0805.1038/thinfilmanalysis.tex
1: \documentclass[12pt,aps,tightenlines]{revtex4}
2: \pdfoutput=1
3: 
4: %\usepackage[a4paper]{geometry}
5: \usepackage[letterpaper]{geometry}
6: \usepackage{amsmath}
7: \usepackage{amsfonts}
8: \usepackage{bm}
9: %\usepackage{euscript}
10: %\usepackage{wasysym}
11: %\usepackage{psfrag}
12: \usepackage{graphicx}
13: \usepackage{subfigure}
14: \usepackage{color}
15: 
16: \graphicspath{{figs_pdf/}}
17: 
18: \newcommand{\Sobo}{W}
19: \newcommand{\needcitn}{$^{\left[\text{citation needed}\right]}$}
20: \newcommand{\Cst}{\kappa}
21: \newcommand{\Small}{\delta}
22: \newcommand{\Stresstensor}{T}
23: \newcommand{\eye}{i}
24: 
25: \newcommand{\pairing}[2]{\left(\!\left(#1\,,\,#2\right)\!\right)}
26: 
27: \newtheorem{theorem}{Theorem}
28: \newtheorem{lemma}{Lemma}
29: 
30: \begin{document}
31: %+Title
32: \title{Nonlinear dynamics of phase separation in ultra-thin films}
33: %
34: \author{Lennon \'O N\'araigh}
35: \affiliation{Department of Mathematics, Imperial College
36:   London, SW7 2AZ, United Kingdom}
37: \author{Jean-Luc Thiffeault}
38: \email{jeanluc@mailaps.org}
39: \affiliation{Department of Mathematics, University of Wisconsin, Madison,
40: WI 53706, USA}
41: \date{7 May 2008}
42: %-Title
43: %
44: \begin{abstract}
45:   We present a long-wavelength approximation to the Navier--Stokes
46:   Cahn--Hilliard equations to describe phase separation in thin films.
47:   The equations we derive underscore the coupled behaviour of
48:   free-surface variations and phase separation.  Since we are
49:   interested in the long-time behaviour of the phase-separating fluid,
50:   we restrict our attention to films that do not rupture.  To do this,
51:   we introduce a regularising Van der Waals potential.  We analyse the
52:   resulting fourth-order equations by constructing a solution as the
53:   limit of a Galerkin approximation, and obtain existence and
54:   regularity results.  In our analysis, we find a nonzero lower bound
55:   for the height of the film, which precludes the possibility of
56:   rupture.  The lower bound depends on the parameters of the problem,
57:   and we compare this dependence with numerical simulations.  We find
58:   that while the theoretical lower bound is crucial to the
59:   construction of a smooth, unique solution to the PDEs, it is
60:   not sufficiently sharp to represent accurately the parametric
61:   dependence of the observed dips in free-surface height.
62: %
63: %
64: \end{abstract}
65: 
66: \maketitle
67: \section{Introduction}
68: \label{sec:intro} 
69: 
70: \noindent Below a certain critical temperature, a well-mixed binary
71: fluid spontaneously separates into its component parts, forming
72: domains of pure liquid.  This process can be characterised by the
73: Cahn--Hilliard equation, and numerous studies describe the physics and
74: mathematics of phase separation.  In this paper we study phase
75: separation in a thin layer, in which the the varying free surface and
76: concentration fields are coupled through a pair of nonlinear evolution
77: equations.
78: 
79: Cahn and Hilliard introduced their eponymous equation
80: in~\cite{CH_orig} to model phase separation in a binary alloy.  Since
81: then, the model has been used in diverse applications: to describe
82: polymeric fluids~\cite{Aarts2005}, fluids with interfacial
83: tension~\cite{LowenTrus}, and self-segregating populations in
84: biology~\cite{Murray1981}.
85: %
86: Analysis of the Cahn--Hilliard (CH) equation was given by Elliott and
87: Zheng in~\cite{Elliott_Zheng}, where they prove the existence,
88: uniqueness, and regularity of solutions: given sufficiently smooth
89: initial data, solutions to the CH equation are bounded for almost all
90: time in the Sobolev space $\Sobo^{4,2}$.  Several authors have
91: developed generalisations of the CH equation: a variable-mobility
92: model was introduced by Elliott and Garcke~\cite{Elliott_varmob},
93: while nonlocal effects were considered by Gajweski and
94: Zacharias~\cite{Gajewski_nonlocal}.  These additional features do not
95: qualitatively change the phase separation, and we therefore turn to
96: one mechanism that does: the coupling of a flow field to the
97: Cahn--Hilliard equation~\cite{Bray_advphys}.
98: % 
99: In this case, the Cahn--Hilliard concentration equation is modified by an
100: advection term, and the flow field is either prescribed or evolves according
101: to some fluid equation.  Ding and co-workers~\cite{Spelt2007} provide
102: a derivation of coupled Navier--Stokes Cahn--Hilliard (NSCH) equations in
103: which the velocity advects the phase-separating concentration field, while
104: concentration gradients modify the velocity through an additional stress
105: term in the momentum equation.  Such models have formed the basis of numerical
106: studies of binary fluids~\cite{Berti2005}, while other studies without
107: this feedback term highlight different regimes of phase separation under
108: flow~\cite{chaos_Berthier, ONaraigh2006, Lacasta1995}.  In this paper, the
109: NSCH equations form the starting point for our asymptotic analysis.
110: 
111: As in other applications involving the Navier--Stokes equations, the
112: complexity of the problem is reduced when the fluid is spread thinly
113: on a substrate, and the upper vertical boundary forms a free
114: surface~\cite{Oron1997}.  Then, provided vertical gradients are small
115: compared to lateral gradients, a long-wavelength approximation is
116: possible, in which the full equations with a moving boundary at the
117: free surface are reduced to a single equation for the free-surface
118: height.  In the present case, the reduction yields two equations: one
119: for the free surface, and one for the Cahn--Hilliard concentration.
120: The resulting thin-film Stokes Cahn--Hilliard equations have already
121: been introduced by the authors in~\cite{ONaraigh2007}, although the
122: focus there was on control of phase separation and numerical
123: simulations in three dimensions.  Here we confine ourselves to the
124: two-dimensional case: we derive the thin-film equations from first
125: principles, present analysis of the resulting equations, and highlight
126: the impossibility of film rupture.
127:  
128: Along with the simplification of the problem that thin-film theory
129: provides, there are many practical reasons for studying phase
130: separation in thin layers.  Thin polymer films are used in the
131: fabrication of semiconductor devices, for which detailed knowledge of
132: film morphology is required~\cite{Karim2002}.  Other industrial
133: applications of polymer films include paints and coatings, which are
134: typically mixtures of polymers.  One potential application of the
135: thin-film Cahn--Hilliard theory is in
136: self-assembly~\cite{Putkaradze2005,Xia2004, Krausch1994}.  Here
137: molecules (usually residing in a thin layer) respond to an
138: energy-minimisation requirement by spontaneously forming large-scale
139: structures.  Equations of Cahn--Hilliard type have been proposed to
140: explain the qualitative features of
141: self-assembly~\cite{Putkaradze2005,Holm2005}, and knowledge of
142: variations in the film height could enhance these models.  Indeed
143: in~\cite{ONaraigh2007} the authors use the present thin-film
144: Cahn--Hilliard model in three dimensions to control phase separation,
145: a useful tool in applications where it is necessary for the molecules
146: in the film to form a given structure.
147: 
148:  
149: The analysis of thin-film equations was given great impetus by Bernis
150: and Friedman in~\cite{Friedman1990}.  They focus 
151: on the basic thin-film equation,
152: %
153: %
154: %
155: \begin{equation}
156: \frac{\partial h}{\partial t}=-\frac{\partial}{\partial x}\left(h^n\frac{\partial^3{h}}{\partial{x^3}}\right),
157: \label{eq:basic_h}
158: \end{equation}
159: %
160: %
161: %
162: with no-flux boundary conditions on a line segment, and smooth
163: nonnegative initial conditions.  For $n=1$ this equation describes a
164: thin bridge between two masses of fluid in a Hele--Shaw cell, for
165: $n<3$ it is used in slip models as $h\rightarrow0$~\cite{Myers1998},
166: while for $n=3$ it gives the evolution of the free surface of a thin
167: film experiencing capillary forces~\cite{Oron1997}.
168: % 
169: Using a decaying free-energy functional, they prove the existence of
170: nonnegative solutions to Eq.~\eqref{eq:basic_h} for $n\geq1$, while
171: for $n\geq 4$, and for strictly positive initial conditions, the solution
172: is unique, strictly positive, and is almost
173: always bounded in the $\Sobo^{3,2}$ norm.  This paper has inspired
174: other work on the subject~\cite{Bertozzi1996,
175:   Bertozzi1998,Laugesen2002}, in which the effect of a Van der Waals
176: term on Eq.~\eqref{eq:basic_h} is investigated.  These works provide
177: results concerning regularity, long-time behaviour, and film rupture
178: in the presence of an attractive Van der Waals force.
179: %
180: More relevant to the present work is the paper by Wieland and
181: Garcke~\cite{Wieland2006}, in which a pair of partial differential
182: equations describes the coupled evolution of free-surface variations
183: and surfactant concentration.  The authors derive the relevant
184: equations using the long-wavelength theory, obtain a decaying energy
185: functional, and prove results concerning the existence and
186: nonnegativity of solutions.  We shall take a similar approach in this
187: paper.
188: 
189: 
190: When the binary fluid forms a thin film on a substrate, we shall show
191: in Sec.~\ref{sec:model} that a long-wave approximation simplifies the
192: Navier--Stokes Cahn--Hilliard equations, which reduce to a pair of
193: coupled evolution equations for the free surface and concentration.
194: If $h(x,t)$ is the scaled free-surface height, and
195: $c(x,t)$ is the binary fluid concentration, then the
196: dimensionless equations take the form
197: %
198: %
199: \begin{subequations}
200: \begin{equation}
201: \frac{\partial h}{\partial t}+\frac{\partial J}{\partial x}=0,\qquad
202: %
203: %
204: \frac{\partial}{\partial t}\left(h c\right)+\frac{\partial}{\partial x}\left(Jc\right)=\frac{\partial}{\partial{x}}\left(h\frac{\partial\mu}{\partial{x}}\right),
205: \end{equation}
206: %
207: %
208: where
209: \begin{equation}
210: J = \tfrac{1}{2} h^2 \frac{\partial\sigma}{\partial{x}}
211: - \tfrac{1}{3} h^3 \bigg\{
212:            \frac{\partial}{\partial{x}}
213:            \left(-\frac{1}{C}\frac{\partial^2{h}}{\partial{x}^2}
214:              + \phi\right) + \frac{r}{h} \frac{\partial}{\partial{x}}
215:            \left[h{\left(\frac{\partial{c}}{\partial{x}}
216:              \right)}^2\right]\bigg\},
217: \end{equation}
218: %
219: %
220: %
221: %
222: \begin{equation}
223: \mu=c^3-c-\frac{C_{\mathrm{n}}^2}{h}\frac{\partial}{\partial{x}}\left(h\frac{\partial{c}}{\partial{x}}\right).
224: \end{equation}%
225: \label{eq:model_intro}%
226: \end{subequations}%
227: %
228: %
229: %
230: Here $C$ is the capillary number, $r$ measures the strength of
231: coupling between the concentration and free-surface variations
232: (backreaction), and $C_{\mathrm{n}}$ is the scaled interfacial
233: thickness.  Additionally, $\sigma$ is the dimensionless,
234: spatially-varying surface tension, and $\phi$ is the body-force
235: potential acting on the film.  In this paper we take
236: $\phi=-|A|h^{-3}$, the repulsive Van der Waals
237: potential~\cite{Book_Parsegian2006}.  This choice stabilises the film
238: and prevents rupture.  Although rupture is in itself an important
239: feature in thin-film equations~\cite{Oron1997, Bertozzi1996,
240:   Bertozzi1998}, in this paper we are interested in late-time phase
241: separation and it is therefore undesirable.
242: 
243: We present the asymptotic analysis that converts the NSCH equations
244: into Eq.~\eqref{eq:model_intro} and prove that the model equations
245: possess smooth solutions that are bounded for almost all time in the
246: $\Sobo^{4,2}$ norm.  The principal tool in this analysis is the
247: construction of a free-energy functional for
248: Eq.~\eqref{eq:model_intro} that is a decaying function of time.  We
249: prove that $h(x,t)>0$ for all time, which is the no-rupture condition.
250: 
251: The paper is organised as follows.  In Sec.~\ref{sec:model} we discuss
252: the Navier--Stokes Cahn--Hilliard equation and the scaling laws that
253: facilitate the passage to the long-wavelength equations, and we derive
254: Eq.~\eqref{eq:model_intro}.  In Sec.~\ref{sec:existence} we analyse
255: these equations by constructing a decaying free-energy functional.  We
256: prove the existence of solutions to Eq.~\eqref{eq:model_intro} and
257: provide regularity results.  We obtain a condition on the minimum
258: value of the free-surface height, and show that this is never zero.  Using
259: numerical studies, we discuss the
260: dependence of the minimum free-surface height on the problem
261: parameters in Sec.~\ref{sec:height_dip}, and compare with the analytic
262: results.  Finally, in Sec.~\ref{sec:conclusions} we present our
263: conclusions.
264: 
265: \section{The Model Equations}
266: \label{sec:model}
267: %
268: \noindent 
269: %
270: In this section we introduce the two-dimensional Navier--Stokes
271: Cahn--Hilliard (NSCH) equation set.
272: %
273: We discuss the assumptions underlying the long-wavelength
274: approximation.
275: %
276: We enumerate the scaling rules necessary to obtain the simplified
277: equations.  
278: %
279: Finally, we arrive at a set of equations that describe phase separation in
280: a thin film subject to arbitrary body forces.
281: 
282: The full NSCH equations describe the coupled effects of phase separation
283: and flow in a binary fluid.  If $\bm{v}$ is the fluid velocity and $c$ is
284: the concentration of the mixture, where $c=\pm1$ indicates total segregation,
285: then these fields evolve as
286: %
287: %
288: \begin{subequations}
289: \begin{gather}
290:         \frac{\partial \bm{v}}{\partial t}+\bm{v}\cdot\nabla\bm{v}=\nabla\cdot\Stresstensor-\frac{1}{\rho}\nabla\phi,\\
291:         \frac{\partial c}{\partial t}+\bm{v}\cdot\nabla c=D \nabla^2\left(c^3-c-\gamma\nabla^2 c\right),\\
292:         \nabla\cdot\bm{v}=0,
293: %
294: \end{gather}%
295: \label{eq:NSCH}%
296: \end{subequations}%
297: %
298: %
299: where
300: \begin{equation}
301:         \Stresstensor_{ij} =-\frac{p}{\rho}\delta_{ij}+\nu\left(\frac{\partial
302:         v_i}{\partial
303:         x_j}+\frac{\partial
304:         v_j}{\partial x_i}\right)-\beta\gamma\frac{\partial c}{\partial x_i}\frac{\partial
305:         c}{\partial x_j}
306: \label{eq:NSCH_tensor}%
307: \end{equation}%
308: %
309: %
310: is the stress tensor, $p$ is the fluid pressure, $\phi$ is the body
311: potential and $\rho$ is the constant density.  The constant $\nu$ is
312: the kinematic viscosity, $\nu=\eta/\rho$, where $\eta$ is the dynamic
313: viscosity.  Additionally, $\beta$ is a constant with units of
314: $[\mathrm{Energy}][\mathrm{Mass}]^{-1}$, $\sqrt{\gamma}$ is a constant
315: that gives the typical width of interdomain transitions, and $D$ a
316: diffusion coefficient with dimensions
317: $[\mathrm{Length}]^2[\mathrm{Time}]^{-1}$.
318: 
319: 
320: If the system has a free surface in the vertical or $z$-direction and has
321: infinite or periodic boundary conditions (BCs) in the lateral or $x$-direction,
322: then the vertical BCs we impose are 
323: %
324: %
325: \begin{subequations}
326: \begin{equation}
327: u = w = c_z = c_{zzz}\text{ on }z=0,
328: \end{equation}
329: %
330: %
331: while on the free surface $z=h(x,t)$ they are
332: %
333: %
334: \begin{equation}
335: \hat{n}_i \hat{n}_j \Stresstensor_{ij} = -\sigma\kappa,\qquad \hat{n}_i \hat{t}_j
336: \Stresstensor_{ij} = -\frac{\partial\sigma}{\partial
337: s},
338: \label{eq:BC_stress}
339: \end{equation}
340: %
341: \begin{equation}
342: w=\frac{\partial h}{\partial t}+u\frac{\partial h}{\partial x},
343: \label{eq:BC_height}
344: \end{equation}
345: %
346: \begin{equation}
347: \hat{n}_i \partial_i c = 0, \qquad \hat{n}_i \partial_i \nabla^2 c = 0,
348: \label{eq:BC_c}
349: \end{equation}%
350: \label{eq:BCs}%
351: \end{subequations}%
352: %
353: %
354: %
355: where $\hat{\bm{n}} =
356: (-\partial_x{h}\,,\,1)/{[{1+(\partial_x{h})^2}]^{1/2}}$ is the unit
357: normal to the surface, $\hat{\bm{t}}$, is the unit vector tangent to
358: the surface, $s$ is the surface coordinate, $\sigma$ is the surface
359: tension, and $\kappa$ is the mean curvature,
360: %
361: %
362: %
363: \begin{equation*}
364: \kappa=\nabla\cdot\hat{\bm{n}}
365: %=\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)
366: =\frac{\partial_{xx}h}{\left[1+\left(\partial_x{h}\right)^2\right]^{\frac{3}{2}}}.
367: \end{equation*}
368: %
369: %
370: %
371: %
372: %
373: This choice of BCs guarantees the conservation of the total mass and volume,
374: %
375: %
376: %
377: %
378: \begin{equation}
379: \text{Mass} = \int_{\text{Dom}(t)} c(\bm{x},t)d^2x,\qquad
380: \text{Volume} = \int_{\text{Dom}(t)}d^2x.
381: \label{eq:integral_quantities}
382: \end{equation}
383: %
384: %
385: %
386: Here 
387: $\text{Dom}(t)%=\{\left(x,y\right)|x\in\left[0,L\right],z\in\left[0,h(x,t)\right]\}
388: $
389: represents the time-dependent domain of integration, owing to the
390: variability of the free surface height.  Note that in view of the
391: concentration BC~\eqref{eq:BC_c}, the stress BC~\eqref{eq:BC_stress}
392: and does not contain $c(\bm{x},t)$ or its derivatives.
393: 
394: These equations simplify considerably if the fluid forms a thin layer
395: of mean thickness $h_0$, for then the scale of lateral variations
396: $\lambda$ is large compared with the scale of vertical variations
397: $h_0$.  Specifically, the parameter $\Small = h_0/\lambda$ is small,
398: and after nondimensionalisation of Eq.~\eqref{eq:NSCH} we expand its
399: solution in terms of this parameter, keeping only the lowest-order
400: terms.  For a review of this method and its applications,
401: see~\cite{Oron1997}.  For simplicity, we shall work in two dimensions,
402: but the generalisation to three dimensions is easily
403: effected~\cite{ONaraigh2007}.
404: 
405: In terms of the small parameter $\Small$, the equations
406: nondimensionalise as follows.  The diffusion time scale is $t_0 =
407: \lambda^2 / D = h_0^2 / \left(\Small^2 D\right)$ and we choose this to
408: be the unit of time.  Then the unit of horizontal velocity is $u_0 =
409: \lambda/ t_0 = \Small D / h_0$ so that $u = \left(\Small D /
410:   h_0\right)U$, where variables in upper case denote dimensionless
411: quantities.  Similarly the vertical velocity is $w = \left(\Small^2 D
412:   / h_0\right)W$.  For the equations of motion to be half-Poiseuille
413: at $O\left(1\right)$ (in the absence of the backreaction) we choose $p
414: = \left(\eta D / h_0^2\right)P$ and $\phi = \left(\eta D /
415:   h_0^2\right)\Phi$.  Using these scaling rules, the dimensionless
416: momentum equations are
417: %
418: %
419: %
420: %
421: %
422: \begin{multline}
423: %
424: %
425: \Small Re \left(\frac{\partial U}{\partial T} + U\frac{\partial
426: U}{\partial X}+ W\frac{\partial U}{\partial Z}\right) 
427: %
428: %
429: = -\frac{\partial}{\partial X}\left(P+\Phi\right)+\Small^2\frac{\partial^2
430: U}{\partial X^2}+\frac{\partial^2 U}{\partial Z^2} \\
431: %
432: %
433: -\tfrac{1}{2}\frac{\beta\gamma}{\nu D}\frac{\partial}{\partial
434: X}\bigg[\Small^2\left(\frac{\partial c}{\partial X}\right)^2
435: %
436: %
437: +\left(\frac{\partial c}{\partial Z}\right)^2\bigg]-\frac{\beta\gamma}{\nu
438: D}\frac{\partial c}{\partial X}\bigg[\Small^2\frac{\partial^2 c}{\partial X^2}+\frac{\partial^2 c}{\partial Z^2}\bigg],
439: %
440: %
441: %
442: %
443: \label{eq:thin_film1}
444: \end{multline}
445: %
446: %
447: %
448: %
449: %
450: %
451: %
452: \begin{multline}
453: %
454: \Small^3 Re \left(\frac{\partial W}{\partial T} + U\frac{\partial
455: W}{\partial X} + W\frac{\partial W}{\partial Z}\right) 
456: %
457: %
458: = -\frac{\partial}{\partial Z}\left(P+\Phi\right)+\Small^4\frac{\partial^2
459: W}{\partial X^2}+\Small^2\frac{\partial^2 W}{\partial Z^2}  \\
460: %
461: %
462: -\tfrac{1}{2}\frac{\beta\gamma}{\nu D}\frac{\partial}{\partial
463: Z}\bigg[\Small^2\left(\frac{\partial c}{\partial X}\right)^2+\left(\frac{\partial
464: c}{\partial Z}\right)^2\bigg]
465: %
466: %
467: -\frac{\beta\gamma}{\nu D}\frac{\partial c}{\partial Z}\bigg[\Small^2\frac{\partial^2
468: c}{\partial X^2}+\frac{\partial^2 c}{\partial Z^2}\bigg],
469: %
470: %
471: %
472: \label{eq:thin_film2}
473: \end{multline}
474: %
475: %
476: where
477: %
478: %
479: \begin{equation}
480: Re = \frac{\Small D}{\nu} = \frac{\left(\Small D / h_0\right)h_0}{\nu} =
481: O\left(1\right).
482: \end{equation}
483: %
484: The choice of ordering for the Reynolds number $Re$ allows us to recover
485: half-Poiseuille flow at $O\left(1\right)$.  We delay choosing the ordering
486: of the dimensionless group $\beta\gamma/D\nu$ until we have examined the
487: concentration equation, which in nondimensional form is
488: %
489: %
490: %
491: %
492: \begin{multline}
493: %
494: %
495: \Small^2 \left(\frac{\partial c}{\partial T} + U\frac{\partial c}{\partial
496: X} + W\frac{\partial c}{\partial Z}\right)\\
497: %
498: %
499: = \Small^2 \frac{\partial^2}{\partial X^2}\left(c^3-c\right)+\frac{\partial^2}{\partial Z^2}\left(c^3-c\right)
500: %
501: %
502: -\Small^4 C_{\mathrm{n}}^2\frac{\partial^4c}{\partial X^4}-C_{\mathrm{n}}^2\frac{\partial^4
503: c}{\partial Z^4}
504: %
505: %
506: -2\Small^2 C_{\mathrm{n}}^2 \frac{\partial^2}{\partial X^2}\frac{\partial
507: c}{\partial Z^2},
508: %
509: %
510: \label{eq:conc0}
511: \end{multline}
512: %
513: %
514: %
515: %
516: where $C_{\mathrm{n}}^2=\gamma/h_0^2$.  By switching off the
517: backreaction in the momentum equations (corresponding to
518: $\beta\gamma/D\nu\rightarrow\infty$), we find the trivial solution to
519: the momentum equations, $U = W = \partial_X\left(P+\Phi\right)
520: = \partial_Z\left(P+\Phi\right) = 0$, $H = 1$.  The concentration
521: boundary conditions are then $c_Z = c_{ZZZ} = 0$ on $Z = 0,1$ which
522: forces $c_Z\equiv0$ so that the Cahn--Hilliard equation is simply
523: %
524: %
525: \[
526: \frac{\partial c}{\partial T} = \frac{\partial^2}{\partial X^2}\left(c^3-c\right)-\Small^2 C_{\mathrm{n}}^2\frac{\partial^4c}{\partial
527: X^4}.
528: \]
529: %
530: To make the lubrication approximation consistent, we take 
531: %
532: \begin{equation}
533: \Small C_{\mathrm{n}}= \tilde{C}_{\mathrm{n}}= \Small\sqrt{\gamma} /
534: h_0 = O\left(1\right).
535: \label{eq:gamma_h}
536: \end{equation}
537: %
538: %
539: %
540: %
541: We now carry out a long-wavelength approximation to
542: Eq.~\eqref{eq:conc0}, writing $U=U_0+O\left(\Small\right)$,
543: $W=W_0+O\left(\Small\right)$, $c=c_0+\Small c_1+\Small^2 c_2+\ldots$.
544: We examine the boundary conditions on $c(\bm{x},t)$ first.
545: They are $\hat{\bm{n}}\cdot\nabla c =
546: \hat{\bm{n}}\cdot\nabla\nabla^2c=0$ on $Z=0,H$; on $Z=0$ these
547: conditions are simply $\partial_Z c = \partial_{ZZZ} c=0$, while on
548: $Z=H$ the surface derivatives are determined by the relations
549: %
550: %
551: \[
552: \hat{\bm{n}}\cdot\nabla\ \propto\ -\Small^2 H_X\partial_X+\partial_Z,
553: \]
554: %
555: %
556: \[
557: \hat{\bm{n}}\cdot\nabla\nabla^2\ \propto\ -\Small^4
558: H_X\partial_{XXX}-\Small^2H_X\partial_X\partial_{ZZ}+\Small^2\partial_{XX}\partial_Z+\partial_{ZZZ}.
559: \]
560: Thus, the BCs on $c_0$ are simply $\partial_Z c_0 = \partial_{ZZZ} c_0=0$
561: on $Z=0,H$, which forces $c_0=c_0\left(X,T\right)$.  Similarly, we
562: find  $c_1=c_1\left(X,T\right)$, and
563: %
564: %
565: \[
566: \frac{\partial c_2}{\partial Z}=Z\frac{H_X}{H}\frac{\partial
567: c_0}{\partial X},\qquad
568: %
569: %
570: \frac{\partial^2 c_2}{\partial Z^2}=\frac{H_X}{H}\frac{\partial
571: c_0}{\partial X}, \qquad \text{for any }Z\in \left[0,H\right].
572: \]
573: %
574: %
575: %
576: In the same manner, we derive the results $\partial_{ZZZZ}c_2=\partial_{ZZZZ}c_3=0$.
577: Using these facts, Eq.~\eqref{eq:conc0} becomes
578: %
579: %
580: %
581: \begin{multline*}
582: %
583: %
584: \frac{\partial c_0}{\partial T} + U_0\frac{\partial c_0}{\partial
585: X} =
586: %
587: %
588: \\
589: \frac{\partial^2}{\partial X^2}\left(c_0^3-c_0\right)
590: %
591: %
592: -\tilde{C}_{\mathrm{n}}^2\frac{\partial^4c}{\partial X^4}
593: %
594: %
595: +\left(3c_0^2-1\right)\frac{H_X}{H}\frac{\partial c_0}{\partial X}
596: %
597: %
598: -2\tilde{C}_{\mathrm{n}}^2\frac{\partial^2}{\partial X^2}\frac{H_X}{H}\frac{\partial
599: c_0}{\partial X}
600: %
601: %
602: -{\tilde{C}_{\mathrm{n}}^2}\frac{\partial^4 c_4}{\partial Z^4}.
603: %
604: %
605: %
606: %
607: %
608: \end{multline*}
609: %
610: %
611: %
612: %
613: We now integrate this equation from $Z=0$ to $H$ and use the boundary conditions
614: %
615: %
616: %
617: \[
618: \frac{\partial^3 c_4}{\partial Z^3}=0\quad\text{on}\quad Z=0,
619: \]
620: %
621: %
622: %
623: \begin{equation*}
624: %
625: %
626: \frac{\partial^3 c_4}{\partial Z^3}=
627: %
628: %
629: H_X\frac{\partial^3 c_0}{\partial X^3}+
630: %
631: %
632: H_X\frac{\partial}{\partial X}\left(\frac{H_X}{H}\frac{\partial c_0}{\partial
633: X}\right)
634: %
635: %
636: -H\frac{\partial^2}{\partial X^2}\left(\frac{H_X}{H}\frac{\partial c_0}{\partial
637: X}\right)\quad\text{on}\quad Z=H.
638: \end{equation*}
639: %
640: %
641: %
642: %
643: After rearrangement, the concentration equation becomes
644: %
645: %
646: %
647: %
648: \begin{multline*}
649: %
650: %
651: H\frac{\partial c_0}{\partial T}+H\langle U_0\rangle\frac{\partial c_0}{\partial
652: X} = 
653: %
654: %
655: \\
656: H\frac{\partial^2}{\partial X^2}\left[c_0^3-c_0-\tilde{C}_{\mathrm{n}}^2\frac{\partial^2
657: c_0}{\partial X^2}-\tilde{C}_{\mathrm{n}}^2\frac{H_X}{H}\frac{\partial c_0}{\partial
658: X}\right]
659: %
660: %
661: +\frac{\partial H}{\partial X}\frac{\partial}{\partial X}\left[c_0^3-c_0-\tilde{C}_{\mathrm{n}}^2\frac{\partial^2
662: c_0}{\partial X^2}-\tilde{C}_{\mathrm{n}}^2\frac{H_X}{H}\frac{\partial c_0}{\partial
663: X}\right],
664: \end{multline*}
665: %
666: %
667: %
668: where
669: %
670: %
671: \[
672: \langle U_0\rangle = \frac{1}{H}\int_0^H U_0\left(X,Z,T\right)dZ
673: \]
674: %
675: %
676: is the vertically-averaged velocity.  Introducing
677: %
678: \[
679: \mu = c_0^3-c_0-\frac{\tilde{C}_{\mathrm{n}}^2}{H}\frac{\partial}{\partial X}\left(H\frac{\partial c_0}{\partial
680: X}\right),
681: \]
682: %
683: %
684: %
685: the thin-film Cahn--Hilliard equation becomes
686: %
687: %
688: %
689: \begin{equation}
690: \frac{\partial c_0}{\partial T}+\langle U_0\rangle\frac{\partial c_0}{\partial
691: X} = \frac{1}{H}\frac{\partial}{\partial X}\left(H\frac{\partial\mu}{\partial
692: X}\right).
693: \label{eq:conc_eqn_ornament}
694: \end{equation}
695: 
696: We are now able to perform the long-wavelength approximation to Eqs.~\eqref{eq:thin_film1}
697: and~\eqref{eq:thin_film2}.  At lowest order, Eq.~\eqref{eq:thin_film2} is~$\partial_ Z\left(P+\Phi\right)=0$,
698: since~$c_0=c_0(X,T)$, and hence
699: %
700: %
701: \[
702: P+\Phi = P_{\mathrm{surf}}+\Phi_{\mathrm{surf}}\equiv P\left(X,h(x,t),T\right)+\Phi\left(X,h(x,t),T\right).
703: \]
704: %
705: %
706: %
707: By introducing the backreaction strength
708: %
709: %
710: \begin{equation}
711: r=\frac{\Small^2\beta\gamma}{D\nu}=O\left(1\right),
712: \end{equation}
713: %
714: %
715: equation~\eqref{eq:thin_film1} becomes
716: %
717: %
718: %
719: \[
720: \frac{\partial^2 U_0}{\partial Z^2}=\frac{\partial}{\partial X}\left(P_{\mathrm{surf}}+\Phi_{\mathrm{surf}}\right)+r\frac{\partial}{\partial
721: X}\left(\frac{\partial c_0}{\partial X}\right)^2+r\frac{\partial c_0}{\partial
722: X}\frac{\partial^2 c_2}{\partial Z^2}.
723: \]
724: %
725: %
726: %
727: Using $\partial_{ZZ}c_2=\left(H_X/H\right)\left(\partial c_0/\partial
728: X\right)$ this becomes
729: %
730: %
731: %
732: \begin{equation}
733: \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
734: X}\left[H\left(\frac{\partial c_0}{\partial X}\right)^2\right].
735: \label{eq:d2U}
736: \end{equation}
737: %
738: %
739: %
740: At lowest order, the BC~\eqref{eq:BC_stress} reduces to
741: %
742: %
743: \[
744: \frac{\partial U_0}{\partial
745:   Z}=\frac{\partial\Sigma}{\partial{X}}\quad\text{on}\quad Z=H,
746: \]
747: %
748: %
749: %
750: which combined with Eq.~\eqref{eq:d2U} yields the relation
751: %
752: %
753: %
754: \[
755: \frac{\partial U_0}{\partial Z}=\frac{\partial\Sigma}{\partial X}+\left(Z-H\right)\bigg\{
756: %
757: %
758: \frac{\partial}{\partial X}\left(P_{\mathrm{surf}}+\Phi_{\mathrm{surf}}\right)+\frac{r}{H}\frac{\partial}{\partial
759: X}\left[H\left(\frac{\partial c_0}{\partial X}\right)^2\right]
760: \bigg\}.
761: %
762: %
763: \]
764: %
765: %
766: %
767: Here $\Sigma$ is the dimensionless, spatially-varying surface tension.  Making
768: use of the BC $U_0=0$ on $Z=0$ and integrating again, we obtain the result
769: %
770: %
771: %
772: \begin{equation}
773: U_0\left(X,Z,T\right)=Z\frac{\partial\Sigma}{\partial X}+\left(\tfrac{1}{2}Z^2-HZ\right)\bigg\{
774: %
775: %
776: \frac{\partial}{\partial X}\left(P_{\mathrm{surf}}+\Phi_{\mathrm{surf}}\right)+\frac{r}{H}\frac{\partial}{\partial
777: X}\left[H\left(\frac{\partial c_0}{\partial X}\right)^2\right]
778: \bigg\}.
779: %
780: %
781: \label{eq:U_0}
782: \end{equation}
783: %
784: %
785: %
786: The vertically-averaged velocity is therefore
787: %
788: %
789: \[
790: \langle U_0\rangle = \tfrac{1}{2}H\frac{\partial\Sigma}{\partial X}-\tfrac{1}{3}H^2\bigg\{
791: %
792: %
793: \frac{\partial}{\partial X}\left(-\frac{1}{C}\frac{\partial^2 H}{\partial X^2}+\Phi_{\mathrm{surf}}\right)+\frac{r}{H}\frac{\partial}{\partial
794: X}\left[H\left(\frac{\partial c_0}{\partial X}\right)^2\right]
795: \bigg\},
796: \]
797: %
798: %
799: where we used the standard Laplace--Young free-surface boundary
800: condition to eliminate the pressure, and
801: %
802: %
803: \begin{equation}
804: C = \frac{\nu\rho D}{h_0\sigma_0\Small^2}=O\left(1\right).
805: \label{eq:C}
806: \end{equation}
807: %
808: %
809: Finally, by integrating the continuity equation in the $Z$-direction, we
810: obtain, in a standard manner, an equation for free-surface variations,
811: %
812: %
813: \begin{equation}
814: \frac{\partial H}{\partial X}+\frac{\partial}{\partial X}\left(H\langle U_0\rangle\right)=0.
815: \label{eq:h_eqn_ornament}
816: \end{equation}
817: %
818: %
819: 
820: 
821: Let us assemble our results, restoring the lower-case fonts and omitting
822: ornamentation over the constants.  The height equation~\eqref{eq:h_eqn_ornament}
823: becomes
824: %
825: %
826: \begin{subequations}
827: %\begin{gather}
828: \begin{equation}
829: \frac{\partial h}{\partial t}+\frac{\partial J}{\partial x}=0,
830: \end{equation}
831: %
832: %
833: %
834: while the concentration equation~\eqref{eq:conc_eqn_ornament} becomes
835: %
836: %
837: %
838: \begin{equation}
839: \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),
840: \end{equation}
841: %
842: %
843: %
844: where
845: %
846: %
847: %
848: \begin{equation}
849: J=\tfrac{1}{2}h^2\frac{\partial\sigma}{\partial{x}}-\tfrac{1}{3}h^3\bigg\{\frac{\partial}{\partial{x}}\left(-\frac{1}{C}\frac{\partial^2{h}}{\partial{x}^2}
850: +\phi\right)+\frac{r}{h}\frac{\partial}{\partial{x}}\left[h\left(\frac{\partial{c}}{\partial{x}}\right)^2\right]\bigg\},
851: \end{equation}
852: %
853: %
854: and
855: %
856: %
857: \begin{equation}
858: \mu=c^3-c-C_{\mathrm{n}}^2\frac{1}{h}\frac{\partial}{\partial{x}}\left(h\frac{\partial{c}}{\partial{x}}\right),
859: \end{equation}%
860: \label{eq:model}%
861: \end{subequations}%
862: %
863: %
864: %
865: and where we have the nondimensional constants
866: %
867: %
868: %
869: \begin{equation}
870: r=\frac{\Small^2\beta\gamma}{D\nu},\qquad C_{\mathrm{n}}=\frac{\Small\sqrt{\gamma}}{h_0},\qquad
871: C=\frac{\nu\rho D}{h_0\sigma_0\Small^2}.
872: \end{equation}
873: %
874: %
875: %
876: %
877: These are the thin-film NSCH equations.  The integral quantities defined in Eq.~\eqref{eq:integral_quantities} are manifestly conserved,
878: while the free surface and concentration are coupled.
879: 
880: 
881: We note that the relation
882: $C_{\mathrm{n}}=\delta\sqrt{\gamma}/h_0=O\left(1\right)$ is the
883: condition that the mean thickness of the film be much smaller than the
884: transition layer thickness.  In experiments involving the smallest
885: film thicknesses attainable ($10^{-8}$ m)~\cite{Sung1996}, this
886: condition is automatically satisfied.  The condition is also realised
887: in ordinary thin films when external effects such as the air-fluid and
888: fluid-substrate interactions do not prefer one binary fluid component
889: or another.  In this case, the vertical extent of the domains becomes
890: comparable to the film thickness at late times, the thin film behaves
891: in a quasi two-dimensional way, and the model equations are
892: applicable.
893: 
894: 
895: The choice of potential $\phi$ determines the behaviour of solutions.
896: If interactions between the fluid and the substrate and air interfaces
897: are important, the potential should take account of the Van der Waals
898: forces present.  A simple model potential is thus
899: \[
900: \phi = Ah^{-n},
901: \]
902: where $A$ is the dimensionless Hamakar constant and typically
903: $n=3$~\cite{Oron1997}.  Here $A$ can be positive or negative, with
904: positivity indicating a net attraction between the fluid and the
905: substrate and negativity indicating a net repulsion.  This choice of
906: potential can also have a regularising effect, preventing a
907: singularity or rupture from occurring in Eq.~\eqref{eq:model}.
908:   
909: For $\phi=-|A|/h^3$ (repulsive Van der Waals interactions), the system of
910: equations~\eqref{eq:model} has a Lyapunov functional $F=F_1+F_2$, where
911: %
912: %
913: \[
914: F_1=\int{dx}\, \left[\frac{1}{2C}\left(\frac{\partial{h}}{\partial{x}}\right)^2+\frac{|A|}{2h^2}\right],\qquad
915: %
916: F_2 = \frac{r}{C_{\mathrm{n}}^2}\int{dx}\,  h\left[\tfrac{1}{4}\left(c^2-1\right)^2+\frac{C_{\mathrm{n}}^2}{2}\left(\frac{\partial{c}}{\partial{x}}\right)^2\right].
917: \]  
918: %
919: %
920: %
921: By differentiating these expressions with respect to time, we obtain
922: the relation
923: %
924: %
925: %
926: \begin{multline}
927: \dot{F}_1+\dot{F}_2
928: %
929: %
930: \\
931: %
932: %
933: =-\tfrac{1}{3} \int{dx}\,  h^3\bigg\{\frac{\partial}{\partial x}\left(\frac{1}{C}\frac{\partial^2h}{\partial
934: x^2}+\frac{|A|}{h^3}\right)-\frac{r}{h}\frac{\partial}{\partial x}\left[h\left(\frac{\partial
935: c}{\partial x}\right)^2\right]  \bigg\}^2
936: %
937: -\int{dx}\,  h\left(\frac{\partial\mu}{\partial x}\right)^2,
938: %
939: %
940: \label{eq:fe_decay}
941: \end{multline}
942: %
943: %
944: which is nonpositive for nonnegative $h$.  This fact is the key to
945: the analytic results of the next section. 
946: 
947: \section{Existence of solutions to the model equations}
948: \label{sec:existence}
949: 
950: \noindent In this section we prove that solutions to the model equations
951: do indeed exist.  We set $C=\tfrac{1}{3}, r=|A|=1$ in Eqs.~\eqref{eq:model}
952: and focus on the resulting equation set
953: %
954: %
955: %
956: %
957: \begin{subequations}
958: \begin{equation}
959: %
960: \frac{\partial h}{\partial t}=
961: %
962: -\frac{\partial}{\partial x}\left[f(h)\frac{\partial^{3}h}{\partial{x^3}}\right]+
963: %
964: \frac{\partial}{\partial x}\left[\frac{1}{g(h)}\frac{\partial h}{\partial{x}}\right]+
965: %
966: \frac{\partial}{\partial x}\bigg\{\frac{f(h)}{g(h)}\frac{\partial}{\partial{x}}\left[g(h)\left(\frac{\partial{c}}{\partial{x}}\right)^2\right]\bigg\},
967: \end{equation}
968: %
969: %
970: %\vskip -0.2in
971: \begin{multline}
972: \frac{\partial}{\partial t}\left(c g(h)\right)=
973: %
974: -\frac{\partial}{\partial x}\left[cf(h)\frac{\partial^{3}h}{\partial{x^3}}\right]+
975: %
976: \frac{\partial}{\partial x}\left[\frac{c}{g(h)}\frac{\partial h}{\partial{x}}\right]+
977: %
978: \frac{\partial}{\partial x}\bigg\{c\frac{f(h)}{g(h)}\frac{\partial}{\partial{x}}\left[g(h)\left(\frac{\partial{c}}{\partial{x}}\right)^2\right]\bigg\}
979: \\
980: +\frac{\partial}{\partial x}\bigg\{g(h)\frac{\partial}{\partial
981: x}\left[c^3-c-\frac{1}{g(h)}\frac{\partial}{\partial
982: x}\left(g(h)\frac{\partial c}{\partial x}\right)\right]\bigg\},
983: \label{eq:model_proof_c}%
984: \end{multline}%
985: \label{eq:model_proof}%
986: \end{subequations}%
987: %
988: %
989: %
990: where
991: %
992: %
993: %
994: \[
995: f(h)=h^3,\qquad g(h)=h.
996: \]
997: %
998: %
999: %
1000: The equations are defined on a periodic domain $\Omega=[0,L]$, while
1001: the initial conditions are
1002: %
1003: %
1004: \begin{equation}
1005: h(x,0)=h_0(x)>0,\qquad c(x,0)=c_0(x),\qquad
1006: h_0(x),\,c_0(x)\in \Sobo^{2,2}(\Omega),
1007: \label{eq:initial_data}
1008: \end{equation}
1009: %
1010: %
1011: We shall prove that the solutions to this equation pair exist in the strong
1012: sense; however, we shall need the definition of weak solutions:
1013: %
1014: %
1015: %
1016: \begin{quote}
1017: A pair $\left(h,c\right)$ is a \emph{weak solution} 
1018: of Eq.~\eqref{eq:model_proof} if the following integral relations hold:
1019: %
1020: %
1021: %
1022: \begin{multline*}
1023: \int_0^{T_0}dt\int_\Omega dx\,\varphi_t h=
1024: \\
1025: \int_0^{T_0}dt\int_\Omega dx\,\varphi_x\bigg\{-f(h)\frac{\partial^{3}h}{\partial{x^3}}+
1026: %
1027: \frac{1}{g(h)}\frac{\partial h}{\partial{x}}+
1028: %
1029: \frac{f(h)}{g(h)}\frac{\partial}{\partial{x}}\left[g(h)\frac{\partial}{\partial{x}}\left(\frac{\partial{c}}{\partial{x}}\right)^2\right]\bigg\},
1030: \end{multline*}
1031: %
1032: and
1033: %
1034: %
1035: \begin{multline}
1036: \int_0^{T_0}dt\int_\Omega dx\,\psi_t cg(h)=
1037: \\
1038: \int_0^{T_0}dt\int_\Omega dx\,\psi_x\bigg\{-cf(h)\frac{\partial^{3}h}{\partial{x^3}}+
1039: %
1040: \frac{c}{g(h)}\frac{\partial h}{\partial{x}}+
1041: %
1042: %
1043: c\frac{f(h)}{g(h)}\frac{\partial}{\partial{x}}\left[g(h)\frac{\partial}{\partial{x}}\left(\frac{\partial{c}}{\partial{x}}\right)^2\right]\bigg\}
1044: %
1045: \\
1046: %
1047: +\int_0^{T_0}dt\int_\Omega dx\,\psi_x \bigg\{g(h)\frac{\partial}{\partial
1048: x}\left[c^3-c-\frac{1}{g(h)}\frac{\partial}{\partial{x}}\left(g(h)\frac{\partial{c}}{\partial{x}}\right)\right]\bigg\},
1049: \label{eq:weak_sln}
1050: \end{multline}
1051: %
1052: %
1053: %
1054: where $T_0>0$ is any time, and $\varphi\left(x,t\right)$ and $\psi\left(x,t\right)$
1055: are arbitrary differentiable test functions that are periodic on $\Omega$
1056: and vanish at $t=0$ and $t=T_0$.  
1057: \end{quote}
1058: %
1059: %
1060: %
1061: In a series of steps in Secs.~\ref{sec:regularization}--\ref{sec:uniqueness},
1062: we prove this result:
1063: %
1064: %
1065: \begin{quote}
1066: %
1067:   \emph{Given the initial data in Eq.~\eqref{eq:initial_data},
1068:     Eqs.~\eqref{eq:model_proof} possess a strong solution endowed with
1069:     the following regularity properties:
1070: %
1071: %
1072: %
1073: \[
1074: \left(h,c\right)\in L^{\infty}\left(0,T_0;\Sobo^{2,2}(\Omega)\right)
1075: \cap L^2\left(0,T_0;\Sobo^{4,2}(\Omega)\right)
1076: \cap C^{\frac{3}{2},\frac{1}{8}}\left(\Omega\times\left[0,T_0\right]\right).
1077: \]
1078: %
1079: %
1080: }
1081: \end{quote}
1082: %
1083: The outline of the proof is as follows: In
1084: Sec.~\ref{sec:regularization} we introduce a regularised version of
1085: Eqs.~\eqref{eq:model_proof}, whose solution we approximate by a
1086: Galerkin sum in Sec.~\ref{sec:galerkin}.  In
1087: Sec.~\ref{sec:a_priori_bds}, we obtain \emph{a priori} bounds on
1088: various norms of the approximate solution.  Crucially, we show that
1089: the free-surface height is always positive.  This enables us to
1090: continue the approximate solution in the time interval
1091: $\left[0,T_0\right]$.  In Secs.~\ref{sec:equicontinuity}
1092: and~\ref{sec:convergence} we show that the Galerkin sum converges to a
1093: solution of the unapproximated equations, in the appropriate limit.
1094: Finally, in Secs.~\ref{sec:regularity} and~\ref{sec:uniqueness} we
1095: discuss the regularity and uniqueness properties of the solution.
1096: 
1097: 
1098: 
1099: \subsection{Regularisation of the problem}
1100: \label{sec:regularization}
1101: 
1102: \noindent We introduce regularised functions
1103: $f_\varepsilon(s)$ and $g_\varepsilon(s)$ such
1104: that $\lim_{\varepsilon\rightarrow0}f_\varepsilon(s)=
1105: f(s)$, and
1106: $\lim_{\varepsilon\rightarrow0,s\geq0}g_\varepsilon(s)={g}(s)$.
1107: For now we do not specify $f_\varepsilon(s)$, although we
1108: mention that a suitable choice of $f_\varepsilon(s)$ will
1109: cure the degeneracy of the fourth-order term in the height equation.
1110: On the other hand, we require that~$g_\varepsilon(s)$ have the properties:
1111: (i) $g_\varepsilon(s)=s+\varepsilon$, for $s\geq0$;
1112: (ii) $g_\varepsilon(s)>0$, for $s<0$;
1113: (iii) $\lim_{s\rightarrow-\infty}g_\varepsilon(s)=\tfrac{1}{2}\varepsilon$;
1114: and (iv) $g_\varepsilon(s)$ is at least $C^3$.
1115: 
1116: From Eqs.~\eqref{eq:model_proof}, the regularised PDEs we study are
1117: %
1118: %
1119: \begin{subequations}
1120: \begin{align}
1121: %\begin{split}
1122: h_t&=-J_{\varepsilon,x}\,,
1123: \\
1124: %
1125: %
1126: \left(cg_\varepsilon(h)\right)_t&=-\left(c J_\varepsilon\right)_x-\left(g_\varepsilon(h)\mu_{\varepsilon,x}\right)_x,
1127: \label{eq:reg_pde_c}%
1128: %\end{split}%
1129: \end{align}%
1130: \label{eq:reg_pde}%
1131: \end{subequations}%
1132: %
1133: %
1134: %
1135: where
1136: %
1137: %
1138: \[
1139: \mu_\varepsilon=c^3-c-\frac{1}{g_\varepsilon(h)}\left(g_\varepsilon(h)c_x\right)_x
1140: \]
1141: and
1142: \[
1143: J_\varepsilon=f_\varepsilon(h)h_{xxx}-\frac{1}{g_\varepsilon(h)}h_x-\frac{f_\varepsilon(h)}{g_\varepsilon(h)}\left(g_\varepsilon(h)c_{x}^2\right)_x\,.
1144: \]
1145: %
1146: %
1147: Equation~\eqref{eq:reg_pde_c} can also be written as 
1148: %
1149: %
1150: \begin{equation}
1151: c_t=-\frac{1}{g_\varepsilon(h)}J_\varepsilon c_x-\frac{1}{g_\varepsilon(h)}\left(g_\varepsilon(h)\mu_{\varepsilon,x}\right)_x-\frac{c}{g_\varepsilon(h)}\left[J_{\varepsilon,x}+g_\varepsilon'(h)h_t\right];
1152: \label{eq:conc_no_flux}
1153: \end{equation}
1154: %
1155: %
1156: this form of the concentration equation will be useful in
1157: Sec.~\ref{sec:uniqueness}.
1158: 
1159: 
1160: \subsection{The Galerkin approximation}
1161: \label{sec:galerkin}
1162: 
1163: \noindent We choose a complete orthonormal basis on the interval
1164: $\Omega$, with periodic boundary conditions.  Let us denote the basis
1165: by $\{\phi_i(x)\}_{i\in\mathbb{N}_0}$.  We consider the
1166: finite vector space $\text{Span}\{\phi_0,\ldots,\phi_n\}$.  For
1167: convenience, let us take the $\phi_i(x)$'s to be the
1168: eigenfunctions of the Laplacian on $\left[0,L\right]$ with periodic
1169: boundary conditions, and corresponding eigenvalues $-\lambda_i^2$.
1170: Let $\phi_0$ be the constant eigenfunction.  We construct
1171: approximate solutions to the PDEs~\eqref{eq:reg_pde} as finite sums,
1172: %
1173: %
1174: \[
1175: h_n(x,t)=\sum_{i=0}^n\eta_{n,i}(t)\phi_i(x),
1176: \qquad c_n(x,t)=\sum_{i=0}^n\gamma_{n,i}(t)\phi_i(x).
1177: \]
1178: %
1179: %
1180: If the (smooth) initial data are given as
1181: %
1182: %
1183: \[
1184: h(x,0)=h_0(x)=\sum_{i=0}^\infty\eta_i^0\phi_i(x)>0,\qquad
1185: c(x,0)=c_0(x)=\sum_{i=0}^\infty\gamma_i^0\phi_i(x),
1186: \]
1187: %
1188: %
1189: %
1190: then the initial data for the Galerkin approximation are
1191: %
1192: %
1193: \[
1194: h_n(x,0)=h_n^0(x)=\sum_{i=0}^n\eta_{i}^0\phi_i(x),\qquad
1195: c_n(x,0)=c_n^0(x)=\sum_{i=0}^n\gamma_{i}^0\phi_i(x),
1196: \]
1197: %
1198: %
1199: %
1200: and the initial data of the Galerkin approximation converge strongly in the
1201: $L^2(\Omega)$ norm to the initial data of the unapproximated problem.
1202: Thus, there is a $n_0\in\mathbb{N}$ such that $h_n^0(x)>0$, everywhere
1203: in $\Omega$, for all $n>n_0$.  Henceforth we work with Galerkin approximations
1204: with $n>n_0$. 
1205: 
1206: Substitution of $h_n=\sum_{i=0}^n\eta_{n,i}\phi_i$  into a weak form of the
1207: $h$-equation yields
1208: %
1209: %
1210: \begin{equation}
1211: \frac{d}{dt}\pairing{h_n}{\phi_j}=\pairing{J_{\varepsilon,n}}{\phi_{j,x}},
1212: \label{eq:weak_h}
1213: \end{equation}
1214: %
1215: %
1216: where
1217: %
1218: %
1219: \[
1220: J_\varepsilon\left(h_n,c_n\right)=f_\varepsilon(h_n)h_{n,xxx}-\frac{1}{g_\varepsilon(h_n)}h_{n,x}-\frac{f_\varepsilon(h_n)}{g_\varepsilon(h_n)}\left(g_\varepsilon\left(h_{n}\right)c_{n,x}^2\right)_x,
1221: \]
1222: is the flux for the regularised $h$-equation, and $\pairing{\varphi(x)}{\psi(x)}$
1223: is the pairing $\int_\Omega\varphi\psi\,{dx}$.
1224: %
1225: %
1226: We recast Eq.~\eqref{eq:weak_h} as
1227: %
1228: %
1229: \[
1230: \frac{d\eta_{n,j}}{dt}=\pairing{J_{\varepsilon,n}}{\phi_{j,x}}=\Phi_{n,j}\left(\eta_n,\gamma_n\right),
1231: \]
1232: %
1233: %
1234: %
1235: where the function $\Phi_n\left(\eta_n,\gamma_n\right)$ depends on
1236: $\eta_n=\left(\eta_{n,0},\ldots,\eta_{n,n}\right)$ and
1237: $\gamma_n=\left(\gamma_{n,0},\ldots,\gamma_{n,n}\right)$, and is
1238: locally Lipschitz in its variables.  This Lipschitz property arises
1239: from the fact that the regularised flux, evaluated at the Galerkin
1240: approximation, is a composition of Lipschitz continuous functions, and
1241: therefore, is itself Lipschitz continuous.
1242: 
1243: Similarly, substitution of $c_n=\sum_{i=0}^n\gamma_{n,i}\phi_i$ into the
1244: weak form of the $c$-equation~\eqref{eq:reg_pde_c} yields
1245: %
1246: %
1247: \begin{equation}
1248: \frac{d}{dt}\pairing{g (h_n)c_n}{\phi_j}=\pairing{K_{\varepsilon,n}}{\phi_{j,x}},
1249: \label{eq:weak_c}
1250: \end{equation}
1251: %
1252: %
1253: where
1254: %
1255: %
1256: \begin{align*}
1257:   K_\varepsilon\left(h_n,c_n\right) &=
1258:   c_nf_\varepsilon(h_n)h_{n,xxx} -
1259:   \frac{c_n}{g_\varepsilon(h_n)}h_x-c_{n}
1260:   \frac{f_\varepsilon(h_n)} {g_\varepsilon(h_n)}
1261:   \left(g_\varepsilon(h_n)
1262:     c_{n,x}^2\right)_x - g (h_n)\mu_{\varepsilon,n,x}\,,\\
1263:   &=
1264:   c_nJ_{\varepsilon,n}-g_\varepsilon(h_n)\mu_{\varepsilon,n,x}\,,
1265: \end{align*}
1266: %
1267: %
1268: %
1269: is the flux for the regularised
1270: $c$-equation~\eqref{eq:reg_pde_c}. 
1271: Rearranging gives
1272: %
1273: %
1274: \[
1275: \pairing{g (h_n)c_{n,t}}{\phi_j}=\pairing{K_{\varepsilon,n}}{\phi_{j,x}}-\pairing{g' (h_n)c_{n}h_{n,t}}{\phi_j},
1276: \]
1277: %
1278: %
1279: %
1280: and the left-hand side can be recast in matrix form as
1281: $\sum_{\eye=0}^nM_{\eye{j}}\dot{\gamma}_{n,\eye}$,
1282: where
1283: %
1284: %
1285: \[
1286: M_{\eye j}=\int_\Omega{dx}\,  g_\varepsilon\left(\sum_\ell \eta_{n,\ell}\phi_\ell\right)\phi_\eye\phi_j,
1287: \] 
1288: %
1289: %
1290: which is manifestly symmetric.  It is positive
1291: definite because given a vector $\left(\xi_0,\ldots,\xi_n\right)$, we
1292: have the relation
1293: %
1294: %
1295: \begin{equation*}
1296: \sum_{\eye,j}\xi_\eye M_{\eye j}\xi_j=\int_\Omega{dx}\,  g_\varepsilon\left(\sum_\ell\eta_{n,\ell}\phi_\ell\right)\sum_{\eye,j}\left(\phi_\eye\xi_\eye\right)\left(\phi_j\xi_j\right)
1297: >0,\quad\text{for }\left(\xi_0,\ldots,\xi_n\right)\neq0,
1298: \end{equation*}
1299: %
1300: %
1301: which follows from the positivity of the regularised function $g_\varepsilon(s)$.
1302:  We therefore have the following equation for $\gamma_{n,j}(t)$,
1303: %
1304: %
1305: %
1306: \begin{equation}
1307: \frac{d\gamma_{n,j}}{dt}=\sum_{\eye=0}^n M_{\eye j}^{-1}\Big[\pairing{K_{\varepsilon,n}}{\phi_{\eye,x}}-\pairing{g_\varepsilon' (h_n)c_{n}h_{n,t}}{\phi_\eye}
1308: \Big].
1309: \label{eq:gamma_dot}
1310: \end{equation}
1311: %
1312: Inspecting the expression for $M_{\eye j}$, $g (h_n)$, and
1313: $K_{\varepsilon,n}$, we see that $\eta_n$ and $\gamma_n$ appear in a
1314: Lipschitz-continuous way in the expression $\sum_{\eye=0}^n M_{\eye
1315:   j}^{-1}\pairing{K_{\varepsilon,n}}{\phi_{\eye,x}}$, while owing to
1316: the imposed smoothness of $g_\varepsilon(s)$, the variables
1317: $\eta_n$, $\gamma_n$ and
1318: $\dot{\eta}_n=\left(\dot{\eta}_{n,0},\ldots,\dot{\eta}_{n,n}\right)$
1319: appear in a Lipschitz-continuous way in the quantity $\sum_{\eye=0}^n
1320: M_{\eye  j}^{-1}\pairing{g_\varepsilon' (h_n)c_{n}h_{n,t}}{\phi_\eye}$.
1321: The vector $\dot{\eta}_n$ can be replaced by the function
1322: $\Phi_n\left(\eta_n,\gamma_n\right)$ and thus we obtain a relation
1323: %
1324: %
1325: \[
1326: \frac{d\gamma_{n,j}}{dt}=\Psi_{n,j}\left(\eta_n,\gamma_n\right),
1327: \]
1328: %
1329: %
1330: %
1331:  in place of Eq.~\eqref{eq:gamma_dot}, where $\Psi_{n,j}\left(\eta_n,\gamma_n\right)$
1332:  is Lipschitz.  We therefore have a system of Lipschitz-continuous equations
1333: %
1334: %
1335: \[
1336: \frac{d\eta_{n,j}}{dt}=\Phi_{n,j}\left(\eta_n,\gamma_n\right),\qquad\frac{d\gamma_{n,j}}{dt}=\Psi_{n,j}\left(\eta_n,\gamma_n\right),
1337: \]
1338: %
1339: %
1340: and thus local existence theory~\cite{DoeringGibbon} guarantees
1341: a solution
1342: for the $\eta_{n,i}$'s
1343: and $\gamma_{n,i}$'s for all times $t$ in a finite interval $0<t<\sigma$.
1344:  This solution is, moreover, unique and continuous.  To continue this approximate
1345:  solution to the PDE problem in Eq.~\eqref{eq:model_proof}
1346:  up to an an arbitrary time $T_0>0$, it is necessary to find \emph{a priori}
1347:  bounds on the approximate local-in-time solution.
1348: 
1349: 
1350: \subsection{\emph{A priori} bounds on the Galerkin approximation}
1351: \label{sec:a_priori_bds}
1352: 
1353: \noindent We identify the free energy
1354: %
1355: %
1356: \[
1357: F = \int_{\Omega}{dx}\,
1358: \left[\tfrac{3}{2}h_x^2+G_\varepsilon(h)\right]
1359: + \int_{\Omega}{dx}\,
1360: g_\varepsilon(h) \left[\tfrac{1}{4} \left(c^2-1\right)^2+\tfrac{1}{2}c_x^2\right],
1361: \]
1362: %
1363: %
1364: where $G_{\varepsilon}''(s)=1/\left[f_\varepsilon(s)g_\varepsilon(s)\right]$.
1365:  Since the Galerkin approximation satisfies the weak form of the PDEs given
1366:  in Eqs.~\eqref{eq:weak_h} and~\eqref{eq:weak_c}, it is possible to obtain
1367:  the free-energy decay law
1368: %
1369: %
1370: \begin{multline}
1371: \frac{dF}{dt}(t)=
1372: -\int_{\Omega-\Omega_-}{dx}\,  f_\varepsilon(h_n)
1373:    \left[-h_{n,xxx}+\frac{h_{n,x}}{g_\varepsilon(h_n)
1374:        h_\varepsilon(h_n)} + \frac{1}{g_\varepsilon
1375:         (h_n)} \left(g_\varepsilon  (h_n)
1376:        c_{n,x}^2\right)_x \right]^2dx \\- \int_\Omega{dx}\,
1377:    g_\varepsilon  (h_n) \mu_{\varepsilon,n,x}^2%\\
1378: + \int_{\Omega_-}{dx}\,[\dots],\qquad
1379:    0\leq t<\sigma,
1380: \label{eq:free_energy_decay}
1381: \end{multline}
1382: %
1383: %
1384: where $\Omega_-(t)=\{x\in\Omega|h_n(x,t)<0\}$.
1385: Now given the time-continuity of $h_n(x,t)$ in
1386: $(0,\sigma)$, and the initial condition $h_n^0(x)>0$ (since
1387: $n>n_0$), there is a time $\sigma_1>0$ such that
1388: $h_n(x,t)>0$ for all $x\in\Omega$ and all
1389: $t\in\left(0,\sigma_1\right)$.  Therefore,
1390: $\Omega_-(t)=\emptyset$ for $t\in\left(0,\sigma_1\right)$, the last
1391: integral in~\eqref{eq:free_energy_decay} vanishes, and hence
1392: %
1393: %
1394: \[
1395: F\left[c_n(x,t),h_n(x,t)\right] \leq
1396: F\left[c_n(x,0),h_n(x,0)\right] \leq
1397: \sup_{\varepsilon,n}F\left[c_n(x,0),h_n(x,0)\right]<\infty,
1398: \]
1399: %
1400: %
1401: for $0< t<\sigma_1$.  Consequently, we obtain the bound $\|h_{n,x}\|_2\leq
1402: k_1$, where $0<t<\sigma_1$, and where $k_1$ depends only on the initial
1403: conditions.  We have Poincar\'e's inequality for $h_{n,x}$,
1404: %
1405: %
1406: %
1407: \[
1408: \|h_n\|_2^2\leq\left[\int_\Omega {dx}\, h_n(x)\right]^2+\left(\frac{L}{2\pi}\right)^2\|h_{n,x}\|_2^2.
1409: \]
1410: %
1411: %
1412: %
1413: Now $\int_\Omega{dx}h_n(x,t)=L\eta_{n,0}(t)$.  Inspection
1414: of Eq.~\eqref{eq:weak_h} shows that $\eta_{n,0}(t)=\eta_{n,0}\left(0\right)=\eta_{0}^0$.
1415:  Thus,
1416: %
1417: %
1418: \[
1419: \|h_n\|_2^2\leq
1420: L^2|\eta_{0}^0|^2+\left(\frac{L}{2\pi}\right)^2k_1\equiv k_2<\infty.
1421: \] 
1422: %
1423: %
1424: Using result~\eqref{eq:result1} from Appendix~\ref{apx:A},
1425: we obtain the bound
1426: %
1427: %
1428: \[
1429: \|h_n\|_\infty\leq\frac{1}{\sqrt{L}}\|h_n\|_2+\sqrt{L}\|h_{n,x}\|_2
1430: \equiv k_3\,.
1431: \]
1432: %
1433: %
1434: %
1435: %
1436: %
1437: Additionally, the following properties hold:
1438: %
1439: %
1440: %
1441: \begin{itemize}
1442: \item The function $h_{n}$ is H\"older continuous in space, with exponent
1443: $\tfrac{1}{2}$\,;
1444: \item $\int_\Omega {dx}\, G_{\varepsilon} (h_n)\leq k_4$\,.
1445: \end{itemize}
1446: %
1447: %
1448: %
1449: These results hold in $0<t<\sigma_1$, and the constants $k_1$,
1450: $k_2$, $k_3$, and $k_4$ are independent of $\varepsilon$, $n$, $\sigma$,
1451: and $\sigma_1$, and in fact depend only on the functions $h_0(x)$
1452: and $c_0(x)$.
1453: 
1454: To continue the estimates to the whole interval
1455: $\left(0,\sigma\right)$, we need to prove that
1456: $h_n\left(\cdot,\sigma_1\right)>0$ almost everywhere (a.e.).  If this
1457: is true, there is a new interval $[\sigma_1,\sigma_2)$,
1458: $\sigma_1<\sigma_2\leq\sigma$, on which $h_n(\cdot,t)>0$
1459: a.e., and we can then provide \emph{a priori} bounds on
1460: $h_n(\cdot,t)$ and $c_n(\cdot,t)$ on the
1461: interval $\left[\sigma_1,\sigma_2\right)$.  It is then possible to
1462: show that $h_n\left(\cdot.,\sigma_2\right)>0$ a.e.\ and thus, by
1463: iteration, we extend the proof to the whole interval
1464: $\left(0,\sigma\right)$, and find that $h_n\left(.,t\right)>0$ a.e.\
1465: on $\left(0,\sigma\right)$.
1466: 
1467: We have the bound
1468: %
1469: %
1470: \begin{equation}
1471: \int_\Omega {dx}\, G_{\varepsilon}(h_n(\cdot,t))\leq k_4,
1472: \label{eq:sigma_1_bound}
1473: \end{equation}
1474: %
1475: %
1476: where $k_4$ depends
1477: only on the initial conditions, and where $0<t<\sigma_1$.  We now
1478: specify $G_\varepsilon(s)$ in more detail.  This function
1479: satisfies the condition
1480: %
1481: %
1482: \[
1483: G_\varepsilon''(s)=\frac{1}{f_\varepsilon(s)\,g_\varepsilon(s)}.
1484: \]
1485: %
1486: %
1487: We take $g_\varepsilon(s)$ to be as defined previously, and we choose a simple
1488: regularisation for $f(s)$:
1489: %
1490: %
1491: \[
1492: f_\varepsilon(s)=g_\varepsilon(s)^3,
1493: \]
1494: %
1495: %
1496: %
1497: which is Lipschitz continuous.
1498: %
1499: %
1500: By defining
1501: %
1502: %
1503: \[
1504: \tilde{G}_\varepsilon(s)=-\int_s^\infty\frac{dr}{f_\varepsilon(r)g_\varepsilon(r)},\qquad
1505: G_\varepsilon(s)=-\int_s^{\infty}{dr}\, \tilde{G}_\varepsilon(r),
1506: \]
1507: %
1508: %
1509: we obtain a function $G_\varepsilon(s)$ that is positive for all
1510: $s\in\left(-\infty,\infty\right)$, and
1511: %
1512: %
1513: %
1514: \[
1515: G_{\varepsilon}(s)=\tfrac{1}{6}\frac{1}{\left(s+\varepsilon\right)^2}\,,
1516: \qquad s\geq0.
1517: \]
1518: %
1519: %
1520: %
1521: Using the boundedness of $G_{\varepsilon}(s)$, and the time-continuity
1522: of $h_n(\cdot,t)$, we apply the Dominated Convergence Theorem,
1523: %
1524: %
1525: \[
1526: \lim_{t\rightarrow\sigma_1}\int_\Omega {dx}\, G_{\varepsilon}(h_n(\cdot,t))=\int_\Omega{dx}\,
1527: \lim_{t\rightarrow\sigma_1}G_{\varepsilon}(h_n(\cdot,t))=\int_{\Omega}{dx}\,
1528: G_{\varepsilon}\left(h_n\left(\cdot,\sigma_1\right)\right)\leq
1529: k_4.
1530: \] 
1531: %
1532: %
1533: Similarly, since the constant $k_1$ in the inequality $\|h_{n,x}\|_2\leq
1534: k_1$, $0\leq t<\sigma_1$ depends
1535: only on the initial data, we extend this last inequality to $t=\sigma_1$,
1536: and thus $h_n(x,\sigma_1)$ is H\"older continuous in space.  
1537: 
1538: In the worst-case scenario, the time $\sigma_1$ is the first time at which
1539: $h_n(x,t)$
1540: touches down to zero, and and we therefore assume for contradiction that $h_n\left(x_0,\sigma_1\right)=0$,
1541: and that $h_n(x,\sigma_1)\geq0$ elsewhere.
1542:  Then, by H\"older continuity, for any $x\in\Omega$ we have the bound
1543:  $0\leq h_n(x,\sigma_1)\leq k_1\left|x-x_0\right|^{\frac{1}{2}}$, and
1544:  thus
1545: %
1546: %
1547: %
1548: \[
1549: \int_\Omega {dx}\, G_{\varepsilon}\left(h_n\left(\cdot,\sigma_1\right)\right)
1550: \geq\frac{k_1}{6}\int_0^L\frac{dx}{|x-x_0|+\varepsilon (2\sqrt{L}+\varepsilon)}.
1551: \]
1552: %
1553: %
1554: Hence,
1555: %
1556: %
1557: %
1558: \begin{multline*}
1559: \frac{6}{k_1}\int_\Omega {dx}\, G_{\varepsilon}\left(h_n\left(\cdot,\sigma_1\right)\right)
1560: \\
1561: \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\}.
1562: \end{multline*}
1563: %
1564: %
1565: %
1566: %
1567: Thus, the integral $\int_\Omega G_\varepsilon(h_n(x,\sigma_1))dx$ can
1568: be made arbitrarily large, which contradicts the
1569: $\varepsilon$-independent bound for this quantity, obtained in
1570: Eq.~\eqref{eq:sigma_1_bound}.  We therefore have the strong condition
1571: that the set on which $h_n\left(\cdot,\sigma_1\right)\leq0$ is empty.
1572: Iterating the argument, we have the important result
1573: %
1574: %
1575: %
1576: \begin{quote}
1577: The set on which $h_n(\cdot,t)\leq0$ is empty, for $0<t<\sigma$.
1578: \end{quote}
1579: %
1580: %
1581: %
1582: Using the same argument, we have an estimate on the minimum value of
1583: $h_n(x,t)$,
1584: %
1585: %
1586: \[
1587: h_{\mathrm{min}}=\min_{x\in\Omega,t\in\left(0,\sigma\right]}h_n(x,t),
1588: \]
1589: %
1590: %
1591: namely,
1592: %
1593: %
1594: %
1595: \[
1596: h_{\mathrm{min}}+\varepsilon\geq -k_1\sqrt{L}+\sqrt{k_1^2L+\frac{k_1^2L}{e^{k_4k_1^2}-1}},
1597: \]
1598: %
1599: %
1600: for all small positive $\varepsilon$.  Thus,
1601: %
1602: %
1603: %
1604: \begin{equation}
1605: h_{\mathrm{min}}\geq M :=-k_1\sqrt{L}+\sqrt{k_1^2L+\frac{k_1^2L}{e^{k_4k_1^2}-1}},
1606: \label{eq:min_h}
1607: \end{equation}
1608: %
1609: %
1610: a lower bound that depends only on the initial data $c_0(x)$ and $h_0(x)$.
1611:  Note
1612: that this result depends on the intimate relationship between the H\"older
1613: continuity of a function and its boundedness in the $W^2$ norm, a relationship
1614: that is true only in one dimension.  Thus, generalisation of this lower bound,
1615: and by extension, long-time existence and uniqueness of solutions, does not
1616: necessarily hold in higher dimensions.
1617: 
1618: 
1619: 
1620: Now, using Eq.~\eqref{eq:min_h} and the boundedness result 
1621: %
1622: %
1623: \[
1624: \int_\Omega{dx}\,
1625: g_\varepsilon(h_n)
1626: \left[\tfrac{1}{4}\left(c_n^2-1\right)^2+\tfrac{1}{2}c_{n,x}^2\right]
1627: \leq k_5\,,
1628: \]
1629: %
1630: %
1631: %
1632: where $k_5$ depends only on the initial data, we obtain an \emph{a priori}
1633: bound on $\|c_{n,x}\|_2^2$,
1634: %
1635: %
1636: \[
1637: \int_\Omega{dx}\,  c_{n,x}^2 \leq \frac{2k_5}{M}\,.
1638: \]
1639: %
1640: %
1641: It is also possible to derive a bound on $\|c_n\|_2^2$.  We have the relation
1642: %
1643: %
1644: \[
1645: \int_\Omega{dx}\,\tfrac{1}{4}\left(c_n^2-1\right)^2\leq \frac{k_5}{M},
1646: \]
1647: %
1648: %
1649: which gives the inequality
1650: %
1651: %
1652: $
1653: \|c_n\|_4^4\leq 2\|c\|_2^2+\left({4k_5}/{M}\right).
1654: $
1655: %
1656: %
1657: Using the H\"older relation $\|c\|_2\leq |\Omega|^{\frac{1}{4}}\|c\|_4$,
1658: we obtain a quadratic inequality in $\|c\|_2^2$,
1659: %
1660: %
1661: %
1662: \[
1663: \|c\|_2^4\leq 2|\Omega|\|c\|_2^2+\frac{4|\Omega|k_5}{M},
1664: \]
1665: %
1666: %
1667: %
1668: with solution
1669: %
1670: %
1671: \[
1672: \|c\|_2^2\leq |\Omega|+\frac{4|\Omega|k_5}{M},
1673: \]
1674: %
1675: %
1676: as required.  From the boundedness of $\|c_{n,x}\|_2$ and $\|c_{n}\|_2$ follows
1677: the relation $\|c_n\|_\infty \leq k_6<\infty$, a result that depends only
1678: on the initial conditions.
1679: %
1680: %
1681: %
1682: Let us recapitulate these results:
1683: %
1684: %
1685: %
1686: \begin{itemize}
1687: \item $\|h_{n,x}\|_2$ is uniformly bounded;
1688: \item $\|h_{n}\|_\infty$ is uniformly bounded;
1689: \item The function $h_{n}$ is H\"older continuous in space, with exponent
1690: $\tfrac{1}{2}$;
1691: \item The function $h_{n}$ is nonzero everywhere and never decreases below
1692: a certain value $M>0$, independent of $n$, $\varepsilon$, and $\sigma$.
1693: \item $\|c_{n,x}\|_2$ is uniformly bounded;
1694: \item $\|c_{n}\|_\infty$ is uniformly bounded;
1695: \item The function $c_{n}$ is H\"older continuous in space, with exponent
1696: $\tfrac{1}{2}$\,.
1697: \end{itemize}
1698: %
1699: %
1700: %
1701: These results are independent of $n$, $\varepsilon$ and $\sigma$, and
1702: hold for $0<t<\sigma$.
1703: %
1704: %
1705: %
1706: %
1707: %
1708: %
1709: \subsection{Equicontinuity and convergence of the Galerkin approximation}
1710: \label{sec:equicontinuity}
1711: 
1712: \noindent  Using Eq.~\eqref{eq:free_energy_decay}, we obtain the bound
1713: %
1714: %
1715: \begin{multline*}
1716: \int_0^t{dt'}\int_\Omega{dx}\, \bigg\{f_\varepsilon(h_n)\left[-h_{n,xxx}+\frac{h_{n,x}}{g_\varepsilon(h_n)f_\varepsilon(h_n)}+\frac{1}{g_\varepsilon(h_n)}\left(g_\varepsilon(h_n)c_{n,x}^2\right)_x\right]^2+g_\varepsilon(h_n)\mu_{\varepsilon,n,x}^2\bigg\}
1717: \\
1718: \leq F\left(0\right),\qquad 0<t<\sigma,
1719: \end{multline*}
1720: %
1721: %
1722: a bound that is independent of $n$, $\sigma$, and $\varepsilon$.  Since the
1723: quantity $\|f_\varepsilon(h_n)\|_\infty=\left(\|h_n\|_\infty+\varepsilon\right)^3$
1724: is bounded above by a constant $A_1$ independent of $n$, $\varepsilon$,
1725: and $\sigma$, we have
1726: %
1727: %
1728: \begin{multline*}
1729: \int_0^tdt'\int_\Omega {dx}\,  J_{\varepsilon,n}^2\\
1730: %
1731: %
1732: \leq A_1\int_0^t{dt'}\int_\Omega{dx}\, 
1733: \bigg\{f_\varepsilon(h_n) \left[-h_{n,xxx} +
1734:   \frac{h_{n,x}}{g_\varepsilon
1735:      (h_n)f_\varepsilon(h_n)} +
1736:   \frac{1}{g_\varepsilon(h_n)}
1737:   \left(g_\varepsilon(h_n) c_{n,x}^2\right)_x\right]^2\bigg\}
1738: %
1739: %
1740: \\
1741: \leq A_1 F\left(0\right)\equiv A_2,
1742: \end{multline*}
1743: %
1744: %
1745: %
1746: and thus
1747: %
1748: %
1749: %
1750: \[
1751: \int_0^tdt'\,\|J_{\varepsilon,n}\|_2^2\leq A_2,\qquad 0< t<\sigma,
1752: \]
1753: %
1754: %
1755: where $A_2$ is independent of $n$, $\varepsilon$, and $\sigma$.  Similarly,
1756: %
1757: %
1758: \[
1759: \int_0^tdt'\,\|g_\varepsilon(h_n)\mu_{\varepsilon,n,x}\|_2^2\leq
1760: A_3,\qquad 0<{t}<\sigma,
1761: \]
1762: %
1763: %
1764: %
1765: and
1766: \[
1767: \int_0^tdt'\,\|c_n J_{\varepsilon,n}\|_2^2\leq A_2\|c_n\|_\infty^2\leq
1768: A_2k_6^2,\qquad 0<t<\sigma,
1769: \]
1770: %
1771: %
1772: %
1773: where $A_2$ and $A_3$ are independent of $n$, $\varepsilon$, and $\sigma$.
1774:  By rewriting the evolution equations as
1775: %
1776: %
1777: \[
1778: h_{n,t}=-J_{\varepsilon,n,x},\qquad \left(g_\varepsilon(h)c\right)_t=-K_{\varepsilon,n,x},
1779: \]
1780: %
1781: %
1782: where $K_{\varepsilon,n}=c_n J_{\varepsilon,n}-g_\varepsilon(h_n)\mu_{\varepsilon,n,x}$,
1783: we see that there are uniform bounds for $\int_0^tdt'\,\|J_{\varepsilon,n}\|_2^2$
1784: and $\int_0^tdt'\,\|K_{\varepsilon,n}\|_2^2$, which depend only on the initial
1785: data $c_0(x)$ and $h_0(x)$.
1786: 
1787: Bernis and Friedman~\cite{Friedman1990} proved the following claim:
1788: %
1789: %
1790: %
1791: \begin{quote}
1792: %
1793: %
1794: %
1795: \emph{Let $\varphi_i\left(x,t\right)$ be a sequence of functions, each of
1796: which weakly satisfies the equation
1797: %
1798: %
1799: \[
1800: \varphi_{i,t}=-J_{i,x},\qquad J_i=J\left(\varphi_i\right).
1801: \]
1802: %
1803: %
1804: If $\varphi_i\left(x,\cdot\right)$ is H\"older continuous (exponent $\tfrac{1}{2}$),
1805: and if the fluxes $J_i$ satisfy
1806: %
1807: %
1808: \[
1809: \int_0^t dt'\,\|J_i\|_2^2\leq B_1,\qquad 0<t<\sigma,
1810: \]
1811: %
1812: %
1813: where $B_1$ is a number independent of the index $i$ and the time $\sigma$,
1814: then there is a constant $B_2$, independent of $i$ and $\sigma$, such that
1815: %
1816: %
1817: \[
1818: \left|\varphi_i\left(\cdot,t_2\right)-\varphi_i\left(\cdot,t_1\right)\right|\leq
1819: B_2\left|t_2-t_1\right|^{\frac{1}{8}},
1820: \]
1821: %
1822: %
1823: for all $t_1$ and $t_2$ in $\left(0,\sigma\right)$.  
1824: }
1825: \end{quote}
1826: %
1827: %
1828: %
1829: We now observe that the fluxes $J_{\varepsilon,n}$, and $K_{\varepsilon,n}$
1830: satisfy the conditions of this theorem, and thus
1831: %
1832: %
1833: \begin{quote}
1834: The functions $h_{n}(\cdot,t)$ and $c_n(\cdot,t)$
1835: are H\"older continuous (exponent $\tfrac{1}{8}$), for $0<t<\sigma$.
1836: \end{quote}
1837: %
1838: %
1839: We therefore have a uniformly bounded and equicontinuous family of
1840: functions $\{\left(h_n,c_n\right)\}_{n=n_0+1}^\infty$.  We also have a
1841: recipe for constructing a uniformly bounded and equicontinuous
1842: approximate solution
1843: $\left(h_n(x,t),c_n(x,t)\right)$, in a small
1844: interval $\left(0,\sigma\right)$.  The recipe can be iterated
1845: step-by-step, and we obtain a uniformly bounded and equicontinuous
1846: family of approximate solutions
1847: $\{\left(h_n,c_n\right)\}_{n=n_0+1}^\infty$, on
1848: $\left(0,T_0\right)\times\Omega$, for an arbitrary time $T_0$.  Then,
1849: using the Arzel\`a--Ascoli theorem, we obtain the convergence result:
1850: %
1851: %
1852: \begin{quote}
1853: There is a subsequence of the family $\{\left(h_n,c_n\right)\}_{n=n_0+1}^\infty$
1854: that converges uniformly to a limit $\left(h,c\right)$, in $\left[0,T_0\right]\times\Omega$.
1855: \end{quote}
1856: %
1857: %
1858: %
1859: %
1860: %
1861: %
1862: We prove several facts about the pair $\left(h,c\right)$.
1863: %
1864: %
1865: %
1866: %
1867: %
1868: \begin{quote}
1869: \emph{Let $\left(h,c\right)$ be the limit of the family of functions
1870: $\{\left(h_n,c_n\right)\}_{n=n_0+1}^\infty$
1871: constructed in Secs.~\ref{sec:regularization}--\ref{sec:equicontinuity}.
1872: Then the following properties hold for this limit:
1873: %
1874: %
1875: %
1876: \begin{enumerate}
1877: \item The functions $h(x,t)$ and $c(x,t)$ are uniformly
1878: H\"older continuous in space (exponent $\tfrac{1}{2}$), and uniformly H\"older
1879: continuous in time (exponent $\tfrac{1}{8}$);
1880: \item The initial condition $\left(h,c\right)\left(x,0\right)=\left(h_0,c_0\right)(x)$
1881: holds;
1882: \item $\left(h,c\right)$ satisfy the boundary conditions of the original
1883: problem (periodic boundary conditions);
1884: \item The derivatives $\left(h,c\right)_t$, $\left(h,c\right)_x$, $\left(h,c\right)_{xx}$,
1885: $\left(h,c\right)_{xxx}$, and $\left(h,c\right)_{xxxx}$ are continuous in
1886: space and time;
1887: \item The function pair $\left(h,c\right)$ satisfy the weak form
1888: of the PDEs,
1889: %
1890: %
1891: \begin{multline*}
1892: \int\int_{Q_{T_0}}dtdx\, h\varphi_t+\int\int_{Q_{T_0}}dtdx\,J_\varepsilon\varphi_x=0,\\
1893: J_\varepsilon=f_\varepsilon(h)h_{xxx}-\frac{1}{g_\varepsilon(h)}h_x-\frac{f_\varepsilon(h)}{g_\varepsilon(h)}\left(g_\varepsilon(h)c_{x}^2\right)_x,
1894: \end{multline*}
1895: %
1896: %
1897: %
1898: \begin{multline*}
1899: \int\int_{Q_{T_0}}dtdx\, g_\varepsilon(h)c\psi_t+\int\int_{Q_{T_0}}dtdx\,K_\varepsilon\psi_x=0,\\
1900: K_\varepsilon= cJ_\varepsilon-g_\varepsilon(h)\left[c^3-c-\frac{1}{g_\varepsilon(h)}\left(g_\varepsilon(h)c_x\right)_x\right],
1901: \end{multline*}
1902: %
1903: %
1904: where $\varphi\left(x,t\right)$ and $\psi\left(x,t\right)$ are suitable test
1905: functions.
1906: %
1907: \label{page:claims45}
1908: \end{enumerate}
1909: }
1910: %
1911: \end{quote}
1912: %
1913: %
1914: %
1915: %
1916: The statements~1,~2, and~3 are obvious.  Now, any pair
1917: $\left(h_n(x,t),c_n(x,t)\right)$ satisfies the
1918: equation set
1919: %
1920: %
1921: \begin{multline*}
1922: \int\int_{Q_{T_0}}dtdx\, h_n\phi_t+\int\int_{Q_{T_0}}dtdx\,J_{\varepsilon,n}\phi_x=0,
1923: \\
1924: J_{\varepsilon,n}=f_{\varepsilon} (h_n)h_{n,xxx}-\frac{1}{g_\varepsilon(h_n)}h_x-\frac{f_\varepsilon(h_n)}{g_\varepsilon(h_n)}\left(g_\varepsilon(h_n)c_{n,x}^2\right)_x,
1925: \end{multline*}
1926: %
1927: %
1928: %
1929: \begin{multline*}
1930: %\hskip -0.2in
1931: \int\int_{Q_{T_0}}dtdx\, g_\varepsilon(h_n)c_n\psi_t+\int\int_{Q_{T_0}}dtdx\,
1932: K_{\varepsilon,n}\psi_x=0,\\
1933: K_{\varepsilon,n}= c_nJ_{\varepsilon,n}-g_\varepsilon
1934: (h_n)\left[c_n^3-c_n-\frac{1}{g_\varepsilon(h_n)}\left(g_\varepsilon
1935:     (h_n) c_{n,x}\right)_x\right],
1936: \end{multline*}
1937: %
1938: %
1939: %
1940: and from the boundedness of the fluxes $J_{\varepsilon,n}$ and $K_{\varepsilon,n}$
1941: in $L^2\left(0,T_0;L^2(\Omega)\right)$, it follows that
1942: %
1943: %
1944: \[
1945: \left(J_{\varepsilon,n}\,,\,K_{\varepsilon,n}\right)\rightharpoonup\left(J_\varepsilon\,,\,K_\varepsilon\right),
1946: \qquad\text{weakly in }L^2\left(0,T_0;L^2(\Omega)\right),
1947: \]
1948: %
1949: %
1950: for a subsequence.  Using the regularity
1951: theory for uniformly parabolic equations and the uniform H\"older continuity
1952: of the $\left(h_n,c_n\right)$'s, it follows that
1953: %
1954: %
1955: \begin{quote}
1956: The derivatives $\left(h_n,c_n\right)_t$, $\left(h_n,c_n\right)_x$,
1957: $\left(h_n,c_n\right)_{xx}$, $\left(h_n,c_n\right)_{xxx}$, and $\left(h_n,c_n\right)_{xxxx}$
1958: are uniformly convergent in any compact subset of $\left(0,T_0\right]\times\Omega$.
1959: \end{quote}
1960: %
1961: %
1962: Thus,
1963: %
1964: %
1965: \[
1966: J_\varepsilon = f_\varepsilon(h)h_{xxx} -
1967: \frac{1}{g_\varepsilon(h)} h_x -
1968: \frac{f_\varepsilon(h)}{g_\varepsilon(h)} \left(g_\varepsilon(h)c_{x}^2\right)_x,
1969: \]
1970: %
1971: %
1972: %
1973: \[
1974: K_\varepsilon=cJ_\varepsilon-g_\varepsilon(h)\left[c^3-c-\frac{1}{g_\varepsilon(h)}\left(g_\varepsilon(h)c_x\right)_x\right],
1975: \]
1976: %
1977: %
1978: %
1979: on $\left(0,T_0\right]\times\Omega$, and therefore, claims~4 and~5 on
1980: p.~\pageref{page:claims45} follow.
1981: 
1982: 
1983: \subsection{Convergence of regularised problem, as $\varepsilon\rightarrow0$}
1984: \label{sec:convergence}
1985: 
1986: \noindent The result in Sec.~\ref{sec:equicontinuity} produced a solution
1987: $\left(h_\varepsilon,c_\varepsilon\right)$ to the regularised problem.  Due
1988: to the result
1989: %
1990: %
1991: %
1992: \begin{equation}
1993: h_{\varepsilon}\left(x,t\right)\geq h_{\mathrm{min}}\geq M=-k_1\sqrt{L}+\sqrt{k_1^2L+\frac{k_1^2L}{e^{k_4k_1^2}-1}}>0,\\
1994: \label{eq:no_rupture}
1995: \end{equation}
1996: %
1997: %
1998: independent of $\varepsilon$, the argument of
1999: Sec.~\ref{sec:equicontinuity} can be recycled to produce a solution
2000: $\left(h,c\right)$ to the unregularised problem.  This solution is
2001: constructed as the limit of a convergent subsequence, formally written as
2002: $\left(h,c\right)=\lim_{\varepsilon\rightarrow0}\left(h_\varepsilon,c_\varepsilon\right)$,
2003: and the results of the theorem in Sec.~\ref{sec:equicontinuity}
2004: apply again to $\left(h,c\right)$.  The result~\eqref{eq:no_rupture}
2005: applies to $h$ constructed as
2006: $h=\lim_{\varepsilon\rightarrow0}h_\varepsilon$, and thus all the
2007: derivatives $\left(h,c\right)_t$, $\left(h,c\right)_x$,
2008: $\left(h,c\right)_{xx}$, $\left(h,c\right)_{xxx}$, and
2009: $\left(h,c\right)_{xxxx}$ are continuous on the whole space
2010: $\left(0,T_0\right]\times\Omega$ and therefore, the weak solution
2011: $\left(h,c\right)$ is in fact a strong one.
2012:  
2013: \subsection{Regularity properties of the solution $\left(h,c\right)$}
2014: \label{sec:regularity}
2015: 
2016: \noindent 
2017: Using a bootstrap argument, we show that the solution $\left(h,c\right)$ belongs to the regularity
2018: classes $L^{\infty}\left(0,T_0;\Sobo^{2,2}(\Omega)\right)$ and $L^2\left(0,T_0;\Sobo^{4,2}(\Omega)\right)$.
2019:  From Sec.~\ref{sec:a_priori_bds} it follows immediately that
2020: %
2021: %
2022: \[
2023: \|h_x\|_2\,,\ \|c_x\|_2 < \infty,
2024: \]
2025: %
2026: %
2027: with time-independent bounds.  Thus, using Poincar\'e's inequality, it follows
2028: that $h,c\in \Sobo^{1,2}(\Omega)$ and, moreover, 
2029: %
2030: %
2031: \[
2032: \sup_{\left[0,T_0\right]}\|h_x\|_2\,,\ \sup_{\left[0,T_0\right]}\|c_x\|_2 <
2033: \infty.
2034: \]
2035: %
2036: % 
2037: From Sec.~\ref{sec:equicontinuity},
2038: it follows that $J$ and $\mu_x$ belong to the regularity class $L^2\left(0,T_0;L^2(\Omega)\right)$,
2039: and hence $J$, $\mu_x\in L^2\left(0,T_0;L^1(\Omega)\right)$. 
2040: The functions $J$, $\mu$, and $\mu_x$ take the form
2041: %
2042: %
2043: \[
2044: J=h^3 h_{xxx}-h_xh^{-1}-h^2\left(h_x c_x^2+2h c_x c_{xx}\right),
2045: \]
2046: %
2047: % 
2048: %
2049: %
2050: \[
2051: \mu = c^3-c - h^{-1}\left(h_x c_x + hc_{xx}\right),
2052: \]
2053: %
2054: %
2055: and
2056: %
2057: %
2058: %
2059: \[
2060: \mu_x= \left(3c^2-1\right)c_x + h^{-2}h_x^2 c_x - h^{-1}h_{xx}c_x - h^{-1}h_xc_{xx}
2061: -c_{xxx},
2062: \]
2063: %
2064: %
2065: respectively.
2066: %
2067: %
2068: We make use of the following observations:
2069: %The following results will help us in our demonstration,
2070: %
2071: %
2072: \begin{itemize}
2073: \item The function $h(x,t)$ is bounded from above and below,
2074: %
2075: %
2076: \[
2077: 0< h_{\mathrm{min}}\leq h(x,t)\leq h_{\mathrm{max}}<\infty;
2078: \]
2079: %
2080: %
2081: %
2082: we also have the boundedness of $c(x,t)$, $\|c\|_\infty
2083: <\infty$. 
2084: \item Since $\mu_x\in L^2\left(0,T_0;L^2(\Omega)\right)$, it follows
2085: that $\mu\in L^2\left(0,T_0; L^2(\Omega)\right)$, by Poincar\'e's
2086: inequality.
2087: \item From this it follows that $c_x h_x + hc_{xx}$ is in the class $L^2\left(0,T_0;L^2(\Omega)\right)$.
2088: \item  Given the inequality 
2089: %
2090: %
2091: \[
2092: \|h\mu c_x\|_2^2\leq h_{\mathrm{max}}\|\mu\|_\infty^2\|c_x\|_2^2\leq
2093: h_{\mathrm{max}}\|c_x\|_2^2\left[\frac{1}{\sqrt{L}}\|\mu\|_2+\sqrt{L}\|\mu_x\|_2\right]^2,
2094: \]
2095: %
2096: we have the result $h_x c_x^2 + h c_x c_{xx}\in L^2\left(0,T_0;L^2(\Omega)\right)$.
2097: %
2098: %
2099: \item Similarly, since $\int_0^{T_0}\|h\mu h_x\|_2^2dt<\infty$, we have the
2100: bound
2101: $h_x^2c_x +hh_x c_{xx}\in  L^2\left(0,T_0;L^2(\Omega)\right)$.
2102: \end{itemize}
2103: %
2104: %
2105: %
2106: Using these facts, and relegating the details to Appendix~\ref{apx:B},
2107: it is possible to show that $c_{xx}$ is in
2108: $L^2\left(0,T_0;L^1(\Omega)\right)$, from which follows the result
2109: $h_{xxx},c_{xxx}\in L^1\left(0,T_0;L^1(\Omega)\right)$.  These results
2110: give rise to further bounds, namely $c_{xx}\in
2111: L^2\left(0,T_0;L^2(\Omega)\right)$, and
2112: $\int_0^{T_0}dt\,\|h_x^2c_x\|_2<\infty$, whence $h_{xx}\in
2113: L^2\left(0,T_0;L^2(\Omega)\right)$.  Using this collection of bounds,
2114: we obtain
2115: %
2116: %
2117: %
2118: %
2119: \[
2120: h,c\in L^1\left(0,T_0;\Sobo^{3,1}(\Omega)\right),
2121: \]
2122: %
2123: %
2124: and hence finally, 
2125: %
2126: %
2127: %
2128: %
2129: \[
2130: h,c\in L^2\left(0,T_0;\Sobo^{3,2}(\Omega)\right).
2131: \]
2132: %
2133: %
2134: Thus, the solution $\left(h,c\right)$ belongs to the following regularity
2135: class:
2136: %
2137: %
2138: \begin{equation}
2139: \left(h,c\right)\in L^{\infty}\left(0,T_0;\Sobo^{1,2}(\Omega)\right)
2140: \cap L^2\left(0,T_0;\Sobo^{3,2}(\Omega)\right)
2141: \cap C^{\frac{1}{2},\frac{1}{8}}\left(\Omega\times\left[0,T_0\right]\right).
2142: \label{eq:regularity0}
2143: \end{equation}
2144: %
2145: 
2146: Extra regularity is obtained by writing the equation pair as
2147: %
2148: %
2149: %
2150: \begin{align*}
2151: \frac{\partial h}{\partial t}+h^3h_{xxxx}&=-3h^2h_xh_{xxx}+\varphi_1+\varphi_2\equiv
2152: \varphi\left(x,t\right),\\
2153: %
2154: %
2155: \frac{\partial c}{\partial t}+c_{xxxx}&=-\frac{2}{h}h_xc_{xxx}+\psi_1+\psi_2\equiv
2156: \psi\left(x,t\right),
2157: \end{align*}
2158: %
2159: %
2160: where 
2161: %
2162: %
2163: %
2164: \begin{align*}
2165: \int_0^{\tau}{dt}\,\|\varphi_1\|_2&\leq \sup_{\left[0,\tau\right]}\|c_{xx}\|_2\int_0^{\tau}{dt}\,|\nu_1|,\qquad
2166: \nu_1\in L^2\left(\left[0,\tau\right]\right),\\
2167: %
2168: %
2169: \int_0^{\tau}{dt}\,\|\psi_1\|_2&\leq \sup_{\left[0,\tau\right]}\|c_{xx}\|_2\int_0^{\tau}{dt}\,|\nu_2|,\qquad
2170: \nu_2\in L^2\left(\left[0,\tau\right]\right),
2171: %
2172: %
2173: \end{align*}
2174: %
2175: %
2176: %
2177: for any $\tau\in \left(0,T_0\right]$, and where $\varphi_2$ and $\psi_2$
2178: belong to
2179: the class $L^2\left(0,T_0;L^2(\Omega)\right)$.  Here we have used
2180: the form of the concentration equation given by Eq.~\eqref{eq:conc_no_flux}.
2181:  By multiplying
2182: the height and concentration equations by $h_{xxxx}$ and $c_{xxxx}$ respectively,
2183: and by integrating over space and time, it is readily shown that
2184: %
2185: %
2186: %
2187: \[
2188: \left(h,c\right)\in L^\infty\left(0,T_0;\Sobo^{2,2}(\Omega)\right),
2189: \]
2190: %
2191: %
2192: and hence
2193: %
2194: %
2195: \[
2196: \left(\varphi,\psi\right)\in L^2\left(0,T_0;L^{2}(\Omega)\right),
2197: \]
2198: %
2199: %
2200: from which follows the regularity result
2201: %
2202: %
2203: %
2204: \begin{equation}
2205: \left(h,c\right)\in L^{\infty}\left(0,T_0;\Sobo^{2,2}(\Omega)\right)
2206: \cap L^2\left(0,T_0;\Sobo^{4,2}(\Omega)\right)
2207: \cap C^{\frac{3}{2},\frac{1}{8}}\left(\Omega\times\left[0,T_0\right]\right).
2208: \label{eq:regularity}
2209: \end{equation}
2210: 
2211: 
2212: 
2213: \subsection{Uniqueness of solutions}
2214: \label{sec:uniqueness}
2215: %
2216: %
2217: %
2218: %
2219: Let us consider two solution pairs $\left(h,c\right)$ and $\left(h',c'\right)$
2220: and form the difference $\left(\delta h,\delta c\right)=\left(h-h',c-c'\right)$.
2221:  Given the initial conditions $\left(\delta c(x,0),\delta h(x,0)\right)=\left(0,0\right)$,
2222:  we show that $\left(\delta h,\delta c \right)=\left(0,0\right)$ for all
2223:  time, that is, that the solution we have constructed is unique.
2224: %
2225: %
2226: %
2227: We observe that that the equation for the difference $\delta c$ can be written
2228: in the form
2229: %
2230: %
2231: \begin{equation}
2232: \frac{\partial}{\partial t}\delta{c}+\frac{\partial^4}{\partial{x}^4}\delta{c}=\delta
2233: \varphi\left(x,t\right),
2234: \label{eq:delta_c_unique}
2235: \end{equation}
2236: %
2237: %
2238: %
2239: where 
2240: %
2241: %
2242: $\delta \varphi\left(x,t\right) \in L^2\left(0,T_0;L^2(\Omega)\right)$,
2243: and where $\delta \varphi\left(\delta c=0\right)=0$.
2244: %
2245: %
2246: %
2247: %
2248: %
2249: Using semigroup theory~\cite{TaylorPDEbook},
2250: we find that Eq.~\eqref{eq:delta_c_unique} has a unique solution.
2251: %
2252: %
2253: Since $\delta{c}=0$ satisfies Eq.~\eqref{eq:delta_c_unique}, and since $\delta{c}\left(x,0\right)=0$,
2254: it follows that $\delta{c}=0$ for all times $t\in\left[0,T_0\right]$.
2255: 
2256: 
2257: 
2258: It is now possible to formulate an equation for the difference $\delta{h}$
2259: by subtracting the evolution equations of $h$ and $h'$ from one another,
2260: mindful that $\delta c=0$.  We multiply the resulting equation by
2261: $\delta{h}_{xx}$, integrate over space, and obtain after using
2262: inequalities (see Appendix~\ref{apx:C})
2263: %
2264: %
2265: \begin{multline*}
2266: 2\kappa\sup_{\tau\in\left[0,T\right]}\|\delta{h}_x\|_2^2\left(\tau\right)
2267: %
2268: %
2269: \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\\
2270: %
2271: %
2272: +\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),
2273: \end{multline*}
2274: %
2275: %
2276: where $\kappa$ and $\kappa_P$ are numerical constants.
2277: Using the results of Sec.~\ref{sec:regularity}, it is
2278: readily shown that $h^{-1}+h'^2c_x^2\in L^2\left(0,T;L^{\infty}(\Omega)\right)$,
2279: and that the functions $h'_{xxx}+4c_xc_{xx}$ and $h_x'+h_xc_x^2$ belong to
2280: the class $L^2\left(0,T;L^2(\Omega)\right)$.  By choosing $T$
2281: sufficiently small, it is possible to impose the inequality
2282: %
2283: %
2284: %
2285: \[
2286: \frac{1}{2\kappa}\left[\int_0^T{dt}\,\|h^{-1}+h'^2c_x^2\|_\infty^2+\kappa_P^2\int_0^T{dt}\left(\|h'_{xxx}+4c_xc_{xx}\|_2^2+\|h_x'+h_xc_x^2\|_2^2\right)\right]<1,
2287: \]
2288: %
2289: %
2290: which in turn forces $\sup_{\tau\in\left[0,T\right]}\|\delta{h}_x\|_2^2=0$,
2291: and hence the solution is unique.
2292: 
2293: 
2294: \section{Parametric dependence of the height dip}
2295: \label{sec:height_dip}
2296: 
2297: \noindent 
2298: In this section we perform numerical simulations of the equations~\eqref{eq:model}
2299: and focus on one feature of the equations: the drop in the free-surface height
2300: at the boundary between binary fluid domains.  Our results in Sec.~\ref{sec:existence}
2301: gave a rigorous upper bound for the magnitude of this height dip, as a function
2302: of the problem parameters, and we compare this estimate with numerical solutions.
2303: 
2304: We perform numerical simulations of the full equations~\eqref{eq:model},
2305: with initial data comprising a perturbation away from the unstable steady
2306: state $\left(h,c\right)=\left(1,0\right)$.
2307: The free surface and concentration evolve to an equilibrium state where the
2308: salient feature is the formation of domains (intervals where $c\approx\pm1$)
2309: that are separated by smooth transition regions, across which the free surface
2310: dips below its average value.  Since we are interested in this characteristic
2311: asymptotic feature of phase separation, we shift our focus instead to the
2312: steady version from Eq.~\eqref{eq:model} obtained by setting $\partial_t=J=\mu_x=0$.
2313:  For this reason also, we consider only the repulsive Van der Waals force,
2314:  for
2315:  which the interaction potential is given by $\phi=-|A|h^{-3}$.
2316: % 
2317: % 
2318: % 
2319: We then solve the boundary-value problem
2320: % 
2321: %
2322: % 
2323: \begin{subequations}
2324: \begin{equation}
2325: \frac{1}{C}\frac{\partial^2 h}{\partial x^2}=|A| C_{\mathrm{n}}^2\left(1-\frac{1}{h^3}\right)
2326: %
2327: +r\left[\tfrac{1}{4}\left(c^2-1\right)^2+\tfrac{1}{2}\left(\frac{\partial
2328: c}{\partial x}\right)^2\right],
2329: \end{equation}
2330: %
2331: \begin{equation}
2332: \frac{\partial^2 c}{\partial x^2}=c^3-c-\frac{1}{h}\frac{\partial
2333: h}{\partial
2334: x}\frac{\partial c}{\partial x},
2335: %
2336: \end{equation}%
2337: \label{eq:eqm}%
2338: \end{subequations}%
2339: %
2340: %
2341: %
2342: % 
2343: where now the domain is infinite and the boundary conditions are $h\left(\pm\infty\right)=1$,
2344: and
2345: $\mu\left(\pm\infty\right)=0$.  We have rescaled lengths by $C_{\mathrm{n}}$.
2346: In Fig.~\ref{fig:hc} we present numerical solutions
2347: %
2348: %
2349: %
2350: %
2351: \begin{figure}[htb]
2352: \subfigure[]{
2353:  \includegraphics[width=.45\textwidth]{h_bvp}
2354: }\hspace{1em}
2355: \subfigure[]{
2356:  \includegraphics[width=.45\textwidth]{c_bvp}
2357: }
2358: \caption{%(Color online)  
2359: Equilibrium solutions of Eq.~\eqref{eq:eqm}
2360: for $C=C_{\mathrm{n}}^2|A|=1$ and $r=0.1,1,10,50$.  In (a) the valley deepens
2361: with increasing $r$ although the film never ruptures, while in (b) the front
2362: steepens with increasing $r$. (From \'O N\'araigh and
2363: Thiffeault~\cite{ONaraigh2007}.)
2364: }
2365: \label{fig:hc}
2366: \end{figure}
2367: %
2368: %
2369: %
2370: %
2371: exhibiting the dependence of the solutions on the parameters in Eq.~\eqref{eq:eqm}.
2372: As before,
2373: the free-surface height possesses peaks and valleys, where the valleys occur
2374: in the concentration field's transition zone.  These profiles are qualitatively
2375: similar to the results obtained in experiments on thin binary films~\cite{WangW2003,
2376: WangH2000, ChungH2004}.  While the valley increases in depth
2377: %
2378: %
2379: %
2380: %
2381: \begin{figure}[htb]
2382: \includegraphics[width=.5\textwidth]{force_balance}
2383: \caption{The forces $F_{\mathrm{cap}}$ and $F_{\mathrm{VdW}}$
2384: for $C=C_{\mathrm{n}}^2=|A|=1$ and $r=50$ are shown to have opposite sign.}
2385: \label{fig:vdw_cap}
2386: \end{figure}
2387: %
2388: %
2389: %
2390: %
2391: %
2392: for large $r$, rupture never takes place, in agreement with the theory in
2393: Sec.~\ref{sec:existence}.  Similar results are obtained by varying $|A|$, with
2394: smaller $|A|$ leading to larger dips.  This behaviour is underlined by the
2395: results in Fig.~\ref{fig:vdw_cap}, where we plot the sign and magnitude of
2396: the backreaction or capillary force $F_{\mathrm{cap}}=-rh^{-1}\left(hc_x^2\right)_x$
2397: and the Van der Waals force $F_{\mathrm{VdW}}=|A|\left(h^{-3}\right)_x$.
2398:  These forces have opposite sign: the Van der Waals force inhibits film rupture,
2399:  while the backreaction promotes film thinning.    The thinning of binary
2400:  films due to capillary or backreaction effects has been documented in experiments~\cite{WangH2000}.
2401:  
2402: We are interested in the magnitude of the dip in the free-surface height,
2403: as a function of the problem parameters, and we plot the dependence of $h_{\mathrm{min}}$
2404: as a function of $|A|$ and $r$ in Fig.~\ref{fig:scaling_r_A}.
2405: %
2406: %
2407: %
2408: \begin{figure}[htb]
2409:         \subfigure[]{
2410:         \includegraphics[width=.45\textwidth]{h_A_aug2}
2411: }\hspace{1em}
2412:         \subfigure[]{
2413:         \includegraphics[width=.45\textwidth]{h_r_aug1}
2414: }
2415: \caption{Dependence of dip magnitude $h_{\mathrm{min}}$ on (a) the parameter
2416: $|A|$; (b) the parameter $r$.  We have set $C=C_{\mathrm{n}}=1$.  In (a) the
2417: dip decreases in magnitude with increasing $|A|$, demonstrating the tendency
2418: of the Van der Waals force to flatten the free surface, while in (b) the
2419: dip increases with increasing $r$, which shows how the backreaction thins
2420: the film across transition zones.}
2421: \label{fig:scaling_r_A}
2422: \end{figure}
2423: %
2424: %
2425: %
2426: We compare these results with the estimate for $h_{\mathrm{min}}$ found in
2427: Sec.~\ref{sec:existence}.  In terms of the physical parameters of the
2428: system, this estimate is
2429: %
2430: %
2431: %
2432: \[
2433: h_{\mathrm{min}}\geq M\equiv\sqrt{2CL(F_0+F_1|A|)}\left(\sqrt{\frac{e^{4C|A|^{-1}\left(F_0+F_1|A|\right)^2}}{e^{4C|A|^{-1}\left(F_0+F_1|A|\right)^2}-1}}-1\right),
2434: \]
2435: %
2436: %
2437: %
2438: where $F_1=\tfrac{1}{2}\int_\Omega{dx}\,\left[h(x,0)\right]^{-2}\neq0$, and
2439: $F_0=F\left(0\right)-F_1$. 
2440: The function $M(|A|,C)$ has no explicit $r$-dependence: although
2441: $F_0$ depends on $r$, it is possible to find initial data to remove this
2442: dependence.
2443: % 
2444: We show a representative plot of $M(|A|,C)$ in Fig.~\ref{fig:M_A}.
2445: Although a comparison between Fig.~\ref{fig:scaling_r_A} and Fig.~\ref{fig:M_A}
2446: is not exact, since the boundary conditions and domains are different in
2447: both cases, we see that the shape of the bound in Fig.~\ref{fig:scaling_r_A}
2448: is different from that in Fig.~\ref{fig:M_A}.  Since the bound in Fig.~\ref{fig:scaling_r_A}
2449: is obtained from numerical simulations, and is intuitively correct, we conclude
2450: that it has the correct shape and that the bound of Fig.~\ref{fig:M_A}, while
2451: mathematically indispensable, is not sharp enough to be useful in determining
2452: the parametric dependence of the dip in free-surface height.
2453: %
2454: %
2455: %
2456: %
2457: \begin{figure}[htb]
2458: \includegraphics[width=.5\textwidth]{M_A}
2459: \caption{A typical plot of $M(|A|,C)$ for $F_0=F_1=\tfrac{1}{2}$
2460: and $C=1$.
2461:  This theoretical lower bound has a different shape from those in Fig.~\ref{fig:scaling_r_A},
2462:  which suggests that while $M(|A|,C)$ plays an important role in
2463:  the analysis of the model equations, it does not capture the physics of
2464:  film thinning.}
2465: \label{fig:M_A}
2466: \end{figure}
2467: %
2468: %
2469: %
2470: %
2471: 
2472: \section{Conclusions}
2473: \label{sec:conclusions}
2474: 
2475: \noindent Starting from the Navier--Stokes Cahn--Hilliard equations, we have
2476: derived a pair of nonlinear parabolic PDEs that describe the coupled 
2477: effects of phase separation and free-surface variations in a thin film of
2478: binary liquid.  Since we are interested in the long-time outcome of the phase
2479: separation, we focused on liquids that experience a repulsive Van der Waals
2480: force, which tends to inhibit film rupture.  Using physical intuition, we
2481: identified a decaying energy functional that facilitated analysis of the
2482: equations.
2483: %
2484: We have shown that given sufficiently smooth initial data, solutions to the
2485: model equations~\eqref{eq:model} exist in a strong sense, and
2486: have certain regularity properties.  Our proof follows the method developed
2487: in~\cite{Friedman1990}.  Central to the analysis is the decaying free energy,
2488: and the derivation of a no-rupture condition, which prevents the film from
2489: touching down to zero.  The no-rupture condition is valid in one spatial
2490: dimension only, and thus existence and uniqueness results will not necessarily
2491: be obtainable in higher dimensions.
2492: 
2493: 
2494: We carried out one-dimensional numerical simulations of the full
2495: equations and found that the free-surface height and concentration
2496: tend to an equilibrium state.  The concentration forms domains; that
2497: is, extended regions where $c\approx\pm1$.  The domains are separated
2498: by narrow zones where the concentration smoothly transitions between
2499: the limiting values $\pm1$.  At the transition zones, the free surface
2500: dips below its mean value, a feature of binary thin-film behaviour
2501: that is observed in experiments.  To study the magnitude of this dip
2502: as a function of the problem parameters, we focused on solving the
2503: equilibrium version of Eq.~\eqref{eq:model} as a boundary-value
2504: problem.  This simplification is carried out without loss of
2505: generality, since we have shown that the system tends asymptotically
2506: to such a state.  We have shown that the magnitude of the dip
2507: decreases by increasing the strength of the repulsive Van der Waals
2508: force, while the dip depth actually increases by increasing the
2509: strength of the backreaction.  Thus, in the absence of the Van der
2510: Waals force, the film would rupture, preventing the occurrence of the
2511: phase separation so characteristic of long-time Cahn--Hilliard
2512: dynamics.  The film-thinning tendency of the backreaction has been
2513: observed experimentally~\cite{WangH2000, WangW2003, ChungH2004}.
2514: %  
2515: Simulations involving two lateral directions have elsewhere been carried
2516: out by the authors~\cite{ONaraigh2007}, and the qualitative features are
2517: similar to those obtained here.
2518: 
2519: \emph{Acknowledgements.} We thank G.~A.~Pavliotis for helpful discussions.
2520: L.\'O.N. was supported by the Irish government and the UK Engineering
2521: and Physical Sciences Research Council. J.-L.T. was supported in part
2522: by the UK EPSRC Grant No. GR/S72931/01.
2523: 
2524: 
2525: \appendix
2526: 
2527: \section{}
2528: \label{apx:A}
2529: % 
2530: % 
2531: \noindent In this appendix we give a list of nonstandard inequalities used in
2532: the paper and provide proofs.
2533: 
2534: 
2535: \begin{enumerate}
2536: \item Let $\phi:\Omega\subset\mathbb{R}\rightarrow\mathbb{R}$ belong
2537: to the class $\Sobo^{1,2}(\Omega)$.  Then the following string of inequalities
2538: holds,
2539: %the following inequality holds,
2540: %
2541: %
2542: %
2543: \begin{equation}
2544: \sup_\Omega |\phi|\leq \frac{1}{L}\|\phi\|_1 + \|\phi_x\|_1
2545: \leq\frac{1}{\sqrt{L}}\,\|\phi\|_2+\sqrt{L}\,\|\phi_x\|_2,
2546: \label{eq:result1}
2547: \end{equation}
2548: %
2549: %
2550: \emph{Proof:} Using the Fundamental Theorem of Calculus, we have
2551: %
2552: %
2553: \[
2554: \phi\left(x\right)=\phi\left(a\right)
2555: + \int_a^x{ds}\,\frac{\partial\phi}{\partial s},
2556: \]
2557: %
2558: %
2559: for any distinct points $x$ and $a$ in $\Omega$.
2560: %
2561: %
2562: Since the function $|\phi(x)|$ is continuous on $\Omega$, it has a maximum
2563: value in $\Omega$, attained at the point $x_{\mathrm{max}}$.
2564: Thus,
2565: \[
2566: |\phi\left(x_{\mathrm{max}}\right)|\equiv\sup_{\Omega}|\phi|\leq |\phi\left(a\right)|+\int_a^{x_{\mathrm{max}}}{ds}\left|\frac{\partial\phi}{\partial
2567: s}\right|\leq  |\phi\left(a\right)|+\int_\Omega{ds}\left|\frac{\partial\phi}{\partial
2568: s}\right|.
2569: \]
2570: %
2571: %
2572: %
2573: Since this is true for any $a\in\Omega$, by integrating over $a$, we obtain
2574: the inequality
2575: %
2576: %
2577: \[
2578: \sup_\Omega|\phi|\leq\frac{1}{L}\|\phi\|_1+\|\phi_x\|_1
2579: \leq \frac{1}{\sqrt{L}}\,\|\phi\|_2+\sqrt{L}\,\|\phi_x\|_2,
2580: \]
2581: %
2582: %
2583: %
2584: where the last inequality follows from the monotonicity of norms.
2585: %
2586: %
2587: %
2588: \item Let $\phi:\Omega\subset\mathbb{R}\rightarrow\mathbb{R}$ belong
2589: to the class $\Sobo^{2,1}(\Omega)$.  Then
2590: %
2591: %
2592: \begin{equation}
2593: \|\phi_{x}\|_2^2\leq L\|\phi_{xx}\|_1^2 + \frac{4}{L}\|u\|_1\|\phi_{xx}\|_1.
2594: \label{eq:result2}
2595: \end{equation}
2596: %
2597: %
2598: %
2599: \emph{Proof:}  We have the identity $\int \phi_x^2{dx}\,=-\int \phi\phi_{xx}$,
2600: true
2601: for any function $\phi$ with periodic boundary conditions.  Using H\"older's
2602: inequality, this becomes
2603: %
2604: %
2605: \[
2606: \|\phi_x\|_2^2\leq \|\phi\|_\infty\|\phi_{xx}\|_1.
2607: \]
2608: %
2609: %
2610: %
2611: Using the relation~\eqref{eq:result1}, this becomes
2612: %
2613: %
2614: \begin{align*}
2615: \|\phi_x\|_2^2&\leq\Bigl(\frac{1}{L}\|\phi\|_1+\|\phi_x\|_1\Bigr)\|\phi_{xx}\|_1,\\
2616: &\leq\Bigl(\frac{1}{L}\|\phi\|_1+\sqrt{L}\,\|\phi_x\|_2\Bigr)\|\phi_{xx}\|_1,
2617: \end{align*}
2618: %
2619: %
2620: which is a quadratic inequality in $\|\phi_x\|_2^2$, with solution
2621: %
2622: %
2623: \[
2624: \|\phi_x\|_2\leq\tfrac{1}{2}\Bigl(\sqrt{L}\,\|\phi_{xx}\|_1+\sqrt{L\|\phi_{xx}\|_1^2+4L^{-1}\|\phi\|_1\|\phi_{xx}\|_1}\Bigr).
2625: \]
2626: %
2627: %
2628: By sacrificing the sharpness of the bound, we obtain a simpler one,
2629: %
2630: %
2631: \[
2632: \|\phi_x\|_2^2\leq {L}\|\phi_{xx}\|_1^2 + \frac{4}{L}\|\phi\|_1\|\phi_{xx}\|_1,
2633: \]
2634: %
2635: %
2636: as required.
2637: \end{enumerate}
2638: 
2639: 
2640: \section{}
2641: \label{apx:B}
2642: % 
2643: % 
2644: 
2645: 
2646: \noindent In this appendix, we fill in the details missing from the discussion
2647: of the regularity of solutions in Sec.~\ref{sec:regularity}
2648: and prove the result
2649: %
2650: %
2651: \[
2652: h,c\in L^2\left(0,T_0;\Sobo^{3,2}(\Omega)\right).
2653: \]
2654: %
2655: %
2656: %
2657: To begin, we notice that Sec.~\ref{sec:a_priori_bds} gives rise to
2658: the result
2659: %
2660: %
2661: \[
2662: \|h_x\|_2,\ \|c_x\|_2 < \infty,
2663: \]
2664: %
2665: %
2666: with time-independent bounds.  Thus, using Poincar\'e's inequality, it follows
2667: that $h,c\in \Sobo^{1,2}(\Omega)$ and, moreover, 
2668: %
2669: %
2670: \[
2671: \sup_{\left[0,T_0\right]}\|h_x\|_2,\ \sup_{\left[0,T_0\right]}\|c_x\|_2 <
2672: \infty.
2673: \]
2674: %
2675: % 
2676: From Sec.~\ref{sec:equicontinuity},
2677: it follows that $J$ and $\mu_x$ belong to the regularity class $L^2\left(0,T_0;L^2(\Omega)\right)$,
2678: and hence $J$, $\mu_x\in L^2\left(0,T_0;L^1(\Omega)\right)$. 
2679: The functions $J$, $\mu$, and $\mu_x$ take the form
2680: %
2681: %
2682: \[
2683: J=h^3 h_{xxx}-h_xh^{-1}-h^2\left(h_x c_x^2+2h c_x c_{xx}\right),
2684: \]
2685: %
2686: % 
2687: %
2688: %
2689: \[
2690: \mu = c^3-c - h^{-1}\left(h_x c_x + hc_{xx}\right),
2691: \]
2692: %
2693: %
2694: and
2695: %
2696: %
2697: %
2698: \begin{equation}
2699: \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}
2700: -c_{xxx},
2701: \label{eq:mux}
2702: \end{equation}
2703: %
2704: %
2705: respectively.
2706: %
2707: %
2708: We make use of the following observations,
2709: %The following results will help us in our demonstration,
2710: %
2711: %
2712: \begin{itemize}
2713: \item The function $h(x,t)$ is bounded from above and below,
2714: %
2715: %
2716: \[
2717: 0< h_{\mathrm{min}}\leq h(x,t)\leq h_{\mathrm{max}}<\infty,
2718: \]
2719: %
2720: %
2721: %
2722: and the boundedness of $c(x,t)$, $\|c\|_\infty <\infty$. 
2723: \item Since $\mu_x\in L^2\left(0,T_0;L^2(\Omega)\right)$, it follows
2724: that $\mu\in L^2\left(0,T_0; L^2(\Omega)\right)$, by Poincar\'e's
2725: inequality.
2726: \item From this it follows that $c_x h_x + hc_{xx}$ is in the class $L^2\left(0,T_0;L^2(\Omega)\right)$.
2727: \item  Given the inequality 
2728: %
2729: %
2730: \[
2731: \|h\mu c_x\|_2^2\leq h_{\mathrm{max}}\|\mu\|_\infty^2\|c_x\|_2^2\leq
2732: h_{\mathrm{max}}\|c_x\|_2^2\left[\frac{1}{\sqrt{L}}\,\|\mu\|_2+\sqrt{L}\,\|\mu_x\|_2\right]^2,
2733: \]
2734: %
2735: we have the result $h_x c_x^2 + h c_x c_{xx}\in L^2\left(0,T_0;L^2(\Omega)\right)$.
2736: %
2737: %
2738: \item Similarly, since $\int_0^{T_0}\|h\mu h_x\|_2^2dt<\infty$, we have the
2739: bound
2740: $h_x^2c_x +hh_x c_{xx}\in  L^2\left(0,T_0;L^2(\Omega)\right)$.
2741: \end{itemize}
2742: 
2743: 
2744: Now inspection of $\mu$ shows that $c_{xx}$ is in $L^2\left(0,T_0;L^1(\Omega)\right)$,
2745: from which follows the result $h_{xxx},c_{xxx}\in L^1\left(0,T_0;L^1(\Omega)\right)$.
2746:  By repeating the same argument, we find that $c_{xx}\in L^2\left(0,T_0;L^2(\Omega)\right)$.
2747:   We also have the result that $\|h_x^2c_x\|_2$ is almost always bounded.
2748:    To show that $h_{xx}\in  L^2\left(0,T_0;L^2(\Omega)\right)$,
2749:    we take the evolution equation for $h(x,t)$, multiply it by
2750:    $h$, and integrate, obtaining
2751: %
2752: %
2753: %
2754: \[
2755: -\tfrac{1}{2}\frac{d}{dt}\int_\Omega{dx}\,h^2-3\int_\Omega{dx}\,h^2h_x^2h_{xx}-\int_\Omega{dx}\,h_xJ_0=\int_\Omega{dx}\,h^3h_{xx}^2,
2756: \]
2757: %
2758: %
2759: %
2760: where $J_0=h_xh^{-1}+2h^3c_xc_{xx}+h^2h_xc_x^2$.   The time-integral
2761: of the first term on the
2762: left-hand side of this equation is obviously bounded in time.  Let us examine
2763: the time-integral of the second term,
2764: %
2765: %
2766: \begin{align*}
2767: \int_0^{T_0}{dt}\int_\Omega{dx}\,h^2h_x^2h_{xx}&\leq h_{\mathrm{max}}^2\sup_{\left[0,T_0\right]}\|h_x\|_2^2\int_0^{T_0}{dt\,}\|h_{xx}\|_\infty\\
2768: &\leq h_{\mathrm{max}}^2\sup_{\left[0,T_0\right]}\|h_x\|_2^2\int_0^{T_0}{dt}\,\left(L^{-1}\|h_{xx}\|_1+\|h_{xxx}\|_1\right)<\infty.
2769: \end{align*}
2770: %
2771: %
2772: The third term on the left-hand side is dispatched in a similar
2773: way, so that $\int_0^{T_0}{dt}\,\|h_{xx}\|_2^2<\infty$.  We have now
2774: shown that~$h,c\in L^2\left(0,T_0;\Sobo^{2,2}(\Omega)\right)$.  Using
2775: this result, together with the previous facts gathered together in
2776: this appendix, it is readily shown that~$h,c\in
2777: L^1\left(0,T_0;\Sobo^{3,1}(\Omega)\right)$, and it follows that
2778: %
2779: %
2780: %
2781: %
2782: \[
2783: h,c\in L^2\left(0,T_0;\Sobo^{3,2}(\Omega)\right).
2784: \]
2785: %
2786: %
2787: To see this more clearly, we show that~$c_{xxx}$ is in the above
2788: class.  For example, consider a typical term in~$\mu_x$,
2789: %
2790: %
2791: %
2792: \begin{align*}
2793: \int_0^{T_0}{dt}\,\int_\Omega{dx}\,h_x^2c_{xx}^2%&\leq \sup_{\left[0,T_0\right]}\|h_x\|_2^2\int_0^T{dt}\,\|c_{xx}\|_\infty^2\\
2794: &\leq \sup_{\left[0,T_0\right]}\|h_x\|_2^2\int_0^{T_0}{dt}\,\left(L^{-1}\|c_{xx}\|_1+\|c_{xxx}\|_1\right)^2<\infty,
2795: \end{align*}
2796: %
2797: %
2798: %
2799: which implies~$h_x c_{xx} \in L^2\left(0,T_0;L^2(\Omega)\right)$.
2800: This bound, together with $h\mu h_x\in
2801: L^2\left(0,T_0;L^2(\Omega)\right)$, implies $h_x^2 c_x \in
2802: L^2\left(0,T_0;L^2(\Omega)\right)$.  Gathering all these results, we have
2803: \[
2804: \mu_x\,,\ h_xc_{xx}\,,\ h_x^2c_x\,,\ h_{xx}c_x \in L^2\left(0,T_0;L^2(\Omega)\right)
2805: \]
2806: From~\eqref{eq:mux}, it follows that $c_{xxx}$ is in this class as
2807: well.  A similar argument holds for $h_{xxx}$.  Thus, the solution
2808: $\left(h,c\right)$ belongs to the regularity class
2809: %
2810: %
2811: \begin{equation*}
2812: \left(h,c\right)\in L^{\infty}\left(0,T_0;\Sobo^{1,2}(\Omega)\right)
2813: \cap L^2\left(0,T_0;\Sobo^{3,2}(\Omega)\right)
2814: \cap C^{\frac{1}{2},\frac{1}{8}}\left(\Omega\times\left[0,T_0\right]\right).
2815: \end{equation*}
2816: %
2817: 
2818: 
2819: \section{}
2820: \label{apx:C}
2821: % 
2822: % 
2823: \noindent
2824: %
2825: %
2826: In this appendix, we describe in full the proof of the
2827: the uniqueness of solutions sketched in Sec.~\ref{sec:uniqueness}.
2828: We consider two solution pairs $\left(h,c\right)$ and $\left(h',c'\right)$
2829: and form the difference $\left(\delta h,\delta c\right)=\left(h-h',c-c'\right)$.
2830:  Given the initial conditions $\left(\delta c(x,0),\delta h(x,0)\right)=\left(0,0\right)$,
2831:  we show that $\left(\delta h,\delta c \right)=\left(0,0\right)$ for all
2832:  time, that is, that the solution we have constructed is unique.
2833: %
2834: %
2835: %
2836: We observe that that the equation for the difference $\delta c$ can be written
2837: in the form
2838: %
2839: %
2840: \begin{equation}
2841: \frac{\partial}{\partial t}\delta{c}+\frac{\partial^4}{\partial{x}^4}\delta{c}=\delta
2842: \varphi\left(x,t\right),
2843: \label{eq:delta_c_unique_appendix}
2844: \end{equation}
2845: %
2846: %
2847: %
2848: where 
2849: %
2850: %
2851: $\delta \varphi\left(x,t\right) \in L^2\left(0,T_0;L^2(\Omega)\right)$,
2852: and where $\delta \varphi\left(\delta c=0\right)=0$.
2853: %
2854: %
2855: %
2856: %
2857: %
2858: As discussed previously, this equation has a unique solution, given smooth
2859: initial data.
2860: %
2861: %
2862: Since $\delta{c}=0$ satisfies Eq.~\eqref{eq:delta_c_unique_appendix}, and
2863: since $\delta{c}\left(x,0\right)=0$, it follows that $\delta{c}=0$ for all
2864: times $t\in\left[0,T_0\right]$.
2865: 
2866: 
2867: 
2868: It is now possible to formulate an equation for the difference $\delta{h}$
2869: by subtracting the evolution equations of $h$ and $h'$ from one another,
2870: mindful that $\delta c=0$.  We multiply the resulting equation by $\delta{h}_{xx}$
2871: and integrate over space to obtain
2872: %
2873: %
2874: %
2875: \begin{multline*}
2876: \tfrac{1}{2}\frac{d}{dt}\int_\Omega{dx}\,\delta h_x^2 +\int_\Omega{dx}\,h^3\delta
2877: h_{xxx}^2=
2878: \int_\Omega{dx}\,\delta h_{xxx}^2\delta h_x\left(h^{-1}+h'^2c_x^2\right)\\
2879: -\int_\Omega{dx}\,\left(h^3-h'^3\right)\left(h'_{xxx}+4c_xc_{xx}\right)\delta
2880: h_{xxx}\\
2881: + \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].
2882: \end{multline*}
2883: %
2884: %
2885: %
2886: Using the lower bound on $h(x,t)\geq h_{\mathrm{min}}>0$ and Young's
2887: first inequality, this equation is transformed into an inequality,
2888: %
2889: %
2890: %
2891: %
2892: %
2893: \begin{multline*}
2894: \tfrac{1}{2}\frac{d}{dt}\int_\Omega{dx}\,\delta h_x^2 +h_{\mathrm{min}}^3\int_\Omega{dx}\,\delta
2895: h_{xxx}^2\leq
2896: %
2897: %
2898: \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\\
2899: %
2900: %
2901: %
2902: %
2903: +\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\\
2904: %
2905: %
2906: +
2907: \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,
2908: \end{multline*}
2909: %
2910: %
2911: %
2912: where $\kappa_1$, $\kappa_2$, and $\kappa_3$ are arbitrary positive constants.
2913:  By choosing $\kappa_1+\kappa_2+\kappa_3=h_{\mathrm{min}}^3$, the inequality
2914:  simplifies to
2915: %
2916: %
2917: \begin{multline*}
2918: 2\kappa\frac{d}{dt}\int_\Omega{dx}\,\delta h_x^2\leq
2919: %
2920: %
2921: \int_\Omega{dx}\,\delta{h}_x^2\left(h^{-1}+h'^2c_x^2\right)^2\\+
2922: %
2923: \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],
2924: %
2925: \end{multline*}
2926: %
2927: %
2928: %
2929: where $\kappa$ is another positive constant.  We integrate over the time
2930: interval $\left[0,T\right]$, and use the fact that $\|\delta{h}_x\|_2\left(0\right)=0$
2931: to obtain
2932: %
2933: %
2934: %
2935: \begin{multline*}
2936: 2\kappa\sup_{\tau\in\left[0,T\right]}\|\delta{h}_x\|_2^2\left(\tau\right)
2937: \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\\
2938: +\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].
2939: \end{multline*}
2940: %
2941: %
2942: %
2943: %
2944: The Poincar\'e inequality can be combined with the one-dimensional differential
2945: inequalities discussed in Appendix A to yield the relation $\|f\|_\infty\leq
2946: \kappa_P\|f_x\|_2$, where $f$ is some mean-zero function and $\kappa_P$ is
2947: an $f$-independent constant.  We therefore arrive at
2948: %
2949: %
2950: %
2951: \begin{multline*}
2952: 2\kappa\sup_{\tau\in\left[0,T\right]}\|\delta{h}_x\|_2^2\left(\tau\right)
2953: %
2954: %
2955: \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\\
2956: %
2957: %
2958: +\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).
2959: \end{multline*}
2960: %
2961: %
2962: %
2963: %
2964: Using the results of Sec.~\ref{sec:regularity}, it is
2965: readily shown that $h^{-1}+h'^2c_x^2\in L^2\left(0,T;L^{\infty}(\Omega)\right)$,
2966: and that the functions $h'_{xxx}+4c_xc_{xx}$ and $h_x'+h_xc_x^2$ belong to
2967: the class $L^2\left(0,T;L^2(\Omega)\right)$.  By choosing $T$
2968: sufficiently small, it is possible to impose
2969: %
2970: %
2971: %
2972: \[
2973: \frac{1}{2\kappa}\left[\int_0^T{dt}\,\|h^{-1}+h'^2c_x^2\|_\infty^2+\kappa_P^2\int_0^T{dt}\left(\|h'_{xxx}+4c_xc_{xx}\|_2^2+\|h_x'+h_xc_x^2\|_2^2\right)\right]<1,
2974: \]
2975: %
2976: %
2977: which in turn forces $\sup_{\tau\in\left[0,T\right]}\|\delta{h}_x\|_2^2=0$,
2978: and hence the solution is unique.
2979: 
2980: %%\bibliographystyle{unsrt}
2981: %\bibliographystyle{plain}
2982: %\bibliography{thin_film_bibliography}
2983:  
2984: \begin{thebibliography}{10}
2985: 
2986: \bibitem{Aarts2005}
2987: D.~G. A.~L. Aarts, R.~P.~A. Dullens, and H.N.W. Lekkerherker.
2988: \newblock Interfacial dynamics in demixing systems with ultralow interfacial
2989:   tension.
2990: \newblock {\em New J. Phys.}, 7:14, 2005.
2991: 
2992: \bibitem{Friedman1990}
2993: F.~Bernis and A.~Friedman.
2994: \newblock Higher order nonlinear degenerate parabolic equations.
2995: \newblock {\em J. Differential Equations}, 83:179--206, 1990.
2996: 
2997: \bibitem{chaos_Berthier}
2998: L.~Berthier, J.~L. Barrat, and J.~Kurchan.
2999: \newblock Phase separation in a chaotic flow.
3000: \newblock {\em Phys. Rev. Lett.}, 86:2014--2017, 2001.
3001: 
3002: \bibitem{Berti2005}
3003: S.~Berti, G.~Boffetta, M.~Cencini, and A.~Vulpiani.
3004: \newblock Turbulence and coarsening in active and passive binary mixtures.
3005: \newblock {\em Phys. Rev. Lett.}, 95:224501, 2005.
3006: 
3007: \bibitem{Bertozzi1996}
3008: A.~L. Bertozzi and M.~Pugh.
3009: \newblock The lubrication approximation for thin viscous films: regularity and
3010:   long-time behaviour of weak solultions.
3011: \newblock {\em Comm. Pure Appl. Math.}, XLIX:85, 1996.
3012: 
3013: \bibitem{Bertozzi1998}
3014: A.~L. Bertozzi and M.~C. Pugh.
3015: \newblock Long-wave instabilities and saturation in thin film equations.
3016: \newblock {\em Comm. Pure Appl. Math.}, LI:0625, 1998.
3017: 
3018: \bibitem{Bray_advphys}
3019: A.~J. Bray.
3020: \newblock Theory of phase-ordering kinetics.
3021: \newblock {\em Adv. Phys.}, 43:357--459, 1994.
3022: 
3023: \bibitem{CH_orig}
3024: J.~W. Cahn and J.~E. Hilliard.
3025: \newblock {Free energy of a nonuniform system. I. Interfacial energy}.
3026: \newblock {\em J. Chem. Phys}, 28:258--267, 1957.
3027: 
3028: \bibitem{ChungH2004}
3029: H.~Chung and R.J. Composto.
3030: \newblock Breakdown of dynamic scaling in thin film binary liquids undergoing
3031:   phase separation.
3032: \newblock {\em Phys. Rev. Lett.}, 92:185704--1, 2004.
3033: 
3034: \bibitem{Murray1981}
3035: D.~S. Cohen and J.~D. Murray.
3036: \newblock A generalized diffusion model for growth and dispersal in a
3037:   population.
3038: \newblock {\em J. Math. Biology}, 12:237, 1981.
3039: 
3040: \bibitem{Spelt2007}
3041: H.~Ding, P.~D.~M. Spelt, and C.~Shu.
3042: \newblock Diffuse interface model for incompressible two-phase flows with large
3043:   density ratios.
3044: \newblock {\em J. Comp. Phys.}, 226:2078, 2007.
3045: 
3046: \bibitem{DoeringGibbon}
3047: C.~R. Doering and J.~D. Gibbon.
3048: \newblock {\em Applied Analysis of the Navier--Stokes Equations}.
3049: \newblock Cambridge University Press, Cambridge, 1995.
3050: 
3051: \bibitem{Elliott_varmob}
3052: C.~M. Elliott and H.~Garcke.
3053: \newblock {The Cahn--Hilliard equation with degenerate mobility}.
3054: \newblock {\em SIAM J. Math. Anal.}, 27:403--423, 1996.
3055: 
3056: \bibitem{Elliott_Zheng}
3057: C.~M. Elliott and S.~Zheng.
3058: \newblock {On the Cahn--Hilliard equation}.
3059: \newblock {\em Arch. Rat. Mech. Anal.}, 96:339--357, 1986.
3060: 
3061: \bibitem{Gajewski_nonlocal}
3062: H.~Gajewski and K.~Zacharias.
3063: \newblock On a nonlocal phase separation model.
3064: \newblock {\em J. Math. Anal. Appl.}, 286:11--31, 2003.
3065: 
3066: \bibitem{Holm2005}
3067: D.~D. Holm and V.~Putkaradze.
3068: \newblock Aggregation of finite size particles with variable mobility.
3069: \newblock {\em Phys. Rev. Lett.}, 95:226106, 2005.
3070: 
3071: \bibitem{Karim2002}
3072: A.~Karim, J.~F. Douglas, L.~P. Sung, and B.~D. Ermi.
3073: \newblock Self-assembly by phase separation in polymer thin films.
3074: \newblock {\em Encyclopedia of Materials: Science and Technology}, page 8319,
3075:   2002.
3076: 
3077: \bibitem{Krausch1994}
3078: G.~Krausch, E.~J. Kramer, M.~H. Rafailovich, and J.~Sokolov.
3079: \newblock Self-assembly of a homopolymer mixture via phase separation.
3080: \newblock {\em Appl. Phys. Lett.}, 64:2655, 1994.
3081: 
3082: \bibitem{Lacasta1995}
3083: A.~M. Lacasta, J.~M. Sancho, and F.~Sagues.
3084: \newblock Phase separation dynamics under stirring.
3085: \newblock {\em Phys. Rev. Lett.}, 75:1791, 1995.
3086: 
3087: \bibitem{Laugesen2002}
3088: R.~S. Laugesen and M.~C. Pugh.
3089: \newblock Heteroclinic orbits, mobility parameters and stability for thin film
3090:   type equations.
3091: \newblock {\em J. Differential Equations}, 95:1--29, 2002.
3092: 
3093: \bibitem{LowenTrus}
3094: J.~Lowengrub and L.~Truskinowsky.
3095: \newblock {Quasi-incompressible Cahn--Hilliard fluids and topological
3096:   transitions}.
3097: \newblock {\em Proc. R. Soc. Lond. A}, 454:2617--2654, 1998.
3098: 
3099: \bibitem{Putkaradze2005}
3100: K.~Mertens, V.~Putkaradze, D.~Xia, and S.R. Brueck.
3101: \newblock Theory and experiment for one-dimensional directed self-assembly of
3102:   nanoparticles.
3103: \newblock {\em J. App. Phys.}, 98:034309, 2005.
3104: 
3105: \bibitem{Myers1998}
3106: T.~G. Myers.
3107: \newblock Thin films with high surface tension.
3108: \newblock {\em SIAM Review}, 40:441, 1998.
3109: 
3110: \bibitem{ONaraigh2006}
3111: L.~{\'O N\'araigh} and J.-L. Thiffeault.
3112: \newblock {Bubbles and Filaments: Stirring a Cahn--Hilliard Fluid}.
3113: \newblock {\em Phys. Rev. E}, 75:016216, 2007.
3114: 
3115: \bibitem{ONaraigh2007}
3116: L.~{\'O N\'araigh} and J.-L. Thiffeault.
3117: \newblock Dynamical effects and phase separation in cooled binary fluid films.
3118: \newblock {\em Phys. Rev. E}, 76:035303(R), 2007.
3119: 
3120: \bibitem{Oron1997}
3121: A.~Oron, S.~H. Davis, and S.~G. Bankoff.
3122: \newblock Long-scale evolution of thin liquid films.
3123: \newblock {\em Rev. Mod. Phys.}, 69:931, 1997.
3124: 
3125: \bibitem{Book_Parsegian2006}
3126: V.~A. Parsegian.
3127: \newblock {\em {Van der Waals Forces}}.
3128: \newblock Cambridge University Press, New York, 2001.
3129: 
3130: \bibitem{Sung1996}
3131: L.~Sung, A.~Karim, J.F. Douglas, and C.C. Han.
3132: \newblock Dimensional crossover in the phase separation kinetics of thin
3133:   polymer blend films.
3134: \newblock {\em Phys. Rev. Lett.}, 76:4368, 1996.
3135: 
3136: \bibitem{TaylorPDEbook}
3137: M.~E. Taylor.
3138: \newblock {\em Partial Differential Equations III}.
3139: \newblock Springer, New York, 1996.
3140: 
3141: \bibitem{WangH2000}
3142: H.~Wang and R.~J. Composto.
3143: \newblock Thin film polymer blends undergoing phase separation and wetting:
3144:   identification of early, intermediate, and late stages.
3145: \newblock {\em J. Chem. Phys.}, 113:10386, 2000.
3146: 
3147: \bibitem{WangW2003}
3148: W.~Wang, T.~Shiwaku, and T.~Hashimoto.
3149: \newblock Phase separation dynamics and pattern formation in thin films of a
3150:   liquid crystalline copolyester in its biphasic region.
3151: \newblock {\em Macromolecules}, 36:8088, 2003.
3152: 
3153: \bibitem{Wieland2006}
3154: S.~Wieland and H.~Garcke.
3155: \newblock {Surfactant spreading on thin viscous films: Nonnegative solutions of
3156:   a coupled degenerate system}.
3157: \newblock {\em SIAM J. Math. Anal.}, 37:2025, 2006.
3158: 
3159: \bibitem{Xia2004}
3160: D.~Xia and S.~Brueck.
3161: \newblock A facile approach to directed assembly of patterns of nanoparticles
3162:   using interference lithography and spin coating.
3163: \newblock {\em Nano Letters}, 4:1295, 2004.
3164: 
3165: \end{thebibliography}
3166: 
3167: \end{document}
3168: