math0611530/i.tex
1: \documentclass[reqno,a4paper]{siamltex}
2: 
3: %\usepackage[notref,notcite]{showkeys}
4: \usepackage{graphicx}
5: \usepackage{color}
6: 
7: \bibliographystyle{siam}
8: 
9: \usepackage{amsmath,amsfonts}
10: \usepackage{mathrsfs}
11: \usepackage{a4wide}
12: \usepackage{lscape}
13: \ifx\theorem\undefined % then we're doing amslatex
14: \newtheorem{theorem}{Theorem}
15: \DeclareMathOperator{\grad}{grad}
16: \DeclareMathOperator\supp{supp}
17: \fi
18: \DeclareMathOperator\tr{tr}
19: 
20: \newcommand{\figref}[1]{Figure~\ref{#1}}
21: 
22: \newcounter{remark}[section]
23: \def\theremark{\thesection.\arabic{remark}.}
24: \newenvironment{remark}%
25:     {\par\medbreak\refstepcounter{remark}%
26:          {\noindent\bf Remark~\theremark\ }}%
27:     {\par\medbreak}
28: %
29: \DeclareMathOperator{\dd}{d\!}
30: \def\NN{\mathbb {N}}
31: \def\RR{\mathbb {R}}
32: \def\ds{\displaystyle}
33: \def\inv{(-\Delta)^{-1}}
34: \def\MP{\mathrm{MP}}
35: \def\wMP{w_\MP^{}}
36: \def\un#1{\,\mathrm{#1}}
37: %
38: \long\def\meta#1{\texttt{#1}}
39: \def\pref#1{(\text{\ref{#1}})}
40: \def\mod#1{\left|#1\right|}
41: \def\Mod#1{\left\|#1\right\|}
42: \def\isdef{:=}
43: %\DeclareMathOperator\supp{supp}
44: %
45: \def\N{\mathbb N}
46: \def\Z{\mathbb Z}
47: \def\R{\mathbb R}
48: \def\M{\mathscr{M}}
49: \def\G{\mathscr{G}}
50: \let\e\varepsilon
51: \let\d\delta
52: \def\blambda{\bar\lambda}
53: \def\limp{P_{\mathrm{imp}}}
54: \def\lambdacr{P_{\mathrm{cr}}}
55: \def\l{\lambda}
56: \def\bw{\overline w}
57: \def\bphi{\overline \phi}
58: \def\Omq{{\Omega_\frac{1}{4}}}
59: \def\Xq{{X_\frac{1}{4}}}
60: \def\bA{\mbox{\boldmath $A$}}
61: \def\bAt{\mbox{\boldmath $\tilde A$}}
62: \def\bAx{{\bA_x^{}}}
63: \def\bAxT{{\bA_x^{T}}}
64: \def\bAy{{\bA_y^{}}}
65: \def\bAyT{{\bA_y^{T}}}
66: \def\bAxx{{\bA_{xx}^{}}}
67: \def\bAyy{{\bA_{yy}^{}}}
68: \def\bAxy{{\bA_{xy}^{}}}
69: \def\bAxyT{{\bA_{xy}^{T}}}
70: \def\bAxyt{{\bAt_{xy}^{}}}
71: \def\bAbih{{\bA_{\Delta^2}^{}}}
72: \def\bAbihinv{{\bA_{\Delta^2}^{-1}}}
73: \def\bAprod{{\bA_{\langle,\rangle}^{}}}
74: \def\bOne{\mbox{\boldmath $1$}}
75: \def\bNull{\mbox{\boldmath $0$}}
76: \def\bC{\mbox{\boldmath $C$}}
77: \def\bCf{{\bC_{\rm f}^{}}}
78: \def\bCb{{\bC_{\rm b}^{}}}
79: \def\bS{\mbox{\boldmath $S$}}
80: \def\bLambda{\mbox{\boldmath $\Lambda$}}
81: \def\bLambdaxx{{\bLambda_{xx}^{}}}
82: \def\bLambdayy{{\bLambda_{yy}^{}}}
83: \def\bLambdaxy{{\bLambda_{xy}^{}}}
84: \def\bLambdabih{{\bLambda_{\Delta^2}^{}}}
85: \def\bD{\mbox{\boldmath $D$}}
86: \def\bB{\mbox{\boldmath $B$}}
87: \def\bG{\mbox{\boldmath $G$}}
88: \def\bGt{\mbox{\boldmath $\tilde G$}}
89: %
90: \def\vKD{Von K\'arm\'an-Donnell}%
91: %  dashed integral
92: %
93: \def\Xint#1{\mathchoice
94:    {\XXint\displaystyle\textstyle{#1}}%
95:    {\XXint\textstyle\scriptstyle{#1}}%
96:    {\XXint\scriptstyle\scriptscriptstyle{#1}}%
97:    {\XXint\scriptscriptstyle\scriptscriptstyle{#1}}%
98:    \!\int}
99: \def\XXint#1#2#3{{\setbox0=\hbox{$#1{#2#3}{\int}$}
100:      \vcenter{\hbox{$#2#3$}}\kern-.5\wd0}}
101: \def\ddashint{\Xint=}
102: \def\dashint{\Xint-}
103: %
104: \def\frparbcenter#1{\framebox(43,60)[t]{\parbox{43mm}{\begin{center}#1\end{center}}}}
105: 
106: \author{Ji\v r\'I Hor\'ak\thanks{Universit\"at K\"oln, Germany}
107: \and Gabriel J. Lord\thanks{Heriot-Watt University, Edinburgh, United Kingdom}
108: \and Mark A. Peletier\thanks{Technische Universiteit Eindhoven, The Netherlands}}
109: \title{Numerical variational methods applied to cylinder buckling}
110: \date{DRAFT \today}
111: 
112: %
113: \begin{document}
114: \maketitle
115: 
116: \begin{abstract}
117: We review and compare different computational variational methods applied to a 
118: system of fourth order equations that arises as a model of cylinder buckling. 
119: We describe both the discretization and implementation, in particular how to 
120: deal with a 1 dimensional null space. We show that we can construct many 
121: different solutions from a complex energy surface.
122: We examine numerically convergence in the spatial discretization and
123: in the domain size.  
124: Finally we give a physical interpretation of some of the solutions found.
125: \end{abstract}
126: 
127: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
128: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
129: \section{Introduction}
130: We describe complementary approaches to finding solutions of
131: systems of fourth order elliptic PDEs. The techniques are applied to a
132: problem that arises in the classic treatment of an isotropic
133: cylindrical shell under axial compression but are also applicable to a
134: wide range of problems such as waves on a suspension bridge
135: \cite{Ho1,HoMcK}, the Fu\v{c}\'{\i}k spectrum of the Laplacian
136: \cite{HoRe}, or the formation of microstructure \cite{SH,FM}.
137: 
138: The cylindrical shell offers a computationally 
139: challenging and physically relevant problem with a complex energy
140: surface.  We take as our model for the shell the \vKD{}
141: equations which can be rescaled \cite{HoLoPe1} to the form
142: \begin{align}
143: \label{eq:main_w2}
144: &  \Delta^2  w +  \l  w_{ x x} 
145:   - \phi_{ x x} - 2\,[ w,\phi] = 0, \\
146: &\Delta^2 \phi +  w_{ x x} + [ w, w] = 0,
147: \label{eq:main_phi2}
148: \end{align}
149: where the bracket is defined as
150: \begin{equation}
151: [u,v] = \frac12 u_{xx}v_{yy} + \frac 12 u_{yy}v_{xx} - u_{xy}v_{xy}.
152: \label{eq:bracket}
153: \end{equation}
154: The function $w$ is a scaled inward radial displacement measured from
155: the unbuckled (fundamental) state, $\phi$ is the Airy stress function,
156: and $\l\in(0,2)$ is a load parameter. The unknowns $w$ and $\phi$ are
157: defined on a two-dimensional spatial domain
158: $\Omega=(-a,a)\times(-b,b)$, where $x\in(-a,a)$ is the axial and
159: $y\in(-b,b)$ is the tangential coordinate. 
160: Since the $y$-domain $(-b,b)$ represents the circumference of the cylinder,
161: the following boundary conditions are prescribed:
162: \begin{subequations}
163: \label{def:BC}
164: \begin{align}
165: &w\text{ is periodic in $y$,\quad  and} \quad  w_x = (\Delta  w)_x = 0
166: \text{ at }x=\pm a,
167: \label{def:BCw} \\
168: &\phi\text{ is periodic in $y$,\quad  and} \quad  \phi_x = (\Delta  \phi)_x = 0
169: \text{ at }x=\pm a,
170: \label{def:BCphi}
171: \end{align}
172: \end{subequations}
173: as shown in Fig.~\ref{fig:geom}~(i), (ii).
174: 
175: \begin{figure}[htbp]
176:   \centering
177:   \setlength{\unitlength}{1mm}
178:   \begin{picture}(140,78)
179:     %\put(0,0){\framebox(140,78){}}
180:     \put(10,45){\includegraphics[width=4.9cm]{figs/geom_cyl}}
181:     \put(67,60){\includegraphics[width=1.35cm]{figs/geom_legend}}
182:     \put(0,0){\includegraphics[width=6.75cm]{figs/geom_full}}
183:     \put(72,0){\includegraphics[width=6.75cm]{figs/geom_quarter}}
184:     \put(83,66){periodic boundary condition for $w,\phi$}
185:     \put(83,59.5){$w_\nu=(\Delta w)_\nu=\phi_\nu=(\Delta\phi)_\nu=0$}
186: 
187:     \put(10,72){(i)}
188:     \put(16,55){\small $w$}
189:     \put(58,72.5){\small $x$}
190:     \put(42,58.5){\small $y$}
191:     \put(16.5,50.5){\small $-a$}
192:     \put(46,65.5){\small $a$}
193: 
194:     \put(-2,36){(ii)}
195:     \put(35,29){\small $w$}
196:     \put(55,29){\small $x$}
197:     \put(8.5,36){\small $y$}
198:     \put(16.5,8){\small $-a$}
199:     \put(46.5,24){\small $a$}
200:     \put(51,4){\small $-b$}
201:     \put(12.5,28){\small $b$}
202:     \put(39,2){\small $\Omega$}
203: 
204:     \put(70,36){(iii)}
205:     \put(107,29){\small $w$}
206:     \put(127,29){\small $x$}
207:     \put(80.5,36){\small $y$}
208:     \put(88.5,8){\small $-a$}
209:     \put(123,4){\small $-b$}
210:     \put(109.7,3){\small $\Omq$}
211:   \end{picture}
212:   \caption{(i) The geometry of the cylinder, (ii) the computational domain and the boundary conditions, (iii) one quarter of the domain and the corresponding boundary conditions.}
213:   \label{fig:geom}
214: \end{figure}
215: 
216: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
217: \subsection{Functional setting}
218: We search for weak solutions $w$, $\phi$ of
219: (\ref{eq:main_w2}--\ref{def:BC}) in the space
220: \[
221: X=\left\{\psi\in H^2(\Omega) : \psi_x(\pm a,\cdot)=0,\ \psi \text{ is
222:     periodic in } y, \text{ and }\,\int_\Omega\psi = 0\right\}
223: \]
224: with norm
225: \[
226:   \Mod{w}^2_X = \int_\Omega\bigl(\,\Delta w^2 +\Delta\phi_1^2\,\bigr),
227: \]
228: where $\phi_1\in H^2(\Omega)$ is the unique solution of
229: \begin{equation}
230:   \label{eq:phi1}
231:   \Delta^2\phi_1 = -w_{xx}, \qquad \phi_1 \text{ satisfies~\pref{def:BCphi}, }
232: \qquad \text{and}\qquad \int_\Omega \phi_1 = 0.  
233: \end{equation}
234: This norm is equivalent to the $H^2$-norm on $X$, and with the
235: appropriate inner product $\langle\cdot,\cdot\rangle_X$ the space $X$
236: is a Hilbert space. Alternatively, if the load parameter $\l\in(0,2)$
237: is fixed, another norm
238: \[
239:   \Mod{w}^2_{X,\l} = \int_\Omega\bigl(\,\Delta w^2 +\Delta\phi_1^2 
240:   - \l w_x^2\,\bigr)
241: \]
242: can be used. Because of the estimate
243: \[
244: \int_\Omega w_x^2 = -\int_\Omega ww_{xx} = \int_\Omega w\Delta^2 \phi_1
245: = \int_\Omega \Delta w\Delta \phi_1 \leq \frac12 \int_\Omega \Delta w  ^2 + \frac12 \int_\Omega \Delta \phi_1^2 
246: = \frac12 \Mod{w}_X^2,
247: \]
248: it is equivalent to $\Mod{\cdot}_X$ and hence also to the $H^2$-norm
249: on $X$. The corresponding inner product will be denoted
250: $\langle\cdot,\cdot\rangle_{X,\l}$.
251: 
252: Equations~(\ref{eq:main_w2}--\ref{eq:main_phi2}) are related to the
253: stored energy $E$, the average axial shortening $S$, and the total
254: potential given by
255: \begin{equation}
256: \label{def:energy}
257: E(w) :=  \frac12\int_\Omega \left( \Delta  w^2 +  \Delta \phi^2\right), \qquad
258: \quad S(w):= \frac12 \int_\Omega  w_{ x}^2, \qquad\quad F_\l=E-\l S.
259: \end{equation}
260: Note that the function $\phi$ in~\pref{def:energy} is determined from
261: $w$ by solving~\pref{eq:main_phi2} with boundary
262: conditions~\pref{def:BCphi}. All the functionals $E$, $S$, and $F_\l$ belong to
263: $C^1(X)$, i.e., are continuously Fr\'echet differentiable.
264: 
265: The fact that~\pref{eq:main_w2} is a reformulation of the stationarity
266: condition $F_\l'=E'-\lambda S'=0$ will be important in
267: Sec.~\ref{sec:discretization}, and we therefore briefly sketch the
268: argument. It is easy to see that
269: \[
270: S'(w)\cdot h=-\int_\Omega w_{xx}h.
271: \]
272: For $E'(w)\cdot h$, let $w,\phi\in X$ solve \pref{eq:main_phi2} and
273: $h,\psi\in X$ solve $ \Delta^2 \psi = -h_{xx}-[h,h]-2[w,h] $. Then,
274: assuming sufficient regularity on $w$,
275: \begin{align}
276: E(w+h)-E(w) &= \int_\Omega\Delta w\Delta h +
277: \frac{1}{2}\int_\Omega(\Delta h)^2 + \int_\Omega\Delta\phi\Delta\psi +
278: \frac{1}{2}\int_\Omega(\Delta\psi)^2 \notag \\
279: &= \int_\Omega\left(h\Delta^2w - h\phi_{xx} - 2[w,h]\phi\right)
280: + \frac{1}{2}\int_\Omega(\Delta h)^2 + \frac{1}{2}\int_\Omega(\Delta\psi)^2
281: - \int_\Omega[h,h]\phi\ , \notag
282: \end{align}
283: where we used integration by parts several times. The last three
284: integrals are $O(\Mod{h}_X^2)$ for $\Mod{h}_X\to 0$ and it can be
285: shown by integration by parts that
286: \begin{equation}
287: \int_\Omega [w,h]\,\phi = \int_\Omega h\,[w,\phi]\ .
288: \label{eq:w_h_phi}
289: \end{equation}
290: Therefore 
291: \[
292: F_\l'(w)\cdot h = E'(w)\cdot h - \lambda S'(w)\cdot h = 
293: \int_\Omega h\Bigl( \Delta^2  w +  \l  w_{ x x} 
294:   - \phi_{ x x} - 2\,[ w,\phi]\Bigr).
295: \]
296: 
297: 
298: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
299: \subsection{Review of some variational numerical methods}\label{sec:var_num}
300: We now describe the variational methods used to find numerical
301: approximations of critical points of the total potential $F_\l$. In
302: our numerical experiments these methods are accompanied by Newton's
303: method and continuation. The advantage of this approach is that it
304: combines the knowledge of global features of the energy landscape with
305: local ones of a neighborhood of a critical point. The details related
306: to spatial discretization will be discussed in
307: Sec.~\ref{sec:discretization}, the Newton-based methods in
308: Sec.~\ref{sec:Newton}.
309: 
310: % % % % % % % % % % % % % % % % % % % % % % % % % % % % % % % % % % 
311: \subsubsection{Steepest descent method (SDM)}\label{sec:SDM}
312: Let the load parameter $\l\in(0,2)$ be fixed; we work in a discretized
313: version of $(X,\langle\cdot,\cdot\rangle_{X,\l})$. We try to minimize
314: the total potential $F_\l$ by following its gradient flow. We solve the initial value
315: problem
316: \[
317: \frac{d}{dt}w(t)=-\nabla_\l F_\l(w(t))\,,\qquad w(0)=w_0\,,
318: \]
319: with a suitable starting point $w_0$ on some interval $(0,T]$. This
320: problem is then discretized in $t$.
321: 
322: In~\cite{HoLoPe1} it was shown that $w=0$ is a local minimizer of
323: $F_\l$. Indeed, if $\Mod{w_0}_{X,\l}$ is small, the numerical solution
324: $w(t)$ converges to zero as $t$ tends to infinity. If, on the other hand,
325: $\Mod{w_0}_{X,\l}$ is large, the numerical solution $w(t)$ stays
326: bounded away from zero. In most of our experiments, the numerical
327: algorithm did not converge for $t\to\infty$ in the large norm case. The only
328: exception for a relatively small value of $\l$ will be mentioned later
329: in Sec.~\ref{sec:domain_size}. Nevertheless, for a sufficiently large
330: computational domain $\Omega$ and a sufficiently large $t>0$ we obtain
331: $F_\l(w(t))<0$. Such a state $w(t)$ is needed for the mountain pass
332: algorithm as explained below. Existence of this state was also proved
333: in~\cite{HoLoPe1}.
334: 
335: % % % % % % % % % % % % % % % % % % % % % % % % % % % % % % % % % % 
336: \subsubsection{Mountain-pass algorithm (MPA)}\label{sec:MPA}
337: %Should there be a brief description of what a MP-point is?
338: The algorithm was first proposed in \cite{ChMcK1} for a second order
339: elliptic problem in 1D and extended in \cite{HoMcK} to a
340: fourth-order problem in 2D. We give a brief description of the
341: algorithm here.
342: 
343: Let the load $\l\in(0,2)$ be fixed; we work again in a
344: discretized version of $(X,\langle\cdot,\cdot\rangle_{X,\l})$. We
345: denote $w_1=0$ the local minimum of $F_\l$ and take a point $w_2$ such
346: that $F_\l(w_2)<0$ (in practice this point is found using the SDM). We
347: take a discretized path $\{z_m\}_{m=0}^{p}$ connecting $z_0=w_1$ with
348: $z_p=w_2$. After finding the point $z_m$ at which $F_\l$ is maximal
349: along the path, this point is moved a small distance in the direction
350: of the steepest descent $-\nabla_\l F_\l(z_m)$. Thus the path has been
351: deformed and the maximum of $F_\l$ lowered. This deforming of the path
352: is repeated until the maximum along the path cannot be lowered any
353: more: a critical point $\wMP$ has been reached. Figure~\ref{fig:mpa}
354: illustrates the main idea of the method.
355: 
356: \begin{figure}[htbp]
357:   \begin{center}
358:     \setlength{\unitlength}{1mm}
359:     \begin{picture}(100,67)
360:       \put(0,0){{\includegraphics[width=10cm]{figs/mpa}}}
361:       \put(4,21){$w_1$}
362:       \put(86,61){$w_2$}
363:       \put(47,45.5){$z_m$}
364:       \put(49,34.5){$z_m^\mathrm{new}$}
365:       \put(59,33.5){$-\nabla_\l F_\l(z_m)$}
366:       \put(53,19){$w_\MP^{}$}
367:       \put(93.5,12){$X$}
368:     \end{picture}
369:   \end{center}
370:   \caption{Deforming the path in the main loop of the mountain pass
371:     algorithm: point $z_m$ is moved a small distance in the direction
372:     $-\nabla_\l F_\l(z_m)$ and becomes $z_m^\mathrm{new}$. This step
373:     is repeated until the mountain pass point $\wMP$ is reached.}
374:   \label{fig:mpa}
375: \end{figure}
376: 
377: The mountain-pass algorithm is local in its nature. The numerical
378: solution $\wMP$ it finds has the mountain-pass property in a certain
379: neighborhood only. The choice of the path endpoint $w_2$ may influence
380: to which critical point the algorithm converges. Different choices of
381: $w_2$ are in turn achieved by choosing different initial points $w_0$
382: in the SDM.
383: 
384: % % % % % % % % % % % % % % % % % % % % % % % % % % % % % % % % % % 
385: \subsubsection{Constrained steepest descent method (CSDM)}
386: We fix the amount of shortening $S$ of the cylinder. This is often
387: considered as what actually occurs in experiments. We work now in a
388: discretized version of $(X,\langle\cdot,\cdot\rangle_X)$. Let $C>0$ be
389: a fixed number and define a set of $w$ with constant shortening
390: \begin{equation}
391:   \label{eq:constraint}
392:   \M=\{w\in X\,:\, S(w)=C\}\,.
393: \end{equation}
394: Critical points of $E$ under this constraint are critical points of
395: $F_\l$, where $\l$ is a Lagrange multiplier. The simplest such points
396: are local minima of the stored energy $E$ on~$\M$. We need to follow the gradient flow
397: of $E$ on~$\M$, hence we solve the initial value problem
398: \[
399: \frac{d}{dt}w(t)=-P_{w(t)}\nabla E(w(t))\,,\qquad w(0)=w_0\in\M\,,
400: \]
401: for $t>0$. $P_w$ denotes the orthogonal projection in $X$ on the tangent space of
402: $\M$ at $w\in\M$:
403: \[
404: P_w u= u - \frac{\langle\nabla S(w),u\rangle_X}{\Mod{\nabla S(w)}_X^2} \nabla S(w)\,.
405: \]
406: The details of the algorithm can be found in~\cite{Ho1}. The initial
407: value problem is solved by repeating the following two steps: given a
408: point $w\in\M$ find $\bar w=w - \Delta t P_w\nabla E(w)$, where
409: $\Delta t>0$ is small, and define $w_{\rm new}=c\bar w$, where the
410: scaling coefficient $c$ is chosen so that $w_{\rm new}\in\M$. The
411: algorithm is stopped when $\Mod{P_w\nabla E(w)}_X$ is smaller than a
412: prescribed tolerance. The corresponding load is given by
413: $$\l=\frac{\langle\nabla S(w),\nabla E(w)\rangle_X}{\Mod{\nabla
414:     S(w)}_X^2}.$$
415: 
416: % % % % % % % % % % % % % % % % % % % % % % % % % % % % % % % % % % 
417: \subsubsection{Constrained mountain-pass algorithm (CMPA)}
418: Let $C>0$ and~$\M$ be the set of $w$ with constant shortening given
419: in~\pref{eq:constraint}. We would like to find mountain-pass points of
420: $E$ on~$\M$. The method has been described in~\cite{Ho1} in detail. We
421: need two local minima $w_1, w_2$ of $E$ on~$\M$ which can be found
422: using the CSDM. The algorithm is then similar to the MPA. We take a
423: discretized path $\{z_m\}_{m=0}^{p}\subset\M$ connecting $z_0=w_1$
424: with $z_p=w_2$. After finding the point $z_m$ at which $E$ is maximal
425: along the path, this point is moved a small distance in the tangent
426: space to~$\M$ at $z_m$ in the direction of the steepest descent
427: $-P_{z_m}\nabla E(z_m)$ and than scaled (as in the CSDM) to come back to
428: $\M$. Thus the path has been deformed on~$\M$ and the maximum of $E$
429: lowered. This deforming of the path is repeated until the maximum
430: along the path cannot be lowered any more: a mountain-pass point of
431: $E$ on~$\M$ has been reached. The load $\l$ is computed as in the CSDM.
432: 
433: The choice of the end points $w_1$ and $w_2$ will in general influence to which
434: critical point the algorithm converges.
435: 
436: 
437: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
438: \subsection{Computational Domains}
439: We consider the problem on the domain $\Omega$
440: (\figref{fig:geom}~(ii)) both without further restraints and under a
441: symmetry assumption, which reduces the computational complexity. In
442: the latter case we assume
443: \begin{equation}
444:   \label{eq:sym}
445:   \begin{array}{c}
446:     w(x,y)=w(-x,y)=w(x,-y) \\
447:     \phi(x,y)=\phi(-x,y)=\phi(x,-y)
448:   \end{array}
449:   \quad \mbox{for}\ (x,y)\in\Omega\ .
450: \end{equation}
451: By looking for solutions $w,\phi\in X$ that satisfy~\pref{eq:sym} the
452: domain is effectively reduced to one quarter:
453: $\Omq=(-a,0)\times(-b,0)$ as shown in~\figref{fig:geom}~(iii). One needs to
454: solve (\ref{eq:main_w2}--\ref{eq:main_phi2}) only on $\Omq$ with the
455: boundary conditions
456: \begin{equation}
457:   \label{def:BCsym}
458:   w_\nu=(\Delta w)_\nu=\phi_\nu=(\Delta\phi)_\nu=0 \quad \mbox{on}\ \partial\Omq\ ,
459: \end{equation}
460: where $\nu$ denotes the outward normal direction to the boundary.
461: Hence we search for weak solutions of
462: (\ref{eq:main_w2}--\ref{eq:main_phi2}), \pref{def:BCsym} in the space
463: \[
464: \Xq=\left\{\psi\in H^2(\Omq) : \psi_\nu=0 \text{ on } \partial\Omq,
465:   \text{ and }\,\int_\Omq\psi = 0\right\}\ .
466: \]
467: We can then use \pref{eq:sym} to extend these functions to the whole
468: $\Omega$.
469: 
470: We have performed numerical experiments both with and without the
471: symmetry assumption. For the sake of simplicity we will give a
472: detailed description of the numerical methods for the second case only
473: where the boundary conditions are the same on all sides of $\Omq$. The
474: first case with periodic conditions on two sides of $\Omega$ is very
475: similar and will be briefly mentioned in~Remark~\ref{rem:full_domain}
476: 
477: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
478: \subsection{Solving the biharmonic equation}
479: In order to obtain $\phi$ for a given $w$, one has to
480: solve~\pref{eq:main_phi2}; to compute the norm of $w$, one has to
481: solve~\pref{eq:phi1}. Both problems are of the form
482: \begin{equation}
483:   \label{eq:biharm}
484:   \Delta^2\psi = f \text{ in } \Omq, \qquad \psi_\nu=(\Delta\psi)_\nu=0
485:   \text{ on } \partial\Omq, \qquad \int_\Omq\psi = 0,
486: \end{equation}
487: where $f\in L^1(\Omq)$ is given. If $\int_\Omq f=0$, then
488: \pref{eq:biharm} has a unique weak solution $\psi$ in $\Xq$. It is a
489: straightforward calculation to verify that the right-hand sides of
490: equations in \pref{eq:main_phi2} and \pref{eq:phi1} have zero average.
491: 
492: In the discretization described below the problem \pref{eq:biharm} is
493: treated as a system:
494: \begin{equation}
495:   \label{eq:biharm_syst}
496:   \begin{array}{c}
497:     -\Delta u = f \\ -\Delta v = u
498:   \end{array}
499:   \text{ in } \Omq, \qquad u_\nu=v_\nu=0
500:   \text{ on } \partial\Omq,
501:   \qquad \int_\Omq u = \int_\Omq v = 0.
502: \end{equation}
503: The system has a unique weak solution $(u,v)\in(H^1(\Omq))^2$. Since
504: the domain $\Omq$ has no reentrant corners, Theorem~1.4.5
505: of~\cite{MazjaNazPlam} guarantees that $v\in H^2(\Omq)$ and therefore
506: that the two formulations are equivalent.
507: 
508: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
509: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
510: \section{Finite difference discretization}\label{sec:discretization}
511: We discretize the domain $\Omq$ by a uniform mesh $(x_m,y_n)\in\Omq$
512: with $M$ points in the $x$-direction and $N$ points in the
513: $y$-direction:
514: \[
515: \begin{array}{ccc}
516:   x_m = -a + (m-\frac{1}{2})\Delta x, & \qquad & m\in\{1\ldots,M\},\\
517:   y_n = -b + (n-\frac{1}{2})\Delta y, & \qquad & n\in\{1\ldots,N\},
518: \end{array}
519: \]
520: where $\Delta x = a/M$, $\Delta y = b/N$. We represent the values of
521: some function $w$ on $\Omq$ at these points by a vector
522: $w=(w_i)_{i=1}^{MN}$, where $w_i=w(x_m,y_n)$ and $i=(n-1)M+m$. In our
523: notation we will not distinguish between $w$ as a function and $w$ as
524: a corresponding vector. The vector $w$ can also be interpreted as a
525: block vector with $N$ blocks, each containing $M$ values of a single
526: row of the mesh. We introduce the following conventions for notation:
527: \begin{itemize}
528: \item For two matrices $A^M=(a_{ij})_{i,j=1}^M$,
529:   $B^N=(b_{k\ell})_{k,\ell=1}^N$ we define $A^M \otimes B^N := (
530:   b_{k\ell}A^M)_{k,\ell=1}^N$, which is an $N\times N$ block matrix,
531:   each block is an $M\times M$ matrix.
532: \item For two vectors $u=(u_i)_{i=1}^{MN}$, $v=(v_i)_{i=1}^{MN}$ we
533:   define $u\odot v = (u_i v_i)_{i=1}^{MN}$, i.e., a product of
534:   the components.
535: \end{itemize}
536: 
537: To discretize second derivatives we use the standard central finite
538: differences (with Neumann boundary conditions \cite{RchtmyrMrtn}).
539: Let $Id^M$ denote the $M\times M$ identity matrix and define another
540: $M\times M$ matrix
541: \[
542: A_2^M=\left[
543:   \begin{array}{rrrrr}
544:     1 & -1 \\
545:     -1 & 2 & -1 \\
546:     & \ddots & \ddots & \ddots \\
547:     & & -1 & 2 & -1 \\
548:     & & & -1 & 1
549:   \end{array}
550: \right].
551: \]
552: The second derivatives $-\partial_{xx}$, $-\partial_{yy}$ and the
553: biharmonic operator $\Delta^2$ with the appropriate boundary
554: conditions are approximated by
555: \[
556:   \bAxx = \frac{1}{\Delta x^2}A_2^M\otimes Id^N, \qquad
557:   \bAyy = \frac{1}{\Delta y^2}Id^M\otimes A_2^N, \qquad
558:   \bAbih = (\bA_{xx} + \bA_{yy})^2,
559: \]
560: respectively.
561: 
562: \subsection{Discretization of $E$, $S$, and the bracket
563:   $[\cdot,\cdot]$}\label{sec:ES_disc}
564: Supposing that we can solve the discretized version of~\pref{eq:main_phi2}
565: \begin{equation}
566:   \label{eq:main_phi_d}
567:   \bAbih\phi-\bAxx w + [w,w]_2^{} = \bNull\,,
568: \end{equation}
569: we can also evaluate the energy $E$ and the shortening $S$:
570: \begin{equation}
571: E(w)=2\left(w^T \bAbih\, w + \phi^T \bAbih\,
572:   \phi\right)\Delta x\,\Delta y, \qquad S(w)=2\left(w^T \bAxx
573:   w\right)\Delta x\,\Delta y.
574: \label{eq:ES_disc}
575: \end{equation}
576: 
577: In order to solve~\pref{eq:main_phi_d} we need to be able to solve the
578: biharmonic equation and we need to choose a discretization of the
579: bracket $[\cdot,\cdot]$. This bracket appears in the equations in two
580: different roles: in equation~\pref{eq:main_phi2} the bracket is part
581: of the mapping $w\mapsto \phi$, and therefore of the definition of the
582: energy $E$; in equation~\pref{eq:main_w2}, which represents the
583: stationarity condition $E'-\lambda S'=0$, the bracket appears as a
584: result of differentiating $E$ with respect to $w$ and applying
585: partial integration. As a result, we need to use two different forms
586: of discretization for the two cases.
587: 
588: In both cases the bracket requires a discretization of the mixed derivative
589: $\partial_{xy}$. One choice is to use one-sided finite
590: differences. Define $M\times M$ matrices
591: \begin{equation}\label{eq:dx_RL}
592: A_{1L}^M=\left[
593:   \begin{array}{rrrr}
594:     0 \\
595:     -1 & 1 \\
596:     & \ddots & \ddots \\
597:     & & -1 & 1
598:   \end{array}
599: \right],\qquad
600: A_{1R}^M=\left[
601:   \begin{array}{rrrr}
602:     -1 & 1 \\
603:     & \ddots & \ddots \\
604:     & & -1 & 1 \\
605:     & & & 0
606:   \end{array}
607: \right].
608: \end{equation}
609: We choose either left or right-sided differences represented by these
610: matrices, respectively, let $A_1^M$ denote our choice
611: (cf.~Sec.~\ref{sec:bias}). The derivatives $\partial_x$, $\partial_y$,
612: and $-\partial_{xy}$ are approximated by
613: \begin{equation}\textstyle\label{eq:d_xy_RL}\textstyle
614: \bAx=\frac{1}{\Delta x}\,A_1^M\otimes Id^N, \qquad
615: \bAy=\frac{1}{\Delta y}\,Id^M\otimes A_1^N, \qquad
616: \bAxy=-\bAx\bAy\,.
617: \end{equation}
618: 
619: For the definition of $\phi$ in terms of $w$
620: (equation~\pref{eq:main_phi2}) we choose
621: \begin{equation}
622:   \label{eq:bracket_phi}
623:   [w,w]_2^{} = (\bAxx w)\odot(\bAyy w) - (\bAxy w)\odot(\bAxy w)\, ,
624: \end{equation}
625: and the corresponding choice for equation~\pref{eq:main_w2} is 
626: \begin{equation}
627:   \label{eq:bracket_w}
628:   \textstyle
629:   [w,\phi]_1^{}=\frac{1}{2}\bAyy\left\{(\bAxx w)\odot\phi\right\} +
630:   \frac{1}{2}\bAxx\left\{(\bAyy w)\odot\phi\right\} -
631:   \bAxyT\left\{(\bAxy w)\odot\phi\right\}\,.
632: \end{equation}
633: These two definitions are related in the sense given
634: in~\pref{eq:w_h_phi}: $[w,h]_2^T\phi= h^T[w,\phi]_1^{}$ for all $h$.
635: 
636: With these definitions the partial derivatives of discretized $E$
637: and $S$ with respect to the components of $w$ are given by
638: \begin{equation}
639:   \label{eq:fr_der}
640:   E'(w)=  \bAbih w +\bAxx\phi -2[w,\phi]_1^{}\,,\qquad
641:   S'(w)=  \bAxx w\,.
642: \end{equation}
643: %(Depending on the point of view, one might view these as the vectors of partial derivatives with respect to the components of $w$, or as the gradient associated with the standard inner product on $\R^{MN}$.)
644: 
645: 
646: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
647: \subsection{Solving the discretized biharmonic equation}
648: Matrix $\bAbih$ is symmetric and has a one-dimensional nullspace:
649: $\bAbih\bOne = \bNull$, where $\bOne$ and $\bNull$ are vectors with
650: $MN$-components which are all one and all zero, respectively. The same is
651: true for $\bAxx$ and $\bAyy$. For a given vector $f$ we would like to
652: solve
653: \begin{equation}
654:   \label{eq:biharm_disc}
655:   \bAbih\psi = f\,,\qquad \bOne^T\psi = 0\,.  
656: \end{equation}
657: A unique solution exists if and only if $f$ has zero average, i.e.,
658: $\bOne^T f=0$. So we must verify that the discretized versions of the
659: right-hand sides in \pref{eq:main_phi2}, \pref{eq:phi1} satisfy this
660: condition. Let $w\in\RR^{MN}$, then
661: \begin{align}
662:   \bOne^T \bAxx w &= 0\,, \notag \\
663:   \bOne^T [w,w]_2^{} &= (\bAxx w)^T (\bAyy w) - (\bAxy w)^T (\bAxy w) 
664:   = w^T \bAxx \bAyy w - w^T \bAxyT \bAxy w = 0\,, \label{eq:int_by_parts}
665: \end{align}
666: where the last equality holds because $\bAxT\bAx=\bAxx$ and
667: $\bAyT\bAy=\bAyy$, and because the $x$- and $y$-matrices commute. We
668: have, in fact, shown that the integration by parts formula from the
669: continuous case holds for our choice of spatial discretization. This is 
670: not true for an arbitrary discretization but is key for a successful scheme.
671: 
672: The inverse of matrix $\bAbih$ on the subspace of vectors with zero
673: average, denoted with a slight abuse of notation by $\bAbihinv$, can
674: be found, for example, using the fast cosine transform described below
675: in Sec.~\ref{sec:Fourier}.
676: 
677: 
678: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
679: \subsection{Computing the gradient}
680: The variational methods of this paper are based on a steepest descent
681: flow and modifications of this algorithm. The direction of the
682: steepest descent of $E$ at a point $w\in X$ is opposite to the
683: gradient of $E$ at $w$. The gradient is the Riesz representative of
684: the Fr\'echet derivative and hence we need to find a vector $u\in X$,
685: such that $E'(w)\cdot v=\langle u,v\rangle$ for all $v\in X$. The
686: inner product is either $\langle\cdot,\cdot\rangle_X$ or
687: $\langle\cdot,\cdot\rangle_{X,\l}$ and hence the gradient depends on
688: the choice of the inner product. We use the notation $u=\nabla E(w)$
689: for the gradient in $(X,\langle\cdot,\cdot\rangle_X)$ and $u=\nabla_\l
690: E(w)$ for the gradient in $(X,\langle\cdot,\cdot\rangle_{X,\l})$. To
691: find the discretized version of the gradient, we first need to
692: discretize the inner product.
693: 
694: Let $u,v\in\R^{MN}$, $\bOne^T u=\bOne^T v =0$. The inner product is
695: evaluated in the following way:
696: \begin{align}
697:   \langle u,v\rangle_{X,\l} &=4 \left(u^T \bAbih v + {\phi_1^u}^T \bAbih
698:     \phi_1^v -\l u^T \bAxx v\right) \Delta x\,\Delta y \notag \\
699:   &=4\Bigl( u^T\left(\bAbih +
700:       \bAxx\bAbihinv\bAxx -\l\bAxx\right)v\Bigr)\Delta x\,\Delta y \,, \notag
701: \end{align}
702: where $\phi_1^u, \phi_1^v$ are solutions of the discretized version
703: of~\pref{eq:phi1} with $w$ replaced by $u$ and $v$, respectively, and
704: we assume that we work on $\Omq$. For $w\in\R^{MN}$, $\bOne^T w=0$ the
705: Riesz representative of $E'(w)$ given in \pref{eq:fr_der} is computed
706: as
707: \begin{equation}\label{eq:grad}
708:   \nabla_\l E(w)=\left(\bAbih + \bAxx\bAbihinv\bAxx -
709:     \l\bAxx \right)^{-1} E'(w)\,.
710: \end{equation}
711: As in the case of $\bAbihinv$ we abused notation here since the
712: inverse only makes sense on a subspace of vectors with zero average.
713: It can be easily verified that $\bOne^T E'(w)=0$. The numerical
714: evaluation of $\nabla_\l S$ and of the $\langle
715: \cdot,\cdot\rangle_X$-gradients is similar.
716: 
717: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
718: \subsection{Fourier coordinates}\label{sec:Fourier}
719: In Fourier coordinates most of the finite difference operators become
720: diagonal matrices. This increases the efficiency of the numerical
721: algorithm and makes it possible to easily find the inverse of matrices
722: like $\bAbih$. See for example \cite{Canuto}.
723: 
724: On a uniform grid, it is a standard procedure to apply some form of
725: the fast Fourier transform to diagonalize finite difference matrices
726: like $A_2^M$ (see, e.g., \cite{Strang}). Due to the Neumann boundary
727: conditions \pref{def:BCsym} we need to employ the fast cosine
728: transform. We define $M\times M$ matrices
729: \[\textstyle
730: C_{\rm f}^M=\frac{1}{\sqrt{2M}}\left(2\cos\frac{(i-1)(2j-1)\pi}{2M}\right)_{i,j=1}^M , \qquad
731: C_{\rm b}^M=\frac{1}{\sqrt{2M}}\left[\left.
732: \begin{array}{c} 1 \\ \vdots \\1 \end{array}\right|
733: \left(2\cos\frac{(2i-1)(j-1)\pi}{2M}\right)_{i=1,j=2}^{M,M} \right],
734: \]
735: which have the following properties:
736: \[\textstyle
737: C_{\rm f}^M C_{\rm b}^M = Id^M, \qquad
738: C_{\rm f}^M A_2^M C_{\rm b}^M = \Lambda^M,
739: \]
740: where $\Lambda^M=\diag(2-2\cos\frac{(m-1)\pi}{M})_{m=1}^M$. Hence they
741: are inverses of each other and diagonalize $A_2^M$.
742: 
743: We further define matrices
744: \[
745: \bCf = C_{\rm f}^M\otimes C_{\rm f}^N, \qquad
746: \bCb = C_{\rm b}^M\otimes C_{\rm b}^N,
747: \]
748: which diagonalize $\bAxx$, $\bAyy$, and $\bAbih$:
749: \begin{equation}\label{eq:diag1}
750: \bCf\bAxx\bCb = \bLambdaxx, \qquad
751: \bCf\bAyy\bCb = \bLambdayy, \qquad
752: \bCf\bAbih\bCb = \bLambdabih,
753: \end{equation}
754: where the diagonal matrices are given by
755: \begin{equation}\label{eq:diag2}\textstyle
756: \bLambdaxx=\frac{1}{\Delta x^2}\Lambda^M\otimes Id^N, \qquad
757: \bLambdayy=\frac{1}{\Delta y^2}Id^M\otimes \Lambda^N, \qquad
758: \bLambdabih=(\bLambdaxx + \bLambdayy)^2.
759: \end{equation}
760: For a vector $w\in\R^{MN}$ we introduce its Fourier coordinates $\hat
761: w$ by
762: \[
763: \hat w = \bCf w \qquad\qquad w = \bCb \hat w \ .
764: \]
765: We note that $\bOne^T w=0$ if and only if the first component of $\hat
766: w$ is zero.
767: 
768: Most of the computations involved in the variational methods described
769: in~Sec.~\ref{sec:var_num} can be done in the Fourier coordinates. The
770: only time one needs to go back to the original coordinates is when
771: evaluating the brackets \pref{eq:bracket_phi} and \pref{eq:bracket_w},
772: because they are nonlinear and involve the discretized mixed
773: derivative operator $\bAxy$.
774: 
775: 
776: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
777: \subsection{Alternative discretization of $-\partial_{xy}$}\label{sec:d_xy}
778: The fast Fourier transform provides us with another discretization of
779: the mixed derivative which is not biased to the left or right. In an
780: analogy to~\pref{eq:diag2} and \pref{eq:diag1} we define
781: \[\textstyle
782: \bLambdaxy=\frac{1}{\Delta x\,\Delta y}\sqrt{\Lambda^M}\otimes
783: \sqrt{\Lambda^N}, \qquad \bAxyt=\bS \bLambdaxy \bCf\ ,
784: \]
785: where $\bS$ is the fast sine transform matrix
786: \[\textstyle
787: \bS = S^M\otimes S^N\,,\quad
788: S^M=\frac{1}{\sqrt{2M}}\left[\left.
789: \begin{array}{c} 0 \\ \vdots \\0 \end{array}\right|
790: \left(2\sin\frac{(2i-1)(j-1)\pi}{2M}\right)_{i=1,j=2}^{M,M} \right]\,.
791: \]
792: Property \pref{eq:int_by_parts} also holds with $\bAxy$ replaced by
793: $\bAxyt$.
794: 
795: 
796: \begin{remark}\label{rem:full_domain}
797:   When discretizing the problem on the full domain $\Omega$ with
798:   boundary conditions~\pref{def:BC}, we need to use different matrices
799:   in the $x$ and $y$-directions. In the $x$-direction we use the
800:   matrices described above, in the $y$-direction to discretize the
801:   second derivatives, for example, we use
802:   \[
803:   A_2=\left[
804:     \begin{array}{rrrrr}
805:       2 & -1 & & & -1 \\
806:       -1 & 2 & -1 \\
807:       & \ddots & \ddots & \ddots \\
808:       & & -1 & 2 & -1 \\
809:       -1 & & & -1 & 2
810:     \end{array}
811:   \right].
812:   \]
813:   In this direction the fast Fourier transform is used instead of the
814:   fast cosine/sine transform.
815: \end{remark}
816: 
817: 
818: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
819: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
820: \section{Newton's method}\label{sec:Newton}
821: We use Newton's method in two different ways. The first is to improve the numerical approximations
822: obtained by the variational numerical methods. Since these
823: are sometimes slow to converge, it is often faster to stop such an
824: algorithm early and use its result as an initial guess for  Newton's
825: method. The second use for Newton's method is as part of a numerical
826: continuation algorithm (see Sec. \ref{sec:continuation}).
827: 
828: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
829: \subsection{Newton's method for given load parameter $\l$}
830: This method can be used to improve solutions obtained by the MPA. Let
831: $\l\in(0,2)$ be given. We are solving
832: \begin{equation}
833:   \label{eq:main_newton_p}
834:   \G(w,\phi)=\left[
835:   \begin{array}{l}
836:     \G_1 \\ \G_2\rule{0pt}{2em}
837:   \end{array}
838:   \right]=
839:   \left[
840:     \begin{array}{l}
841:       \bAbih w- \l\bAxx w+ \bAxx\phi -2[w,\phi]_1 \\
842:       -\bAbih\phi + \bAxx w - [w,w]_2\rule{0pt}{2em}
843:     \end{array}
844:   \right]=
845:   \left[
846:     \begin{array}{l}
847:       \bNull \\ \bNull\rule{0pt}{2em}
848:     \end{array}
849:   \right]
850: \end{equation}
851: for $w$ and $\phi$ with zero average using Newton's method. The
852: matrix we need to invert is
853: \begin{equation}
854:   \label{eq:newton_matrix}
855:   \G'(w,\phi)=\left[
856:     \begin{array}{ll}
857:       \frac{\partial\G_1}{\partial w} &
858:       \frac{\partial\G_1}{\partial \phi} \\
859:       \frac{\partial\G_2}{\partial w} &
860:       \frac{\partial\G_2}{\partial \phi}\rule{0pt}{2em}
861:     \end{array}
862:   \right]=
863:   \left[
864:     \begin{array}{ll}
865:       \bAbih-\l\bAxx -2\bB_1 \quad & \bAxx -2\bB_2 \\
866:       \bAxx -2\bB_2^T & -\bAbih \rule{0pt}{2em}
867:     \end{array}
868:   \right]\,,
869: \end{equation}
870: where
871: \begin{align}
872:   \bB_1 = \frac{\partial}{\partial w}[w,\phi]_1 &=
873:   \frac{1}{2}\bAxx(\diag\phi)\bAyy + \frac{1}{2}\bAyy(\diag\phi)\bAxx -
874:   \bAxyT(\diag\phi)\bAxy\,, \notag \\
875:   \bB_2 = \frac{\partial}{\partial\phi}[w,\phi]_1 &=
876:   \frac{1}{2}\left(\frac{\partial}{\partial w}[w,w]_2\right)^T \notag \\
877:   &= \frac{1}{2}\bAxx(\diag\bAyy w) + \frac{1}{2}\bAyy(\diag\bAxx w) -
878:   \bAxyT(\diag\bAxy w)\,. \notag
879: \end{align}
880: From the properties of $\bAxx, \bAyy, \bAxy, \bAbih$ it follows that
881: matrix $\G'(w,\phi)$ is symmetric and singular, its nullspace is
882: spanned by $ \left[
883:   \begin{array}{c}
884:     \bOne\\ \bNull
885:   \end{array}
886: \right]
887: $, $
888: \left[
889:   \begin{array}{c}
890:     \bNull \\ \bOne
891:   \end{array}
892: \right]
893: $.
894: 
895: To describe how to find the inverse of $\G'(w,\phi)$ on a subspace
896: orthogonal to its nullspace, we introduce a new notation for the four
897: blocks of $\G'(w,\phi)$ from~\pref{eq:newton_matrix}:
898: \[
899: \G'(w,\phi)=\left[
900:   \begin{array}{cc}
901:     \bG_{11}^{} & \bG_{12}^{} \\ \bG_{12}^T & \bG_{22}^{}
902:   \end{array}
903: \right]
904: \]
905: For given vectors $u$, $\eta$ with zero average we need to find
906: vectors $v$, $\zeta$ with zero average such that
907: \begin{equation}\label{eq:newt_lin_syst}
908: \left[
909:   \begin{array}{cc}
910:     \bG_{11}^{} & \bG_{12}^{} \\ \bG_{12}^T & \bG_{22}^{}
911:   \end{array}
912: \right]
913: \left[
914:   \begin{array}{c}
915:     v \\ \zeta
916:   \end{array}
917: \right]=
918: \left[
919:   \begin{array}{c}
920:     u \\ \eta
921:   \end{array}
922: \right]\,.
923: \end{equation}
924: Let the tilde denote the block of the first $MN-1$ rows and columns of a
925: matrix and $MN-1$ components of a vector. The matrix $\left[
926:   \begin{array}{cc}
927:     \bGt_{11}^{} & \bGt_{12}^{} \\ \bGt_{12}^T & \bGt_{22}^{}
928:   \end{array}
929: \right]$ is symmetric, nonsingular, and sparse. It can be inverted by
930: a linear sparse solver. System~\pref{eq:newt_lin_syst} is then
931: solved in the following steps:
932: \[
933: \left[
934:   \begin{array}{c}
935:     r \\ \rho
936:   \end{array}
937: \right] := 
938: \left[
939:   \begin{array}{cc}
940:     \bGt_{11}^{} & \bGt_{12}^{} \\ \bGt_{12}^T & \bGt_{22}^{}
941:   \end{array}
942: \right]^{-1}
943: \left[
944:   \begin{array}{c}
945:     \tilde u \\ \tilde \eta
946:   \end{array}
947: \right]\,,\qquad
948: \begin{array}{ll}
949:   s :=-\frac{1}{MN}\tilde\bOne^T r\,, \qquad &
950:   v=\left[
951:     \begin{array}{c}
952:       r + s\tilde\bOne \\ s
953:     \end{array}
954:   \right]\,, \\
955:   \sigma := -\frac{1}{MN}\tilde\bOne^T \rho\,, &
956:   \eta=\left[
957:     \begin{array}{c}
958:       \rho + \sigma\tilde\bOne \\ \sigma
959:     \end{array}
960:   \right]\,.\rule{0pt}{2em}
961: \end{array}
962: \]
963: 
964: 
965: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
966: \subsection{Newton's method for given $S$}
967: This method can be used to improve solutions obtained by the CSDM and
968: the CMPA. Let $C>0$ be given. We are looking for numerical solutions
969: of~(\ref{eq:main_w2}--\ref{eq:main_phi2}) in the set~$\M$ defined
970: by~\pref{eq:constraint}. Hence we are solving
971: \begin{equation}
972:   \label{eq:main_newton_C}
973:   \left[
974:     \begin{array}{l}
975:       \bAbih w- \l\bAxx w+ \bAxx\phi -2[w,\phi]_1 \\
976:       -\bAbih\phi + \bAxx w - [w,w]_2\rule{0pt}{2em} \\
977:       -\frac{1}{2}w^T\bAxx w + C/(4\Delta x\Delta y)\rule{0pt}{2em}
978:     \end{array}
979:   \right]=
980:   \left[
981:     \begin{array}{c}
982:       \bNull \\ \bNull\rule{0pt}{2em} \\ 0\rule{0pt}{2em}
983:     \end{array}
984:   \right]
985: \end{equation}
986: for $w$ and $\phi$ with zero average and $\l$ using Newton's method.
987: The approach is very similar to that described in the previous
988: section, the resulting matrix is symmetric, has just one more row and
989: column than the matrix in~\pref{eq:newton_matrix}.
990: 
991: 
992: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
993: \subsection{Continuation}\label{sec:continuation}
994: To follow branches of solutions $(\l,w)$ of~\pref{eq:main_newton_p} we
995: adopt a continuation method described in~\cite{Keller}. We introduce a
996: parameter $s\in\R$ by adding a constraint---pseudo-arclength
997: normalization (in the $(\l,\Mod{w}_X)$-plane). For a given value of
998: $s$ we are solving
999: \begin{equation}
1000:   \label{eq:main_cont}
1001:   \bar\G(w,\phi,\l)=\left[
1002:   \begin{array}{l}
1003:     \G_1 \\ \G_2\rule{0pt}{2em} \\ \G_3\rule{0pt}{2em}
1004:   \end{array}
1005:   \right]=
1006:   \left[
1007:     \begin{array}{l}
1008:       \bAbih w- \l\bAxx w+ \bAxx\phi -2[w,\phi]_1 \\
1009:       -\bAbih\phi + \bAxx w - [w,w]_2\rule{0pt}{2em} \\
1010:       \theta \langle \dot w_0,w-w_0\rangle_X + (1-\theta)
1011:       \dot\l_0 (\l-\l_0) - (s-s_0)\rule{0pt}{2em}
1012:     \end{array}
1013:   \right]=
1014:   \left[
1015:     \begin{array}{c}
1016:       \bNull \\ \bNull\rule{0pt}{2em} \\ 0\rule{0pt}{2em}
1017:     \end{array}
1018:   \right]
1019: \end{equation}
1020: for $w$, $\phi$ with zero average, and the load $\l$, where the value
1021: of $\theta\in(0,1)$ is fixed (e.g., $\theta=\frac{1}{2}$). We assume
1022: that we are given a value $s_0$, an initial point $(\l_0,w_0)$ on the
1023: branch, and an approximate direction $(\dot\l_0,\dot w_0)$ of the
1024: branch at this point (an approximation of the derivative
1025: $\frac{d}{ds}(\l(s),w(s))|_{s=s_0}$).
1026: 
1027: System \pref{eq:main_cont} is solved for a discrete set of values of
1028: $s$ in some interval $(s_0,s_1)$ by Newton's method. Then, a new
1029: initial point on the branch is defined by setting $w_0=w(s_1)$,
1030: $\l_0=\l(s_1)$, $s_0=s_1$, a new direction $(\dot\l_0,\dot w_0)$ at
1031: this point is computed and the process is repeated. The matrix we need
1032: to invert in Newton's method is
1033: \begin{equation}
1034:   \label{eq:cont_matrix}
1035:   \bar \G'(w,\phi,\l)=
1036:   \left[
1037:     \begin{array}{cc}
1038:       \G'(w,\phi,\l) & g \\
1039:       h^T & d %\rule{0pt}{2em}
1040:     \end{array}
1041:   \right],\quad
1042:   g=\left[
1043:     \begin{array}{c}
1044:       -\bAxx w \\ \bNull
1045:     \end{array}
1046:   \right],\ 
1047:   h=\left[
1048:     \begin{array}{c}
1049:       4\theta \bAprod \dot w_0\,\Delta x\Delta y \\ \bNull
1050:     \end{array}
1051:   \right],\ 
1052:   d= (1-\theta)\dot\l_0,
1053: \end{equation}
1054: where $\bAprod=\bAbih+\bAxx\bAbihinv\bAxx$.
1055: 
1056: Solving a linear system with this matrix amounts to solving system
1057: \pref{eq:newt_lin_syst} for two right-hand sides. For a given
1058: $u\in\R^{2MN}$ with $[\bOne^T\ \bOne^T]u=0$ and a given $\eta\in\R$ we
1059: want to solve
1060: \begin{equation}\label{eq:cont_lin_syst}
1061: \left[
1062:   \begin{array}{cc}
1063:     \G'(w,\phi,\l) & g \\ h^T & d
1064:   \end{array}
1065: \right]
1066: \left[
1067:   \begin{array}{c}
1068:     v \\ \zeta
1069:   \end{array}
1070: \right]=
1071: \left[
1072:   \begin{array}{c}
1073:     u \\ \eta
1074:   \end{array}
1075: \right]\,.
1076: \end{equation}
1077: for $v$ with $[\bOne^T\ \bOne^T]v=0$ and $\zeta$. System
1078: \pref{eq:cont_lin_syst} is solved in the following steps:
1079: \[
1080: \begin{array}{ll}
1081:   {\rm solve}: & \G'(w,\phi,\l)v_1=g\,, \\
1082:   & \G'(w,\phi,\l)v_2=u\,,
1083: \end{array}\qquad
1084: \zeta = \frac{\eta-h^Tv_2}{d-h^Tv_1}\,,\qquad
1085: v=v_2-\zeta v_1\,.
1086: \]
1087: 
1088: \begin{remark}
1089: Note that in this implementation we simply follow a solution of the
1090: equation, there is no guarantee that this remains a local minimum, a
1091: MP-solution or constrained MP-solution (cf.~Fig.~\ref{fig:cont_100}).
1092: \end{remark}
1093: 
1094: \begin{remark}
1095:   Newton's method and continuation have been implemented only using
1096:   a one-sided discretization of the mixed derivative $\partial_{xy}$
1097:   and only on the domain $\Omq$ assuming symmetry~\pref{eq:sym}.
1098:   The alternative discretization of $\partial_{xy}$ described in
1099:   Sec.~\ref{sec:d_xy} uses the fast cosine/sine transform. The
1100:   resulting matrix $\bAxyt$ is not sparse and therefore we would
1101:   obtain a dense block $\bG_{12}^{}$ in system~\pref{eq:newt_lin_syst}
1102:   which would prevent us from using a sparse solver.
1103:   
1104:   On the full domain $\Omega$ we assume periodicity of $w$ and $\phi$
1105:   in the $y$-direction. Hence for a discretization with a small step
1106:   $\Delta y$ the matrix we invert when solving \pref{eq:newt_lin_syst}
1107:   would become close to singular. The shift in the $y$ direction is
1108:   prevented by assuming the symmetry $w(x,y)=w(x,-y)$,
1109:   $\phi(x,y)=\phi(x,-y)$.
1110: \end{remark}
1111:   
1112: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1113: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1114: \section{Numerical solutions}
1115: We fix the size of the domain and the size of the space step for the
1116: following numerical computations: $a=b=100$, $\Delta x=\Delta y=0.5$.
1117: We obtain solutions using the variational techniques SDM, MPA, CSDM,
1118: and CMPA.
1119:   Table~\ref{tab:methods} provides a summary of which discretization
1120:   was used in which algorithm.
1121: 
1122: \begin{table}[htbp]
1123:   \centering
1124:   \renewcommand{\arraystretch}{1.2}
1125:   \begin{tabular}{l|p{15mm}|p{15mm}|p{15mm}|p{15mm}|}
1126:     \cline{2-5}
1127:     & \multicolumn{2}{|c|}{mixed derivative $\partial_{xy}$} &
1128:     \multicolumn{2}{|c|}{computational domain} \tabularnewline
1129:     \cline{2-5}
1130:     & \centering one-sided & \centering Fourier & \centering full &
1131:     \centering $1/4$ \tabularnewline
1132:     \hline
1133:     \multicolumn{1}{|l|}{variational methods} &
1134:     \centering $\surd$ & \centering $\surd$ & \centering $\surd$ &
1135:     \centering $\surd$ \tabularnewline
1136:     \hline
1137:     \multicolumn{1}{|l|}{Newton/continuation} &
1138:     \centering $\surd$ & & & \centering $\surd$ \tabularnewline
1139:     \hline
1140:   \end{tabular}
1141:   \vspace{3mm}
1142:   \caption{Summary of which spatial discretization was used in the
1143:     different numerical methods.}
1144:   \label{tab:methods}
1145: \end{table}
1146: 
1147: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1148: \subsection{A mountain-pass solution on the full domain $\Omega$}\label{sec:num_mp}
1149: %start with MPA on the full domain using Fourier, observe symmetry
1150: %then MPA on a quarter with one-sided, CSDM, continuation, verify MP
1151: %for some points on the curve
1152: 
1153: The first numerical experiments are done on the full domain $\Omega$, without symmetry restrictions, and with the unbiased (Fourier) discretization of $\partial_{xy}$ (Sec.~\ref{sec:d_xy}). For a fixed load $\l=1.4$ we computed a mountain-pass solution using 
1154: the MPA (Sec.~\ref{sec:MPA}). As end points were taken $w_1=0$ and a second point $w_2$ obtained by the SDM (here the initial point for the SDM was chosen to have a single peak centered
1155: at $x=y=0)$. The graph of the
1156: solution  $\wMP$ is shown in Fig.~\ref{fig:num_wMP}
1157: (left). The figure on the right shows $\wMP$ rendered on a cylinder 
1158: and we see it has the form of a single dimple. The value of shortening 
1159: for this solution is $S(\wMP)=14.93529$.
1160: 
1161: Alternatively, if we apply the CSDM with $S=14.93529$ and use a function
1162: with a single peak in the center of the domain as the initial
1163: condition $w_0$ we also obtain the same solution $\wMP$, this time as 
1164: a local minimizer of $E$ under constrained $S$.
1165: 
1166: We remark that although the MPA and the CSDM have a local character we have 
1167: not found any numerical mountain-pass solution with the total
1168: potential $F_\l$ smaller than $F_\l(\wMP)$ for $\l=1.4$. Similarly,
1169: using the CSDM we have not found any solution with energy $E$ smaller than
1170: $E(\wMP)$ under the constraint $S=14.93529$.
1171: We briefly discuss the physical relevance of this solution
1172: in~Sec.~\ref{sec:conc}.
1173: 
1174: \begin{figure}[htbp]
1175: \begin{center}
1176: \color[rgb]{.5,.5,.5}
1177: \fbox{\includegraphics[height=35mm]{numerfigs/graph_01fulldom}
1178:   \includegraphics[height=35mm]{numerfigs/cyl_01fulldom}}
1179: \color{black}
1180: \end{center}
1181: \caption{Mountain-pass solution for $\l=1.4$ found using the MPA on the full
1182:   domain $\Omega$ with $\partial_{xy}$ discretized using the fast
1183:   Fourier transform.}
1184: \label{fig:num_wMP}
1185: \end{figure}
1186: 
1187: 
1188: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1189: \subsection{Solutions under symmetry restrictions}
1190: The solution $\wMP$ of Fig.~\ref{fig:num_wMP} satisfies the symmetry
1191: property~\pref{eq:sym}. In the computations described below we
1192: enforced this symmetry and worked on the quarter domain $\Omq$, thus
1193: reducing the complexity of the problem. In order to improve the
1194: variational methods by combining them with Newton's method we also
1195: discretized the mixed derivative $\partial_{xy}$ using left-sided
1196: finite differences. The influence of this choice on the numerical
1197: solution is described in~Sec.~\ref{sec:bias}.
1198: 
1199: % % % % % % % % % % % % % % % % % % % % % % % % % % % % % % % % % % 
1200: \subsubsection{Constrained steepest descent method}
1201: We first fixed $S=40$ and used the CSDM to obtain constrained local
1202: minimizers of $E$ described in~Table~\ref{tab:sol_csdm}. They are
1203: ordered according to the increasing value of stored energy $E$. Their
1204: graphs and renderings on a cylinder are shown
1205: in~Fig.~\ref{fig:sol_csdm}. 
1206: %When we compare MPA solutions 1.1--1.7
1207: %to CSDM solutions 2.1--2.7, respectively, we find that although the
1208: %corresponding values of $\l$, $S$, $E$, and $F_\l$ differ, the main
1209: %features of their shape are the same. 
1210: Solution 1.1 is similar to the single
1211: dimple solution $\wMP$ described above 
1212: and according to Table~\ref{tab:sol_csdm} it has, indeed,  the
1213: smallest value of $E$.
1214: 
1215: \begin{table}[htbp]
1216:   \centering
1217:   \begin{tabular}{|c||c|c|c|c|c|}
1218:     \hline
1219:     {\bf CSDM} & $\l$ & $S$ & $E$ & $F_\l$ & same shape as {\bf MPA}\tabularnewline
1220:     \hline
1221:     1.1 & 1.108121 & 40 & 56.85636 & 12.53151 & 2.1 \tabularnewline %cat#01
1222:     \hline
1223:     1.2 & 1.299143 & 40 & 62.76150 & 10.79577 & 2.2 \tabularnewline %cat#11
1224:     \hline
1225:     1.3 & 1.316146 & 40 & 63.21646 & 10.57063 & 2.3 \tabularnewline %cat#17
1226:     \hline
1227:     1.4 & 1.311687 & 40 & 63.64083 & 11.17334 & 2.4 \tabularnewline %cat#21
1228:     \hline
1229:     1.5 & 1.309586 & 40 & 63.70623 & 11.32278 & 2.5 \tabularnewline %cat#05
1230:     \hline
1231:     1.6 & 1.328997 & 40 & 64.00875 & 10.84889 & 2.6 \tabularnewline %cat#02
1232:     \hline
1233:     1.7 & 1.344898 & 40 & 64.52244 & 10.72651 & 2.7 \tabularnewline %cat#03
1234:     \hline
1235:   \end{tabular}
1236:   \vspace{3mm}
1237:   \caption{Numerical solutions obtained by the CSDM on $\Omq$ with $\partial_{xy}$ discretized using left-sided finite differences. Graphs are shown in~Fig.~\ref{fig:sol_csdm}.}
1238:   \label{tab:sol_csdm}
1239: \end{table}
1240: 
1241: \begin{landscape}
1242:   \begin{figure}[htbp]
1243:     \centering
1244:     \setlength{\unitlength}{1mm}
1245:     \begin{picture}(181,124)
1246:       \color[rgb]{.5,.5,.5}
1247:       \put(0,64){\frparbcenter{\includegraphics[width=4.233cm]{numerfigs/graph_01}\\{\includegraphics[width=4.064cm]{numerfigs/cyl_01}}}}
1248:       \put(46,64){\frparbcenter{\includegraphics[width=4.233cm]{numerfigs/graph_11}\\{\includegraphics[width=4.064cm]{numerfigs/cyl_11}}}}
1249:       \put(92,64){\frparbcenter{\includegraphics[width=4.233cm]{numerfigs/graph_17}\\{\includegraphics[width=4.064cm]{numerfigs/cyl_17}}}}
1250:       \put(138,64){\frparbcenter{\includegraphics[width=4.233cm]{numerfigs/graph_21}\\{\includegraphics[width=4.064cm]{numerfigs/cyl_21}}}}
1251:       \put(0,0){\frparbcenter{\includegraphics[width=4.233cm]{numerfigs/graph_05}\\{\includegraphics[width=4.064cm]{numerfigs/cyl_05}}}}
1252:       \put(46,0){\frparbcenter{\includegraphics[width=4.233cm]{numerfigs/graph_02}\\{\includegraphics[width=4.064cm]{numerfigs/cyl_02}}}}
1253:       \put(92,0){\frparbcenter{\includegraphics[width=4.233cm]{numerfigs/graph_03}\\{\includegraphics[width=4.064cm]{numerfigs/cyl_03}}}}
1254:       \color{black}
1255:       \put(31,65){\makebox(11,5)[rb]{(1.1)}}
1256:       \put(77,65){\makebox(11,5)[rb]{(1.2)}}
1257:       \put(123,65){\makebox(11,5)[rb]{(1.3)}}
1258:       \put(169,65){\makebox(11,5)[rb]{(1.4)}}
1259:       \put(31,1){\makebox(11,5)[rb]{(1.5)}}
1260:       \put(77,1){\makebox(11,5)[rb]{(1.6)}}
1261:       \put(123,1){\makebox(11,5)[rb]{(1.7)}}
1262:     \end{picture}
1263:   \caption{Numerical solutions found using the CSDM with axial end
1264:     shortening $S=40$. More details are in Table~\ref{tab:sol_csdm}.}
1265:   \label{fig:sol_csdm}
1266: \end{figure}
1267: \end{landscape}
1268: 
1269: % % % % % % % % % % % % % % % % % % % % % % % % % % % % % % % % % % 
1270: \subsubsection{Mountain-pass algorithm}
1271: We then used the MPA for fixed $\l=1.4$ and various choices of $w_2$ to
1272: obtain the local mountain-pass points of $F_\l$ described
1273: in~Table~\ref{tab:sol_mpa}. They are ordered according to the
1274: increasing value of the total potential $F_\l$. The shape of their
1275: graph is very similar to that of the CSDM solutions discussed above
1276: and depicted in Fig.~\ref{fig:sol_csdm} and we do not show
1277: their graphs here. Solution 2.1 is again the single dimple solution and the
1278: table shows that it has the smallest value of $F_\l$.  
1279: 
1280: \begin{table}[htbp]
1281:   \centering
1282:   \begin{tabular}{|c||c|c|c|c|c|}
1283:     \hline
1284:     {\bf MPA} & $\l$ & $S$ & $E$ & $F_\l$ & same shape as {\bf CSDM}\tabularnewline
1285:     \hline
1286:     2.1 & 1.4 & 17.73822 & 29.42997 & 4.596460 & 1.1 \tabularnewline %cat#01
1287:     \hline
1288:     2.2 & 1.4 & 29.85121 & 49.08882 & 7.297132 & 1.2 \tabularnewline %cat#11
1289:     \hline
1290:     2.3 & 1.4 & 31.28849 & 51.39952 & 7.595635 & 1.3 \tabularnewline %cat#17
1291:     \hline
1292:     2.4 & 1.4 & 31.41723 & 52.01893 & 8.034809 & 1.4 \tabularnewline %cat#21
1293:     \hline
1294:     2.5 & 1.4 & 31.22992 & 51.84074 & 8.118852 & 1.5 \tabularnewline %cat#05
1295:     \hline
1296:     2.6 & 1.4 & 32.77491 & 54.15818 & 8.273314 & 1.6 \tabularnewline %cat#02
1297:     \hline
1298:     2.7 & 1.4 & 34.19888 & 56.56472 & 8.686284 & 1.7 \tabularnewline %cat#03
1299:     \hline
1300:   \end{tabular}
1301:   \vspace{3mm}
1302:   \caption{Numerical solutions obtained by the MPA on $\Omq$ with $\partial_{xy}$ discretized using left-sided finite differences.}
1303:   \label{tab:sol_mpa}
1304: \end{table}
1305: 
1306: 
1307: % % % % % % % % % % % % % % % % % % % % % % % % % % % % % % % % % % 
1308: \subsubsection{Constrained mountain-pass algorithm}
1309: We then fixed $S=40$  and applied the CMPA to obtain constrained local
1310: mountain passes of $E$ described in~Table~\ref{tab:sol_cmpa}. They are
1311: again ordered according to the increasing value of stored energy $E$.
1312: Their graphs and renderings on a cylinder are shown
1313: in~Figs.~\ref{fig:sol_cmpa1} and~\ref{fig:sol_cmpa2}. As endpoints
1314: $w_1$, $w_2$ of the path in the CMPA we used the constrained local
1315: minimizers 1.1--1.7.
1316: 
1317: There are 21 possible pairs $(w_1,w_2)$ to be used but only 19
1318: solutions in Table~\ref{tab:sol_cmpa}. The algorithm did not converge
1319: for the following three pairs: (1.1, 1.3), (1.5, 1.6), and (1.4, 1.6),
1320: most likely due to the complicated nature of the energy landscape
1321: between these endpoints. On the other hand, two choices of pairs
1322: denoted by $\ast$ and $\dagger$ in the Table yielded two solutions
1323: each. When the path is deformed it sometimes comes close to another
1324: critical point of $E$ which is not a constrained mountain pass. In
1325: that case the algorithm slows down and one can apply Newton's method
1326: to such a point. It is a matter of luck whether Newton's method
1327: converges. The CMPA then runs further and might converge to another point,
1328: this time a constrained mountain-pass point. And finally, two choices
1329: of $(w_1,w_2)$ yielded the same solution 3.3.
1330: 
1331: \begin{table}[htbp]
1332:   \centering
1333:   \begin{tabular}{|c||c|c|c|c|c|}
1334:     \hline
1335:     {\bf CMPA} & $\l$ & $S$ & $E$ & $F_\l$ & end points \tabularnewline
1336:     \hline
1337:      3.1 & 1.310815 & 40 & 63.98996 & 11.55737 & 1.2, 1.4
1338:      \tabularnewline %cat#26
1339:     \hline
1340:      3.2 & 1.332112 & 40 & 64.38609 & 11.10161 & 1.3, 1.6
1341:      \tabularnewline %cat#04
1342:     \hline
1343:      3.3 & 1.447626 & 40 & 66.49032 & 8.585294 & 1.2, 1.3 or 1.3, 1.4
1344:      \tabularnewline %cat#20
1345:     \hline
1346:      3.4 & 1.440841 & 40 & 66.72079 & 9.087129 & 1.1, 1.4
1347:      \tabularnewline %cat#22
1348:     \hline
1349:      3.5 & 1.447594 & 40 & 66.97057 & 9.066810 & 1.1, 1.6
1350:      \tabularnewline %cat#06
1351:     \hline
1352:      3.6 & 1.484790 & 40 & 68.20637 & 8.814758 & 1.3, 1.5
1353:      \tabularnewline %cat#19
1354:     \hline
1355:      3.7 & 1.477769 & 40 & 68.23274 & 9.121955 & 1.1, 1.5$^\ast$
1356:      \tabularnewline %cat#09
1357:     \hline
1358:      3.8 & 1.482261 & 40 & 68.41086 & 9.120428 & 1.1, 1.7
1359:      \tabularnewline %cat#07
1360:     \hline
1361:      3.9 & 1.413917 & 40 & 68.56697 & 12.01028 & 1.1, 1.2
1362:      \tabularnewline %cat#12
1363:     \hline
1364:      3.10 & 1.520975 & 40 & 68.83818 & 7.999162 & 1.2, 1.5
1365:      \tabularnewline %cat#13
1366:     \hline
1367:      3.11 & 1.475705 & 40 & 69.00087 & 9.972652 & 1.1, 1.5$^\ast$
1368:      \tabularnewline %cat#08
1369:     \hline
1370:      3.12 & 1.532000 & 40 & 69.27379 & 7.993781 & 1.4, 1.5$^\dagger$
1371:      \tabularnewline %cat#24
1372:     \hline
1373:      3.13 & 1.527108 & 40 & 69.35834 & 8.274019 & 1.2, 1.6
1374:      \tabularnewline %cat#14
1375:     \hline
1376:      3.14 & 1.551762 & 40 & 69.47838 & 7.407904 & 1.4, 1.5$^\dagger$
1377:      \tabularnewline %cat#25
1378:     \hline
1379:      3.15 & 1.547955 & 40 & 69.68292 & 7.764712 & 1.6, 1.7
1380:      \tabularnewline %cat#10
1381:     \hline
1382:      3.16 & 1.539785 & 40 & 69.78487 & 8.193480 & 1.5, 1.7
1383:      \tabularnewline %cat#15
1384:     \hline
1385:      3.17 & 1.546480 & 40 & 69.85900 & 7.999795 & 1.3, 1.7
1386:      \tabularnewline %cat#18
1387:     \hline
1388:      3.18 & 1.549780 & 40 & 70.16253 & 8.171339 & 1.2, 1.7
1389:      \tabularnewline %cat#16
1390:     \hline
1391:      3.19 & 1.561117 & 40 & 70.74117 & 8.296474 & 1.4, 1.7
1392:      \tabularnewline %cat#23
1393:     \hline
1394:   \end{tabular}
1395:   \vspace{3mm}
1396:   \caption{Numerical solutions obtained by the CMPA/Newton on $\Omq$ with $\partial_{xy}$ discretized using left-sided finite differences. Graphs are shown in Figs.~\ref{fig:sol_cmpa1},~\ref{fig:sol_cmpa2}.}
1397:   \label{tab:sol_cmpa}
1398: \end{table}
1399: 
1400: \begin{landscape}
1401: \begin{figure}[!ht]
1402: \begin{center}
1403: \setlength{\unitlength}{1mm}
1404: \begin{picture}(227,124)
1405: \color[rgb]{.5,.5,.5}
1406: \put(0,64){\frparbcenter{\includegraphics[width=4.233cm]{numerfigs/graph_26}\\\includegraphics[width=4.064cm]{numerfigs/cyl_26}}}
1407: \put(46,64){\frparbcenter{\includegraphics[width=4.233cm]{numerfigs/graph_04}\\\includegraphics[width=4.064cm]{numerfigs/cyl_04}}}
1408: \put(92,64){\frparbcenter{\includegraphics[width=4.233cm]{numerfigs/graph_20}\\\includegraphics[width=4.064cm]{numerfigs/cyl_20}}}
1409: \put(138,64){\frparbcenter{\includegraphics[width=4.233cm]{numerfigs/graph_22}\\\includegraphics[width=4.064cm]{numerfigs/cyl_22}}}
1410: \put(184,64){\frparbcenter{\includegraphics[width=4.233cm]{numerfigs/graph_06}\\\includegraphics[width=4.064cm]{numerfigs/cyl_06}}}
1411: \put(0,0){\frparbcenter{\includegraphics[width=4.233cm]{numerfigs/graph_19}\\\includegraphics[width=4.064cm]{numerfigs/cyl_19}}}
1412: \put(46,0){\frparbcenter{\includegraphics[width=4.233cm]{numerfigs/graph_09}\\\includegraphics[width=4.064cm]{numerfigs/cyl_09}}}
1413: \put(92,0){\frparbcenter{\includegraphics[width=4.233cm]{numerfigs/graph_07}\\\includegraphics[width=4.064cm]{numerfigs/cyl_07}}}
1414: \put(138,0){\frparbcenter{\includegraphics[width=4.233cm]{numerfigs/graph_12}\\\includegraphics[width=4.064cm]{numerfigs/cyl_12}}}
1415: \put(184,0){\frparbcenter{\includegraphics[width=4.233cm]{numerfigs/graph_13}\\\includegraphics[width=4.064cm]{numerfigs/cyl_13}}}
1416: \color{black}
1417: \put(31,65){\makebox(11,5)[rb]{(3.1)}}
1418: \put(77,65){\makebox(11,5)[rb]{(3.2)}}
1419: \put(123,65){\makebox(11,5)[rb]{(3.3)}}
1420: \put(169,65){\makebox(11,5)[rb]{(3.4)}}
1421: \put(215,65){\makebox(11,5)[rb]{(3.5)}}
1422: \put(31,1){\makebox(11,5)[rb]{(3.6)}}
1423: \put(77,1){\makebox(11,5)[rb]{(3.7)}}
1424: \put(123,1){\makebox(11,5)[rb]{(3.8)}}
1425: \put(169,1){\makebox(11,5)[rb]{(3.9)}}
1426: \put(215,1){\makebox(11,5)[rb]{(3.10)}}
1427: \end{picture}
1428: \end{center}
1429: \caption{Numerical solutions found using the CMPA/Newton with $S=40$. More details are in Table~\ref{tab:sol_cmpa}.}
1430: \label{fig:sol_cmpa1}
1431: \end{figure}
1432: \end{landscape}
1433: 
1434: \begin{landscape}
1435: \begin{figure}[!ht]
1436: \begin{center}
1437: \setlength{\unitlength}{1mm}
1438: \begin{picture}(227,124)
1439: \color[rgb]{.5,.5,.5}
1440: \put(0,64){\frparbcenter{\includegraphics[width=4.233cm]{numerfigs/graph_08}\\\includegraphics[width=4.064cm]{numerfigs/cyl_08}}}
1441: \put(46,64){\frparbcenter{\includegraphics[width=4.233cm]{numerfigs/graph_24}\\\includegraphics[width=4.064cm]{numerfigs/cyl_24}}}
1442: \put(92,64){\frparbcenter{\includegraphics[width=4.233cm]{numerfigs/graph_14}\\\includegraphics[width=4.064cm]{numerfigs/cyl_14}}}
1443: \put(138,64){\frparbcenter{\includegraphics[width=4.233cm]{numerfigs/graph_25}\\\includegraphics[width=4.064cm]{numerfigs/cyl_25}}}
1444: \put(184,64){\frparbcenter{\includegraphics[width=4.233cm]{numerfigs/graph_10}\\\includegraphics[width=4.064cm]{numerfigs/cyl_10}}}
1445: \put(0,0){\frparbcenter{\includegraphics[width=4.233cm]{numerfigs/graph_15}\\\includegraphics[width=4.064cm]{numerfigs/cyl_15}}}
1446: \put(46,0){\frparbcenter{\includegraphics[width=4.233cm]{numerfigs/graph_18}\\\includegraphics[width=4.064cm]{numerfigs/cyl_18}}}
1447: \put(92,0){\frparbcenter{\includegraphics[width=4.233cm]{numerfigs/graph_16}\\\includegraphics[width=4.064cm]{numerfigs/cyl_16}}}
1448: \put(138,0){\frparbcenter{\includegraphics[width=4.233cm]{numerfigs/graph_23}\\\includegraphics[width=4.064cm]{numerfigs/cyl_23}}}
1449: \color{black}
1450: \put(31,65){\makebox(11,5)[rb]{(3.11)}}
1451: \put(77,65){\makebox(11,5)[rb]{(3.12)}}
1452: \put(123,65){\makebox(11,5)[rb]{(3.13)}}
1453: \put(169,65){\makebox(11,5)[rb]{(3.14)}}
1454: \put(215,65){\makebox(11,5)[rb]{(3.15)}}
1455: \put(31,1){\makebox(11,5)[rb]{(3.16)}}
1456: \put(77,1){\makebox(11,5)[rb]{(3.17)}}
1457: \put(123,1){\makebox(11,5)[rb]{(3.18)}}
1458: \put(169,1){\makebox(11,5)[rb]{(3.19)}}
1459: \end{picture}
1460: \end{center}
1461: \caption{Numerical solutions found using the CMPA/Newton with $S=40$. More details are in Table~\ref{tab:sol_cmpa}.}
1462: \label{fig:sol_cmpa2}
1463: \end{figure}
1464: \end{landscape}
1465: 
1466: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1467: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1468: \section{Remarks on the numerics}
1469: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1470: \subsection{Bias in the discretization of $\partial_{xy}$}
1471: \label{sec:bias}
1472: In this section we examine the influence of the
1473: discretization of the mixed derivative $\partial_{xy}$ on the numerical
1474: solution. We recall that the mixed derivative $\partial_{xy}$ can be
1475: discretized using left/right-sided finite differences
1476: \pref{eq:dx_RL}, \pref{eq:d_xy_RL} or using the fast
1477: Fourier transform (Section~\ref{sec:d_xy}). For 
1478: comparison we use the single-dimple solution on
1479: $\Omega=(-100,100)^2$ at load $\l=1.4$ obtained by the MPA. 
1480: 
1481: Let $\Delta x=\Delta y=0.5$. Table~\ref{tab:mpa_xy} gives a list of
1482: numerical experiments together with the values of shortening and
1483: energy. Figure~\ref{fig:profiles_xy} shows a profile of the numerical
1484: solutions in the circumferential direction at $x=0$.
1485: 
1486: \begin{table}[htbp]
1487:   \centering\setlength{\unitlength}{1mm}
1488:   \begin{tabular}{|c|c|c|c|c|c|c|}
1489:     \hline
1490:     domain & discretization of $\partial_{xy}$ & $\l$ & $S$ & $E$ &
1491:     $F_\l$ & Figure~\ref{fig:profiles_xy} \tabularnewline
1492:     \hline
1493:     $\Omega$ or $\Omq$ & Fourier & 1.4 & 14.93529 & 24.71825 & 3.808850 &
1494:     \begin{picture}(11.25,2)
1495:       \linethickness{.5pt}
1496:       \put(0,1){\line(1,0){11.25}}
1497:     \end{picture} \tabularnewline
1498:     \hline
1499:     $\Omega$ & left/right-sided & 1.4 & 14.93617 & 24.70828 & 3.797636 &
1500:     \begin{picture}(11.25,2)
1501:       \linethickness{.1pt}
1502:       \put(0,1){\line(1,0){11.25}}
1503:     \end{picture} \tabularnewline
1504:     \hline
1505:     $\Omq$ & left-sided & 1.4 & 17.73822 & 29.42997 & 4.596460 &
1506:     \begin{picture}(11.25,2)
1507:       \linethickness{.15mm}
1508:       \multiput(0,1)(1,0){12}{\line(1,0){.15}}
1509:     \end{picture} \tabularnewline
1510:     \hline
1511:     $\Omq$ & right-sided & 1.4 & 12.81205 & 21.16342 & 3.226549 &
1512:     \begin{picture}(11.25,2)
1513:       \linethickness{.25pt}
1514:       \multiput(0,1)(2.5,0){5}{\line(1,0){1.25}}
1515:     \end{picture} \tabularnewline
1516:     \hline
1517:   \end{tabular}
1518:   \vspace{3mm}
1519:   \caption{Single-dimple numerical solution obtained by the MPA with and
1520:     without the symmetry assumption \pref{eq:sym} and with various kinds
1521:     of discretization of $\partial_{xy}$.}
1522:   \label{tab:mpa_xy}
1523: \end{table}
1524: 
1525: \begin{figure}[htbp]
1526:   \centering\setlength{\unitlength}{1mm}
1527:   \begin{picture}(120,57)
1528:     \put(0,0){\includegraphics[width=120mm]{numerfigs/d_xy_profiles}}
1529:     \linethickness{.5pt}
1530:     \put(82,52){\line(1,0){3.755}}
1531:     \linethickness{.1pt}
1532:     \put(82,48){\line(1,0){3.75}}
1533:     \linethickness{.15mm}
1534:     \multiput(82.3,44)(1,0){4}{\line(1,0){.15}}
1535:     \linethickness{.25pt}
1536:     \multiput(82,40)(2.5,0){2}{\line(1,0){1.25}}
1537:     \put(88,51){\scriptsize Fourier $\partial_{xy}$}
1538:     \put(88,47){\scriptsize $\Omega$, left-sided $\partial_{xy}$}
1539:     \put(88,43){\scriptsize $\Omq$, left-sided $\partial_{xy}$}
1540:     \put(88,39){\scriptsize $\Omq$, right-sided $\partial_{xy}$}
1541:   \end{picture}
1542:   \caption{Profile of the single-dimple numerical solution $w_\MP$ at
1543:     $x=0$ obtained by the MPA with and without the symmetry assumption
1544:     \pref{eq:sym} and with various kinds of discretization of
1545:     $\partial_{xy}$.}
1546:   \label{fig:profiles_xy}
1547: \end{figure}
1548: 
1549: On the full domain $\Omega$ with no assumption on symmetry of
1550: solutions the discretization of $\partial_{xy}$ using the
1551: left/right-sided finite differences~\pref{eq:dx_RL} provides a
1552: numerical solution that is slightly asymmetric
1553: (Fig.~\ref{fig:profiles_xy}, thin solid line).  The Fourier transform
1554: provides a symmetric solution (Fig.~\ref{fig:profiles_xy}, thick solid
1555: line). The same numerical solution can be obtained on $\Omq$ under the
1556: symmetry assumption \pref{eq:sym} with $\partial_{xy}$ discretized
1557: using the fast cosine/sine transform.
1558: 
1559: On $\Omq$ the symmetry of numerical solutions is guaranteed by
1560: assumption~\pref{eq:sym}. The use of left/right-sided discretization
1561: of $\partial_{xy}$ does, however, have an influence on the shape of
1562: the numerical solution, as Fig.~\ref{fig:profiles_xy} shows (the
1563: dotted and the dashed line).
1564: 
1565: \subsection{Convergence}
1566: We now turn to the influence of the size of the space step $\Delta x$,
1567: $\Delta y$ on the numerical solution. We run the MPA on $\Omq$ under the
1568: symmetry assumption~\pref{eq:sym} with $\partial_{xy}$ discretized
1569: using (a) the fast cosine/sine transform, (b) left-sided finite
1570: differences, (c) right-sided finite differences. We consider $\Delta
1571: x=\Delta y = 0.5, 0.4, 0.3, 0.2, 0.1$, i.e., we take 200, 250, 333,
1572: 500, and 1000 points in both axis directions, respectively.
1573: Figure~\ref{fig:various_dxdy} illustrates convergence 
1574: as $\Delta x, \Delta y\to
1575: 0$ of the numerical solutions obtained by various types of discretization
1576: of $\partial_{xy}$.
1577: 
1578: \begin{figure}[htbp]
1579:   \centering\setlength{\unitlength}{1mm}
1580:   \begin{picture}(140,57)
1581:     \put(0,0){\includegraphics[width=65mm]{numerfigs/d_xy_infty_norm}}
1582:     \put(70,0){\includegraphics[width=65mm]{numerfigs/d_xy_S}}
1583:     \put(37,47){\scriptsize $\Mod{w^{}_L-w^{}_{CS}}_\infty$}
1584:     \put(44,25){\scriptsize $\Mod{w^{}_R-w^{}_{CS}}_\infty$}
1585:     \put(95,49){\scriptsize $S(w^{}_L)$}
1586:     \put(95,15){\scriptsize $S(w^{}_R)$}
1587:     \put(110,27){\scriptsize $S(w^{}_{CS})$}
1588:   \end{picture}
1589:   \caption{Influence of the size of the space step $\Delta x$, $\Delta y$
1590:     on the numerical solution $w_\MP$ obtained by the MPA for three
1591:     different kinds of discretization of $\partial_{xy}$. Let
1592:     $w^{}_L$, $w^{}_R$ denote the numerical solutions obtained using
1593:     the left and right-sided discretization of $\partial_{xy}$,
1594:     respectively, $w^{}_{CS}$ using the fast cosine/sine transform.
1595:     Left: comparison of the solutions in the maximum norm; right: the
1596:     value of shortening $S$.}
1597:   \label{fig:various_dxdy}  
1598: \end{figure}
1599: 
1600: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1601: \subsection{Dependence on the size of the domain}\label{sec:domain_size}
1602: As observed in~\cite{HoLoPe1}, the localized nature of the solutions
1603: suggests that they should be independent of domain size, in the sense
1604: that for a sequence of domains of increasing size the solutions
1605: converge (for instance uniformly on compact subsets). Such a
1606: convergence would also imply convergence of the associated energy
1607: levels. Similarly, we would expect that the aspect ratio of the domain
1608: is of little importance in the limit of large domains.
1609: 
1610: We tested these hypotheses by computing the single-dimple solution on
1611: domains of different sizes and aspect ratios. In all the computations
1612: the space step $\Delta x = \Delta y = 0.5$ is fixed. In order 
1613: to use the continuation method of Sec.~\ref{sec:continuation},
1614: we discretized $\partial_{xy}$ using the left-sided finite differences.
1615: We also assumed symmetry of solutions given by~\pref{eq:sym} and worked
1616: on $\Omq$. We recall the notation of computational domains,
1617: $\Omega=(-a,a) \times (-b,b)$, $\Omq=(-a,0)\times(-b,0)$.
1618: 
1619: Figure~\ref{fig:w_domains} shows the results for load $\l=1.4$. First
1620: we notice that the central dimple has almost the same shape in all the
1621: shown cases. But there seems to be a difference in the slope of the
1622: ``flat'' part leading to this dimple. On domains with small $a$ (short
1623: cylinder) the derivative in the circumferential $y$-direction in this
1624: part is larger than on domains with larger $a$ (longer cylinder). The
1625: circumferential length $b$ seems to be less important for the shape of
1626: the solution: for example, the cases (200,50) and (200,100) look like
1627: restrictions of the case (200,200) to smaller domains.
1628: 
1629: We
1630: take a closer look at domains of sizes $(a,b) = (100,100)$,
1631: $(100,200)$, $(200,100)$, and $(200,200)$ and the corresponding
1632: solutions $w_{100,100}$, $w_{100,200}$, $w_{200,100}$, and
1633: $w_{200,200}$ shown in the figure. We compare the first three with the
1634: last one, respectively. It does not make sense to compare the values
1635: of $w$ itself since the energy functional $F_\l$ depends on
1636: derivatives of $w$ only. We choose to compare $w_{xx}$ and
1637: $w_{yy}$. Table~\ref{tab:w_100_200} gives the infinity norm of the
1638: relative differences. Fig.~\ref{fig:w_100_200} shows graphs of the
1639: difference $w_{100,100}-w_{200,200}$ and of the second derivatives
1640: $(w_{200,200}-w_{100,100})_{xx}$, $(w_{200,200}-w_{100,100})_{yy}$ on
1641: the subdomain $(-100,0)^2$.
1642: 
1643: We conclude that solutions on different domains compare well; the
1644: maximal difference in the second derivatives of $w$ is three orders of
1645: magnitude smaller than the supremum norm of the same derivative. We
1646: also observe that varying the length parameter $a$ while keeping
1647: the circumference parameter $b$ fixed causes larger changes in the numerical
1648: solution than varying the cylinder circumference while keeping the length
1649: fixed.
1650: 
1651: \begin{figure}[htbp]
1652:  \centering\setlength{\unitlength}{1mm}
1653:  \begin{picture}(139,115)
1654: %   \put(0,0){\framebox(139,115){}}
1655:    \put(0,80){\includegraphics[width=45mm]{numerfigs/graph_01_50x50}}
1656:    \put(47,80){\includegraphics[width=45mm]{numerfigs/graph_01_50x100}}
1657:    \put(94,80){\includegraphics[width=45mm]{numerfigs/graph_01_50x200}}
1658:    \put(5,110){\scriptsize $(a,b)=(50,50)$}
1659:    \put(52,110){\scriptsize $(a,b)=(50,100)$}
1660:    \put(99,110){\scriptsize $(a,b)=(50,200)$}
1661:    \put(0,40){\includegraphics[width=45mm]{numerfigs/graph_01_100x50}}
1662:    \put(47,40){\includegraphics[width=45mm]{numerfigs/graph_01_100x100}}
1663:    \put(94,40){\includegraphics[width=45mm]{numerfigs/graph_01_100x200}}
1664:    \put(5,70){\scriptsize $(a,b)=(100,50)$}
1665:    \put(52,70){\scriptsize $(a,b)=(100,100)$}
1666:    \put(99,70){\scriptsize $(a,b)=(100,200)$}
1667:    \put(0,0){\includegraphics[width=45mm]{numerfigs/graph_01_200x50}}
1668:    \put(47,0){\includegraphics[width=45mm]{numerfigs/graph_01_200x100}}
1669:    \put(94,0){\includegraphics[width=45mm]{numerfigs/graph_01_200x200}}
1670:    \put(5,30){\scriptsize $(a,b)=(200,50)$}
1671:    \put(52,30){\scriptsize $(a,b)=(200,100)$}
1672:    \put(99,30){\scriptsize $(a,b)=(200,200)$}
1673:  \end{picture}
1674:  \caption{The single-dimple mountain-pass solution
1675:    with $\l=1.4$ computed under the assumption of
1676:    symmetry~\pref{eq:sym} with left-sided discretization of
1677:    $\partial_{xy}$ for various domain sizes.
1678:    }
1679:  \label{fig:w_domains}
1680: \end{figure}
1681: 
1682: \begin{figure}[htbp]
1683:   \centering\setlength{\unitlength}{1mm}
1684:   \begin{picture}(139,35)
1685: %    \put(0,0){\framebox(139,35){}}
1686:     \put(0,0){\includegraphics[width=45mm]{numerfigs/w100-w200}}
1687:     \put(47,0){\includegraphics[width=45mm]{numerfigs/w100-w200_mxx}}
1688:     \put(94,0){\includegraphics[width=45mm]{numerfigs/w100-w200_myy}}
1689:     \put(5,32){\scriptsize $w_{100,100}-w_{200,200}$}
1690:     \put(52,32){\scriptsize $(w_{200,200}-w_{100,100})_{xx}$}
1691:     \put(99,32){\scriptsize $(w_{200,200}-w_{100,100})_{yy}$}
1692:   \end{picture}
1693:   \caption{Comparison of solutions $w_{100,100}$, $w_{200,200}$
1694:     from Fig.~\ref{fig:w_domains} obtained on
1695:     square domains with $a=b=100$ and $a=b=200$, respectively, and their
1696:     second derivatives. For a reference, we note that
1697:     $\Mod{(w_{200,200})_{xx}}_\infty = 1.064522$ and
1698:     $\Mod{(w_{200,200})_{yy}}_\infty = 0.8242491$.}
1699:   \label{fig:w_100_200}
1700: \end{figure}
1701: 
1702: \begin{table}[htbp]
1703:   \centering
1704:   \renewcommand{\arraystretch}{1.2}
1705:   \begin{tabular}{c|c|c|}
1706:     \cline{2-3}
1707:     & $\frac{\Mod{(w-w_{200,200})_{xx}}_\infty}{\Mod{(w_{200,200})_{xx}}_\infty}$
1708:     & $\frac{\Mod{(w-w_{200,200})_{yy}}_\infty}{\Mod{(w_{200,200})_{yy}}_\infty}$
1709:     \tabularnewline
1710:     \hline
1711:     \multicolumn{1}{|c|}{$w=w_{100,100}$} & $2.835\cdot 10^{-3}$ &
1712:     $5.313\cdot 10^{-3}$ \tabularnewline
1713:     \hline
1714:     \multicolumn{1}{|c|}{$w=w_{100,200}$} & $1.943\cdot 10^{-3}$ &
1715:     $4.917\cdot 10^{-3}$ \tabularnewline
1716:     \hline
1717:     \multicolumn{1}{|c|}{$w=w_{200,100}$} & $1.827\cdot 10^{-4}$ &
1718:     $9.638\cdot 10^{-4}$ \tabularnewline
1719:     \hline
1720:   \end{tabular}
1721:   \renewcommand{\arraystretch}{1}
1722:   \vspace{3mm}
1723:   \caption{Comparison of the second derivatives of solutions
1724:     from Fig.~\ref{fig:w_domains}
1725:     computed on domains with $(a,b)=(100,100)$,
1726:     $(100,200)$, $(200,100)$, and $(200,200)$.}
1727:   \label{tab:w_100_200}
1728: \end{table}
1729: 
1730: Another way of studying the influence of the domain size on the
1731: numerical solution is comparing solution branches obtained by
1732: continuation as described in
1733: Sec.~\ref{sec:continuation}. We start with
1734: the mountain-pass solution for $\l=1.4$ shown in
1735: Fig.~\ref{fig:w_domains} and continue it for both $\l>1.4$ and
1736: $\l<1.4$. The results are presented in Fig.~\ref{fig:cont}. We observe
1737: that the branches corresponding to the considered domains do not
1738: differ much for the range of $\l$ between approximately $0.71$ and
1739: $2$. Below $\l\approx 0.71$ the size of the domain, particularly the
1740: length of the cylinder described by $a$, has a strong influence. The
1741: graph on the right shows that the larger (longer) the domain $\Omega$
1742: the smaller the value of $\l$ at which the norm $\Mod{w}_X$ starts to
1743: rapidly increase for decreasing $\l$. The graph on the left shows the
1744: energy $F_\l(w)$ along a solution branch. The data shown here
1745: correspond to the ones in the graph on the right marked by a solid
1746: line. The dashed line in the right graph shows also some data after
1747: the first limit point is passed.
1748: 
1749: \begin{figure}[htbp]
1750:   \centering\setlength{\unitlength}{1mm}
1751:   \begin{picture}(140,50)(0,-1)
1752: %    \put(0,-1){\framebox(140,50){}}
1753:     \put(0,0){\includegraphics[width=67mm]{numerfigs/vlambda}}
1754:     \put(75,0){\includegraphics[width=65mm]{numerfigs/normlambda_dashed}}
1755: 
1756:     \put(73,20.5){\rotatebox{90}{\scriptsize $\Mod{w}_X^2$}}
1757:     \put(135,-1){\scriptsize $\l$}
1758: 
1759:     \put(93.5,38){\scriptsize $a=200$}
1760:     \put(88,37){\line(1,0){15}}
1761:     \put(88,34){\scriptsize $b=50$}
1762:     \put(91,30){\scriptsize $b=100$}
1763:     \put(94,26){\scriptsize $b=200$}
1764: 
1765:     \put(116,21){\scriptsize $a=100$}
1766:     \put(110.5,20.5){\line(1,0){15}}
1767:     \put(110.5,18){\scriptsize $b=50$}
1768:     \put(113,16){\scriptsize $b=100$}
1769:     \put(116.5,14){\scriptsize $b=200$}
1770: 
1771:     \put(132,12){\tiny $a=50$}
1772:     \put(122.5,11.5){\line(1,0){16.5}}
1773:     \put(122.5,9.5){\tiny $b=50,100,200$}
1774: 
1775:     \put(-2.5,20.5){\rotatebox{90}{\scriptsize $F_\l(w)$}}
1776:     \put(61,-1){\scriptsize $\l$}
1777: 
1778:     \put(24.5,37){\scriptsize $a=200$}
1779:     \put(23,33){\line(0,1){9.5}}
1780:     \put(11.5,41){\scriptsize $b=50$}
1781:     \put(12,36){\scriptsize $b=100$}
1782:     \put(12.5,33){\scriptsize $b=200$}
1783: 
1784:     \put(28,23){\scriptsize $a=100$}
1785:     \put(27,20.5){\line(0,1){6.5}}
1786:     \put(15,25.5){\scriptsize $b=50$}
1787:     \put(15.5,23){\scriptsize $b=100$}
1788:     \put(16,20.5){\scriptsize $b=200$}
1789: 
1790:     \put(20,14.5){\scriptsize $b=50,100,200$, $a=50$}
1791:   \end{picture}
1792:   \caption{Continuation of the single-dimple solution found as a numerical
1793:     mountain pass for $\l=1.4$ on domains of various sizes for a range
1794:     of values $\l$. Left: $F_\l(w)$ as a function of $\l$, right:
1795:     $\Mod{w}_{X}$ as a function of $\l$.}
1796:   \label{fig:cont}
1797: \end{figure}
1798: 
1799: Figure~\ref{fig:cont_100} shows how the graph of $w(x,y)$ changes as a
1800: solution branch is followed. Here we chose a square domain with
1801: $a=b=100$ and plotted the solution for four values of $\lambda$ (note
1802: that Figs.~\ref{fig:cont_100}(c), \ref{fig:w_domains}(100,100), and
1803: \ref{fig:profiles_xy}(dotted line) show the same numerical solution).
1804: We observe that with decreasing $\l$ the height of the central dimple
1805: increases, the dimple becomes wider, and the ripples (present at $\l$
1806: close to 2) disappear. In Fig.~\ref{fig:cont_100}(a) we observe that
1807: new dimples are being formed next to the central dimple.
1808: 
1809: \begin{figure}[htbp]
1810:   \centering\setlength{\unitlength}{1mm}
1811:   \begin{picture}(139,75)
1812:     %\put(0,0){\framebox(139,75){}}
1813:     \put(0,40){\includegraphics[width=45mm]{numerfigs/graph_01_l0593475}}
1814:     \put(47,40){\includegraphics[width=45mm]{numerfigs/graph_01_l061}}
1815:     \put(94,40){\includegraphics[width=45mm]{numerfigs/graph_01_l140}}
1816:     \put(5,70){\scriptsize (a) $\l\approx0.593$}
1817:     \put(52,70){\scriptsize (b) $\l=0.61$ (mountain pass)}
1818:     \put(99,70){\scriptsize (c) $\l=1.4$ (mountain pass)}
1819:     \put(94,0){\includegraphics[width=45mm]{numerfigs/graph_01_l195}}
1820:     \put(9.5,0){\includegraphics[width=70mm]{numerfigs/normlambda_100x100}}
1821:     \put(99,30){\scriptsize (d) $\l=1.95$ (mountain pass)}
1822:     \put(73,-1){\scriptsize $\l$}
1823:     \put(6,22){\rotatebox{90}{\scriptsize $\Mod{w}_X^2$}}
1824:     \put(40,35){\scriptsize $a=b=100$}
1825:     \put(54.5,27){\scriptsize mountain pass}
1826:     \put(54.5,24){\scriptsize local minimizer}
1827:     \put(54.5,21){\scriptsize (of $F_\l$)}
1828:     \put(24,25){\scriptsize (a)}
1829:     \put(25,8){\scriptsize (b)}
1830:     \put(50.5,9){\scriptsize (c)}
1831:     \put(69.2,8.3){\scriptsize (d)}
1832:   \end{picture}
1833:   \caption{A detailed look at the continuation of the single-dimple solution on the domain with $a=b=100$.}
1834:   \label{fig:cont_100}
1835: \end{figure}
1836: 
1837: It should be remarked that although we started the continuation at
1838: $\l=1.4$ at a mountain-pass point, not all the points along a
1839: continuation branch are mountain passes. Since it is not feasible to
1840: use the MPA to verify this for each point, we chose just a few. Still
1841: on the example of the domain with $a=b=100$ in
1842: Fig.~\ref{fig:cont_100}, the circles on the continuation branch mark
1843: those points that have also been found by the MPA (for $\l=0.61, 0.65,
1844: 0.8, 1.0, 1.2, 1.4, 1.6, 1.8, 1.95$). As described in
1845: Sec.~\ref{sec:MPA}, in order to start the MP-algorithm a point $w_2$
1846: is needed such that $F_\l(w_2)<0$. The analysis in~\cite{HoLoPe1}
1847: shows that for a given $\l$ such a point exists provided the domain
1848: $\Omega$ is large enough and in practice it is found by the SDM of
1849: Sec.~\ref{sec:SDM}. This was, indeed, the case for all the chosen
1850: values of $\l$ except for $\l=0.61$. In this case, starting from some
1851: $w_0$ with a large norm, the SDM provides a trajectory $w(t)$ such
1852: that $F_\l(w(t))>0$ for all $t>0$. In fact, the steepest descent
1853: method converges to a local minimizer $w_\mathrm{M}^{}$ with
1854: $F_\l(w_\mathrm{M}^{})\approx 76.1$. This is hence no mountain pass
1855: but, nevertheless, lies on the same continuation branch and is marked
1856: by a triangle in the figure. Despite $F_\l(w_\mathrm{M}^{})>0$ we can
1857: still try to run the MPA with $w_2=w_\mathrm{M}^{}$. It converges and
1858: yields $\wMP$ with $F_\l(\wMP)\approx94.8$ (marked by a circle at
1859: $\l=0.61$ and shown in graph (b)).
1860: 
1861: The comparison of solutions computed on different domains and their
1862: respective energies suggests that for each $\l$ we are indeed dealing
1863: with a single, localized function defined on $\R^2$, of which our
1864: computed solutions are finite-domain adaptations. Based on this
1865: suggestion and the above discussion of the mountain-pass solutions we
1866: could, for example, conclude that the mountain-pass energy
1867: \[
1868: V(\l,\Omega) := \inf_{w_2} \bigl\{\, F_\l\bigl(\wMP(\l,\Omega,w_2)\bigr):
1869: F_\l(w_2)<0\, \bigr\}
1870: \]
1871: is a finite-domain approximation of a function $V(\l)$, whose graph
1872: almost coincides with that of $V(\l,\Omega)$ for $\l$ not too small
1873: (cf.~Fig.~\ref{fig:cont} left).
1874: 
1875: \section{Discussion}
1876: \label{sec:conc}
1877: 
1878: \subsection{Variational numerical methods}
1879: We have seen that given a complex energy surface many solutions may be 
1880: found using
1881: these variational techniques. For example, for a fixed end shortening 
1882: of $S=40$,~Fig.~\ref{fig:sol_csdm}, Fig.~\ref{fig:sol_cmpa1}
1883: and Fig.~\ref{fig:sol_cmpa2} are all solutions. Which of these solutions is of
1884: greatest relevance depends on the question that is being asked. 
1885: 
1886: 
1887: 
1888: In the context of the cylinder (and similar structural applications)
1889: the mountain-pass solution from the unbuckled state ($w_1=0$) with 
1890: minimal energy is of physical interest. Often the experimental buckling 
1891: load may be at $20$--$30\%$ of the linear prediction from a bifucation 
1892: analysis (in our scaling this corresponds to $\lambda=2$). 
1893: This uncertainty in the buckling
1894: load is a drawback for design and so ``knockdown'' factors have been
1895: introduced, based on experimental data. It was argued in \cite{HoLoPe1}
1896: that the energy of the mountain-pass solution $\wMP$ in fact provides a
1897: lower bound on the energy required to buckle the cylinder and so these
1898: solutions provide bounds on the (observed) buckling load of the
1899: cylinder. 
1900: 
1901: This example illustrates an important aspect of the (constrained)
1902: mountain-pass algorithm: its explicit non-locality. The algorithm
1903: produces a saddle point which has an additional property: it is the
1904: separating point (and level) between the basins of attraction of the
1905: end points $w_1$ and $w_2$.
1906: 
1907: Another technique to investigate a complex energy surface is to perform 
1908: a simulated annealing computation, essentially to solve the SDM (or the CSDM) 
1909: problem with additive stochastic forcing. The aim in these techniques is 
1910: often to find a global minimizer (if it exists) where there are a large 
1911: number of local minimizers. Here by either the MPA or the CMPA we find the solution 
1912: between two such minima and so get an estimate on the surplus energy needed to 
1913: change between local minima.
1914:   
1915: \subsection{Numerical issues}
1916: 
1917: The numerical issues that we encountered are of two types. First there
1918: are the requirements that are related to the specific problem of the
1919: \vKD\ equations, such as the discretization of the mixed derivative
1920: and the bracket, and the fact that the solutions are symmetric and
1921: highly localized.
1922: 
1923: For other difficulties it is less clear. For smaller values of $\l$
1924: each of the variational methods converged remarkably slowly. Newton's
1925: method provides a way of improving the convergence, but the question
1926: is relevant whether this slow convergence is typical for a whole class
1927: of variational problems. It would be interesting to connect the rate
1928: of convergence of, for instance, the SDM to certain easily measurable
1929: features of the energy landscape.
1930: 
1931: \bibliography{refs2}
1932: 
1933: \end{document}
1934: 
1935: