\tfrac{1}{4}\left(c^2-1\right)^2+\frac{C_{\mathrm{n}}^2}{2}\left(\frac{\partial
835: % %c}{\partial x}\right)^2\right]dx
836: % %+\int h\left(c^3-c+C_{\mathrm{n}}^2\frac{\partial c}{\partial x}\frac{\partial}{\partial
837: % %x}\right)\frac{\partial c}{\partial t} dx,\\
838: % %
839: % %
840: % %
841: % %&=&
842: % %\int uh\left[c^3-c+C_{\mathrm{n}}^2\frac{\partial^2
843: % %c}{\partial x^2}\right]\frac{\partial c}{\partial x}dx
844: % %+\int h\left(c^3-c+C_{\mathrm{n}}^2\frac{\partial c}{\partial x}\frac{\partial}{\partial
845: % %x}\right)\frac{\partial c}{\partial t} dx,\\
846: % %
847: % %
848: % %&=&
849: % %\int uh\left[c^3-c+C_{\mathrm{n}}^2\frac{\partial^2
850: % %c}{\partial x^2}\right]\frac{\partial c}{\partial x}dx
851: % %+\int h\left(c^3-c\right)\frac{\partial c}{\partial t}dx+\int C_{\mathrm{n}}^2\frac{\partial
852: % %c}{\partial x}\frac{\partial}{\partial
853: % %x}\frac{\partial c}{\partial t} dx,\\
854: % %
855: % %
856: % %
857: % %&=&\int uh\left[c^3-c+C_{\mathrm{n}}^2\frac{\partial^2
858: % %c}{\partial x^2}\right]\frac{\partial c}{\partial x}dx
859: % %
860: % %+\int h\underbrace{\left[c^3-c-C_{\mathrm{n}}^2\frac{1}{h}\frac{\partial}{\partial
861: % %x}\left(h\frac{\partial c}{\partial x}\right)\right]}_{=\mu}\frac{\partial
862: % %c}{\partial t}dx,\\
863: % %
864: % %
865: % %&=&\int uh\left[c^3-c+C_{\mathrm{n}}^2\frac{\partial^2
866: % %c}{\partial x^2}\right]\frac{\partial c}{\partial x}dx
867: % %+\int h\mu\left[-u\frac{\partial c}{\partial x}+\frac{1}{h}\frac{\partial}{\partial
868: % %x}\left(h\frac{\partial\mu}{\partial x}\right)\right]dx,\\
869: % %
870: % %
871: % %
872: % %&=&
873: % %\int uh\left(c^3-c\right)\frac{\partial c}{\partial x}dx + C_{\mathrm{n}}^2\int
874: % %uh\frac{\partial c}{\partial x}\frac{\partial^2c}{\partial x^2}dx-\int
875: % uh\left(c^3-c\right)\frac{\partial
876: % %c}{\partial x}dx\\
877: % %
878: % %&&+C_{\mathrm{n}}^2\int uh\left[\frac{1}{h}\frac{\partial}{\partial x}\left(\frac{1}{h}\frac{\partial
879: % %c}{\partial x}\right)\right]\frac{\partial c}{\partial x}dx
880: % %
881: % %+\int\mu\frac{\partial}{\partial x}\left(h\frac{\partial\mu}{\partial
882: % x}\right)dx,\\
883: % %
884: % %
885: % %
886: % %&=&C_{\mathrm{n}}^2\int uh\left[2\frac{\partial c}{\partial x}\frac{\partial^2
887: % %c}{\partial x^2}+\frac{1}{h}\frac{\partial h}{\partial x}\left(\frac{\partial
888: % %c}{\partial x}\right)^2\right]dx
889: % %
890: % %-\int h\left(\frac{\partial\mu}{\partial
891: % %x}\right)^2dx,\\
892: % %
893: % %
894: % %
895: % r\int uh\left[\frac{1}{h}\frac{\partial}{\partial x}h\left(\frac{\partial
896: % c}{\partial x}\right)^2\right]dx
897: % %
898: % -\frac{r}{C_{\mathrm{n}}^2}\int h\left(\frac{\partial\mu}{\partial x}\right)^2dx,\\
899: % \end{eqnarray*}
900: % %
901: % %
902: % while
903: % %
904: % %
905: % %
906: % %
907: % \[
908: % \dot{F}_1=-\int\left(\frac{1}{C}\frac{\partial^2 h}{\partial x^2}+\frac{A}{h^3}\right)\frac{\partial
909: % h}{\partial t}dx,
910: % =-\int uh\frac{\partial}{\partial x}\left(\frac{1}{C}\frac{\partial^2 h}{\partial
911: % x^2}+\frac{A}{h^3}\right) dx.
912: %\]
913: %
914: %
915: %Hence,
916: %
917: %
918: \begin{multline}
919: \dot{F}_1+\dot{F}_2
920: %
921: %
922: \\
923: %
924: %
925: =-\tfrac{1}{3}\int_\Omega{dx}\,  h^3\bigg\{\frac{\partial}{\partial x}\left(\frac{1}{C}\frac{\partial^2h}{\partial
926: x^2}+\frac{|A|}{h^3}\right)-\frac{r}{h}\frac{\partial}{\partial x}\left[h\left(\frac{\partial
927: c}{\partial x}\right)^2\right]  \bigg\}^2
928: %
929: -\int_\Omega{dx}\,  h\left(\frac{\partial\mu}{\partial x}\right)^2,
930: %
931: %
932: \label{eq:fe_decay}
933: \end{multline}
934: %
935: %
936: which is nonpositive for nonnegative $h$.  This fact holds the key to
937: the analytic results obtained in the next section.
938: %
939: %
940: %
941: %
942: %
943: %
944: %
945: %
946: %
947: %
948: %
949: %
950: %
951: %
952: %
953: %
954: %
955: %
956: %
957: %
958: %
959: \section{Existence and regularity of solutions}
960: \label{sec:analysis_thin_films:existence}
961: %
962: %
963: In this section we prove that solutions to the model equations do indeed
964: exist.  We focus on the one-dimensional equation set
965: %
966: %
967: %
968: %
969: \[
970: %
971: \frac{\partial h}{\partial t}=
972: %
973: -\frac{\partial}{\partial x}\left[f\left(h\right)\frac{\partial^{3}h}{\partial{x^3}}\right]+
974: %
975: \frac{\partial}{\partial x}\left[\frac{1}{g\left(h\right)}\frac{\partial
976: h}{\partial{x}}\right]+
977: %
978: \frac{\partial}{\partial x}\bigg\{\frac{f\left(h\right)}{g\left(h\right)}\frac{\partial}{\partial{x}}\left[g\left(h\right)\left(\frac{\partial{c}}{\partial{x}}\right)^2\right]\bigg\},\nonumber\\
979: \]
980: %
981: %
982: %
983: \vskip -0.2in
984: \begin{multline}
985: \frac{\partial}{\partial t}\left(c g\left(h\right)\right)=
986: %
987: -\frac{\partial}{\partial x}\left[cf\left(h\right)\frac{\partial^{3}h}{\partial{x^3}}\right]+
988: %
989: \frac{\partial}{\partial x}\left[\frac{c}{g\left(h\right)}\frac{\partial
990: h}{\partial{x}}\right]+
991: %
992: \frac{\partial}{\partial x}\bigg\{c\frac{f\left(h\right)}{g\left(h\right)}\frac{\partial}{\partial{x}}\left[g\left(h\right)\left(\frac{\partial{c}}{\partial{x}}\right)^2\right]\bigg\}
993: \\
994: +\frac{\partial}{\partial x}\bigg\{g\left(h\right)\frac{\partial}{\partial
995: x}\left[c^3-c-\frac{1}{g\left(h\right)}\frac{\partial}{\partial
996: x}\left(g\left(h\right)\frac{\partial c}{\partial x}\right)\right]\bigg\},
997: \label{eq:model_proof}
998: \end{multline}
999: %
1000: %
1001: %
1002: where
1003: %
1004: %
1005: %
1006: \[
1007: f\left(h\right)=h^3,\qquad g\left(h\right)=h.
1008: \]
1009: %
1010: %
1011: %
1012: The application we have in mind involves a periodic domain $\Omega=[0,L]$.
1013:  We shall prove our result for smooth initial conditions,
1014: %
1015: %
1016: %
1017: \begin{multline}
1018: h\left(x,0\right)=h_0\left(x\right)>0,\qquad c\left(x,0\right)=c_0\left(x\right),\qquad
1019: \left(h_0\left(x\right),c_0\left(x\right)\right)\in H^{2,2}\left(\Omega\right).
1020: %c_0\left(x\right)\in H^{2,2}\left(\Omega\right).
1021: \label{eq:initial_data}
1022: \end{multline}
1023: %
1024: %
1025: %
1026: We shall prove that the solutions to this equation pair exist in the strong
1027: sense; however, we shall need the following definition of weak solutions:
1028: %
1029: %
1030: %
1031: \begin{quote}
1032: A pair $\left(h,c\right)$ is a weak solution of Eq.~\eqref{eq:model_proof}
1033: if the following integral relations hold,
1034: %
1035: %
1036: %
1037: \begin{multline*}
1038: \int_0^{T_0}dt\int_\Omega {dx}\, \varphi_t h=
1039: \\
1040: \int_0^{T_0}dt\int_\Omega {dx}\, \varphi_x\bigg\{-f\left(h\right)\frac{\partial^{3}h}{\partial{x^3}}+
1041: %
1042: \frac{1}{g\left(h\right)}\frac{\partial h}{\partial{x}}+
1043: %
1044: \frac{f\left(h\right)}{g\left(h\right)}\frac{\partial}{\partial{x}}\left[g\left(h\right)\frac{\partial}{\partial{x}}\left(\frac{\partial{c}}{\partial{x}}\right)^2\right]\bigg\},
1045: \end{multline*}
1046: %
1047: %
1048: %
1049: \begin{multline}
1050: \int_0^{T_0}dt\int_\Omega {dx}\, \psi_t cg\left(h\right)=
1051: \\
1052: \int_0^{T_0}dt\int_\Omega {dx}\, \psi_x\bigg\{-cf\left(h\right)\frac{\partial^{3}h}{\partial{x^3}}+
1053: %
1054: \frac{c}{g\left(h\right)}\frac{\partial h}{\partial{x}}+
1055: %
1056: %
1057: c\frac{f\left(h\right)}{g\left(h\right)}\frac{\partial}{\partial{x}}\left[g\left(h\right)\frac{\partial}{\partial{x}}\left(\frac{\partial{c}}{\partial{x}}\right)^2\right]\bigg\}
1058: %
1059: \\
1060: %
1061: +\int_0^{T_0}dt\int_\Omega {dx}\, \psi_x \bigg\{g\left(h\right)\frac{\partial}{\partial
1062: x}\left[c^3-c-\frac{1}{g\left(h\right)}\frac{\partial}{\partial{x}}\left(g\left(h\right)\frac{\partial{c}}{\partial{x}}\right)\right]\bigg\},
1063: \label{eq:weak_sln}
1064: \end{multline}
1065: %
1066: %
1067: %
1068: where $T_0>0$ is any time, and $\varphi\left(x,t\right)$ and $\psi\left(x,t\right)$
1069: are arbitrary differentiable test functions that are periodic on $\Omega$
1070: and vanish at $t=0$ and $t=T_0$. 
1071: \end{quote}
1072: %
1073: %
1074: We address the following statement,
1075: %
1076: %
1077: \begin{quote}
1078: \textit{Given the initial data in Eq.~\eqref{eq:initial_data}, the equations~\eqref{eq:model_proof}
1079: possess a unique strong solution endowed with the following regularity properties:
1080: %
1081: %
1082: %
1083: %
1084: \[
1085: \left(h,c\right)\in L^\infty\left(0,T_0,H^{2,2}\left(\Omega\right)\right)
1086: \cap L^2\left(0,T_0;H^{4,2}\left(\Omega\right)\right)
1087: \cap C^{\frac{3}{2},\frac{1}{8}}\left(\Omega\times\left[0,T_0\right]\right).
1088: \]
1089: %
1090: %
1091: }
1092: \end{quote}
1093: %
1094: %
1095: %
1096: Recall that the function $f$ resides in the class $H^{q,p}\left(\Omega\right)$
1097: if
1098: %
1099: %
1100: \[
1101: \Big(\|f\|_p^p+\|\partial_x f\|_p^p+...+\|\partial^q_x f\|_p^p\Big)^{\frac{1}{p}}<\infty.
1102: \]
1103: %
1104: %
1105: %
1106: %
1107: %
1108: The proof is given in the following steps.
1109: 
1110: % the outline of which is as follows:
1111: % in Sec.~\ref{sec:regularization}
1112: % we introduce a regularized version of Eqs.~\eqref{eq:model_proof}, whose
1113: % solution we approximate by a Galerkin sum in Sec.~\ref{sec:galerkin}. 
1114: %  In
1115: % Sec.~\ref{sec:a_priori_bds}, we obtain \emph{a priori} bounds on various
1116: % norms of the approximate solution.  Crucially, we show that the free-surface
1117: % height is
1118: % always positive.  This enables us to continue the approximate solution
1119: % in
1120: % the time interval $\left[0,T_0\right]$.  In Secs.~\ref{sec:equicontinuity}
1121: % and~\ref{sec:convergence} we show
1122: % that the Galerkin sum converges to a solution of the unapproximated equations,
1123: % in the appropriate limit.  Finally, in Secs.~\ref{sec:regularity} and~\ref{sec:uniqueness}
1124: % we discuss the regularity and uniqueness properties of the solution.
1125: 
1126: \subsection{Regularization of the problem}
1127: \label{sec:regularization}
1128: 
1129: \noindent We introduce regularized functions $f_\varepsilon\left(s\right)$
1130: and $g_\varepsilon\left(s\right)$ such that 
1131: %
1132: %
1133: %
1134: \[
1135: \lim_{\varepsilon\rightarrow0} f_\varepsilon\left(s\right)= f\left(s\right),\qquad
1136: \lim_{\varepsilon\rightarrow0,s\geq0} g_\varepsilon\left(s\right)={g}\left(s\right).
1137: \]
1138: %
1139: %
1140: %
1141: For now we do not specify $f_\varepsilon\left(s\right)$,
1142: although we mention that a suitable choice of $f_\varepsilon\left(s\right)$
1143: will cure the degeneracy of the fourth-order term in the height equation.
1144:  On the other hand, we require that the function $g_\varepsilon\left(s\right)$
1145:  have the following properties:
1146: %
1147: \begin{itemize}
1148: \item $g_\varepsilon\left(s\right)=s+\varepsilon$, for $s\geq0$,
1149: \item $g_\varepsilon\left(s\right)>0$, for $s<0$,
1150: \item $\lim_{s\rightarrow-\infty}g_\varepsilon\left(s\right)=\tfrac{1}{2}\varepsilon$,
1151: \item $g_\varepsilon\left(s\right)$ has as many derivatives as necessary.
1152: \end{itemize}
1153: %
1154: One way of obtaining such a function is to define 
1155: %
1156: %
1157: %
1158: \begin{equation}
1159: g_\varepsilon\left(s\right)=\Bigg\{\begin{array}{cc}
1160: \frac{\varepsilon}{2}, &s\leq-\varepsilon, \\
1161: s+\frac{s^2}{2\varepsilon}-\frac{\varepsilon}{4\pi^2}\cos\left(\frac{2\pi{s}}{\varepsilon}+\pi\right)+\varepsilon-\frac{\varepsilon}{4\pi^2},
1162: & -\varepsilon\leq s\leq 0,\\
1163: \varepsilon+s, & s\geq 0,
1164: \end{array}
1165: \label{eq:regularized_g}
1166: \end{equation}
1167: %
1168: %
1169: %
1170: which is $C^{3}$ in the variable $s$, and has Lipschitz third derivative; inspection of Eq.~\eqref{eq:model_proof}
1171: shows that this degree of smoothness is sufficient to regularize the equations.
1172:  A sketch of this regularization is shown in Fig.~\ref{fig:regularized_g}.
1173: %
1174: %
1175: %
1176: %
1177: \begin{figure}
1178: \centering
1179: \includegraphics[width=0.35\textwidth]{regularized_g}
1180: \caption{A sketch of the regularized function $g_\varepsilon\left(s\right)$
1181: obtained in Eq.~\eqref{eq:regularized_g}.}
1182: \label{fig:regularized_g}
1183: \end{figure}
1184: 
1185: The regularized PDEs we study are thus
1186: %
1187: %
1188: \begin{eqnarray}
1189: h_t&=&-J_{\varepsilon,x},\qquad 
1190: J_\varepsilon=f_\varepsilon\left(h\right)h_{xxx}-\frac{1}{g_\varepsilon\left(h\right)}h_x-\frac{f_\varepsilon\left(h\right)}{g_\varepsilon\left(h\right)}\left(g_\varepsilon\left(h\right)c_{x}^2\right)_x,
1191: \nonumber\\
1192: %
1193: %
1194: \left(cg_\varepsilon\left(h\right)\right)_t&=&-\left(c J_\varepsilon\right)_x-\left(g_\varepsilon\left(h\right)\mu_{\varepsilon,x}\right)_x,
1195: \label{eq:reg_pde}
1196: \end{eqnarray}
1197: %
1198: %
1199: %
1200: where
1201: %
1202: %
1203: \[
1204: \mu_\varepsilon=c^3-c-\frac{1}{g_\varepsilon\left(h\right)}\left(g_\varepsilon\left(h\right)c_x\right)_x.
1205: \]
1206: %
1207: %
1208: The $c$-equation can also be written as
1209: %
1210: %
1211: \[
1212: c_t=-\frac{1}{g_\varepsilon\left(h\right)}J_\varepsilon c_x+\frac{1}{g_\varepsilon\left(h\right)}\left(g_\varepsilon\left(h\right)\mu_{\varepsilon,x}\right)_x-\frac{c}{g_\varepsilon\left(h\right)}\left[J_{\varepsilon,x}+g_\varepsilon'\left(h\right)h_t\right].
1213: \]
1214: %
1215: %
1216: 
1217: \subsection{The Galerkin approximation}
1218: \label{sec:galerkin}
1219: 
1220: \noindent We choose a complete orthonormal basis on the interval $\Omega$,
1221: with periodic boundary conditions.  Let us denote the basis by $\{\phi_i\left(x\right)\}_{i\in\mathbb{N}_0}$.
1222:  We consider the finite-dimensional vector space $\text{span}\{\phi_0,...\phi_\ord\}$.
1223:   For convenience, let us take the $\phi_i\left(x\right)$'s to be
1224:   the
1225:   eigenfunctions of the Laplacian on $\left[0,L\right]$ with periodic boundary
1226:   conditions, with corresponding eigenvalues $-\lambda_i^2$.  Moreover, let
1227:   $\phi_0$ be the constant eigenfunction.  We construct approximate solutions
1228:   to the PDEs~\eqref{eq:reg_pde} as finite sums,
1229: %
1230: %
1231: \[
1232: h_\ord\left(x,t\right)=\sum_{i=0}^\ord\eta_{\ord,i}\left(t\right)\phi_i\left(x\right),
1233: \qquad c_\ord\left(x,t\right)=\sum_{i=0}^\ord\gamma_{\ord,i}\left(t\right)\phi_i\left(x\right).
1234: \]
1235: %
1236: %
1237: If the (smooth) initial data are given as
1238: %
1239: %
1240: \[
1241: h\left(x,0\right)=h_0\left(x\right)=\sum_{i=0}^\infty\eta_i^0\phi_i\left(x\right)>0,\qquad
1242: c\left(x,0\right)=c_0\left(x\right)=\sum_{i=0}^\infty\gamma_i^0\phi_i\left(x\right),
1243: \]
1244: %
1245: %
1246: %
1247: then the initial data for the Galerkin approximation are
1248: %
1249: %
1250: \[
1251: h_\ord\left(x,0\right)=h_\ord^0\left(x\right)=\sum_{i=0}^\ord\eta_{i}^0\phi_i\left(x\right),\qquad
1252: c_\ord\left(x,0\right)=c_\ord^0\left(x\right)=\sum_{i=0}^\ord\gamma_{i}^0\phi_i\left(x\right),
1253: \]
1254: %
1255: %
1256: %
1257: and the initial data of the Galerkin approximation converge strongly in the
1258: $L^2\left(\Omega\right)$ norm to the initial data of the unapproximated problem.
1259:  Thus, there is a $\ord_0\in\mathbb{N}$ such that $h_\ord^0\left(x\right)>0$,
1260:  everywhere in $\Omega$, for all $\ord>\ord_0$.  Henceforth we work with
1261:  Galerkin approximations with $\ord>\ord_0$.
1262:  
1263: Substitution of $h_\ord=\sum_{i=0}^\ord\eta_{\ord,i}\phi_i$  into a weak
1264: form of the $h$-equation yields
1265: %
1266: %
1267: \begin{equation}
1268: \frac{d}{dt}\left(\left(h_\ord,\phi_j\right)\right)=\left(\left(J_{\varepsilon,\ord},\phi_{j,x}\right)\right),
1269: \label{eq:weak_h}
1270: \end{equation}
1271: %
1272: %
1273: where
1274: %
1275: %
1276: \[
1277: J_\varepsilon\left(h_\ord,c_\ord\right)=f_\varepsilon\left(h_\ord\right)h_{\ord,xxx}-\frac{1}{g_\varepsilon\left(h_\ord\right)}h_{\ord,x}-\frac{f_\varepsilon\left(h_\ord\right)}{g_\varepsilon\left(h_\ord\right)}\left(g_\varepsilon\left(h_{\ord}\right)c_{\ord,x}^2\right)_x,
1278: \]
1279: %
1280: %
1281: is the flux for the regularized $h$-equation, and in this chapter, $\left(\left(\phi,\psi\right)\right)$
1282: denotes the pairing $\int_\Omega\phi^{*}\psi{dx}$.
1283: %
1284: %
1285: This is recast as
1286: %
1287: %
1288: \[
1289: \frac{d\eta_{\ord,j}}{dt}=\left(\left(J_{\varepsilon,\ord},\phi_{j,x}\right)\right)=\Phi_{\ord,j}\left(\eta_\ord,\gamma_\ord\right),
1290: \]
1291: %
1292: %
1293: %
1294: where the function $\Phi_\ord\left(\eta_\ord,\gamma_\ord\right)$ depends
1295: on
1296: %
1297: %
1298: \[
1299: \eta_\ord=\left(\eta_{\ord,0},...,\eta_{\ord,\ord}\right),\qquad
1300: \gamma_\ord=\left(\gamma_{\ord,0},...,\gamma_{\ord,\ord}\right),
1301: \]
1302: %
1303: %
1304: %
1305: and is locally Lipschitz in its variables.  This Lipschitz property arises
1306: from the fact that the regularized flux, evaluated at the Galerkin approximation,
1307: is a composition of Lipschitz continuous functions, and therefore, is itself
1308: Lipschitz continuous.
1309: 
1310: % The function $\Phi_\ord\left(\eta_\ord,\gamma_\ord\right)$ depends on $\eta_\ord=\left(\eta_{\ord,0},...,\eta_{\ord,\ord}\right)$
1311: % and $\gamma_\ord=\left(\gamma_{\ord,0},...,\gamma_{\ord,\ord}\right)$ and
1312: % is a combination of functions that are (locally) Lipschitz in $\gamma_\ord$
1313: % and $\eta_\ord$;
1314: % inspection of the definition of the regularized flux $J_{\varepsilon,\ord}$
1315: % shows that $\Phi_\ord\left(\eta_\ord,\gamma_\ord\right)$ is Lipschitz.
1316: 
1317: Similarly, substitution of $c_\ord=\sum_{i=0}^\ord\gamma_{\ord,i}\phi_i$
1318: into the weak form of the $c$-equation yields
1319: %
1320: %
1321: \begin{equation}
1322: \frac{d}{dt}\left(\left(g\left(h_\ord\right)c_\ord,\phi_j\right)\right)=\left(\left(K_{\varepsilon,\ord},\phi_{j,x}\right)\right),
1323: \label{eq:weak_c}
1324: \end{equation}
1325: %
1326: %
1327: where
1328: %
1329: %
1330: \begin{eqnarray*}
1331: K_\varepsilon\left(h_\ord,c_\ord\right)&=&c_{\ord}f_\varepsilon\left(h_\ord\right)h_{\ord,xxx}-\frac{c_\ord}{g_\varepsilon\left(h_\ord\right)}h_{\ord,x}\\
1332: %
1333: %
1334: &\phantom{a}&\phantom{aaaaaaaaaaaaaaaaaa}
1335: -c_{\ord}\frac{f_\varepsilon\left(h_{\ord}\right)}{g_\varepsilon\left(h_\ord\right)}\left(g_\varepsilon\left(h_{\ord}\right)c_{\ord,x}^2\right)_x-g\left(h_\ord\right)\mu_{\varepsilon,\ord,x},\\
1336: &=&c_{\ord}J_{\varepsilon,\ord}-g_\varepsilon\left(h_\ord\right)\mu_{\varepsilon,\ord,x},
1337: \end{eqnarray*}
1338: %
1339: %
1340: is the flux for the regularized $c$-equation.   Rearranging gives
1341: %
1342: %
1343: \[
1344: \left(\left(g\left(h_\ord\right)c_{\ord,t},\phi_j\right)\right)=\left(\left(K_{\varepsilon,\ord},\phi_{j,x}\right)\right)-\left(\left(g'\left(h_\ord\right)c_{\ord}h_{\ord,t},\phi_j\right)\right),
1345: \]
1346: %
1347: %
1348: and the left-hand side can be recast in matrix form as $\sum_{\alpha=0}^{\ord}M_{\alpha{j}}\dot{\gamma}_{\ord,\alpha}$,
1349: where
1350: %
1351: %
1352: \[
1353: M_{\alpha j}=\int_\Omega{dx}\,  g_\varepsilon\left(\sum_i \eta_{\ord,i}\phi_i\right)\phi_\alpha\phi_j,
1354: \] 
1355: %
1356: %
1357: which is manifestly symmetric.  It is positive definite because given a nonzero vector
1358: $\left(\xi_0,...,\xi_\ord\right)$, we have the relation
1359: %
1360: %
1361: \begin{eqnarray*}
1362: \sum_{\alpha,j}\xi_\alpha M_{\alpha j}\xi_j&=&
1363: \int_\Omega {dx}\, g_\varepsilon\left(\sum_i\eta_{\ord,i}\phi_i\right)\sum_{\alpha{j}}\left(\phi_\alpha\xi_\alpha\right)\left(\phi_j\xi_j\right)\\
1364: &=&\int_\Omega {dx}\, g_\varepsilon\left(\sum_i\eta_{\ord,i}\phi_i\right)\left(\sum_{j}\phi_j\xi_j\right)^2\\
1365: &>&0\qquad\text{for }\left(\xi_0,...,\xi_\ord\right)\neq0,
1366: \end{eqnarray*}
1367: %
1368: %
1369: which follows from the positivity of the regularized function $g_\varepsilon\left(s\right)$.
1370:  We therefore have the following equation for $\gamma_{\ord,i}\left(t\right)$,
1371: %
1372: %
1373: %
1374: \begin{equation}
1375: \frac{d\gamma_{\ord,j}}{dt}=\sum_{\alpha=0}^{\ord} M_{\alpha j}^{-1}\Big[\left(\left(K_{\varepsilon,\ord},\phi_{\alpha,x}\right)\right)-\left(\left(g_\varepsilon'\left(h_\ord\right)c_{n}h_{\ord,t},\phi_\alpha\right)\right)
1376: \Big].
1377: \label{eq:gamma_dot}
1378: \end{equation}
1379: %
1380: Inspecting the expression for $M_{\alpha j}$, $g\left(h_\ord\right)$, and
1381: $K_{\varepsilon,\ord}$, we see that $\eta_\ord$ and $\gamma_\ord$ appear
1382: in a (locally) Lipschitz-continuous way in the
1383: expression $\sum_{\alpha=0}^n M_{\alpha j}^{-1}\left(\left(K_{\varepsilon,\ord},\phi_{\alpha,x}\right)\right)$,
1384: while owing to the imposed smoothness of $g_\varepsilon\left(s\right)$, the
1385: variables $\eta_\ord$, $\gamma_\ord$ and $\dot{\eta}_{\ord}=\left(\dot{\eta}_{\ord,0},...,\dot{\eta}_{\ord,\ord}\right)$
1386: appear in a Lipschitz-continuous way in the quantity 
1387: %
1388: %
1389: \[
1390: \sum_{\alpha=0}^{\ord}M_{\alpha j}^{-1}\left(\left(g_\varepsilon'\left(h_\ord\right)c_{\ord}h_{\ord,t},\phi_\alpha\right)\right).
1391: \]
1392: %
1393: %
1394: %
1395: The vector $\dot{\eta}_\ord$ can be replaced by the function $\Phi_\ord\left(\eta_\ord,\gamma_\ord\right)$
1396: and thus we obtain a relation
1397: %
1398: %
1399: \[
1400: \frac{d\gamma_{\ord,j}}{dt}=\Psi_{\ord,j}\left(\eta_\ord,\gamma_\ord\right),
1401: \]
1402: %
1403: %
1404: %
1405:  in place of Eq.~\eqref{eq:gamma_dot}, where $\Psi_{\ord,j}\left(\eta_\ord,\gamma_\ord\right)$
1406:  is Lipschitz.  We therefore have a system of Lipschitz-continuous equations
1407: %
1408: %
1409: \[
1410: \frac{d\eta_{\ord,j}}{dt}=\Phi_{\ord,j}\left(\eta_\ord,\gamma_\ord\right),\qquad\frac{d\gamma_{\ord,j}}{dt}=\Psi_{\ord,j}\left(\eta_\ord,\gamma_\ord\right),
1411: \]
1412: %
1413: %
1414: and thus local existence theory guarantees a solution for the $\eta_{\ord,i}$'s
1415: and $\gamma_{\ord,i}$'s for all times $t$ in a finite interval $0<t<\sigma$.
1416:  This solution is, moreover, unique and continuous.
1417: 
1418: 
1419: \subsection{\emph{A priori} bounds on the Galerkin approximation}
1420: \label{sec:a_priori_bds}
1421: 
1422: \noindent We identify the free energy
1423: %
1424: %
1425: \[
1426: F=\int_{\Omega}{dx}\, \left[\tfrac{3}{2}h_x^2+G_\varepsilon\left(h\right)\right]
1427: +\int_{\Omega}{dx}\, g_\varepsilon\left(h\right)\left[\tfrac{1}{4}\left(c^2-1\right)^2+\tfrac{1}{2}c_x^2\right],
1428: \]
1429: %
1430: %
1431: where $G_{\varepsilon}''\left(s\right)=1/\left[f_\varepsilon\left(s\right)g_\varepsilon\left(s\right)\right]$.
1432:  Since the Galerkin approximation satisfies the weak form of the PDEs given
1433:  in Eqs.~\eqref{eq:weak_h} and~\eqref{eq:weak_c}, it is possible to obtain
1434:  the following free-energy decay law,
1435: %
1436: %
1437: %
1438: %\begin{small}
1439: \begin{multline}
1440: \frac{dF}{dt}\left(t\right)=
1441: \\
1442: -\int_{\Omega-\Omega_-}{dx}\,  f_\varepsilon\left(h_\ord\right)\left[-h_{\ord,xxx}+\frac{h_{\ord,x}}{f_\varepsilon\left(h_\ord\right)g_\varepsilon\left(h_\ord\right)}+\frac{1}{g_\varepsilon\left(h_\ord\right)}\left(g_\varepsilon\left(h_\ord\right)c_{\ord,x}^2\right)_x\right]^2
1443: %
1444: %
1445: %
1446: \\
1447: %
1448: %
1449: -\int_\Omega{dx}\,  g_\varepsilon\left(h_\ord\right)\mu_{\varepsilon,\ord,x}^2
1450: %
1451: %
1452: +\int_{\Omega_-}{dx}\, J_{\varepsilon,\ord}\left[-h_{\ord,xxx}+\frac{h_{\ord,x}}{f_\varepsilon\left(h_\ord\right)g_\varepsilon\left(h_\ord\right)}\right]
1453: %
1454: %
1455: %
1456: \\
1457: +\int_{\Omega_-}{dx}\, \Big\{g_\varepsilon'\left(h_\ord\right)J_{\varepsilon,\ord}\left(c^3-c-c_{\ord,xx}\right)
1458: %
1459: %
1460: %
1461: %-g_\varepsilon\left(h_\ord\right)\mu_{\varepsilon,\ord}\frac{c_\ord}{g_\varepsilon\left(h_\ord\right)}
1462: -\mu_{\varepsilon,\ord}c_\ord
1463: %
1464: %
1465: \left[g_\varepsilon'\left(h_\ord\right)+J_{\varepsilon,\ord,x}\right]-J_{\varepsilon,\ord}\mu_{\varepsilon,\ord}c_{\ord,x}\Big\},
1466: \\
1467: 0\leq t<\sigma,
1468: \label{eq:free_energy_decay}
1469: \end{multline}
1470: %\end{small}
1471: %
1472: %
1473: %
1474: where $\Omega_-\left(t\right)=\{x\in\Omega|h_\ord\left(x,t\right)<0\}$. 
1475: %
1476: %
1477: Now given the time continuity of $h_\ord\left(x,t\right)$ in $(0,\sigma)$,
1478: and the initial condition $h_\ord^0\left(x\right)>0$ (since $\ord>\ord_0$),
1479: there is a time $\sigma_1$, such that $0<\sigma_1\leq\sigma$, and such that
1480: $h_\ord\left(x,t\right)>0$ for all $x\in\Omega$ and all $t\in\left(0,\sigma_1\right)$.
1481: %
1482: % 
1483: Therefore, $\Omega_-\left(t\right)=\emptyset$ for $t\in\left(0,\sigma_1\right)$,
1484: and hence
1485: %
1486: %
1487: \[
1488: F\left[c_\ord\left(x,t\right),h_\ord\left(x,t\right)\right]\leq F\left[c_\ord\left(x,0\right),h_\ord\left(x,0\right)\right]\leq\sup_{\varepsilon,\ord\in\left[0,\infty\right)}F\left[c_\ord\left(x,0\right),h_\ord\left(x,0\right)\right]<\infty,
1489: \]
1490: %
1491: %
1492: for $0< t<\sigma_1$.  Consequently, we obtain the bound $\|h_{\ord,x}\|_2\leq
1493: k_1$, where $0<t<\sigma_1$, and where $k_1$ depends only on the initial
1494: conditions.  We have Poincar\'e's inequality for $h_{\ord,x}$,
1495: %
1496: %
1497: %
1498: \[
1499: \|h_\ord\|_2^2\leq\left[\int_\Omega{dx}\,  h_\ord\left(x\right)\right]^2+\left(\frac{L}{2\pi}\right)^2\|h_{\ord,x}\|_2^2.
1500: \]
1501: %
1502: %
1503: %
1504: Now $\int_\Omega{dx}\,h_\ord\left(x,t\right)=L\eta_{\ord,0}\left(t\right)$.
1505:  Inspection of Eq.~\eqref{eq:weak_h} shows that $\eta_{\ord,0}\left(t\right)=\eta_{\ord,0}\left(0\right)=\eta_{0}^0$.
1506:   Thus,
1507: %
1508: %
1509: \[
1510: \|h_\ord\|_2^2\leq L^2|\eta_{0}^0|^2+\left(\frac{L}{2\pi}\right)^2k_1\equiv
1511: k_2<\infty.
1512: \] 
1513: %
1514: %
1515: Using the following inequality from Sec.~\ref{sec:background:inequalities},
1516: %
1517: %
1518: \[
1519: \|\varphi\|_\infty\leq \frac{1}{L}\|\varphi\|_1 + \|\varphi_x\|_1\leq\frac{1}{\sqrt{L}}\|\varphi\|_2+\sqrt{L}\|\varphi_x\|_2,
1520: \]
1521: %
1522: %
1523: we obtain the bound
1524: %
1525: %
1526: \[
1527: \|h_\ord\|_\infty\leq\frac{1}{\sqrt{L}}\|h_\ord\|_2+\sqrt{L}\|h_{\ord,x}\|_2,
1528: \]
1529: %
1530: %
1531: and hence $\|h_{\ord}\|_\infty\leq k_3$.
1532: %
1533: %
1534: %
1535: Additionally, the following properties hold:
1536: %
1537: %
1538: %
1539: \begin{itemize}
1540: \item The function $h_{\ord}$ is H\"older continuous in space, with exponent
1541: $\tfrac{1}{2}$,
1542: \item $\int_\Omega{dx}\,  G_{\varepsilon}\left(h_\ord\right)\leq k_4$,
1543: \end{itemize}
1544: %
1545: %
1546: %
1547: where these results hold in $0<t<\sigma_1$, and where the constants $k_1$,
1548: $k_2$, $k_3$, and $k_4$ are independent of $\varepsilon$, $\ord$, $\sigma$,
1549: and $\sigma_1$, and in fact depend only on the functions $h_0\left(x\right)$
1550: and $c_0\left(x\right)$.
1551: 
1552: In order to continue the estimates to the whole interval $\left(0,\sigma\right)$,
1553: we need to prove that $h_\ord\left(\cdot,\sigma_1\right)>0$ almost everywhere
1554: (a.e.).  If this is
1555: true, there is a new interval $\left[\sigma_1,\sigma_2\right)$, $\sigma_1<\sigma_2\leq\sigma$,
1556: on which $h_\ord\left(\cdot,t\right)>0$ a.e., and we can then provide \emph{a
1557: priori} bounds on $h_\ord\left(\cdot,t\right)$ and $c_\ord\left(\cdot,t\right)$
1558: on
1559: the interval $\left[\sigma_1,\sigma_2\right)$.  It is then possible to show
1560: that $h_\ord\left(\cdot.,\sigma_2\right)>0$ a.e. and thus, by iteration,
1561: we extend
1562: the proof to the whole interval $\left(0,\sigma\right)$, and find that $h_\ord\left(.,t\right)>0$
1563: a.e. on $\left(0,\sigma\right)$.
1564: 
1565: We have the bound
1566: %
1567: %
1568: \begin{equation}
1569: \int_\Omega{dx}\,  G_{\varepsilon}\left(h_\ord\left(\cdot,t\right)\right)\leq
1570: k_4,
1571: \label{eq:sigma_1_bound}
1572: \end{equation}
1573: %
1574: %
1575: where $k_4$ depends only on the initial conditions,
1576: and where $0<t<\sigma_1$.  We now specify $G_\varepsilon\left(s\right)$,
1577: in more detail.  This function satisfies the condition
1578: %
1579: %
1580: \[
1581: G_\varepsilon''\left(s\right)=\frac{1}{f_\varepsilon\left(s\right)g_\varepsilon\left(s\right)}.
1582: \]
1583: %
1584: %
1585: We take $g_\varepsilon\left(s\right)$ to be as defined previously, while
1586: $f_\varepsilon\left(s\right)$ can be regularized as $g_\varepsilon\left(s\right)^3$,
1587: which is Lipschitz continuous.
1588: %
1589: %
1590: By defining
1591: %
1592: %
1593: \[
1594: \tilde{G}_\varepsilon\left(s\right)=-\int_s^\infty\frac{dr}{f_\varepsilon\left(r\right)g_\varepsilon\left(r\right)},\qquad
1595: G_\varepsilon\left(s\right)=-\int_s^{\infty}{dr}\, \tilde{G}_\varepsilon\left(r\right),
1596: \]
1597: %
1598: %
1599: we obtain a function $G_\varepsilon\left(s\right)$ that is positive for all
1600: $s\in\left(-\infty,\infty\right)$, and
1601: %
1602: %
1603: %\begin{multline*}
1604: %G_\varepsilon\left(s=0,\varepsilon\right)=\tfrac{5}{6}\frac{\log\varepsilon}{1-\varepsilon^2}-\frac{1}{1-\varepsilon^2}\left[\tfrac{1}{4}\log\tfrac{4}{3}+1\right]+\frac{\varepsilon^{\frac{2}{3}}}{1-\varepsilon^2}\left[\tfrac{1}{4}\log\tfrac{4}{3}+\tfrac{1}{3}-\tfrac{\sqrt{3}\pi}{18}\right]-\tfrac{1}{6}\frac{\varepsilon^{\frac{2}{3}}\log\varepsilon}{1-\varepsilon^2}\\
1605: %\sim\tfrac{5}{6}\log\varepsilon,\qquad\text{as }\varepsilon\rightarrow0.
1606: %\end{multline*}
1607: %
1608: %
1609: %
1610: %
1611: \[
1612: G_{\varepsilon}\left(s\right)=\tfrac{1}{6}\frac{1}{\left(s+\varepsilon\right)^2},\qquad
1613: s\geq0.
1614: \]
1615: %
1616: %
1617: %
1618: Using the boundedness of $G_{\varepsilon}\left(s\right)$,
1619: and the time continuity of $h_n\left(\cdot,t\right)$, we employ the Dominated
1620: Convergence Theorem,
1621: %
1622: %
1623: \[
1624: \lim_{t\rightarrow\sigma_1}\int_\Omega {dx}\, G_{\varepsilon}\left(h_\ord\left(\cdot,t\right)\right)=
1625: \int_\Omega {dx}\, \lim_{t\rightarrow\sigma_1}G_{\varepsilon}\left(h_\ord\left(\cdot,t\right)\right)=
1626: \int_{\Omega}{dx}\, G_{\varepsilon}\left(h_\ord\left(\cdot,\sigma_1\right)\right)
1627: \leq k_4.
1628: \] 
1629: %
1630: %
1631: %
1632: %
1633: Similarly, since the constant $k_1$ in the inequality $\|h_{\ord,x}\|_2\leq
1634: k_1$, $0\leq t<\sigma_1$ depends
1635: only on the initial data, we extend this last inequality to $t=\sigma_1$,
1636: and thus $h_\ord\left(x,\sigma_1\right)$ is H\"older continuous in space.
1637: %
1638: %
1639: 
1640: % Note that the last limit definitely holds if $\sigma_1<\sigma$.  If $\sigma_1=\sigma$,
1641: % we assume that $h$ touches down to zero at time $\sigma$.  Then $\Omega_-\left(\sigma\right)$
1642: % is empty and the free energy is decreasing.  Thus, there is a bound on
1643: % $\|h_\ord\|_2^2$ etc., and this bound is independent of time.  It can therefore
1644: % be extended to $t=\sigma$, and the Galerkin vectors $\eta$ and $\gamma$
1645: % can be continuously extended to $t=\sigma$.
1646: 
1647: In the worst-case scenario, the time $\sigma_1$ is the first time at which
1648: $h_\ord\left(x,t\right)$
1649: touches down to zero, and and we therefore assume for contradiction that
1650: $h_\ord\left(x_0,\sigma_1\right)=0$,
1651: and that $h_\ord\left(x,\sigma_1\right)\geq0$ elsewhere.
1652:  Then by the H\"older continuity, for any $x\in\Omega$ we have the bound
1653:  $0\leq h_\ord\left(x,\sigma_1\right)\leq k_1\left|x-x_0\right|^{\frac{1}{2}}$,
1654:  and thus
1655: %
1656: %
1657: %
1658: \begin{multline*}
1659: \int_\Omega {dx}\, G_{\varepsilon}\left(h_\ord\left(\cdot,\sigma_1\right)\right)=\tfrac{1}{6}\int_\Omega\frac{dx}{\left[h_\ord\left(x,\sigma_1\right)+\varepsilon\right]^2}
1660: \\
1661: \geq\frac{k_1}{6}\int_0^L\frac{dx}{\left[|x-x_0|^{\frac{1}{2}}+\varepsilon\right]^2}\geq\frac{k_1}{6}\int_0^L\frac{dx}{|x-x_0|+\varepsilon\left(2\sqrt{L}+\varepsilon\right)}.
1662: \end{multline*}
1663: %
1664: %
1665: Hence,
1666: %
1667: %
1668: %
1669: \begin{multline*}
1670: \frac{6}{k_1}\int_\Omega{dx}\, G_{\varepsilon}\left(h_\ord\left(\cdot,\sigma_1\right)\right)
1671: \\
1672: \geq-2\log\left[\varepsilon\left(2\sqrt{L}+\varepsilon\right)\right]+\log\bigg\{\left[L-x_0+\left(2\sqrt{L}+\varepsilon\right)\varepsilon\right]\left[x_0+\left(2\sqrt{L}+\varepsilon\right)\varepsilon\right]\bigg\}.
1673: \end{multline*}
1674: %
1675: %
1676: %
1677: %
1678: Thus, the integral $\int_\Omega G_\varepsilon\left(h_\ord\left(x,\sigma_1\right)\right)dx$
1679: can be made arbitrarily large, which contradicts the $\varepsilon$-independent
1680: bound for this quantity, obtained in Eq.~\eqref{eq:sigma_1_bound}.
1681: We therefore have the strong condition that set on which $h_\ord\left(\cdot,\sigma_1\right)\leq0$
1682: is empty.  Iterating the argument, we have the following important result:
1683: %
1684: %
1685: %
1686: %
1687: \begin{quote}
1688: The set on which $h_\ord\left(\cdot,t\right)\leq0$ is empty, for $0<t<\sigma$.
1689: \end{quote}
1690: %
1691: %
1692: %
1693: Using the same argument, we have an estimate on the minimum value of $h_\ord\left(x,t\right)$,
1694: %
1695: %
1696: \[
1697: h_{\mathrm{min}}=\min_{x\in\Omega,t\in\left(0,\sigma\right]}h_\ord\left(x,t\right),
1698: \]
1699: %
1700: %
1701: namely,
1702: %
1703: %
1704: %
1705: \[
1706: h_{\mathrm{min}}+\varepsilon\geq -k_1\sqrt{L}+\sqrt{k_1^2L+\frac{k_1^2L}{e^{k_4k_1^2}-1}},
1707: \]
1708: %
1709: %
1710: % By taking $\varepsilon\leq\tfrac{1}{2}\left[-k_1L+\sqrt{k_1^2L^2+Lk_1^2\left(e^{k_4k_h^2}-1\right)^{-1}}\right]$,
1711: % we obtain the $\varepsilon$-independent bound
1712: % 
1713: % 
1714: for all small positive $\varepsilon$.  Thus,
1715: %
1716: %
1717: %
1718: \begin{equation}
1719: h_{\mathrm{min}}\geq B :=-k_1\sqrt{L}+\sqrt{k_1^2L+\frac{k_1^2L}{e^{k_4k_1^2}-1}},
1720: \label{eq:min_h}
1721: \end{equation}
1722: %
1723: %
1724: a result that depends only on the initial data $c_0\left(x\right)$ and $h_0\left(x\right)$.
1725: Now, using Eq.~\eqref{eq:min_h} and the boundedness result 
1726: %
1727: %
1728: \[
1729: \int_\Omega{dx}\, g_\varepsilon\left(h_\ord\right)\left[\tfrac{1}{4}\left(c_\ord^2-1\right)^2+\tfrac{1}{2}c_{\ord,x}^2\right]\leq
1730: k_5
1731: \]
1732: %
1733: %
1734: %
1735: where $k_5$ depends only on the initial data, we obtain \emph{a priori}
1736: bounds on $\|c_{\ord,x}\|_2^2$,
1737: %
1738: %
1739: \[
1740: \int_\Omega{dx}\, c_{\ord,x}^2 \leq \frac{2k_5}{B},
1741: \]
1742: %
1743: %
1744: It is also possible to derive a bound on $\|c_\ord\|_2^2$.  We have the relation
1745: %
1746: %
1747: \[
1748: \int_\Omega{dx}\,\tfrac{1}{4}\left(c_\ord^2-1\right)^2\leq \frac{k_5}{B},
1749: \]
1750: %
1751: %
1752: which gives the inequality
1753: %
1754: %
1755: $
1756: \|c_\ord\|_4^4\leq 2\|c_\ord\|_2^2+\left({4k_5}/{B}\right).
1757: $
1758: %
1759: %
1760: Using the H\"older relation $\|c_\ord\|_2\leq |\Omega|^{\frac{1}{4}}\|c_\ord\|_4$,
1761: we obtain a quadratic inequality in $\|c_\ord\|_2^2$,
1762: %
1763: %
1764: %
1765: \[
1766: \|c_\ord\|_2^4\leq 2|\Omega|\|c_\ord\|_2^2+\frac{4|\Omega|k_5}{B},
1767: \]
1768: %
1769: %
1770: %
1771: with solution
1772: %
1773: %
1774: \[
1775: \|c_\ord\|_2^2\leq |\Omega|+\frac{4|\Omega|k_5}{B},
1776: \]
1777: %
1778: %
1779: as required.  From the boundedness of $\|c_{\ord,x}\|_2$ and $\|c_{\ord}\|_2$
1780: follows
1781: the relation $\|c_\ord\|_\infty \leq k_6<\infty$, a result that depends only
1782: on the initial conditions.
1783: %
1784: %
1785: %
1786: Let us recapitulate these results:
1787: %
1788: %
1789: %
1790: \begin{itemize}
1791: \item $\|h_{\ord,x}\|_2$, is uniformly bounded;
1792: \item $\|h_{\ord}\|_\infty$ is uniformly bounded;
1793: \item The function $h_{\ord}$ is H\"older continuous in space, with exponent
1794: $\tfrac{1}{2}$;
1795: \item The function $h_{\ord}$ is nonzero everywhere and never decreases
1796: below a certain value $B>0$, independent of $\ord$, $\varepsilon$, and $\sigma$.
1797: \item $\|c_{\ord,x}\|_2$ is uniformly bounded;
1798: \item $\|c_{\ord}\|_\infty$ is uniformly bounded;
1799: \item The function $c_{\ord}$ is H\"older continuous in space, with exponent
1800: $\tfrac{1}{2}$,
1801: \end{itemize}
1802: %
1803: %
1804: %
1805: where these results are independent of $\ord$, $\varepsilon$ and $\sigma$,
1806: and hold for $0<t<\sigma$.
1807: %
1808: %
1809: %
1810: %
1811: %
1812: %
1813: \subsection{Equicontinuity of the Galerkin approximation; convergence
1814: of Galerkin approximation}
1815: \label{sec:equicontinuity}
1816: 
1817: \noindent  Using Eq.~\eqref{eq:free_energy_decay}, we obtain the bound
1818: %
1819: %
1820: %
1821: \begin{small}
1822: \begin{multline*}
1823: \int_0^t{dt'}\int_\Omega{dx}\,\bigg\{f_\varepsilon\left(h_\ord\right)\left[-h_{\ord,xxx}+\frac{h_{\ord,x}}{g_\varepsilon\left(h_\ord\right)f_\varepsilon\left(h_\ord\right)}+\frac{1}{g_\varepsilon\left(h_\ord\right)}\left(g_\varepsilon\left(h_\ord\right)c_{\ord,x}^2\right)_x\right]^2+g_\varepsilon\left(h_\ord\right)\mu_{\varepsilon,\ord,x}^2\bigg\}
1824: \\
1825: \leq F\left(0\right),\qquad 0<t<\sigma,
1826: \end{multline*}
1827: \end{small}
1828: %
1829: %
1830: %
1831: a bound that is independent of $\ord$, $\sigma$, and $\varepsilon$.  Since
1832: the
1833: quantity $\|f_\varepsilon\left(h_\ord\right)\|_\infty=\left(\|h_\ord\|_\infty+\varepsilon\right)^3$
1834: is bounded above by a constant ${A}_1$ independent of $\ord$, $\varepsilon$,
1835: and $\sigma$, we have the following string of inequalities,
1836: %
1837: %
1838: %
1839: %
1840: \begin{small}
1841: \begin{multline*}
1842: \int_0^tdt'\int_\Omega {dx}\, J_{\varepsilon,\ord}^2\\
1843: =\int_0^t{dt'}\int_\Omega{dx}\,\bigg\{f_\varepsilon\left(h_\ord\right)\left[-h_{\ord,xxx}+\frac{h_{\ord,x}}{g_\varepsilon\left(h_\ord\right)f_\varepsilon\left(h_n\right)}+\frac{1}{g_\varepsilon\left(h_n\right)}\left(g_\varepsilon\left(h_\ord\right)c_{\ord,x}^2\right)_x\right]\bigg\}^2\\
1844: %
1845: %
1846: \leq\int_0^t{dt'}\|f_\varepsilon\left(h_\ord\right)\|_\infty\int_\Omega{dx}\,\bigg\{f_\varepsilon\left(h_\ord\right)\left[-h_{\ord,xxx}+\frac{h_{\ord,x}}{g_\varepsilon\left(h_\ord\right)f_\varepsilon\left(h_\ord\right)}+\frac{1}{g_\varepsilon\left(h_\ord\right)}\left(g_\varepsilon\left(h_\ord\right)c_{\ord,x}^2\right)_x\right]^2\bigg\}\\
1847: \leq {A}_1\int_0^t{dt'}\int_\Omega{dx}\,
1848: \bigg\{f_\varepsilon\left(h_\ord\right)\left[-h_{\ord,xxx}+\frac{h_{\ord,x}}{g_\varepsilon\left(h_\ord\right)f_\varepsilon\left(h_\ord\right)}+\frac{1}{g_\varepsilon\left(h_\ord\right)}\left(g_\varepsilon\left(h_\ord\right)c_{\ord,x}^2\right)_x\right]^2\bigg\}
1849: %
1850: %
1851: \\
1852: \leq {A}_1 F\left(0\right)\equiv A_2,
1853: \end{multline*}
1854: \end{small}
1855: %
1856: %
1857: and thus
1858: %
1859: %
1860: %
1861: \[
1862: \int_0^t{dt'}\,\|J_{\varepsilon,\ord}\|_2^2\leq A_2,\qquad 0< t<\sigma,
1863: \]
1864: %
1865: %
1866: where $A_2$ is independent of $\ord$, $\varepsilon$, and $\sigma$.  Similarly,
1867: %
1868: %
1869: \[
1870: \int_0^t{dt'}\,\|g_\varepsilon\left(h_\ord\right)\mu_{\varepsilon,\ord,x}\|_2^2\leq
1871: A_3,\qquad 0<{t}<\sigma,
1872: \]
1873: %
1874: %
1875: %
1876: and
1877: \[
1878: \int_0^t{dt'}\,\|c_\ord J_{\varepsilon,\ord}\|_2^2\leq A_2\|c_\ord\|_\infty^2\leq
1879: A_2k_6^2,\qquad 0<t<\sigma,
1880: \]
1881: %
1882: %
1883: %
1884: where $A_2$ and $A_3$ are independent of $\ord$, $\varepsilon$, and $\sigma$.
1885:  By rewriting the evolution equations as
1886: %
1887: %
1888: \[
1889: h_{\ord,t}=-J_{\varepsilon,\ord,x},\qquad \left(g_\varepsilon\left(h_\ord\right)c\right)_t=-K_{\varepsilon,\ord,x},
1890: \]
1891: %
1892: %
1893: where $K_{\varepsilon,\ord}=c_\ord J_{\varepsilon,n}-g_\varepsilon\left(h_\ord\right)\mu_{\varepsilon,\ord,x}$,
1894: we see that there are uniform bounds for $\int_0^t{dt'}\,\|J_{\varepsilon,\ord}\|_2^2$
1895: and $\int_0^t{dt'}\,\|K_{\varepsilon,\ord}\|_2^2$, which depend only on the
1896: initial data $c_0\left(x\right)$ and $h_0\left(x\right)$.
1897: %
1898: %
1899: %
1900: We mention the following result due Bernis and Friedman~\cite{Friedman1990}.
1901: %
1902: %
1903: \begin{quote}
1904: \textit{[Bernis and Friedman, 1990] Let $\varphi_i\left(x,t\right)$ be a sequence
1905: of functions that weakly satisfy the equation
1906: %
1907: %
1908: \[
1909: \varphi_{i,t}=-J_{i,x},\qquad J_i=J\left(\varphi_i\right).
1910: \]
1911: %
1912: %
1913: If $\varphi_i\left(x,\cdot\right)$ is H\"older continuous (exponent $\tfrac{1}{2}$),
1914: and if the fluxes $J_i$ satisfy
1915: %
1916: %
1917: \[
1918: \int_0^t dt'\,\|J_i\|_2^2\leq A_1,\qquad 0<t<\sigma,
1919: \]
1920: %
1921: %
1922: where $A_1$ is a number independent of the index $i$ and the time $\sigma$,
1923: then there is a constant $A_2$, independent of $i$ and $\sigma$, such that
1924: %
1925: %
1926: \[
1927: \left|\varphi_i\left(\cdot,t_2\right)-\varphi_i\left(\cdot,t_1\right)\right|\leq
1928: A_2\left|t_2-t_1\right|^{\frac{1}{8}},
1929: \]
1930: %
1931: %
1932: for all $t_1$ and $t_2$ in $\left(0,\sigma\right)$.  
1933: }
1934: \end{quote}
1935: %
1936: %
1937: %
1938: We now observe that the fluxes $J_{\varepsilon,\ord}$, and $K_{\varepsilon,\ord}$
1939: satisfy the conditions of this theorem, and thus
1940: %
1941: %
1942: %
1943: \begin{quote}
1944: The functions $h_{\ord}\left(\cdot,t\right)$ and $c_\ord\left(\cdot,t\right)$
1945: are H\"older continuous (exponent $\tfrac{1}{8}$), for $0<t<\sigma$.
1946: \end{quote}
1947: %
1948: %
1949: We therefore have a uniformly bounded and equicontinuous family of functions
1950: $\{\left(h_\ord,c_\ord\right)\}_{\ord=\ord_0+1}^\infty$.
1951: We also have a recipe for constructing a uniformly bounded and equicontinuous
1952: approximate solution $\left(h_\ord\left(x,t\right),c_\ord\left(x,t\right)\right)$,
1953: in
1954: a small interval $\left(0,\sigma\right)$.  The recipe can be iterated step-by-step,
1955: and we obtain a uniformly bounded and equicontinuous family of approximate
1956: solutions $\{\left(h_\ord,c_\ord\right)\}_{\ord=\ord_0+1}^\infty$, on $\left(0,T_0\right]\times\Omega$,
1957: for an arbitrary time $T_0$.
1958: Then, using the Arzel\`a--Ascoli theorem, we obtain the following convergence
1959: result:
1960: %
1961: %
1962: %
1963: \begin{quote}
1964: There is a subsequence of the family $\{\left(h_\ord,c_\ord\right)\}_{\ord=\ord_0+1}^\infty$
1965: that converges uniformly to a limit $\left(h,c\right)$, in $\left(0,T_0\right]\times\Omega$.
1966: \end{quote}
1967: %
1968: %
1969: %
1970: %
1971: %
1972: %
1973: In contrast to Ch.~\ref{ch:background}, this convergence result is uniform
1974: and true everywhere in $x\in \Omega$ (uniform strong convergence).  We prove
1975: several facts about the pair $\left(h,c\right)$.
1976: %
1977: %
1978: %
1979: \begin{quote}
1980: \textit{Let $\left(h,c\right)$ be the limit of the family of functions $\{\left(h_\ord,c_\ord\right)\}_{\ord=\ord_0+1}^\infty$
1981: constructed in Steps~\eqref{sec:regularization}--\eqref{sec:equicontinuity}.
1982:  Then the following properties hold for this limit:
1983: %
1984: %
1985: %
1986: \begin{enumerate}
1987: \item The functions $h\left(x,t\right)$ and $c\left(x,t\right)$ are uniformly
1988: H\"older continuous in space (exponent $\tfrac{1}{2}$), and uniformly H\"older
1989: continuous in time (exponent $\tfrac{1}{8}$);
1990: \item The initial condition $\left(h,c\right)\left(x,0\right)=\left(h_0,c_0\right)\left(x\right)$
1991: holds;
1992: \item $\left(h,c\right)$ satisfy the boundary conditions of the original
1993: problem (periodic boundary conditions);
1994: \item The derivatives $\left(h,c\right)_t$, $\left(h,c\right)_x$, $\left(h,c\right)_{xx}$,
1995: $\left(h,c\right)_{xxx}$, and $\left(h,c\right)_{xxxx}$ are continuous in
1996: the set $\left(0,T_0\right]\times\Omega$;
1997: \item The function pair $\left(h,c\right)$ satisfy the following weak form
1998: of the PDEs,
1999: %
2000: %
2001: %
2002: \begin{multline*}
2003: \int\int_{Q_{T_0}}dt{dx}\, h\varphi_t+\int\int_{Q_{T_0}}dt{dx}\,J_\varepsilon\varphi_x=0,\\
2004: J_\varepsilon=f_\varepsilon\left(h\right)h_{xxx}-\frac{1}{g_\varepsilon\left(h\right)}h_x-\frac{f_\varepsilon\left(h\right)}{g_\varepsilon\left(h\right)}\left(g_\varepsilon\left(h\right)c_{x}^2\right)_x,
2005: \end{multline*}
2006: %
2007: %
2008: \vskip -0.5in
2009: %
2010: %
2011: \begin{multline*}
2012: \int\int_{Q_{T_0}}dt{dx}\, g_\varepsilon\left(h\right)c\psi_t+\int\int_{Q_{T_0}}dt{dx}\,K_\varepsilon\psi_x=0,\\
2013: K_\varepsilon= cJ_\varepsilon-g_\varepsilon\left(h\right)\left[c^3-c-\frac{1}{g_\varepsilon\left(h\right)}\left(g_\varepsilon\left(h\right)c_x\right)_x\right],
2014: \end{multline*}
2015: %
2016: %
2017: where $\varphi\left(x,t\right)$ and $\psi\left(x,t\right)$ are suitable test
2018: functions.
2019: %
2020: %
2021: %\item We have the convergence result $\left(h,c\right)_x\left(x,t\right)\rightarrow\left(h_{0,x},c_{0,x}\right)\left(x\right)$,
2022: %strongly in $L^2\left(\Omega\right)$, as $t\rightarrow0$;
2023: \end{enumerate}
2024: }
2025: \end{quote}
2026: %
2027: %
2028: %
2029: %
2030: %
2031: The statements~(1),~(2), and~(3) are obvious.  Now, any pair $\left(h_\ord\left(x,t\right),c_\ord\left(x,t\right)\right)$
2032: satisfies the equation set
2033: %
2034: %
2035: %
2036: \begin{multline*}
2037: \int\int_{Q_{T_0}}dt{dx}\, h_\ord\varphi_t+\int\int_{Q_{T_0}}dt{dx}\, J_{\varepsilon,\ord}\varphi_x=0,
2038: \\
2039: J_{\varepsilon,\ord}=f_{\varepsilon}\left(h_\ord\right)h_{\ord,xxx}-\frac{1}{g_\varepsilon\left(h_\ord\right)}h_{\ord,x}-\frac{f_\varepsilon\left(h_\ord\right)}{g_\varepsilon\left(h_\ord\right)}\left(g_\varepsilon\left(h_{\ord}\right)c_{\ord,x}^2\right)_x,
2040: \end{multline*}
2041: %
2042: %
2043: \begin{multline*}
2044: %\hskip -0.2in
2045: \int\int_{Q_{T_0}}dt{dx}\, g_\varepsilon\left(h_\ord\right)c_\ord\psi_t+\int\int_{Q_{T_0}}dt{dx}\,
2046: K_{\varepsilon,\ord}\psi_x=0,\\
2047: K_{\varepsilon,\ord}= c_{\ord}J_{\varepsilon,\ord}-g_\varepsilon\left(h_\ord\right)\left[c_\ord^3-c_\ord-\frac{1}{g_\varepsilon\left(h_\ord\right)}\left(g_\varepsilon\left(h_\ord\right)c_{\ord,x}\right)_x\right],
2048: \end{multline*}
2049: %
2050: %
2051: %
2052: and from the boundedness of the fluxes $J_{\varepsilon,\ord}$ and $K_{\varepsilon,\ord}$
2053: in $L^2\left(0,T_0,L^2\left(\Omega\right)\right)$, it follows that
2054: %
2055: %
2056: \[
2057: \left(J_{\varepsilon,\ord},K_{\varepsilon,\ord}\right)\rightharpoonup\left(J_\varepsilon,K_\varepsilon\right),\qquad\text{weakly
2058: in }L^2\left(0,T_0,L^2\left(\Omega\right)\right),
2059: \]
2060: %
2061: %
2062: for a subsequence.  Using the regularity
2063: theory for uniformly parabolic equations and the uniform H\"older continuity
2064: of the $\left(h_{\ord},c_{\ord}\right)$'s, it follows that
2065: %
2066: %
2067: %
2068: \begin{quote}
2069: The derivatives 
2070: %
2071: %
2072: $\left(h_\ord,c_\ord\right)_t$, 
2073: %
2074: %
2075: $\left(h_\ord,c_\ord\right)_x$, $\left(h_\ord,c_\ord\right)_{xx}$, $\left(h_\ord,c_\ord\right)_{xxx}$
2076: %
2077: and $\left(h_\ord,c_\ord\right)_{xxxx}$ 
2078: %
2079: %
2080: are uniformly convergent in any compact
2081: subset of $\left(0,T_0\right]\times\Omega$.
2082: %
2083: %
2084: %
2085: \end{quote}
2086: %
2087: %
2088: Thus,
2089: %
2090: %
2091: \[
2092: J_\varepsilon=f_\varepsilon\left(h\right)h_{xxx}-\frac{1}{g_\varepsilon\left(h\right)}h_x-\frac{f_\varepsilon\left(h\right)}{g_\varepsilon\left(h\right)}\left(g_\varepsilon\left(h\right)c_{x}^2\right)_x,
2093: \]
2094: %
2095: %
2096: \[
2097: K_\varepsilon=cJ_\varepsilon-g_\varepsilon\left(h\right)\left[c^3-c-\frac{1}{g_\varepsilon\left(h\right)}\left(g_\varepsilon\left(h\right)c_x\right)_x\right],
2098: \]
2099: %
2100: %
2101: %
2102: on $\left(0,T_0\right]\times\Omega$, and therefore Claims~(4) and~(5)
2103: follow. 
2104: 
2105: 
2106: \subsection{Convergence of regularized problem, as $\varepsilon\rightarrow0$}
2107: \label{sec:convergence}
2108: 
2109: \noindent The result in Step~\eqref{sec:equicontinuity} produced a solution
2110: $\left(h_\varepsilon,c_\varepsilon\right)$ to the regularized problem.  Due
2111: to the result
2112: %
2113: %
2114: %
2115: \begin{equation}
2116: h_{\varepsilon}\left(x,t\right)\geq h_{\mathrm{min}}\geq B=-k_1\sqrt{L}+\sqrt{k_1^2L+\frac{k_1^2L}{e^{k_4k_1^2}-1}}>0,\\
2117: \label{eq:no_rupture}
2118: \end{equation}
2119: %
2120: %
2121: independent of $\varepsilon$, the argument of Step~\eqref{sec:equicontinuity}
2122: can be recycled to produce a solution $\left(h,c\right)$ to the unregularized
2123: problem.  
2124: This solution is constructed as a limit $\left(h,c\right)=\lim_{\varepsilon\rightarrow0}\left(h_\varepsilon,c_\varepsilon\right)$,
2125: and the results of the theorem in Step~\eqref{sec:equicontinuity} apply again
2126: to $\left(h,c\right)$.
2127:  The result~\eqref{eq:no_rupture} applies to $h$ constructed as $h=\lim_{\varepsilon\rightarrow0}h_\varepsilon$,
2128:  and thus all the derivatives $\left(h,c\right)_t$, $\left(h,c\right)_x$,
2129:  $\left(h,c\right)_{xx}$, $\left(h,c\right)_{xxx}$, and $\left(h,c\right)_{xxxx}$
2130:  are continuous on the whole space $\left(0,T_0\right]\times\Omega$ and
2131:  therefore, the weak solution $\left(h,c\right)$ is in fact a strong one.
2132:  
2133: \subsection{Regularity properties of the solution $\left(h,c\right)$}
2134: \label{sec:analysis_thin_films:regularity}
2135: 
2136: Using a bootstrap argument, we show that the solution $\left(h,c\right)$
2137: belongs to the regularity
2138: classes $L^{\infty}\left(0,T_0;H^{2,2}\left(\Omega\right)\right)$ and $L^2\left(0,T_0;H^{4,2}\left(\Omega\right)\right)$.
2139:  From Step~\eqref{sec:a_priori_bds} it follows immediately that
2140: %
2141: %
2142: \[
2143: \|h_x\|_2, \|c_x\|_2 < \infty,
2144: \]
2145: %
2146: %
2147: with time-independent bounds.  Thus, using Poincar\'e's inequality, it follows
2148: that $h,c\in H^{1,2}\left(\Omega\right)$ and, moreover, 
2149: %
2150: %
2151: \[
2152: \sup_{\left[0,T_0\right]}\|h_x\|_2, \sup_{\left[0,T_0\right]}\|c_x\|_2 <
2153: \infty.
2154: \]
2155: %
2156: % 
2157: From Step~\eqref{sec:equicontinuity},
2158: it follows that $J$ and $\mu_x$ belong to the regularity class $L^2\left(0,T_0;L^2\left(\Omega\right)\right)$,
2159: and hence $J$, $\mu_x\in L^2\left(0,T_0;L^1\left(\Omega\right)\right)$. 
2160: The functions $J$, $\mu$, and $\mu_x$ take the form
2161: %
2162: %
2163: \[
2164: J=h^3 h_{xxx}-h_xh^{-1}-h^2\left(h_x c_x^2+2h c_x c_{xx}\right),
2165: \]
2166: %
2167: % 
2168: %
2169: %
2170: \[
2171: \mu = c^3-c - h^{-1}\left(h_x c_x + hc_{xx}\right),
2172: \]
2173: %
2174: %
2175: and
2176: %
2177: %
2178: %
2179: \[
2180: \mu_x= \left(3c^2-1\right)c_x + h^{-2}h_x^2 c_x - h^{-1}h_{xx}c_x - h^{-1}h_xc_{xx}
2181: -c_{xxx},
2182: \]
2183: %
2184: %
2185: respectively.
2186: %
2187: %
2188: The following results will help us in our demonstration,
2189: %
2190: %
2191: \begin{itemize}
2192: \item We shall use the boundedness of $h\left(x,t\right)$,
2193: %
2194: %
2195: \[
2196: 0< h_{\mathrm{min}}\leq h\left(x,t\right)\leq h_{\mathrm{max}}<\infty,
2197: \]
2198: %
2199: %
2200: %
2201: and the boundedness of $c\left(x,t\right)$, $\|c\|_\infty <\infty$. 
2202: \item Since $\mu_x\in L^2\left(0,T_0;L^2\left(\Omega\right)\right)$, it follows
2203: that $\mu\in L^2\left(0,T_0; L^2\left(\Omega\right)\right)$, by Poincar\'e's
2204: inequality.
2205: \item From this it follows that $c_x h_x + hc_{xx}$ is in the class $L^2\left(0,T_0;L^2\left(\Omega\right)\right)$.
2206: \item  We have the inequality 
2207: %
2208: %
2209: \[
2210: \|h\mu c_x\|_2^2\leq h_{\mathrm{max}}\|\mu\|_\infty^2\|c_x\|_2^2\leq
2211: h_{\mathrm{max}}\|c_x\|_2^2\left[\frac{1}{\sqrt{L}}\|\mu\|_2+\sqrt{L}\|\mu_x\|_2\right]^2,
2212: \]
2213: %
2214: and thus $h_x c_x^2 + h c_x c_{xx}\in L^2\left(0,T_0;L^2\left(\Omega\right)\right)$.
2215: \item Similarly, since $\int_0^{T_0}{dt}\,\|h\mu h_x\|_2^2<\infty$, we have the result
2216: $h_x^2c_x +hh_x c_{xx}\in  L^2\left(0,T_0;L^2\left(\Omega\right)\right)$.
2217: \end{itemize}
2218: 
2219: Now inspection of $\mu$ shows that $c_{xx}$ is in $L^2\left(0,T_0;L^1\left(\Omega\right)\right)$,
2220: from which follows the result $h_{xxx},c_{xxx}\in L^1\left(0,T_0;L^1\left(\Omega\right)\right)$.
2221:  By repeating the same argument, we find that $c_{xx}\in L^2\left(0,T_0;L^2\left(\Omega\right)\right)$.
2222:   We also have the result that $\|h_x^2c_x\|_2$ is almost always bounded.
2223:    To show that $h_{xx}\in  L^2\left(0,T_0;L^2\left(\Omega\right)\right)$,
2224:    we take the evolution equation for $h\left(x,t\right)$, multiply it by
2225:    $h$, and integrate, obtaining
2226: %
2227: %
2228: \[
2229: -\tfrac{1}{2}\frac{d}{dt}\int_\Omega{dx}\,h^2-3\int_\Omega{dx}\,h^2h_x^2h_{xx}-\int_\Omega{dx}\,h_xJ_0=\int_\Omega{dx}\,h^3h_{xx}^2,
2230: \]
2231: %
2232: %
2233: where $J_0=h_xh^{-1}+2h^3c_xc_{xx}+h^2h_xc_x^2$.   The time integral of the
2234: first term on the
2235: left-hand side of this equation is obviously bounded in time.  Let us examine
2236: the time integral of the second term,
2237: %
2238: %
2239: \begin{eqnarray*}
2240: \int_0^{T_0}{dt}\int_\Omega{dx}\,h^2h_x^2h_{xx}&\leq& h_{\mathrm{max}}^2\left(\sup_{\left[0,T_0\right]}\|h_x\|_2^2\right)\int_0^{T_0}{dt}\,\|h_{xx}\|_\infty\\
2241: &\leq& h_{\mathrm{max}}^2\left(\sup_{\left[0,T_0\right]}\|h_x\|_2^2\right)\int_0^{T_0}{dt}\,\left(L^{-1}\|h_{xx}\|_1+\|h_{xxx}\|_1\right)<\infty.
2242: \end{eqnarray*}
2243: %
2244: %
2245: The third term on the left-hand side is despatched with in a similar way,
2246: so that $\int_0^{T_0}{dt}\|h_{xx}\|_2^2<\infty$.  We have now shown that
2247: %
2248: %
2249: \[
2250: h,c\in L^2\left(0,T_0;H^{2,2}\left(\Omega\right)\right).
2251: \]
2252: %
2253: %
2254: Using this result, together with the previous facts gathered together in
2255: this section, it is readily shown that
2256: %
2257: %
2258: \[
2259: h,c\in L^2\left(0,T_0;H^{3,1}\left(\Omega\right)\right),
2260: \]
2261: %
2262: %
2263: And finally, using this result, it follows that
2264: %
2265: %
2266: %
2267: %
2268: \[
2269: h,c\in L^2\left(0,T_0;H^{3,2}\left(\Omega\right)\right).
2270: \]
2271: %
2272: %
2273: For example, 
2274: %
2275: %
2276: %
2277: \begin{eqnarray*}
2278: \int_0^{T_0}{dt}\int_\Omega{dx}\,h_x^2c_{xx}^2&\leq& \left(\sup_{\left[0,T_0\right]}\|h_x\|_2^2\right)\int_0^{T_0}{dt}\,\|c_{xx}\|_\infty^2\\
2279: &\leq&\left(\sup_{\left[0,T_0\right]}\|h_x\|_2^2\right)\int_0^{T_0}{dt}\,\left(L^{-1}\|c_{xx}\|_1+\|c_{xxx}\|_1\right)^2<\infty.
2280: \end{eqnarray*}
2281: %
2282: %
2283: %
2284: This bound, together with the result $h\mu h_x\in L^2\left(0,T_0;L^2\left(\Omega\right)\right)$,
2285: implies that 
2286: %
2287: %
2288: %
2289: \[
2290: h_x^2 c_x\in L^2\left(0,T_0;L^2\left(\Omega\right)\right).
2291: \]
2292: %
2293: %
2294: Similarly,
2295: %
2296: %
2297: \[
2298: \int_0^{T_0}{dt}\int_\Omega{dx}\, h_x^2c_{xx}^2\leq\left(\sup_{\left[0,T_0\right]}\|h_x\|_2^2\right)\int_0^{T_0}{dt}\,\left(L^{-1}\|c_{xx}\|_1+\|c_{xxx}\|_1\right)^2,
2299: \]
2300: %
2301: %
2302: Hence $h_x c_{xx}\in   L^2\left(0,T_0;L^2\left(\Omega\right)\right)$, and
2303: by
2304: symmetry, $c_{x} h_{xx} \in  L^2\left(0,T_0;L^2\left(\Omega\right)\right)$.
2305: Since
2306: %
2307: %
2308: \[
2309: \mu_x= \left(3c^2-1\right)c_x + h^{-2}h_x^2 c_x - h^{-1}h_{xx}c_x - h^{-1}h_xc_{xx}
2310: -c_{xxx},
2311: \]
2312: %
2313: %
2314: is in $L^2\left(0,T_0;L^2\left(\Omega\right)\right)$, it follows that $c_{xxx}$
2315: is in this class too.  A similar argument holds for $h_{xxx}$.  
2316: Thus, the solution $\left(h,c\right)$ belongs to the following regularity
2317: class:
2318: %
2319: %
2320: \begin{equation}
2321: \left(h,c\right)\in L^{\infty}\left(0,T_0;H^{1,2}\left(\Omega\right)\right)
2322: \cap L^2\left(0,T_0;H^{3,2}\left(\Omega\right)\right)
2323: \cap C^{\frac{1}{2},\frac{1}{8}}\left(\Omega\times\left[0,T_0\right]\right).
2324: \label{eq:regularity0}
2325: \end{equation}
2326: %
2327: %
2328: %
2329: %
2330: %
2331: %
2332: %
2333: %
2334: %
2335: %
2336: %
2337: %
2338: %
2339: %
2340: %
2341: % Inspecting the form of $J$, we see that to prove $h_{xxx}\in L^1\left(0,T_0;L^1\left(\Omega\right)\right)$,
2342: % it suffices to show that $h_x c_x^2$ and $c_x c_{xxx}$ belong to this class:
2343: % %
2344: % %
2345: % \begin{itemize}
2346: % \item
2347: % Since $\mu_x\in L^2\left(0,T_0;L^1\left(\Omega\right)\right)$, it follows
2348: % that $\mu\in L^2\left(0,T_0; L^1\left(\Omega\right)\right)$, by Poincar\'e's
2349: % inequality.  Thus, $\int_0^{T_0}{dt}\|h_x c_x + h c_{xx}\|_1^2 <\infty$,
2350: % and
2351: % since $\|h_x c_x\|_1\leq \|h_x\|_2\|c_x\|_2 < \infty$, it follows that
2352: % %
2353: % %
2354: % \[
2355: % \int_0^{T_0}{dt}\|c_{xx}\|_1^2 <\infty.
2356: % \]
2357: % %
2358: % %
2359: % %
2360: % \item Now $\|c_x c_{xx}\|_1\leq \sup_\Omega|c_x|\|c_{xx}\|_1$, and using
2361: % the inequality
2362: % %
2363: % %
2364: % \[
2365: % \sup_\Omega |\phi|\leq \frac{1}{L}\|\phi\|_1 + \|\phi_x\|_1,
2366: % \]
2367: % %
2368: % %
2369: % it follows that $\|c_x c_{xx}\|_1\leq L^{-1}\|c_x\|_1\|c_{xx}\|_1 + \|c_{xx}\|_1^2$,
2370: % and thus
2371: % %
2372: % %
2373: % \begin{equation}
2374: % \int_0^{T_0}{dt}\|c_x c_{xx}\|_1 \leq \frac{1}{L} \left(\sup_{\left[0,T_0\right]}\|c_x\|_1\right)
2375: % \int_0^{T_0}{dt}\|c_{xx}\|_1 + \int_0^{T_0}{dt}\|c_{xx}\|_1^2 < \infty.
2376: % \label{eq:ineq1}
2377: % \end{equation}
2378: % %
2379: % %
2380: % \item We have the inequality 
2381: % %
2382: % %
2383: % \[
2384: % \|h\mu c_x\|_1^2\leq h_{\mathrm{max}}\|\mu\|_2^2\|c_x\|_2^2\leq
2385: % h_{\mathrm{max}}\|c_x\|_2^2\left[\left(\int_\Omega{dx}\,\mu\right)^2+\left(\frac{L}{2\pi}\right)^2\|\mu_x\|_2^2\right],
2386: % \]
2387: % %
2388: % %
2389: % %
2390: % and thus
2391: % %
2392: % %
2393: % \begin{multline*}
2394: % \int_0^{T_0}{dt}\|h\mu c_x\|_1^2 \leq h_{\mathrm{max}}^2\left(\sup_{\left[0,T_0\right]}\|c_x\|_2^2\right)\left(\|c\|_\infty^3+\|c\|_\infty+h_{\mathrm{min}}^{-1}\|h_x\|_2\|c_x\|_2\right)^2T_0\\
2395: % +\left(\frac{h_{\mathrm{max}}L}{2\pi}\right)^2\sup_{\left[0,T_0\right]}\|c_x\|_2^2\int_0^{T_0}{dt}\|\mu_x\|_2^2
2396: % < \infty.
2397: % \end{multline*}
2398: % %
2399: % %
2400: % %
2401: % Since $h\mu c_x = h c_x\left(c^3-c\right)-h_x c_x^2 - h c_x c_{xx}$ and
2402: % $c_x
2403: % c_{xx}$ are bounded in $L^1\left(0,T_0;L^1\left(\Omega\right)\right)$,
2404: % it
2405: % follows that 
2406: % %
2407: % %
2408: % %
2409: % \begin{equation}
2410: % \int _0^{T_0}{dt}\|h_x c_x^2\|_1 < \infty.
2411: % \label{eq:ineq2}
2412: % \end{equation}
2413: % \end{itemize}
2414: % %
2415: % %
2416: % From Eqs.~\eqref{eq:ineq1} and~\eqref{eq:ineq2}, it follows that $h_{xxx}\in
2417: % L^1\left(0,T_0;L^1\left(\Omega\right)\right)$.
2418: % 
2419: % 
2420: % 
2421: % 
2422: % 
2423: % To prove the boundedness of $\int_0^{T_0}{dt}\|c_{xxx}\|_1$, we inspect
2424: % the form
2425: % of $\mu_x$: it is
2426: % %
2427: % %
2428: % \[
2429: % \mu_{x} = \left(3c^2-1\right)c_x + h^{-2}h_x^2c_x - h^{-1}h_{xx}c_x-h^{-1}h_xc_{xx}-c_{xxx}.
2430: % \]
2431: % %
2432: % %
2433: % If we can show the boundedness of all but the $c_{xxx}$-term in this expression
2434: % (in the $L^1\left(0,T_0;L^{1}\left(\Omega\right)\right)$ sense),
2435: % it follows that $c_{xxx}$ will be bounded in this sense.
2436: % %
2437: % %
2438: % %
2439: % \begin{itemize}
2440: % \item The first term is obviously bounded in $L^1\left(0,T_0;L^1\left(\Omega\right)\right)$.
2441: %  In this section by `bounded' we shall mean bounded in this sense.
2442: % \item We have the inequality
2443: % %
2444: % %
2445: % \[
2446: % \|h_x^2 c_{x}\|\leq\sup_\Omega|c_x|\|h_x\|_2^2\leq
2447: % \|h_x\|_2^2\left(L^{-1}\|c_x\|_1+\|c_{xx}\|_1\right),\]
2448: % %
2449: % %
2450: % which provides a bound on the second term.
2451: % %
2452: % %
2453: % %
2454: % \item The third term requires that $\|h_{xx}c_x\|_1$ be bounded.  We have
2455: % the inequality $\|h_{xx}c_x\|_1\leq \|h_{xx}\|_2\|c_{x}\|_2$.  Using the
2456: % following relation from Appendix A,
2457: % %
2458: % %
2459: % \[
2460: % \|u_{x}\|_2\leq |\Omega|^{\frac{1}{2}}\|u_{xx}\|_1 + \frac{2}{\sqrt{L}}\|u\|_1^{\frac{1}{2}}\|u_{xx}\|_1^{\frac{1}{2}},
2461: % \]
2462: % %
2463: % %
2464: % it follows that
2465: % %
2466: % %
2467: % \[
2468: % \int_0^{T_0}{dt}\|h_{xx}\|_2\leq|\Omega|^{\frac{1}{2}}\int_0^{T_0}{dt}\|h_{xxx}\|_1+\frac{2}{\sqrt{L}}\sup_{\left[0,T_0\right]}\|h_x\|_1^{\frac{1}{2}}\left[\int_0^{T_0}{dt}\|h_{xxx}\|_1\right]^{\frac{1}{2}},
2469: % \]
2470: % %
2471: % %
2472: % and the boundedness of $\|h_{xx}c_x\|_1$ follows.
2473: % %
2474: % %
2475: % %
2476: % \item Since $h\mu h_x = h h_x\left(c^3-c\right)-h_x^2 c_x - h h_x c_{xx}$
2477: % is bounded, it follows that 
2478: % %
2479: % %
2480: % %
2481: % \begin{equation}
2482: % \int _0^{T_0}{dt}\|h_x c_{xx}\|_1 < \infty.
2483: % \label{eq:ineq2}
2484: % \end{equation}
2485: % \end{itemize}
2486: % %
2487: % %
2488: % %
2489: % From this list of inequalities, it follows that $c_{xxx}\in
2490: % L^1\left(0,T_0;L^1\left(\Omega\right)\right)$.  The continuity $\left(h,c\right)\in
2491: % C^{\frac{1}{8},\frac{1}{2}}\left(\left[0,T_0\right]\times\Omega\right)$
2492: % follows
2493: % from Sec.~\eqref{sec:equicontinuity}. 
2494: % 
2495: % 
2496: %
2497: %
2498: %
2499: 
2500: Extra regularity is obtained by writing the equation pair as
2501: %
2502: %
2503: %
2504: \begin{eqnarray*}
2505: \frac{\partial h}{\partial t}+h^3h_{xxxx}&=&-3h^2h_xh_{xxx}+\varphi_1+\varphi_2\equiv
2506: \varphi\left(x,t\right),\\
2507: %
2508: %
2509: \frac{\partial c}{\partial t}+c_{xxxx}&=&-\frac{2}{h}h_xc_{xxx}+\psi_1+\psi_2\equiv
2510: \psi\left(x,t\right),
2511: \end{eqnarray*}
2512: %
2513: %
2514: where 
2515: %
2516: %
2517: %
2518: \begin{eqnarray*}
2519: \int_0^{\tau}{dt}\,\|\varphi_1\|_2&\leq& \left(\sup_{\left[0,\tau\right]}\|c_{xx}\|_2\right)\int_0^{\tau}{dt}\,|\nu_1|,\qquad
2520: \nu_1\in L^2\left(\left[0,\tau\right]\right),\\
2521: %
2522: %
2523: \int_0^{\tau}{dt}\,\|\psi_1\|_2&\leq& \left(\sup_{\left[0,\tau\right]}\|c_{xx}\|_2\right)\int_0^{\tau}{dt}\,|\nu_2|,\qquad
2524: \nu_2\in L^2\left(\left[0,\tau\right]\right),
2525: %
2526: %
2527: \end{eqnarray*}
2528: %
2529: %
2530: %
2531: for any $\tau\in \left(0,T_0\right]$, and where $\varphi_2$ and $\psi_2$
2532: belong to
2533: the class $L^2\left(0,T_0;L^2\left(\Omega\right)\right)$.  By multiplying
2534: the height and concentration equations by $h_{xxxx}$ and $c_{xxxx}$ respectively,
2535: and by integrating over space and time, it is readily shown that
2536: %
2537: %
2538: %
2539: \[
2540: \left(h,c\right)\in L^\infty\left(0,T_0,H^{2,2}\left(\Omega\right)\right),
2541: \]
2542: %
2543: %
2544: and hence
2545: %
2546: %
2547: \[
2548: \left(\varphi,\psi\right)\in L^2\left(0,T_0,L^{2}\left(\Omega\right)\right),
2549: \]
2550: %
2551: %
2552: from which follows the regularity result
2553: %
2554: %
2555: %
2556: \begin{equation}
2557: \left(h,c\right)\in L^{\infty}\left(0,T_0;H^{2,2}\left(\Omega\right)\right)
2558: \cap L^2\left(0,T_0;H^{4,2}\left(\Omega\right)\right)
2559: \cap C^{\frac{3}{2},\frac{1}{8}}\left(\Omega\times\left[0,T_0\right]\right).
2560: \label{eq:regularity}
2561: \end{equation}
2562: %
2563: %
2564: %
2565: %
2566: %
2567: \subsection{Uniqueness of solutions}
2568: \label{sec:analysis:thin_films:uniqueness}
2569: %
2570: %
2571: %
2572: %
2573: Let us consider two solution pairs $\left(h,c\right)$ and $\left(h',c'\right)$
2574: and form the difference $\left(\delta h,\delta c\right)=\left(h-h',c-c'\right)$.
2575:  Given the initial conditions $\left(\delta c\left(x,0\right),\delta h\left(x,0\right)\right)=\left(0,0\right)$,
2576:  we show that $\left(\delta h,\delta c \right)=\left(0,0\right)$ for all
2577:  time, that is, that the solution we have constructed is unique.
2578: %
2579: %
2580: %
2581: We observe that that the equation for the difference $\delta c$ can be written
2582: in the form
2583: %
2584: %
2585: \begin{equation}
2586: \frac{\partial}{\partial t}\delta{c}+\frac{\partial^4}{\partial{x}^4}\delta{c}=\delta
2587: \varphi\left(x,t\right),
2588: \label{eq:delta_c_unique}
2589: \end{equation}
2590: %
2591: %
2592: %
2593: where 
2594: %
2595: %
2596: %
2597: % \begin{multline*}
2598: % \delta f\left(x,t\right)=f_0\left(x,t\right)\delta{c}+f_1\left(x,t\right)\delta{c}_x+f_2\left(x,t\right)\delta{c}_{xx}+f_3\left(x,t\right)\delta{c}_{xxx}\in
2599: % L^1\left(0,T;L^2\left(\Omega\right)\right),\\
2600: % 0<T\leq T_0,
2601: % \end{multline*}
2602: %
2603: %
2604: %
2605: %
2606: %
2607: %
2608: %
2609: % and where the $f_i$'s are determined by the the smooth functions $\left(h,c\right)$,
2610: % $\left(h',c'\right)$, and their derivatives.  
2611: % 
2612: $\delta \varphi\left(x,t\right) \in L^2\left(0,T_0,L^2\left(\Omega\right)\right)$,
2613: and where $\delta \varphi\left(\delta c=0\right)=0$.
2614: %
2615: %
2616: %
2617: %
2618: %
2619: Thus, using the semigroup theory of Sec.~\ref{sec:background:Galerkin},
2620: we find that Eq.~\eqref{eq:delta_c_unique} has a unique solution.
2621: %
2622: %
2623: %
2624: %
2625: % \[
2626: % \delta c\left(x,t\right)=\sum_{j=1}^{\infty}\phi_j\int_0^t{ds}\,\left(\delta
2627: % g\left(x,s\right),\phi_j\right)e^{-\nu_j\left(t-s\right)},
2628: % \]
2629: %
2630: %
2631: % where $\Delta^2\phi_j=\nu_j\phi_j$ solves the eigenvalue problem for the
2632: % hyperdiffusion operator and $\left(\phi_i,\phi_j\right)=\delta_{ij}$. 
2633: % The eigenvalues $\nu_j$ are positive, except when $\phi_i$ is the constant
2634: % eigenfunction, in which case the corresponding eigenvalue is zero.
2635: Since $\delta{c}=0$ satisfies Eq.~\eqref{eq:delta_c_unique}, and since $\delta{c}\left(x,0\right)=0$,
2636: it follows that $\delta{c}=0$ for all times $t\in\left[0,T_0\right]$.
2637: 
2638: It is now possible to formulate an equation for the difference $\delta{h}$
2639: by subtracting the evolution equations of $h$ and $h'$ from one another,
2640: mindful that $\delta c=0$.  We multiply the resulting equation by $\delta{h}_{xx}$
2641: and integrate over space, obtaining the result
2642: %
2643: %
2644: %
2645: \begin{multline*}
2646: \tfrac{1}{2}\frac{d}{dt}\int_\Omega{dx}\,\delta h_x^2 +\int_\Omega{dx}\,h^3\delta
2647: h_{xxx}^2=
2648: %
2649: %
2650: \int_\Omega{dx}\,\delta h_{xxx}^2\delta h_x\left(h^{-1}+h'^2c_x^2\right)\\
2651: %
2652: %
2653: %+\int_\Omega{dx}\,\delta h_{xxx}\delta c_x\left(2h'^3c_{xx}+h'^2h'_x\left(c_x+c'_x\right)\right)
2654: %
2655: %
2656: -\int_\Omega{dx}\,\left(h^3-h'^3\right)\left(h'_{xxx}+4c_xc_{xx}\right)\delta
2657: h_{xxx}\\
2658: %
2659: %
2660: %+2\int_\Omega{dx}\,\delta h_{xxx}\delta c_{xx} h'^3c'_x
2661: %
2662: %
2663: +
2664: \int_\Omega{dx}\,\delta h_{xxx}\left[h'_x\left(h^{-1}-h'^{-1}\right)+h_xc_x^2\left(h^2-h'^2\right)\right].
2665: \end{multline*}
2666: %
2667: %
2668: %
2669: Using the lower bound on $h\left(x,t\right)\geq h_{\mathrm{min}}>0$ and Young's
2670: first inequality, this equation is transformed into an inequality,
2671: %
2672: %
2673: %
2674: %
2675: %
2676: \begin{multline*}
2677: \tfrac{1}{2}\frac{d}{dt}\int_\Omega{dx}\,\delta h_x^2 +h_{\mathrm{min}}^3\int_\Omega{dx}\,\delta
2678: h_{xxx}^2\leq
2679: %
2680: %
2681: \kappa_1\int_\Omega{dx}\,\delta h_{xxx}^2+\frac{1}{4\kappa_1}\int_\Omega{dx}\,\delta{h}_x^2\left(h^{-1}+h'^2c_x^2\right)^2\\
2682: %
2683: %
2684: %
2685: %
2686: +\kappa_2\int_\Omega{dx}\,\delta{h}_{xxx}^2+\frac{1}{4\kappa_2}\int_\Omega{dx}\,\left(h^3-h'^3\right)^2\left(h'_{xxx}+4c_xc_{xx}\right)^2\\
2687: %
2688: %
2689: +
2690: \kappa_3\int_\Omega{dx}\,\delta h_{xxx}^2+\frac{1}{4\kappa_3}\int_\Omega{dx}\,\left[h'_x\left(h^{-1}-h'^{-1}\right)+h_xc_x^2\left(h^2-h'^2\right)\right]^2,
2691: \end{multline*}
2692: %
2693: %
2694: %
2695: where $\kappa_1$, $\kappa_2$, and $\kappa_3$ are arbitrary positive constants.
2696:  By choosing $\kappa_1+\kappa_2+\kappa_3=h_{\mathrm{min}}^3$, the inequality
2697:  simplifies,
2698: %
2699: %
2700: %
2701: \begin{multline*}
2702: \tfrac{1}{2}\frac{d}{dt}\int_\Omega{dx}\,\delta h_x^2\leq
2703: %
2704: %
2705: \frac{1}{4\kappa_1}\int_\Omega{dx}\,\delta{h}_x^2\left(h^{-1}+h'^2c_x^2\right)^2\\
2706: %
2707: %
2708: %
2709: +\frac{1}{4\kappa_2}\int_\Omega{dx}\,\left(h^3-h'^3\right)^2\left(h'_{xxx}+4c_xc_{xx}\right)^2\\
2710: %
2711: %
2712: +
2713: \frac{1}{4\kappa_3}\int_\Omega{dx}\,\left[h'_x\left(h^{-1}-h'^{-1}\right)+h_xc_x^2\left(h^2-h'^2\right)\right]^2,
2714: \end{multline*}
2715: %
2716: %
2717: %
2718: which in turn reduces to
2719: %
2720: %
2721: %
2722: \begin{multline*}
2723: 2\kappa\frac{d}{dt}\int_\Omega{dx}\,\delta h_x^2\leq
2724: %
2725: %
2726: \int_\Omega{dx}\,\delta{h}_x^2\left(h^{-1}+h'^2c_x^2\right)^2\\+
2727: %
2728: %
2729: %
2730: \int_\Omega{dx}\,\delta{h}^2\left[\left(h'_{xxx}+4c_xc_{xx}\right)^2+\left(h_x'+h_xc_x^2\right)^2\right],
2731: %
2732: %
2733: %
2734: %+\int_\Omega{dx}\,\left[h'_x\left(h^{-1}-h'^{-1}\right)+h_xc_x^2\left(h^2-h'^2\right)\right]^2,
2735: \end{multline*}
2736: %
2737: %
2738: %
2739: where $\kappa$ is another positive constant.  We integrate over the time
2740: interval $\left[0,T\right]$, and use the fact that $\|\delta{h}_x\|_2\left(0\right)=0$
2741: to obtain the string of inequalities
2742: %
2743: %
2744: %
2745: % \begin{small}
2746: % \begin{eqnarray*}
2747: % &&2\kappa\sup_{\tau\in\left[0,T\right]}\|\delta{h}_x\|_2^2\left(\tau\right)\\
2748: % %
2749: % %
2750: % &\leq&\int_0^T{dt}\int_\Omega{dx}\,\delta{h}_x^2\left(h^{-1}+h'^2c_x^2\right)^2+
2751: % %
2752: % %
2753: % \int_0^T{dt}\int_\Omega{dx}\,\delta{h}^2\left[\left(h'_{xxx}+4c_xc_{xx}\right)^2+\left(h_x'+h_xc_x^2\right)^2\right],\\
2754: % %
2755: % %
2756: % %
2757: % &\leq&\int_0^T{dt}\|\delta{h}_x\|_2^2\|h^{-1}+h'^2c_x^2\|_\infty^2+\int_0^T{dt}\|\delta{h}\|_\infty^2\int_\Omega{dx}\,\left[\left(h'_{xxx}+4c_xc_{xx}\right)^2+\left(h_x'+h_xc_x^2\right)^2\right],\\
2758: % %
2759: % %
2760: % %
2761: % &\leq&\sup_{\tau\in\left[0,T\right]}\|\delta{h}_x\|_2^2\left(\tau\right)\int_0^T{dt}\|h^{-1}+h'^2c_x^2\|_\infty^2\\
2762: % %
2763: % %
2764: % %
2765: % &\phantom{a}&\phantom{aaaaaaaaaaaaa}
2766: % +\sup_{\tau\in\left[0,T\right]}\|\delta{h}\|_\infty^2\left(\tau\right)\int_0^T{dt}\int_\Omega{dx}\,\left[\left(h'_{xxx}+4c_xc_{xx}\right)^2+\left(h_x'+h_xc_x^2\right)^2\right].
2767: % \end{eqnarray*}
2768: % \end{small}
2769: %
2770: %
2771: %
2772: %
2773: %
2774: %
2775: %
2776: %
2777: \begin{multline*}
2778: 2\kappa\sup_{\tau\in\left[0,T\right]}\|\delta{h}_x\|_2^2\left(\tau\right)\\
2779: %
2780: %
2781: \leq\int_0^T{dt}\int_\Omega{dx}\,\delta{h}_x^2\left(h^{-1}+h'^2c_x^2\right)^2\phantom{aaaaaaaaaaaaaaaaaaddddddddaaaaaaaaaaaa}\\
2782: %
2783: %
2784: +\int_0^T{dt}\int_\Omega{dx}\,\delta{h}^2\left[\left(h'_{xxx}+4c_xc_{xx}\right)^2+\left(h_x'+h_xc_x^2\right)^2\right],
2785: \end{multline*}
2786: %
2787: %
2788: %
2789: \vskip -0.7in
2790: %
2791: %
2792: \begin{multline*}
2793: \leq\int_0^T{dt}\|\delta{h}_x\|_2^2\|h^{-1}+h'^2c_x^2\|_\infty^2\\
2794: %
2795: %
2796: +\int_0^T{dt}\|\delta{h}\|_\infty^2\int_\Omega{dx}\,\left[\left(h'_{xxx}+4c_xc_{xx}\right)^2+\left(h_x'+h_xc_x^2\right)^2\right],
2797: \end{multline*}
2798: %
2799: %
2800: %
2801: \vskip -0.7in
2802: %
2803: %
2804: \begin{multline*}
2805: \leq\sup_{\tau\in\left[0,T\right]}\|\delta{h}_x\|_2^2\left(\tau\right)\int_0^T{dt}\|h^{-1}+h'^2c_x^2\|_\infty^2\\
2806: %
2807: %
2808: %
2809: +\sup_{\tau\in\left[0,T\right]}\|\delta{h}\|_\infty^2\left(\tau\right)\int_0^T{dt}\int_\Omega{dx}\,\left[\left(h'_{xxx}+4c_xc_{xx}\right)^2+\left(h_x'+h_xc_x^2\right)^2\right].
2810: \end{multline*}
2811: %
2812: %
2813: %
2814: %
2815: %
2816: %
2817: %
2818: %
2819: %
2820: %
2821: %
2822: The Poincar\'e inequality can be combined with the one-dimensional differential
2823: inequalities discussed in Ch.~\ref{ch:background} to yield the relation $\|f\|_\infty\leq
2824: \kappa_P\|f_x\|_2$, where $f$ is some mean-zero function and $\kappa_P$ is
2825: an $f$-independent constant.  We therefore arrive at the inequality
2826: %
2827: %
2828: %
2829: \begin{multline*}
2830: 2\kappa\sup_{\tau\in\left[0,T\right]}\|\delta{h}_x\|_2^2\left(\tau\right)
2831: %
2832: %
2833: \leq\sup_{\tau\in\left[0,T\right]}\|\delta{h}_x\|_2^2\left(\tau\right)\int_0^T{dt}\|h^{-1}+h'^2c_x^2\|_\infty^2\\
2834: %
2835: %
2836: +\kappa_P^2\sup_{\tau\in\left[0,T\right]}\|\delta{h}_x\|_2^2\left(\tau\right)\int_0^T{dt}\left(\|h'_{xxx}+4c_xc_{xx}\|_2^2+\|h_x'+h_xc_x^2\|_2^2\right).
2837: \end{multline*}
2838: %
2839: %
2840: %
2841: %
2842: Using the results of Sec.~\ref{sec:analysis_thin_films:regularity}, it is
2843: readily shown that $h^{-1}+h'^2c_x^2\in L^2\left(0,T;L^{\infty}\left(\Omega\right)\right)$,
2844: and that the functions $h'_{xxx}+4c_xc_{xx}$ and $h_x'+h_xc_x^2$ belong to
2845: the class $L^2\left(0,T;L^2\left(\Omega\right)\right)$.  By choosing $T$
2846: sufficiently small, it is possible to impose the inequality
2847: %
2848: %
2849: %
2850: \[
2851: \frac{1}{2\kappa}\left[\int_0^T{dt}\|h^{-1}+h'^2c_x^2\|_\infty^2+\kappa_P^2\int_0^T{dt}\left(\|h'_{xxx}+4c_xc_{xx}\|_2^2+\|h_x'+h_xc_x^2\|_2^2\right)\right]<1,
2852: \]
2853: %
2854: %
2855: which in turn forces $\sup_{\tau\in\left[0,T\right]}\|\delta{h}_x\|_2^2=0$,
2856: and hence the solution is unique.
2857: 
2858: \section{Summary}
2859: \label{sec:analysis_thin_films:conclusionos}
2860: 
2861: In this chapter we have switched focus from the passive to the active tracer.
2862:  The jump in complexity is significant, and involves the Navier--Stokes Cahn--Hilliard
2863:  equations.  We focus on the case where the active system forms a thin layer
2864:  on a substrate, and obtain simplified thin-film Stokes Cahn--Hilliard equations
2865:  that are valid when the system's vertical gradients are small compared to
2866:  the those in the lateral directions.  We analyze these equations in one
2867:  lateral dimension, using the tools developed in Ch.~\ref{ch:background},
2868:  and prove existence, regularity, and uniqueness results for the equations.
2869:   While this
2870:  analysis yields many precise
2871:  results, it is unable to provide qualitative information about the solutions,
2872:  in particular the effect of the backreaction on the concentration and free
2873:  surface height, and we therefore turn to numerical simulations in the next
2874:  chapter.
2875: 
2876: 
2877: