0809.1834/GL.tex
1: \documentclass[11pt,reqno]{amsart}      %use amsart package with 11pt font
2:                                        %size and reqno (equation number on
3:                                         %the right ) options
4: %\usepackage[first,bottomafter,light]{draftcopy}
5: %\usepackage[none,bottom,light]{draftcopy}
6: \usepackage{cite}                       %Makes citations fancier
7: \usepackage{amssymb,amsthm}             %get some extra symbols  and
8:                                         %environments defined 
9: %\usepackage{a4}                         %fix pagesize
10: %\usepackage[notcite]{showkeys}         %show labels but not cite 
11: %fix to get the right \Scri:
12: %\usepackage{eucal}                      %make mathcal use Euler font
13:                                         %now \CMcal is the same as the old 
14:                                         %\mathcal
15: %\newcommand{\EUcal}{\mathcal}           %now EUcal gives Euler family
16:                                         %in this document use \CMcal and 
17:                                         %\EUcal instead of \mathcal 
18: %end of fix
19: 
20: %\usepackage{maple2e}
21: 
22: \usepackage{calc}                       %graphics packages (calc is used by
23:                                         %graphicx) 
24: \usepackage{graphicx}
25: %\usepackage{epic,eepic}                 
26: 
27: \numberwithin{figure}{section}          %tie figure numbering to
28:                                         %section counter
29: 
30: \numberwithin{equation}{section}        %tie equation numbering to
31:                                         %section counter 
32: %general definitions
33: \newcommand{\mnote}[1]{\marginpar{\raggedright\tiny\em
34: $\!\!\!\!\!\!\,\bullet$ #1}}%Marginal Note
35: %\newcommand{\mnote}[1]{} %no marginal notes
36: \renewcommand{\Re}{{\mathbb R}}         %the real numbers
37: \newcommand{\Compl}{\mathbb C}          %the complex numbers
38: \newcommand{\mathbfr}{\mathbf R}        %bibtex fix to get  bf R
39: \newcommand{\rmix}{\text{\rm IX}}       %bibtex fix to get IX
40: \newcommand{\NatNum}{{\mathbb N}}       %the natural numbers
41: \newcommand{\la}{\langle}               %left angle bracket
42: \newcommand{\ra}{\rangle}               %right angle bracket
43: \newcommand{\half}{\frac{1}{2}}         %half
44: \newcommand{\third}{\frac{1}{3}}        %third
45: \newcommand{\quart}{\frac{1}{4}}        %quart
46: \newcommand{\sixth}{\frac{1}{6}}        %sixth
47: \newcommand{\nth}{\frac{1}{n}}          %nth
48: \newcommand{\abs}[1]{\lvert#1\rvert}    %absolute value
49: \newcommand{\norm}[1]{\lVert#1\rVert}   %norm
50: \DeclareMathOperator*\esssup {ess \, sup}  %essential supremum
51: %
52: %local definitions
53: \newcommand{\hpi}{\hat \pi}             %''tracefree part'' of \pi
54: \newcommand{\agamma}{{\bar \gamma}}     %ambient background metric tensor 
55: \newcommand{\Weyl}[1]{\overset{(#1)}{W}}%differentiated Weyl tensor
56: \newcommand{\so}{\mathfrak{so}}         %Lie algebra of SO
57: \newcommand{\iso}{\mathfrak{iso}}       %Lie algebra of ISO
58: \newcommand{\gl}{\mathfrak{gl}}         %Lie algebra of GL
59: \newcommand{\semidir}{\ltimes}          %semidirect product             
60: \newcommand{\tens}{\otimes}             %tensorproduct
61: \newcommand{\Teich}{\CMcal T}           %Teichmuller space
62: \newcommand{\EE}{\CMcal E}              %energy
63: \newcommand{\OO}{\CMcal O}              %open cover
64: \newcommand{\rmO}{\mathrm O}            %orthogonal group
65: \newcommand{\PP}{\CMcal P}              %positive definite matrices
66: \newcommand{\U}{\text{\rm U}}           %unitary group
67: \newcommand{\SU}{\text{\rm SU}}         %special unitary group
68: \newcommand{\ISO}{\text{\rm ISO}}       %inhomogeneous special orthogonal group
69: \newcommand{\SO}{\text{\rm SO}}         %special orthogonal group
70: \newcommand{\Ad}{\text{\rm Ad}}         %Adjoint representation
71: \newcommand{\vect}{\text{\rm vec}}       %vector representation
72: \newcommand{\Riem}{\text{\rm Riem}}     %Riemann tensor
73: \newcommand{\Ric}{\text{\rm Ric}}       %Ricci tensor
74: \newcommand{\Scal}{\text{\rm Scal}}     %Scalar curvature
75: %
76: \newcommand{\BR}{Q}                     %Bel-Robinson Tensor
77: \newcommand{\EBR}{{\CMcal Q}}         %Bel-Robinson Tensor
78: \newcommand{\FF}{\CMcal F}            %junk terms
79: \newcommand{\GG}{\CMcal G}            %junk terms
80: \newcommand{\RR}{\CMcal R}            %curvature junk
81: \newcommand{\LL}{\CMcal L}            %Lagrangian
82: \newcommand{\DD}{\CMcal D}            %domain
83: 
84: %ambient definitions
85: \newcommand{\aM}{{\bar M}}              %ambient manifold
86: \newcommand{\anabla}{{\bar \nabla}}     %ambient covariant derivative
87: \newcommand{\ame}{{\bar g}}             %ambient metric tensor 
88: \newcommand{\aGamma}{{\bar \Gamma}}      %ambient Christoffel
89: \newcommand{\ah}{{\bar h}}              %ambient perturbation
90: \newcommand{\aR}{{\bar R}}              %ambient Riemann tensor
91: \newcommand{\aRiem}{\overline{\text{Riem}}} %ambient Riemann tensor
92: \newcommand{\aRic}{\overline{\text{Ric}}}   %ambient Riemann tensor
93: \newcommand{\avol}{\mu(\ame)}           %ambient volume element
94: \newcommand{\vol}{\mu(g)}               %volume element w.r.t.  g
95: \newcommand{\hvol}{\mu(\hme)}           %volume element w.r.t. \hat g
96: \newcommand{\he}{\hat e}                %background frame element
97: %conformal definitions
98: \newcommand{\tM}{\tilde M}
99: \newcommand{\tme}{\tilde g}             %conformal metric
100: \newcommand{\tnabla}{\tilde \nabla}     %conformal covariant derivative 
101: \newcommand{\tR}{\tilde R}              %conformal curvature
102: \newcommand{\Scri}{\EUcal I}    %Scri
103: 
104: \newcommand{\supp}{\text{\rm supp}}     %support 
105: \newcommand{\bm}{\text{\bf \rm m}}      %null vector
106: \newcommand{\bl}{\text{\bf \rm l}}      %null vector
107: 
108: \newcommand{\CC}{{\CMcal C}}          %constraint set
109: \newcommand{\TT}{\text{\sc TT}}         %transverse traceless
110: \newcommand{\Vol}{\text{\rm Vol}}       %volume
111: \newcommand{\inj}{\text{\rm inj}}       %injectivity radius
112: \newcommand{\diam}{\text{\rm diam}}     %diameter
113: \newcommand{\ii}{ii}                    %double contraction
114: \newcommand{\bfi}{\text{\bf i}}         %imbedding
115: \newcommand{\Eps}{\epsilon}             %alternating tensor
116: \newcommand{\eps}{\epsilon}             %epsilon or sign 
117: \newcommand{\Lie}{{\CMcal L}}         %Lie derivative
118: \newcommand{\modLie}{\Hat{\Lie}}        %modified Lie derivative
119: \newcommand{\tr}{{\text{\rm tr}}}       %trace
120: \newcommand{\diag}{{\text{\rm diag}}}   %diag
121: \renewcommand{\div}{\text{\rm div}}     %divergence (on tensors)
122: \newcommand{\curl}{\text{\rm curl}}     %curl (on tensors)
123: \newcommand{\Lapse}{N}                  %Lapse function
124: \newcommand{\hme}{\hat g}               %background metric
125: \newcommand{\hnabla}{\hat \nabla}       %background covariant derivative
126: \newcommand{\hGamma}{\hat \Gamma}       %background Christoffel
127: \newcommand{\Shift}{X}                  %Shift Vectorfield
128: \newcommand{\Orbit}{\CMcal O}         %Orbit
129: \newcommand{\Def}[1]{{^{(#1)}\pi}}      %deformation tensor
130: \newcommand{\hDef}[1]{{^{(#1)}\hat\pi}} %tracefree part of deformation tensor
131: \newcommand{\Proj}{\Pi}                 %Projection to hypersurface
132: \newcommand{\Smallset}{\CMcal B}         %smallness condition
133: \newcommand{\Frob}{\text{Frob}}         %Frobenius
134: \newcommand{\Id}{\text{Id}}             %identity map
135: \newcommand{\del}{\delta}               %gauge quantity
136: 
137: 
138: \newcommand{\bfH}{\mathbf H}
139: \newcommand{\bfE}{\mathbf E}
140: \newcommand{\bfR}{\mathbf R}
141: \newcommand{\tSL}{\widetilde{\text{SL}}}
142: \newcommand{\frakg}{\mathfrak g}
143: 
144: 
145: 
146: %constants
147: 
148: %%%%%%%%theorems etc.%%%%%%%%%%%%%%%%%%%%%%
149: \theoremstyle{plain}
150: \newtheorem{thm}{Theorem}[section]
151: \newtheorem{cor}[thm]{Corollary}
152: \newtheorem{conj}[thm]{Conjecture}
153: \newtheorem{lemma}[thm]{Lemma}
154: %\newtheorem{example}[thm]{Example}
155: \newtheorem{definition}[thm]{Definition}
156: \newtheorem{prop}[thm]{Proposition}
157: \newtheorem{remark}[thm]{Remark}
158: %\newtheorem{lemma}{Lemma}
159: %%%%%%%%%%Title etc%%%%%%%%%%%%%%%%%%%%%%%%
160: \title[Optimal Control for Ginzburg-Landau]{Convergence Rates for an
161:   Optimally Controlled Ginzburg-Landau equation}
162: \subjclass[2000]{49M29, 65M12}
163: \keywords{Ginzburg-Landau Equation, Optimal Control, Hamilton-Jacobi,
164:   Error Estimates, Stochastic Partial Differential Equation,
165:   Symplectic Euler}
166: \author{Mattias Sandberg}
167: \address{CMA, University of Oslo, P.O. Box 1053 Blindern, 0316 Oslo, Norway
168: }
169: \email{mattias.sandberg@cma.uio.no}
170: %\date{\today }
171: %\ {\em File: \jobname{.tex}}
172: %\date{November 9, 1999}
173: 
174: 
175: \setcounter{tocdepth}{2}
176: 
177: \begin{document}
178: \begin{abstract}
179: An optimal control problem related to the probability of transition
180: between stable states for a thermally driven Ginzburg-Landau equation
181: is considered. The value function for  the optimal control problem
182: with a spatial discretization is shown to converge quadratically to
183: the value function for the original problem. This is done by using that the
184: value functions solve similar Hamilton-Jacobi equations, the equation for the
185: original problem being defined on an infinite dimensional Hilbert
186: space. Time discretization is performed using the Symplectic Euler
187: method. Imposing a reasonable condition this method is shown to be
188: convergent of order one in time, with a constant independent of the spatial discretization.
189: \end{abstract}
190: \maketitle
191: 
192: \tableofcontents
193: %\listoffigures
194: \section{Introduction}
195: We shall in this paper study the convergence of the Symplectic Euler
196: scheme for approximating optimal control of the real Ginzburg-Landau
197: equation. This follows the work developed in \cite{Sandberg-Szepessy}, where a
198: convergence result for the value function to an optimally controlled
199: ODE is obtained using the corresponding Hamilton-Jacobi equation. As
200: there exists a rigorous theory also for infinite-dimensional
201: Hamilton-Jacobi equations, developed by M.\ Crandall and P.-L. Lions
202: \cite{Crandall-Lions1,Crandall-Lions2,Crandall-Lions3,Crandall-Lions4,Crandall-Lions5,Crandall-Lions6,Crandall-Lions7}, it is possible to perform a convergence analysis
203: for a spatial discretization of an optimally controlled PDE, using
204: that the value function is a viscosity solution to an
205: infinite-dimensional Hamilton-Jacobi equation. In this paper the
206: analysis is performed for the specific problem at hand, but hopefully
207: the analysis is clear enough to make adaptations to other circumstances
208: (fairly) easy.
209: 
210: Consider the stochastic PDE
211: \begin{equation}\label{eq:SPDE}
212: \varphi_t=\delta\varphi_{xx}-\delta^{-1}V'(\varphi)+\sqrt{\varepsilon}\eta,
213: \quad \text{in } [0,T]\times[0,1],
214: \end{equation}
215: where $\delta$ is a positive number and $\eta$ is white noise in two dimensions; this means that $\eta$
216: is a random Gaussian distribution with zero mean and covariance
217: \begin{equation*}
218: E\big(\eta(x,t),\eta(x',t')\big)=\bar\delta(x-x')\bar\delta(t-t'),
219: \end{equation*}
220: where $E$ denotes the expectation and $\bar\delta$ the Dirac delta distribution. The ``state'' variable $\varphi$
221: satisfies the Dirichlet boundary conditions
222: \begin{equation*}
223: \varphi(t,0)=\varphi(t,1)=0,
224: \end{equation*}
225: and $V$ is the ``double-well'' potential
226: \begin{equation*}
227: V(\varphi)=\quart (\varphi^2-1)^2;
228: \end{equation*}
229: see Figure \ref{fig:V}. In one space dimension the stochastic PDE
230: (\ref{eq:SPDE}) makes sense, as existence and uniqueness of solutions
231: can be proved. Taking $\varepsilon=0$, the solutions to
232: (\ref{eq:SPDE}) generically approach one of the two stable critical
233: points, $\varphi_+$ or $\varphi_-$, (see Figure \ref{fig:phiplusphiminus}), which constitute
234: minima to the energy
235: \begin{equation*}
236: \int_0^1 \big( \frac{\delta}{2}\varphi_x^2 + \delta^{-1} V(\varphi)\big)dx.
237: \end{equation*}
238: \begin{figure}
239: \centering
240: \includegraphics[width=0.5\textwidth]{phiplusphiminus.eps}
241: \caption{Upper curve: $\varphi_+$. Lower curve: $\varphi_-$.}
242: \label{fig:phiplusphiminus}
243: \end{figure}
244: With a small $\varepsilon$, the solutions to (\ref{eq:SPDE}) spend
245: most of the time in the vicinity of either $\varphi_+$ or $\varphi_-$, but
246: as rare events make the transition from one to the other. The equation
247: (\ref{eq:SPDE}) may therefore be taken as a model for thermally driven
248: phase transitions, nucleations, etc.
249: 
250: The probability of jumping from $\varphi_+$ to $\varphi_-$ in the finite
251: time $T$ is related to the action functional
252: \begin{equation}\label{eq:actionfunctional}
253: I(\varphi)=\half\int_0^T\int_0^1 \big(\varphi_t - \delta\varphi_{xx} +
254: \delta^{-1} V'(\varphi)\big)^2 dx\, dt.
255: \end{equation}
256: Introduce the probability $P_T$ that a solution $\varphi$ to
257: (\ref{eq:SPDE}), with $\varphi(0)=\varphi_+$, satisfies $\varphi(T)\in S$, where
258: $S$ is an open subset of the set of continuous functions on the
259: spatial interval
260: $[0,1]$. Theory of large deviations  in \cite{Faris-Jona-Lasinio} gives that 
261: \begin{align*}
262: & -I(S) \leq \liminf_{\varepsilon \rightarrow 0}
263:  \varepsilon\log P_T \\
264: \intertext{and}
265: &\limsup_{\varepsilon \rightarrow 0}\varepsilon\log P_T \leq
266:  -I(\bar S)\\
267: \intertext{where}
268: &I(S)=\inf I(\varphi),
269: \end{align*}
270: with the infimum in the last equality taken over all continuous
271: functions $\varphi$ in $[0,T]\times[0,1]$ starting at $\varphi_+$ and ending
272: in $S$, and where $\bar S$ is the closure of $S$. By taking $S$ a small neighborhood of $\varphi_-$ we can
273: for small $\varepsilon$ approximate the probability of transition from
274: $\varphi_+$ to $\varphi_-$ with
275: \begin{equation*}
276: P_T \approx e^{-I(S)/\varepsilon}.
277: \end{equation*}
278: 
279: In \cite{Weinan-Ren-Vanden-Eijnden} the minimization of (\ref{eq:actionfunctional}) is
280: performed for $\varphi(0)=\varphi_+$ and $\varphi(T)=\varphi_-$ using optimization
281: of a finite difference approximation.
282: In this paper $\varphi(T)$ will not be held fixed, but
283: instead a penalty cost at the final time is added to the functional
284: (\ref{eq:actionfunctional}) in order to force the solution to end up
285: close to $\varphi_-$. The optimal control problem which will be
286: considered here is the following. Minimize, over all $\alpha \in
287: L^2\big(0,T;L^2(0,1)\big)$, the value $v_{\varphi_+,0}(\alpha)$, where
288: the functional $v$ is 
289: defined by
290: \begin{equation}\label{eq:costfunctional}
291: v_{\varphi_0,t_0}(\alpha)=\int_{t_0}^T h\big(\alpha(t)\big)dt+g\big(\varphi(T)\big),
292: \end{equation}
293: and where $\varphi$ is a mild solution to 
294: \begin{equation}\label{eq:flow}
295: \varphi_t=\delta \varphi_{xx}-\delta^{-1} V'(\varphi)+\alpha, \quad \varphi(t_0)=\varphi_0.
296: \end{equation}
297: In order to define mild solutions we denote by $S(t)$ the contraction
298: semigroup of bounded linear operators on $L^2(0,1)$ generated by
299: $\delta\, d^2/dx^2$ defined on $H^1_0(0,1) \cap H^2(0,1)$. A mild solution
300: to \eqref{eq:flow} is a function 
301: $\varphi \in C(t_0,T;L^2)$ such
302: that, for all $t_0 \leq t \leq T$,
303: \begin{equation}\label{eq:phimild}
304: \varphi(t)=S(t-t_0)\varphi_0 + \int_{t_0}^T S(t-s) \big(
305: -\delta^{-1}V'(\varphi(s))+\alpha(s) \big) ds.
306: \end{equation}
307: In the appendix existence and uniqueness of weak solutions in
308: $C(t_0,T;H_0^1)$ of \eqref{eq:flow} is proved when the starting
309: position $\varphi_0 \in H_0^1(0,1)$. Such weak solutions are
310: also mild solutions (this can be seen by using  e.g.\ the calculation on page 105 in \cite{Pazy}). Furthermore,
311: with $\alpha$ bounded in $L^2(t_0,T;L^2)$, the weak solution is
312: bounded in $C(t_0,T;H_0^1)$. Hence, the potential $V$ may be changed
313: outside an interval $[-s,s]$ without changing the result of the
314: optimal control problem. For simplicity, we shall henceforth use the
315: potential $\tilde V$ in Figure~ \ref{fig:V} and quickly change
316: notation, so that we let $V \equiv \tilde V$, i.e.\ $V$ is given by
317: the dashed line. Letting the transition from the interval $[-s,s]$
318: to the outside be a smooth one we can assume that arbitrarily many
319: derivatives of $V$ are bounded. 
320: \begin{figure}
321: \centering
322: \includegraphics[width=0.5\textwidth]{V.eps}
323: \caption{The original potential $V$ is drawn with the solid line. The
324:   modified potential $\tilde V$ coincides with $V$ in the interval
325:   $[-s,s]$ and is drawn with dashed lines outside that interval.}
326: \label{fig:V}
327: \end{figure}
328: When the function $V'$ is bounded in supremum-norm and the control,
329: $\alpha$, is bounded in $L^2(t_0,T;L^2)$, uniqueness of mild solutions
330: to \eqref{eq:flow} holds; see \cite{Cannarsa-Frankowska1}. For
331: starting positions $\varphi_0 \in H_0^1(0,1)$  it therefore holds that
332: mild solutions and weak solutions are the same, and for the analysis
333: either concept of solution may be used.
334: 
335: The running cost, $h$, corresponds to the action functional
336: (\ref{eq:actionfunctional}) as
337: \begin{equation}\label{eq:runningcost}
338: h(\alpha)=||\alpha||_{L^2(0,1)}^2/2,
339: \end{equation}
340: and the final cost is the squared $L^2$ distance from $\varphi_-$,
341: \begin{equation}\label{eq:finalcost}
342: g(\varphi)=K||\varphi-\varphi_-||_{L^2(0,1)}^2,
343: \end{equation}
344: where $K$ is a constant large enough to force $\varphi(T)\approx
345: \varphi_-$. We denote by $u$ the \emph{value function}, i.e.\ the best
346: possible value of (\ref{eq:costfunctional}) for each starting position
347: $(\varphi_0,t_0)$:
348: \begin{equation}\label{eq:valuefunction}
349: u(\varphi_0,t_0)=\inf \big\{v_{\varphi_0,t_0}(\alpha)\ :\ \alpha \in L^2\big(t_0,T;L^2(0,1)\big)\big\}
350: \end{equation}
351: 
352: \emph{Notation:} We henceforth let $||\cdot ||$ and $(\cdot,\cdot)$ be
353: the $L^2$ norm and inner product on $(0,1)$, and $|\cdot|$ be the
354: supremum norm on $\Re$.
355: The Dirac delta distribution will be denoted $\bar\delta$, as $\delta$
356: is used as the diffusivity constant.
357: 
358: \emph{Outline:} Section \ref{sec:Preliminaries} contains some facts
359: regarding the value function, which are applied in Section
360: \ref{sec:spatialdiscr} when the error from the spatial discretization
361: is established. In Section \ref{sec:timediscr} convergence of the time
362: discretization using the Symplectic Euler method is examined. Under a
363: reasonable condition, this method is shown to be convergent of order
364: one, with a constant independent of the spatial
365: discretization. Numerical results with examples of the convergence
366: rate for discretization in both space and time  is given in Section~ \ref{sec:NumRes}.  
367: \section{Preliminaries}\label{sec:Preliminaries}
368: This section contains some results which will be needed when the spatial
369: discretization error bound is established in Section
370: \ref{sec:spatialdiscr}. We start with a theorem about the boundedness
371: of optimal controls. 
372: \begin{thm}\label{thm:boundedcontrol}
373: For all $\varphi_0 \in H_0^1(0,1)$ and $0\leq t_0 \leq T$ the value
374: function $u$ satisfies
375: \begin{equation*}
376: u(\varphi_0,t_0)=\inf \big\{v_{\varphi_0,t_0}(\alpha)\ :\ ||\alpha||_{L^\infty(t_0,T;L^2)} \leq E ||\varphi_0||
377: + F\big\}
378: \end{equation*}
379: where the constants $E$ and $F$ depend on $\delta$, $K$, $\varphi_-$,
380: $T$, $|V'|$ and $|V''|$, but not on $\varphi_0$ and $t_0$.
381: \end{thm}
382: \begin{proof}
383: It is first shown that with $\alpha(t)= 0$, for all $t$, the state variable at
384: the terminal time, $\varphi(T)$, is bounded in $L^2$ by a constant which 
385: depends on the starting point $\varphi_0$. This can be done by taking the inner product
386: with $\varphi$ in (\ref{eq:flow}), using $||\varphi_x||^2 \geq
387: 0$, and noting that the function $t \mapsto ||\varphi(t)||^2$ is
388: absolutely continuous with 
389: $(\varphi,\varphi_t)=\frac{d}{dt}||\varphi||^2/2$ almost everywhere in
390: $[t_0,T]$. Hence
391: \begin{equation*}
392: \frac{d}{dt}||\varphi||^2/2 \leq -\delta^{-1} \big(V'(\varphi),\varphi\big)
393: \leq \delta^{-1}|V'| \cdot ||\varphi||,
394: \end{equation*} 
395: almost everywhere in $[t_0,T]$, and thereby
396: \begin{equation}\label{eq:fnd}
397: \frac{d}{dt}||\varphi|| \leq \delta^{-1}|V'|.
398: \end{equation}
399: By the fact that $\varphi_t$ is bounded in $L^2(t_0,T;L^2(0,1))$ (see
400: \cite{Evans}), it follows that the function $t \mapsto ||\varphi(t)||$
401: is absolutely continuous, and therefore \eqref{eq:fnd} implies that 
402: $||\varphi(T)||$ is bounded by $||\varphi_0|| + \delta^{-1}|V'|T$. Hence the final
403: cost, $g\big(\varphi(T)\big)$, is bounded in terms of the starting position:
404: \begin{multline*}
405: g\big(\varphi(T)\big) =K ||\varphi(T)-\varphi_-||^2 \leq 2K \big( ||\varphi(T)||^2
406: + ||\varphi_-||^2 \big) \\
407: \leq 4K||\varphi_0||^2 + 4K\delta^{-2}|V'|^2 T^2 +
408: 2K||\varphi_-||^2 =: M.
409: \end{multline*} 
410: It therefore holds that
411: \begin{equation*}
412: u(\varphi_0,t_0)=\inf \big\{v_{\varphi_0,t_0}(\alpha)\ :\
413: ||\alpha||_{L^2(t_0,T;L^2)}^2 \leq 2M \big\}.
414: \end{equation*}
415: For all $\alpha$ bounded by $\sqrt{2M}$ in $L^2(t_0,T;L^2)$ we have
416: that $\varphi(T)$ is bounded, again by taking the inner product with
417: $\varphi$ in (\ref{eq:flow}):
418: \begin{equation*}
419: \half \frac{d}{dt}||\varphi||^2 \leq \delta^{-1}|V'| \cdot ||\varphi|| +
420: ||\alpha||\cdot ||\varphi||,
421: \end{equation*}
422: which implies
423: $\frac{d}{dt} ||\varphi|| \leq \delta^{-1}|V'| + ||\alpha||$
424: and so
425: \begin{multline*}
426: ||\varphi(T)|| \leq ||\varphi_0|| + \delta^{-1}|V'|T + \int_{t_0}^T
427:   ||\alpha||dt \\
428: \leq ||\varphi_0|| + \delta^{-1}|V'|T +
429:   \sqrt{T}||\alpha||_{L^2(t_0,T;L^2)} \\
430: \leq ||\varphi_0|| +
431:   \delta^{-1}|V'|T +\sqrt{2T}\sqrt{M} \leq E ||\varphi_0|| +F,
432: \end{multline*}
433: for some constants $E$ and $F$ which do not depend on $\varphi_0$ and $t_0$.
434: 
435: 
436: It shall now  be proved that changing the control $\alpha$ a small
437: amount changes the state $\varphi$ a small amount. We shall therefore
438: compare two solutions, $\varphi^1$ and $\varphi^2$, both starting at
439: $(\varphi_0,t_0)$, such that $\varphi^1$ solves (\ref{eq:flow}) with control
440: $\alpha^1$ and $\varphi^2$ with control $\alpha^2$. Subtract the two
441: evolution equations and take the inner product with $\varphi^1-\varphi^2$ to
442: obtain
443: \begin{multline*}
444: \half\frac{d}{dt}||\varphi^1-\varphi^2||^2+\delta||\varphi^1_x-\varphi^2_x||^2
445: = \\
446: \delta^{-1}(-V'(\varphi^1)+V'(\varphi^2),\varphi^1-\varphi^2)+(\alpha^1-\alpha^2,\varphi^1-\varphi^2)
447: \end{multline*}
448: which, by the boundedness of $V''$, entails
449: \begin{equation*}
450: \frac{d}{dt}||\varphi^1-\varphi^2|| \leq \delta^{-1}|V''|\cdot ||\varphi^1-\varphi^2||+||\alpha^1-\alpha^2||.
451: \end{equation*}
452: By Gr\"onwall's lemma we therefore have that
453: \begin{equation*}
454: ||\varphi^1(T)-\varphi^2(T)|| \leq \exp (\delta^{-1} |V''|T)
455:   \int_{t_0}^T ||\alpha^1-\alpha^2||dt,
456: \end{equation*}
457: so, provided $\alpha^1$ and $\alpha^2$ are both bounded by $\sqrt{2M}$
458: in $L^2(t_0,T;L^2)$, the difference in terminal cost has the following bound:
459: \begin{equation}\label{eq:gdifference}
460: \begin{split}
461: |g\big(\varphi^1(T)\big)-g\big(\varphi^2(T)\big)| &= K \big|
462:  ||\varphi^1(T)-\varphi_-||^2-||\varphi^2(T)-\varphi_-||^2\big| \\
463: &=K\big|(\varphi^1(T)+\varphi^2(T)-2\varphi_-,\varphi^1(T)-\varphi^2(T))\big| \\
464: & \leq 2K(E||\varphi_0||+F+||\varphi_-||) \cdot ||\varphi^1(T)-\varphi^2(T)|| \\
465: & \leq R \int_{t_0}^T ||\alpha^1-\alpha^2||dt, \\
466: \intertext{where}
467: R&=2K \exp (\delta^{-1} |V''|T)(E||\varphi_0||+F+||\varphi_-||) \\
468: &=:
469: E'||\varphi_0|| +F',
470: \end{split}
471: \end{equation}
472: with the constants $E'$ and $F'$ independent of $\varphi_0$ and $t_0$.
473: Let now $\alpha^1$ be a control bounded by $\sqrt{2M}$ in $L^2(t_0,T;L^2)$ and
474: let $\alpha^2$ be a modification of $\alpha^1$:
475: \begin{equation*}
476: \alpha^2(t)=
477: \begin{cases}
478: \alpha^1(t), & \text{for all $t$ such that $||\alpha^1(t)|| \leq 2R$}, \\
479: 0, & \text{otherwise.}  
480: \end{cases}
481: \end{equation*}
482: The difference in the terminal cost thus has the  bound
483: \begin{equation*}
484: |g(\varphi^1(T))-g(\varphi^2(T))| \leq R \int_{\{t:||\alpha^1(t)||>2R \}}
485:  ||\alpha^1|| dt,
486: \end{equation*}
487: while the difference in running cost is 
488: \begin{equation*}
489: \int_{\{t:||\alpha^1(t)||>2R \}} \frac{||\alpha^1||^2}{2} dt \geq
490: R\int_{\{t:||\alpha^1(t)||>2R \}} ||\alpha^1||dt,
491: \end{equation*}
492: so
493: \begin{equation*}
494: v_{\varphi_0,t_0}(\alpha^2) \leq v_{\varphi_0,t_0}(\alpha^1).
495: \end{equation*}
496: Hence for any control bounded in $L^2(t_0,T;L^2)$ there is another
497: control, bounded by $2E'||\varphi_0||+2F'$ in $L^\infty(t_0,T;L^2)$, which gives a smaller or equal value functional.
498: \end{proof}
499: With Theorem \ref{thm:boundedcontrol} the theory in
500: \cite{Cannarsa-Frankowska1} may be used, which establishes existence
501: of optimal controls to $u$ in \eqref{eq:valuefunction}. We state this
502: in a corollary.
503: \begin{cor}\label{cor:existence}
504: For each $(\varphi_0,t_0) \in H_0^1(0,1) \times [0,T]$ there exists a
505: minimizer $\alpha$, bounded in $L^\infty(t_0,T;L^2)$, in \eqref{eq:valuefunction}.
506: \end{cor}  
507: \begin{proof}
508: Use Theorem \ref{thm:boundedcontrol} and \cite{Cannarsa-Frankowska1}.
509: \end{proof}
510: Theorem \ref{thm:boundedcontrol} is also used when proving Theorem
511: \ref{thm:semiconcave} about semiconcavity. In
512: \cite{Cannarsa-Frankowska2} a theorem on semiconcavity on $L^2(0,1)
513: \times [0,T)$ is established. This result could have been used in this
514:     paper, but as only the weaker result of semiconcavity on
515:     $H^1_0(0,1) \times [0,T]$ is needed for our purposes, an easier
516:     proof is given for this case.
517: \begin{thm}\label{thm:semiconcave}
518: The restriction of the value function, $u$,  to $H_0^1\times[0,T]$ is semiconcave.
519: \end{thm}
520: \begin{proof}
521: It will be shown that for every constant $B$, every closed interval $I
522: \subset [0,T)$, and all starting positions
523: $(\varphi_0^1,t^1)$ and $(\varphi_0^2,t^2)$ with
524: $||\varphi_0^1||_{H_0^1(0,1)} +||\varphi_0^2||_{H_0^1(0,1)} \leq B$
525:   and $t^1,t^2 \in I$, there is a constant C such that 
526: \begin{equation*}
527: u(\varphi^1_0,t^1)+u(\varphi^2_0,t^2)-2u\Big(\frac{\varphi^1_0+\varphi^2_0}{2},\frac{t^1+t^2}{2}\Big)
528: \leq C (||\varphi^1_0-\varphi^2_0||^2_{H_0^1}+|t^1-t^2|^2).
529: \end{equation*}
530: In order to
531: keep  constants simple we use that $u$ may be defined in $H_0^1(0,1)
532: \times (-\infty, T]$, so that we may set $t^1=h$ and $t^2=-h$, and realize that the
533: result for other times follows analogously. In this proof we let $C$
534: be any constant which may depend on $B$.
535: 
536: Let $\alpha:[0,T]\rightarrow L^2$ be an optimal control for the cost
537: functional $v_{\frac{\varphi^1_0+\varphi^2_0}{2},0}$ defined in
538: \eqref{eq:costfunctional}, and let $\varphi^3:[0,T]\rightarrow H_0^1$ be
539: the corresponding state. 
540: Define controls for solutions starting in $(\varphi^1_0,h)$ and
541: $(\varphi^2_0,-h)$ by dilations of $\alpha$ as
542: \begin{align*}
543: \alpha^1(t)&=\alpha\Big(T\frac{t-h}{T-h}\Big), \\
544: \alpha^2(t)&=\alpha\Big(T\frac{t+h}{T+h}\Big), \\
545: \end{align*}
546: and let the corresponding states be denoted $\varphi^1:[h,T]\rightarrow H_0^1$ with $\varphi^1(h)=\varphi^1_0$ and
547: $\varphi^2:[-h,T]\rightarrow H_0^1$ with $\varphi^2(-h)=\varphi^2_0$.
548: The evolution equation \eqref{eq:flow} for $\varphi^1$ and $\varphi^2$ is
549: now transformed to the interval $[0,T]$. The following equations are
550: thereby obtained:
551: \begin{align*}
552: \varphi^1_t &= \frac{T-h}{T}(\delta \varphi^1_{xx}-\delta^{-1} V'(\varphi^1)+\alpha), \quad
553: \varphi^1(0)=\varphi^1_0, \\
554: \varphi^2_t &= \frac{T+h}{T}(\delta\varphi^2_{xx}-\delta^{-1}V'(\varphi^2)+\alpha), \quad
555: \varphi^2(0)=\varphi^2_0, \\
556: \varphi^3_t &= \delta\varphi^3_{xx}-\delta^{-1}V'(\varphi^3)+\alpha, \quad
557: \varphi^3(0)=\frac{\varphi^1_0+\varphi^2_0}{2}. \\
558: \end{align*}
559: 
560: The function 
561: \begin{equation*}
562: z(t)=\varphi^1(t)+\varphi^2(t)-2\varphi^3(t)
563: \end{equation*}
564: is now introduced. We will obtain a bound for $||z(T)||$. 
565: The equation solved by $z$ is
566: \begin{multline}\label{eq:z}
567: z_t=\delta z_{xx}-\delta^{-1}\big(
568: V'(\varphi^1)+V'(\varphi^2)-2V'(\varphi^3)\big)\\
569: +\frac{\delta h}{T}(\varphi^2-\varphi^1)_{xx}
570: +\frac{\delta h}{T}\big( V'(\varphi^1)-V'(\varphi^2)\big).
571: \end{multline}
572: It is therefore necessary to find a bound for $\varphi^1-\varphi^2$. The
573: evolution equation for $\varphi^1-\varphi^2$ is
574: \begin{multline}\label{eq:phidiff}
575: (\varphi^1-\varphi^2)_t=\delta (\varphi^1-\varphi^2)_{xx}-\delta^{-1}\big(V'(\varphi^1)-V'(\varphi^2)\big)\\
576: -\frac{\delta h}{T}(\varphi^1+\varphi^2)_{xx}
577:   + \frac{\delta h}{T}\big(V'(\varphi^1)+V'(\varphi^2)\big) -\frac{2h}{T}\alpha.
578: \end{multline}
579: After the inner product is taken with $\varphi^1-\varphi^2$ the following
580: inequality is obtained:
581: \begin{multline*}
582: \half \frac{d}{dt}||\varphi^1-\varphi^2||^2 \leq \delta^{-1}|V''| \cdot
583: ||\varphi^1-\varphi^2||^2 + \frac{\delta h}{T} ||\varphi^1_{xx}+\varphi^2_{xx}||\cdot
584: ||\varphi^1-\varphi^2|| \\
585: + \frac{\delta^{-1} h}{T} ||V'(\varphi^1)+V'(\varphi^2)|| \cdot
586: ||\varphi^1-\varphi^2|| + \frac{2h}{T} ||\alpha||\cdot ||\varphi^1-\varphi^2||,
587: \end{multline*}
588: and hence
589: \begin{multline*}
590: \frac{d}{dt}||\varphi^1-\varphi^2|| \leq \delta^{-1}|V''| \cdot
591: ||\varphi^1-\varphi^2|| + \frac{\delta h}{T}
592: ||\varphi^1_{xx}+\varphi^2_{xx}|| \\
593: + \frac{\delta^{-1} h}{T} ||V'(\varphi^1)+V'(\varphi^2)|| + \frac{2h}{T} ||\alpha||.
594: \end{multline*}
595: Thus, by Gr\"onwall's Lemma,
596: \begin{multline*}
597: ||\varphi^1(t)-\varphi^2(t)|| \leq
598:   e^{\delta^{-1}|V''|T}||\varphi^1(0)-\varphi^2(0)||+\\
599: e^{\delta^{-1}|V''|T}\frac{h}{T} \int_0^T
600:   \big(\delta||\varphi^1_{xx} + \varphi^2_{xx}|| + \delta^{-1}||V'(\varphi^1)+V'(\varphi^2)|| + 2||\alpha||\big)dt.
601: \end{multline*}
602: Since $||\varphi_0^1||_{H_0^1(0,1)}+||\varphi_0^2||_{H_0^1(0,1)} \leq B$ it follows that 
603: $\varphi^1$ and $\varphi^2$ are bounded by a constant $C$ in
604: $L^2(0,T;H^2)$;
605: see \cite{Evans}. Together with the fact that $V'$ is bounded this
606: implies that 
607: \begin{equation}\label{eq:phidrift}
608: ||\varphi^1(t)-\varphi^2(t)|| \leq C(||\varphi^1_0-\varphi^2_0|| + h), \quad
609:   \text{for all } 0 \leq t \leq T.
610: \end{equation}
611: Equation \eqref{eq:phidiff} is now used once again
612: together with the fact that $|V'(\varphi^1)-V'(\varphi^2)| \leq
613: |V''|\cdot|\varphi^1-\varphi^2|$ and Theorem 5 on page 360 in \cite{Evans}, to
614: draw the conclusion that
615: \begin{equation}\label{eq:H1difference}
616: \esssup_{0 \leq t \leq T}||\varphi^1(t)-\varphi^2(t)||_{H_0^1} + ||\varphi^1_{xx}-\varphi^2_{xx}||_{L^2(0,T;L^2)} \leq
617:   C(||\varphi^1_0-\varphi^2_0||_{H_0^1} + h).
618: \end{equation}
619: There is also the term $V'(\varphi^1)+V'(\varphi^2)-2V'(\varphi^3)$ in
620: \eqref{eq:z}. This can be handled as
621: \begin{multline}\label{eq:Vsplit}
622: |V'(\varphi^1)+V'(\varphi^2)-2V'(\varphi^3)|\\
623:  \leq
624:  |V'(\varphi^1)+V'(\varphi^2)-2V'(\frac{\varphi^1+\varphi^2}{2})| +
625:  2|V'(\frac{\varphi^1+\varphi^2}{2}) - V'(\varphi^3)| \\
626: \leq \frac{|V'''|}{2}
627:  |\varphi^1-\varphi^2|^2 + |V''|\cdot |z|.
628: \end{multline}
629: We are now ready to take the inner product with $z$ in \eqref{eq:z} to
630: obtain
631: \begin{multline*}
632: \half\frac{d}{dt} ||z||^2 \leq \frac{\delta^{-1}|V'''|}{2} \int_0^1
633: (\varphi^1-\varphi^2)^2|z|dx + \delta^{-1}|V''|\cdot||z||^2 \\
634: +
635: \frac{\delta h}{T}||\varphi^1_{xx}-\varphi^2_{xx}|| \cdot ||z|| +
636: |V''| \frac{\delta^{-1} h}{T}
637: ||\varphi^1-\varphi^2|| \cdot ||z||,
638: \end{multline*}
639: which implies
640: \begin{multline*}
641: \frac{d}{dt} ||z|| \leq \delta^{-1}|V''|\cdot||z|| +
642: \frac{\delta^{-1}|V'''|}{2}||\varphi^1-\varphi^2||_{L^4(0,1)}^2 \\
643: + \frac{\delta
644:   h}{T}||\varphi^1_{xx}-\varphi^2_{xx}||  + |V''| \frac{\delta^{-1} h}{T}
645: ||\varphi^1-\varphi^2||.
646: \end{multline*}
647: By Gr\"onwall's Lemma
648: \begin{multline}\label{eq:zdrift}
649: ||z(T)|| \leq e^{\delta^{-1}|V''|T} \int_0^T
650: \big(\frac{\delta^{-1}|V'''|}{2}||\varphi^1-\varphi^2||_{L^4(0,1)}^2 \\
651: + \frac{\delta h}{T}||\varphi^1_{xx}-\varphi^2_{xx}||  + |V''| \frac{\delta^{-1}h}{T}
652: ||\varphi^1-\varphi^2||)dt. 
653: \end{multline}
654: Sobolev's inequality gives that $||\varphi^1-\varphi^2||_{L^4(0,1)}
655: \leq C||\varphi^1-\varphi^2||_{H^1_0(0,1)}$, so \eqref{eq:H1difference}
656: together with \eqref{eq:zdrift}
657: implies that 
658: \begin{equation*}
659: ||z(T)|| \leq C \big( ||\varphi^1_0-\varphi^2_0||^2_{H^1_0(0,1)} + h^2 \big).
660: \end{equation*}
661: This fact is now used to show that 
662: \begin{equation}\label{eq:vconcave}
663: v_{\varphi^1_0,h}(\alpha^1)+v_{\varphi^2_0,-h}(\alpha^2)-2v_{\frac{\varphi^1_0+\varphi^2_0}{2},0}(\alpha)
664: \leq C \big(||\varphi^1_0-\varphi^2_0||^2 + h^2 \big),
665: \end{equation}
666: The terminal cost is treated first. We use
667: the notation $\varphi_T \equiv \varphi(T)$ and perform a simple rearrangement:
668: \begin{multline}\label{eq:terminalcosts}
669: |g(\varphi^1_T)+g(\varphi^2_T)-2g(\varphi^3_T)| \\ 
670: = 
671: \big| \frac{K}{2}||\varphi^1_T -\varphi^2_T||^2 + K \big(\frac{\varphi^1_T+\varphi^2_T}{2}+
672: \varphi^3_T -2 \varphi_-, \varphi^1_T+\varphi^2_T-2\varphi^3_T\big) \big| \\
673: \leq C \big(||\varphi^1_0-\varphi^2_0||^2_{H^1_0} + h^2 \big),
674: \end{multline}
675: where \eqref{eq:phidrift}, \eqref{eq:zdrift}, and the fact that
676: $\varphi^1_T$, $\varphi^2_T$ and $\varphi^3_T$, are bounded are used. The
677: running costs must also be treated. A simple calculation shows that
678: \begin{equation}\label{eq:runningcosts}
679: \int_h^T ||\alpha^1||^2 dt + \int_{-h}^T ||\alpha^2||^2 dt -2 \int_0^T
680: ||\alpha||^2 dt =0.
681: \end{equation}
682: The desired result \eqref{eq:vconcave} follows from
683: \eqref{eq:terminalcosts} and \eqref{eq:runningcosts}.
684: \end{proof}
685: \section{Discretization in space}\label{sec:spatialdiscr}
686: We shall compare the value functions associated with our original
687: problem and a finite element approximation. The value function we want
688: to approximate is
689: $u$ defined in \eqref{eq:valuefunction}.
690: %\begin{equation}\label{eq:exactu}
691: %u(\varphi_0,t)=\inf_{\alpha \in L^2(t,T;L^2(0,1))}\big( g(\varphi(T))+
692: %\int_t^T h(\alpha)ds \big), \quad \varphi(t)=\varphi_0,
693: %\end{equation}
694: %where $\varphi$ is the mild solution to 
695: %\begin{equation}\label{eq:phievol}
696: %\varphi_t=\varphi_{xx}-V'(\varphi)+\alpha, \quad \varphi(t,0)=\varphi(t,1)=0.
697: %\end{equation}
698: The approximate value function is, similarly as in \eqref{eq:valuefunction},
699: \begin{equation}\label{eq:approxu}
700: \bar u (\bar\varphi_0,t_0)=\inf_{\bar\alpha \in L^2(t_0,T;V)} \Big\{ g(\bar \varphi(T))+
701: \int_t^T h(\bar \alpha)ds\ :\   \bar\varphi(t_0)=\bar\varphi_0\Big\},
702: \end{equation}
703: where $\bar\varphi \in C(t_0,T;V)$ solves
704: \begin{equation}\label{eq:approxphievol}
705: (\bar\varphi_t,v)=-\delta (\bar\varphi_x,v_x)+(-\delta^{-1}
706:   V'(\bar\varphi)+\bar\alpha,v), \quad
707:   \text{for all } v \in V,
708: \end{equation}
709: and $V$ is the space of continuous piecewise linear functions on $[0,1]$ which
710: are zero at $0$ and $1$ and linear on the intervals $(0,\Delta x)$,
711: $(\Delta x, 2\Delta x)$, and so on. We note that the infima in
712: (\ref{eq:valuefunction}) and (\ref{eq:approxu}) are attained, using
713: Corollary \ref{cor:existence} for the original problem
714: \eqref{eq:valuefunction} and the easier theory in
715: \cite{Cannarsa-Sinestrari} for the approximation problem
716: \eqref{eq:approxu}. Therefore  we
717: can replace the infima with minima. 
718: The same sort of convergence analysis which is presented here is
719: performed for problems of optimal design in \cite{Carlsson-Sandberg-Szepessy}.
720: 
721: We now introduce some
722: notation needed in Theorem \ref{thm:valueerror}. We denote by
723: $P$ the $L^2$
724: projection from $L^2(0,1)$ to $V$ or from $L^2(0,1)\times\Re$ to $V
725: \times \Re$. Let $\Omega$ be an open subset of a Hilbert space $X$,
726: and $z:\Omega \rightarrow \Re$. For any $x_0 \in \Omega$ the
727: superdifferential $D^+ z(x_0)$ is defined as follows:
728: \begin{equation*}
729: D^+ z(x_0) = \Big\{ p \in X \big| \limsup_{x \rightarrow x_0}
730:   \frac{z(x)-z(x_0)-(p,x-x_0)_X}{|x-x_0|_X} \leq 0 \Big\}.
731: \end{equation*} 
732: The Hamiltonian, $H$, for the optimal
733: control problem \eqref{eq:valuefunction} is given by
734: \begin{equation}\label{eq:Hamiltonian}
735: H(\lambda,\varphi)=-\delta(\lambda_x,\varphi_x) -\delta^{-1}\big(\lambda,V'(\varphi)\big)-||\lambda||^2/2,
736: \end{equation} 
737: for all $\lambda$, $\varphi \in H_0^1(0,1)$. The restrictions of $u$ to
738: the subspaces $V \times [0,T]$ and $H_0^1 \times [0,T]$ are denoted
739: $u_V$ and $u_H$.
740: \begin{thm}\label{thm:valueerror}
741: Let $\varphi_0 \in V$. 
742: Denote an optimal pair
743: (control and state) for $u(\varphi_0,t_0)$ by $\alpha:[t_0,T]\rightarrow
744: L^2$ and $\varphi:[t_0,T]\rightarrow L^2$ and an optimal pair
745: for $\bar u(\varphi_0,t_0)$ by $\bar\alpha:[t_0,T]\rightarrow V$ and
746: $\bar\varphi:[t_0,T]\rightarrow V$.
747: Then
748: \begin{multline}\label{eq:errorrepr}
749: \int_{t_0}^T \Big(p^*_t(s)+H\big(P p^*_\varphi(s), \bar\varphi(s)\big)\Big)ds \\
750: \leq \bar u(\varphi_0,t_0)-u(\varphi_0,t_0) \\
751: \leq g\big(P \varphi(T)\big)-g\big(\varphi(T)\big)+
752: \int_{t_0}^T \Big(H\big(p^\#_\varphi(s),P \varphi(s)\big)-H\big(p^\#_\varphi(s),\varphi(s)\big)\Big)ds
753: \end{multline}
754: where $p^*(s)=\big(p^*_\varphi(s),p^*_t(s)\big) \in L^2(0,1)\times \Re$
755: is any measurable function with values in $D^+ u_H(\bar\varphi(s),s)$, and
756: $p^\#(s)=\big(p^\#_\varphi(s),p^\#_t(s)\big) \in V\times \Re$ is any
757: measurable function with values  in
758: $D^+ \bar u(P \varphi(s),s)$.
759: \end{thm}
760: \begin{proof}
761: We divide the proof into two steps: In \emph{Step 1} we obtain a lower bound for
762: $\bar u(\varphi_0,t_0)-u(\varphi_0,t_0)$, and in \emph{Step 2} we do
763: likewise for $u(\varphi_0,t_0)-\bar u(\varphi_0,t_0)$. 
764: 
765: \emph{Step 1.} \quad Using the definitions (\ref{eq:valuefunction}) and
766: (\ref{eq:approxu}) for $u$ and $\bar u$, the fact that
767: $\bar u \big(\bar\varphi(T),T\big)=g\big(\bar\varphi(T)\big)$, and that $u_H$ is the
768: restriction of $u$ to $H_0^1 \times [0,T]$, we can write 
769: \begin{equation}\label{eq:udiff1}
770: \begin{split}
771: \bar u(\varphi_0,t_0)-u(\varphi_0,t_0)&=u_H\big(\bar\varphi(T),T\big)-u_H\big(\bar\varphi(t_0),t_0\big)+\int_{t_0}^T
772: h(\bar\alpha)ds \\
773: &=\int_t^T \Big(\frac{d}{ds} u_H\big(\bar\varphi(s),s\big)+h\big(\bar\alpha(s)\big)\Big)ds,
774: \end{split}
775: \end{equation}
776: since $u_H(\bar\varphi(s),s)$ is absolutely continuous as $u$ is
777: locally Lipschitz continuous (see \cite{Cannarsa-Frankowska2}) and $\bar\varphi$ is absolutely continuous
778: as a function of $s$.
779: 
780: We now use that $u_H$ is a semiconcave function (with linear modulus),
781: so that for every $p \in D^+ u_H(z_0)$ there exists a constant $K$
782: such that
783: \begin{equation}\label{eq:semiconmodulus}
784: u_H(z)-u_H(z_0)- (p,z-z_0 ) \leq K |z-z_0|^2
785: \end{equation}
786: for all $z$ in a neighborhood of $z_0 \in H_0^1(0,1) \times (0,T)$; see \cite{Cannarsa}. Let now $p^*(s)=\big(p^*_\varphi(s),p^*_t(s)\big)$ be any element
787: in $D^+u_H\big(\bar\varphi(s),s\big) \cap \big(L^2(0,1)\times \Re\big)$.
788: Consider a point $s$ where the derivative
789: $\bar\varphi_t(s)$ exists. 
790: A lower bound for the backward derivative of $u_H\big(\bar\varphi(s),s\big)$ will
791: now be obtained. We split the difference quotient approximating the
792: backward derivative at $s$:
793: \begin{align*}
794: &\frac{u_H(\bar\varphi(s),s)-u_H(\bar\varphi(s-\Delta s),s-\Delta s)}{\Delta s}
795: \\
796: =& -\big[ u_H\big(\bar\varphi(s-\Delta s),s-\Delta
797:   s\big)-u_H\big(\bar\varphi(s),s\big)\\
798: &-p^*_t(s)(-\Delta s) -
799:   \big(p^*_\varphi(s),\bar\varphi(s-\Delta s)-\bar\varphi(s)\big)\big]/{\Delta s}\\
800: & +p^*_t(s) + \Big(p^*_\varphi(s),\frac{\bar\varphi(s)-\bar\varphi(s-\Delta s)}{\Delta
801: s}\Big).
802: \end{align*}
803: If equation \eqref{eq:semiconmodulus} is used together with the
804: fact that $\bar\varphi$ is differentiable at $s$ it can be deduced that
805: the quotient involving the square bracket in the above equation is greater than or equal to $-K'
806: \Delta s$, for some constant $K'$. Letting $\Delta s \rightarrow 0$ we
807: see that 
808: \begin{equation*}
809: \frac{d}{ds}u_H \big(\bar\varphi(s),s\big) \geq p^*_t(s) +\big(p^*_\varphi(s),\bar\varphi_t(s)\big),
810: \end{equation*} 
811: where (temporarily) $d/ds$ denotes the backward derivative.
812: In order to be able to apply (\ref{eq:approxphievol}) we
813: note that $\bar\varphi_t \in V$ implies
814: \begin{equation*}
815: (p^*_\varphi,\bar\varphi_t)=(P p^*_\varphi,\bar\varphi_t).
816: \end{equation*}
817: Thus the integrand in
818: (\ref{eq:udiff1}), using the backward derivative,  can be bounded from
819: below as follows:
820: \begin{multline*}
821: \frac{d}{ds} u_H\big(\bar\varphi(s),s\big)+h\big(\bar\alpha(s)\big) 
822: \geq p^*_t(s)+\big(\bar\varphi_t(s),P p^*_\varphi(s)\big)+h\big(\bar\alpha(s)\big) \\
823: = p^*_t(s)-\delta\big(\bar\varphi_{x}(s),(P
824: p^*_\varphi(s))_x\big)-\delta^{-1}\Big(V'\big(\bar\varphi(s)\big),P p^*(s)\Big) \\
825: +\big(\bar\alpha(s), P p^*_\varphi(s)\big) +
826: \half ||\bar\alpha(s)||^2 \\
827: \geq p^*_t(s)+ H\big(P p^*_\varphi(s),\bar\varphi(s)\big),
828: \end{multline*}
829: since
830: \begin{equation*}
831: H(\lambda,\varphi)=\min_{\alpha \in L^2(0,1)} \Big(
832: -\delta(\varphi_x,\lambda_x)
833: -\delta^{-1}\big(V'(\varphi),\lambda\big)+(\alpha,\lambda) + \half ||\alpha||^2\Big).
834: \end{equation*}
835: The double sided and the
836: backward time derivatives of $u_H(\bar\varphi(s),s)$ differ on a set
837: of measure zero, so there is no problem in using the backward
838: derivative in (\ref{eq:udiff1}).
839: %Since $u$ is semiconcave $u_V$ is semiconcave as well, and thus
840: %its directional derivative exists in every
841: %direction at every point. 
842: %Consider a point $s$ where the derivative $\bar\varphi_t(s)$ exists. We
843: %split the difference quotient approximating the backward derivative at
844: %$s$:
845: %\begin{align*}
846: %& - \frac{u_V(\bar\varphi(s-\Delta s),s-\Delta
847: %  s)-u_V(\bar\varphi(s),s)}{\Delta s} \\ 
848: %& \quad= -\frac{u_V(\bar\varphi(s-\Delta
849: %  s),s-\Delta s)-u_V(\bar\varphi(s)-\Delta s \bar\varphi_t(s),s-\Delta
850: %  s)}{\Delta s}  \\
851: %& \quad \ \ \ - \frac{u_V(\bar\varphi(s)-\Delta s \bar\varphi_t(s),s-\Delta
852: %  s)-u_V(\bar\varphi(s),s)}{\Delta s} \equiv I+II.
853: %\end{align*}
854: %Part $I$ in the above equation converges to zero when $\Delta s
855: %\rightarrow 0$ since $u_V$ is locally Lipschitz continuous. Part $II$
856: %converges to the negative directional derivative in the direction 
857: %$(-\bar\varphi_t,-1)$. If we, temporarily, let $d/ds$ denote the backward
858: %derivative we therefore have that
859: %\begin{equation*}
860: %\frac{u\arrowvert_V(\bar\varphi(s+\Delta s),s+\Delta
861: %  s)-u\arrowvert_V(\bar\varphi(s),s)}{\Delta s} = \partial
862: %  u\arrowvert_V(\bar\varphi(s),s;( \frac{\bar\varphi(s+\Delta s)-\bar\varphi(s)}{\Delta s},
863: %  1)) + o(1),
864: %\end{equation*}
865: %where $\partial u\arrowvert_V$ denotes the directional
866: %derivative. Since $u\arrowvert_V$ is locally Lipschitz we therefore
867: %obtain the forward derivative, when $\bar\varphi_t$ exists, as
868: %\begin{multline}\label{eq:forwardder}
869: %\frac{d}{ds} u_V(\bar\varphi(s),s) =- \partial
870: %  u_V(\bar\varphi(s),s;(-\bar\varphi_t,-1)) \\
871: %=- \min_{p \in D^+
872: %  u_V(\bar\varphi(s),s)}(p,(-\bar\varphi_t(s),-1))
873: %=\max_{p \in D^+
874: %  u_V(\bar\varphi(s),s)}(p,(\bar\varphi_t(s),1))\\
875: % \geq (Pp^*(s),(\bar\varphi_t(s),1) =(p^*(s),(\bar\varphi_t(s),1))
876: %\end{multline}
877: %where the second equality  follows from (\cite{Cannarsa-Sinestrari}), the inequality
878: %by Lemma \ref{lem:projection} and the last
879: %equality by $\bar\varphi_t \in V$. The double sided and the
880: %backward time derivatives of $u\arrowvert_V(\bar\varphi(s),s)$ differ on a set
881: %of measure zero, so there is no problem in using the backward
882: %derivative in (\ref{eq:udiff1}).
883: %
884: %We let $p^*=(p^*_\varphi,p^*_t)$ (i.e.\ $p^*_\varphi$ is the ``spatial'' component
885: %of $p$, and $p^*_t$ is the time component) be the element in $D^+
886: %u(\bar\varphi(s),s)$ for which the minimum in (\ref{eq:forwardder})  is
887: %attainded. In order to be able to apply (\ref{eq:approxphievol}) we
888: %note that 
889: %\begin{equation*}
890: %(\bar\varphi_t,p^*_\varphi)=(\bar\varphi_t,P p^*_\varphi).
891: %\end{equation*}
892: %Thus the integrand in
893: %(\ref{eq:udiff1}), using the backward derivative,  becomes
894: %\begin{multline*}
895: %p^*_t-\delta(\bar\varphi_{x},(P
896: %p^*_\varphi(s))_x)-\delta^{-1}(V'(\bar\varphi),P p^*(s))+(\bar\alpha, P p^*_\varphi(s)) +
897: %\half ||\bar\alpha||^2 \\
898: %\geq H(P p^*_\varphi(s),\bar\varphi)-H(p^*_\varphi(s), \bar\varphi),
899: %\end{multline*}
900: %since by Corollary 5.7 in \cite{Cannarsa-Frankowska2} 
901: %\begin{equation*}
902: %p_t + H(p_\varphi,\varphi) = 0, \quad \text{for all } (p_t,p_\varphi)\in D^*u(\varphi_0,t_0),
903: %\end{equation*}
904: %when $t_0 \in [0,T]$ and $\varphi_0 \in H^1_0$, and 
905: %\begin{equation*}
906: %H(\lambda,\varphi)=\min_{\alpha \in L^2(0,1)} \big(
907: %-\delta(\lambda_x,\varphi_x)
908: %-\delta^{-1}(\lambda,V'(\varphi))+(\lambda,\alpha) + \half ||\alpha||^2\big).
909: %\end{equation*}
910: 
911: \emph{Step 2.} Lower bound for $u(\varphi_0,t_0)-\bar u(\varphi_0,t_0)$. It is
912: now assumed that $\varphi_0 \in V$. Similarly as in \emph{Step 1} we
913: write, noting that $\bar u$ is only defined on $V\times [0,T]$,
914: \begin{equation}\label{eq:udiffg}
915: \begin{split}
916: & u(\varphi_0,t_0)-\bar u(\varphi_0,t_0) \\
917: &= g\big(\varphi(T)\big)-g\big(P \varphi(T)\big) +
918: \bar u\big(P \varphi(T),T\big)-\bar u\big(P \varphi(t_0),t_0\big) +
919: \int_{t_0}^T h\big(\alpha(s)\big)ds \\
920: &=g\big(\varphi(T)\big)-g\big(P \varphi(T)\big) +\int_{t_0}^T \Big(\frac{d}{ds} \bar
921: u\big(P \varphi(s),s\big) +h\big(\alpha(s)\big)\Big)ds =: I + II.
922: \end{split}
923: \end{equation}
924: A lower bound for part $II$ is  obtained by splitting the
925: difference quotient approximating the backward derivative at $s$: 
926: \begin{equation}\label{eq:bdsplit}
927: \begin{split}
928: &\frac{\bar u\big(P \varphi(s),s\big)-\bar u\big(P \varphi(s-\Delta s),s-\Delta s\big)}{\Delta s}
929: \\
930: =& -\big[\bar u\big(P \varphi(s-\Delta s),s-\Delta
931:   s\big)-\bar u\big(P \varphi(s),s\big)\\
932: &-p^\#_t(s)(-\Delta s) -
933:   \big(p^\#_\varphi(s),P \varphi(s-\Delta s)-P \varphi(s)\big)\big]/{\Delta s}\\
934: & +p^\#_t(s) + \Big(p^\#_\varphi(s),\frac{P \varphi(s)-P \varphi(s-\Delta s)}{\Delta
935: s}\Big).
936: \end{split}
937: \end{equation}
938: The derivative $\varphi_t(s)$ exists for $t_0 < s <T$ by the theory in
939: e.g.\ Chapter 3 in \cite{Henry}, where we have used also that the
940: control, $\alpha=-\lambda$, solves an adjoint backward parabolic PDE,
941: and therefore is H\"older continuous. It is now used that $||P x||
942: \leq ||x||$, $\bar u$ is semiconcave (see e.g.\ \cite{Cannarsa-Sinestrari}), and that
943: \begin{equation*}
944: \Big(p^\#_\varphi(s),\frac{P \varphi(s)-P \varphi(s-\Delta s)}{\Delta
945: s}\Big) = \Big(p^\#_\varphi(s),\frac{\varphi(s)-\varphi(s-\Delta s)}{\Delta
946: s}\Big)
947: \end{equation*}
948: in  equation \eqref{eq:bdsplit}, so that we have, similarly as in \emph{Step 1}, that 
949: \begin{equation*}
950: \frac{d}{ds}\bar u (P \varphi(s),s) \geq p^\#_t(s) +(p^\#_\varphi(s),\bar\varphi_t(s)).
951: \end{equation*}
952: By further using Chapter 3 in \cite{Henry} it is known that equation
953: \eqref{eq:flow} is satisfied in the $L^2$ sense, with $\varphi(s) \in
954: H^2(0,1) \cap H_0^1(0,1)$, for $t_0 < s <T$. Similarly as in
955: \emph{Step 1}, the integrand in \eqref{eq:udiffg}, using the backward
956: derivative can be bounded from below:
957: \begin{equation*}
958: \frac{d}{ds} \bar u(P \varphi(s),s) + h(\alpha(s)) \geq 
959: p^\#_t(s) + H(p^\#_\varphi(s),\varphi(s)).
960: \end{equation*}
961: As $\bar u$ is a viscosity solution to the Hamilton-Jacobi equation
962: for the discrete value function it holds that
963: \begin{equation*}
964: p^\#_t(s) + H(p^\#_\varphi(s),P \varphi(s)) \geq 0,
965: \end{equation*}
966: which proves the second inequality in \eqref{eq:errorrepr}.
967: %
968: %The semiconcavity of
969: %$\bar u$ implies that the directional derivative to $\bar u$ exists in
970: %every direction at every point. As in \emph{Step 1} we therefore have
971: %\begin{multline*}
972: %\frac{\bar u(\text{Proj}_V \varphi(s+\Delta s),s+\Delta
973: %  s)-\bar u(\text{Proj}_V \bar\varphi(s),s)}{\Delta s} \\
974: %= \partial
975: %  \bar u(\text{Proj}_V \varphi(s),s;( \frac{\text{Proj}_V \varphi(s+\Delta
976: %  s)-\text{Proj}_V \varphi(s)}{\Delta s},
977: %  1)) + o(1)\\
978: %=\min_{p \in D^+ \bar u} \big(p_t +(p_\varphi,\frac{\text{Proj}_V \varphi(s+\Delta
979: %  s)- \text{Proj}_V \varphi(s)}{\Delta s})\big) + o(1) \\
980: %=\min_{p \in D^+ \bar u} \big(p_t +(p_\varphi,\frac{\varphi(s+\Delta
981: %  s)- \varphi(s)}{\Delta s})\big) + o(1).
982: %\end{multline*}
983: %It can now be used that 
984: %\begin{equation*}
985: %\varphi \in L^2(t,T;H^2(0,1)) \cap C(t,T;H^1_0(0,1))
986: %\end{equation*}
987: %so that 
988: %\begin{align*}
989: %&(p_\varphi,\frac{\varphi(s+\Delta
990: %  s)- \varphi(s)}{\Delta s}) \\
991: %&= \frac{1}{\Delta s}\int_s^{s+\Delta s}\big(
992: %  -(\partial_x p_\varphi,\varphi_x)+(p_\varphi,-V'(\varphi)+\alpha)\big)ds
993: %  \rightarrow -(\partial_x p_\varphi,\varphi_x(t))+(p_\varphi,-V'(\varphi(t))+\alpha(t))
994: %\end{align*}
995: %We thus see that
996: %\begin{equation*}
997: %\frac{d}{ds} \bar
998: %u(\text{Proj}_V \varphi(s),s) +h(\alpha(s) = p_t +H(p^\#_\varphi(s),\varphi(s))
999: %\geq -H(p^\#_\varphi(s),\text{Proj}_V \varphi(s)) + H(p^\#_\varphi(s),\varphi(s))
1000: %\end{equation*} 
1001: %which gives us the desired lower bound.
1002: \end{proof}
1003: Theorem \ref{thm:valueerror} will be used when the error between the
1004: original and the approximate value functions is computed. For this to
1005: work some knowledge about the superdifferential $D^+ u_H$ is
1006: needed. The dual equation
1007: \begin{subequations}\label{eq:l}
1008: \begin{align}
1009: -\lambda_t &= \delta \lambda_{xx} - \delta^{-1}\lambda V''(\varphi), \label{eq:l1}\\
1010:  \lambda(T)&=2K \big( \varphi(T) - \varphi^-\big), \label{eq:l2}
1011: \end{align}
1012: \end{subequations}
1013: is introduced. Let $\alpha$ and $\varphi$ be optimal pairs as in Theorem
1014: \ref{thm:valueerror}. According to Theorem \ref{thm:boundedcontrol} it
1015: is possible to choose a bounded control. For the mild solution $\lambda$  to \eqref{eq:l} 
1016: there exists, according to Theorem 3.1 in \cite{Cannarsa-Frankowska1}, a
1017: subset $\mathcal{L} \subset [t_0, T]$, of full measure, such that, for
1018: all $t \in \mathcal{L}$,
1019: \begin{equation}\label{eq:valuesinDplus}
1020: \varphi(t) \in H_0^1(0,1) \cap H^2(0,1) \implies
1021: \Big( \lambda(t), -H\big(\lambda(t), \varphi(t)\big) \Big) \in D^+ u \big(\varphi(t),t\big).
1022: \end{equation} 
1023: By the same theorem, it holds that for almost every $t \in [t_0,T]$,
1024: \begin{equation}\label{eq:lambdaalpha}
1025: \big( \lambda(t),\alpha(t)\big) + \frac{||\alpha||^2}{2} =
1026: \min_{\substack{a \in L^2(0,1) \\ ||a|| \leq L}} \Big(
1027:   \big(\lambda(t),a \big) + \frac{||a||^2}{2} \Big),
1028: \end{equation}
1029: where $L$ is the bound on the control from Theorem
1030: \ref{thm:boundedcontrol}. This bound is included in order to be able
1031: to use the aforementioned Theorem 3.1 in \cite{Cannarsa-Frankowska1},
1032: but since $\alpha$ and $\lambda$ correspond to the original problem
1033: \eqref{eq:valuefunction} with no bound we could also have used any
1034: constant greater than $L$ in \eqref{eq:lambdaalpha}. Hence we see that
1035: \eqref{eq:lambdaalpha} holds also without the requirement that $a$ is
1036: bounded. From this we draw the conclusion that $\lambda(t) =
1037: -\alpha(t)$ a.e. The mild solutions $\varphi$ and $\lambda$ therefore
1038: satisfy the system
1039: \begin{subequations}\label{eq:mildsols}
1040: \begin{align}
1041: \varphi(t) &= S(t-t_0)\varphi_0 + \int_{t_0}^t S(t-s) \big(-\delta^{-1}
1042: V'(\varphi(s)) -\lambda(s) \big)ds, \label{eq:mildsols1} \\
1043: \varphi(t_0) &=\varphi_0, \label{eq:mildsols2} \\
1044: \lambda(t) &= S(T-t)\lambda(T) - \delta^{-1} \int_{t}^T S(s-t) \big(
1045: \lambda(s)V''(\varphi(s)) \big)ds, \label{eq:mildsols3} \\
1046: \lambda(T) &= 2K(\varphi(T)-\varphi^-), \label{eq:mildsols4}
1047: \end{align}
1048: \end{subequations}
1049: where $S(t)$ is the contraction semigroup of linear operators
1050: generated by $\delta \frac{d^2}{dx^2}$, \eqref{eq:mildsols1} is
1051: equation \eqref{eq:phimild} with $\lambda = -\alpha$, and
1052: \eqref{eq:mildsols3} is the equation for mild solutions to
1053: \eqref{eq:l1}; see e.g.\ Theorem 3.1 in \cite{Cannarsa-Frankowska1}. 
1054: Following the notation in \cite{Henry} we introduce
1055: \begin{equation*}
1056: A \equiv -\delta \frac{d^2}{dx^2}.
1057: \end{equation*}
1058: The operator $A$ has eigenvalues $k_n = \delta \pi^2 n^2$, $(n=1,2,3,
1059: \ldots )$ with corresponding eigenfunctions $\psi_n(x) = \sqrt{2} \sin
1060: (n\pi x)$. Fractional powers of $A$ may be defined using this:
1061: \begin{equation}\label{eq:Agamma}
1062: A^\gamma \varphi = \sum_{n=1}^\infty k_n^\gamma(\psi_n,\varphi)\psi_n,
1063: \end{equation}
1064: for $\gamma \geq 0$. The domain for $A^\gamma$ is given by
1065: \begin{equation}\label{eq:Agammadomain}
1066: D(A^\gamma) = \Big\{ \varphi \in L^2(0,1): \ \sum_{n=1}^\infty k_n^{2\gamma}
1067: (\psi_n,\varphi)^2 < \infty \Big\}.
1068: \end{equation}
1069: For $\gamma=1$ and $\gamma=1/2$ we have that $||A \varphi|| = \delta ||\varphi_{xx}||$ and
1070: $||A^{1/2}\varphi|| = \sqrt{\delta} ||\varphi_x||$.
1071: We state a few useful properties of the fractional powers of $A$,
1072: which may be found in e.g.\ \cite{Henry}.
1073: For any $K>0$ and all $0 < \gamma < K$ there exists a constant $C$
1074: such that 
1075: \begin{subequations}\label{eq:Aproperties}
1076: \begin{align}
1077: ||A^\gamma S(t)|| &\leq C t^{-\gamma}, \quad \text{for $t>0$,}
1078:    \label{eq:Aproperties1}\\
1079: \intertext{and if $0 < \gamma \leq 1, \varphi \in D(A^\gamma)$,}
1080: ||(S(t)-I)\varphi|| & \leq \frac{1}{\gamma}C t^\gamma ||A^\gamma
1081: \varphi||. \label{eq:Aproperties2} \\
1082: \intertext{It also holds that}
1083: A^{\gamma^1}A^{\gamma^2}&=A^{\gamma^2}A^{\gamma^1}=A^{\gamma^1+\gamma^2}
1084: \ \text{ on $D(A^{\gamma^1+\gamma^2})$ when $\gamma^1, \gamma^2 \geq
1085:     0$,} \label{eq:Aproperties3} \\
1086: A^\gamma S(t)&=S(t) A^\gamma \ \text{ on $D(A^\gamma)$, $t>0$.} \label{eq:Aproperties4}
1087: \end{align}
1088: \end{subequations}
1089: 
1090: 
1091: In  the following Theorem it is shown how an element in $D^+
1092: u_H(\bar\varphi(s),s)$ can be obtained, which is needed according to Theorem \ref{thm:valueerror}.
1093: \begin{thm}\label{thm:HJsatisfied}
1094: When $\varphi_0 \in V$ and $0 \leq s \leq T$ one element in $D^+
1095: u_H(\varphi_0,s)$ is given by
1096: \begin{equation*}
1097: \Big( \lambda(s), -H\big(\lambda(s),\varphi_0\big) \Big),
1098: \end{equation*}
1099: where $\lambda$ is a mild solution to  \eqref{eq:l} 
1100: and $\varphi$ is an optimal solution to \eqref{eq:valuefunction} with
1101: $t_0=s$.
1102: %We may therefore take 
1103: %\begin{equation*}
1104: %\int_{t_0}^T \big( -H(\lambda(s),\bar\varphi(s))+H(P
1105: %\lambda(s),\bar\varphi(s)) \big)ds
1106: %\end{equation*}
1107: %as the first integral in \eqref{eq:errorrepr}. 
1108: \end{thm} 
1109: \begin{proof}
1110: %According to Theorem 3.1 in \cite{Cannarsa-Frankowska1} there exists a
1111: %subset $\mathcal{L} \subset [t_0, T]$, of full measure, such that, for
1112: %all $t \in \mathcal{L}$,
1113: %\begin{equation*}
1114: %\varphi(t) \in H_0^1(0,1) \cap H^2(0,1) \implies
1115: %\big( \lambda(t), -H(\lambda(t), \varphi(t)) \big) \in D^+ u (\varphi(t),t).
1116: %\end{equation*} 
1117: Using Chapter 3 in \cite{Henry} we have that 
1118: $\varphi(t) \in H_0^1(0,1) \cap H^2(0,1)$ for $t_0 < t < T$, so by
1119: \eqref{eq:valuesinDplus} 
1120: \begin{equation*}
1121: \big( \lambda(t), -H(\lambda(t), \varphi(t)) \big) \in D^+ u (\varphi(t),t),
1122: \quad \text{a.e.}
1123: \end{equation*} 
1124: It follows from the definition that then also
1125: \begin{equation*}
1126: \big( \lambda(t), -H(\lambda(t), \varphi(t)) \big) \in D^+ u_H (\varphi(t),t),
1127: \quad \text{a.e.}
1128: \end{equation*}
1129: We shall verify that the semiconcavity of $u_H$ implies that
1130: \begin{equation}\label{eq:uppersemicontinuous}
1131: z_n \rightarrow z_0,\ D^+u_H(z_n) \ni p_n \rightarrow p \implies 
1132: p \in D^+ u_H(z_0).
1133: \end{equation}
1134: In order to prove this we use \eqref{eq:semiconmodulus} at the points
1135: $z_n$. We thereby have that 
1136: \begin{equation*}
1137: \begin{split}
1138: &u_H(z) - u_H(z_0) - (p,z-z_0)\\ 
1139: = & u_H(z)-u_H(z_n) -(p_n,z-z_n) \\
1140: & \quad+  u_H(z_n) - u_H(z_0) + (p_n-p,z-z_0) +
1141: (p_n,z - z_n) \\
1142:  \leq & K |z-z_0|^2 + \varepsilon,
1143: \end{split}
1144: \end{equation*}
1145: where $\varepsilon$ can be made arbitrarily small by using the
1146: convergence $z_n \rightarrow z_0$, $p_n \rightarrow p$, and that $u_H$
1147: is continuous (it is even locally Lipschitz continuous, see
1148: \cite{Cannarsa-Frankowska2}). Hence
1149: \begin{equation*}
1150: u_H(z) - u_H(z_0) - (p,z-z_0) \leq  K |z-z_0|^2,
1151: \end{equation*}
1152: which implies that $p \in D^+u_H(z_0)$, and
1153: \eqref{eq:uppersemicontinuous} holds. 
1154: (As can be seen in the above argument it suffices that $p_n
1155: \rightharpoonup p$ weakly,
1156: but we will not need this here.)
1157: 
1158: Since the Hamiltonian $H:H_0^1(0,1) \times H_0^1(0,1) \rightarrow \Re$
1159: is locally Lipschitz continuous and \eqref{eq:valuesinDplus} holds,
1160: what remains is to prove that $\varphi$ and $\lambda$ are continuous as
1161: functions of time with values in $H_0^1(0,1)$ at
1162: $t_0$. By equation \eqref{eq:mildsols1} we
1163: have that
1164: \begin{multline}\label{eq:Ahalfevol}
1165: A^{1/2} \big(\varphi(t)-\varphi_0 \big) = \big( S(t-t_0) -I \big)
1166: A^{1/2}\varphi_0 \\
1167: + \int_{t_0}^t A^{1/2} S(t-s) \big(
1168: -\delta^{-1}V'(\varphi(s)) -\lambda(s)\big)ds,
1169: \end{multline} 
1170: where passing $A^{1/2}$ under the integral sign is justified by the
1171: fact that $A^{1/2}$ is a closed operator. By \eqref{eq:Agammadomain}
1172: it is a straightforward calculation to confirm that $V \subset D(A^\gamma)$ for
1173: $\gamma < 3/4$. Since $\varphi_0 \in V$, \eqref{eq:Aproperties2}  may be
1174: used to get a bound for the first term in the right hand side of
1175: \eqref{eq:Ahalfevol}:
1176: \begin{equation*}
1177: ||\big(S(t-t_0)-I \big) A^{1/2}\varphi_0 || \leq 10 C t^{1/10} ||A^{3/5} \varphi_0||.
1178: \end{equation*} 
1179: The norm of the integral in \eqref{eq:Ahalfevol} converges to zero as
1180: $t \rightarrow t_0$ by \eqref{eq:Aproperties1} and the fact that $V'$
1181: and $\lambda$ (since it equals $-\alpha$) are bounded. Hence 
1182: \begin{equation*}
1183: ||A^{1/2} \big(\varphi(t) - \varphi_0 \big)|| \rightarrow 0, \quad \text{as
1184:   $t \searrow t_0$.}
1185: \end{equation*} 
1186: The function $\lambda$ is also   continuous as a function with values in $H_0^1(0,1)$
1187: when $t \searrow t_0$, as, by Theorem 3.5.2 in \cite{Henry},
1188: $||A^\gamma \lambda_t||$ exists when $t<T$ and $\gamma <1$, e.g.\ $\gamma=1/2$.
1189: \end{proof}
1190: In order to be able to use Theorem \ref{thm:valueerror} and
1191: \ref{thm:HJsatisfied} a few results about the regularity for the state
1192: and the dual is established. The original setting, without
1193: discretization in space, is considered first.
1194: %\begin{lemma}\label{lem:projection}
1195: %\begin{equation*}
1196: %D\arrowvert_V^+ u(t,x)=\text{Proj}_V D^+ u(t,x), \quad \text{when } x
1197: %\in V.
1198: %\end{equation*}
1199: %\end{lemma}
1200: %\begin{proof}
1201: %It is a direct consequence of the definition that $\text{Proj}_V D^+
1202: %u(t,x) \subset D\arrowvert_V^+ u(t,x)$. Hence we show that 
1203: %$D\arrowvert_V^+ u(t,x)\subset \text{Proj}_V D^+ u(t,x)$. This is done by
1204: %proving that $D\arrowvert_V^* u(t,x)\subset \text{Proj}_V D^* u(t,x)$. Let
1205: %\begin{equation*}
1206: %p=\lim_{i \to \infty} D\arrowvert_V u (t_i,x_i) \in D\arrowvert_V^* u(t,x),
1207: %\end{equation*}
1208: %where $x_i \in V$ for all $i$. Since $D^+ u$ is upper semicontinuous
1209: %there exists points $(s_i,y_i)$ where $y_i \in L^2$ and where u is
1210: %differentiable such that 
1211: %\begin{equation}\label{eq:uppersemicont}
1212: %||(t_i,x_i)-(s_i,y_i)|| < \frac{1}{i}, \quad Du(s_i,y_i) \in
1213: %  D^+u(t_i,x_i) + \frac{1}{i}B.
1214: %\end{equation}
1215: %The fact that 
1216: %\begin{equation*}
1217: %\text{Proj}_V Du(s_i,y_i)= D\arrowvert_V u(s_i,y_i)
1218: %\end{equation*}
1219: %together with (\ref{eq:uppersemicont}) thus entails
1220: %\begin{equation}\label{eq:projconvergence}
1221: %\lim_{i \to \infty} \text{Proj}_V Du(s_i,y_i) = p.
1222: %\end{equation}
1223: %The set $\{Du(s_i,y_i)\}$ is bounded in $L^2$, so there exists a
1224: %subsequence, for which the subscript $i$ is also used, which converges
1225: %weakly in $L^2$ to  an element $q \in D^* u(t,x)$. This fact together
1226: %with (\ref{eq:projconvergence}) gives us that 
1227: %\begin{equation*}
1228: %\text{Proj}_V q=p.
1229: %\end{equation*} 
1230: %\end{proof}
1231: \begin{thm}\label{thm:infinitedimregularity}
1232: For every $C>0$ and all starting positions $(\varphi_0,t_0)$ satisfying
1233: $||\varphi_0||_{H_0^1(0,1)} \leq C$, $0 \leq t_0 \leq T$, there exists a $D>0$ and an optimal
1234: state $\varphi$ to problem \eqref{eq:valuefunction}, with corresponding
1235: dual $\lambda$, solving \eqref{eq:l}, such that for all $t_0 \leq t
1236: \leq T$,
1237: \begin{subequations}\label{eq:infinitedimregularity}
1238: \begin{align}
1239: ||\varphi(t)||_{H_0^1(0,1)} & \leq D, \label{eq:infinitedimregularity1}
1240:   \\
1241: ||\varphi(t)||_{H^2(0,1)} & \leq D(t-t_0)^{-1/2}, \label{eq:infinitedimregularity2}
1242:   \\
1243: ||\lambda(t)||_{H_0^1(0,1)} & \leq D, \label{eq:infinitedimregularity3}
1244:   \\
1245: ||\lambda(t)||_{H^2(0,1)} & \leq
1246:   D(T-t_0)^{-1/2}. \label{eq:infinitedimregularity4} \\
1247: \intertext{If $\varphi_0$ satisfies the higher regularity $||A^{5/7}
1248:   \varphi_0|| \leq C$ it further holds that}
1249: ||\varphi(t)||_{H^2(0,1)} &\leq D(t-t_0)^{-2/7},  \label{eq:infinitedimregularity5} \\
1250: ||\lambda(t)||_{H^2(0,1)} &\leq D(T-t_0)^{-2/7}.  \label{eq:infinitedimregularity6} 
1251: \end{align}
1252: \end{subequations}
1253: \end{thm}
1254: \begin{proof}
1255: In the proof, we will write $D$ for any constant which may depend on
1256: $C$, but not on $t_0$.
1257: 
1258: \emph{Step 1.} By theorem \ref{thm:boundedcontrol} it is possible to
1259: choose an optimal control $\alpha$ such that $||\alpha(t)|| \leq L$,
1260: for some constant $L$ which only depends on $C$. Since
1261: $\lambda=-\alpha$ the same holds for $\lambda$.
1262: 
1263: \emph{Step 2.} Since $A^\gamma$ and $S(t)$ commute (see
1264: \eqref{eq:Aproperties4}) we can operate with $A^{1/2}$ on equation
1265: \eqref{eq:mildsols1} to obtain
1266: \begin{equation*}
1267: A^{1/2}\varphi(t)=S(t-t_0)A^{1/2}\varphi_0 + \int_{t_0}^t A^{1/2} S(t-s)
1268: \big(-\delta^{-1} V'(\varphi(s))-\lambda(s) \big)ds.
1269: \end{equation*} 
1270: As $S(t)$ is a contraction semigroup and by the boundedness of $V'$
1271: and $||\lambda(s)||$ together with \eqref{eq:Aproperties1} it therefore holds that 
1272: \begin{equation*}
1273: ||A^{1/2}\varphi(t)|| \leq ||A^{1/2}\varphi_0|| + D \int_{t_0}^t
1274:   (t-s)^{-1/2} ds,
1275: \end{equation*}
1276: and hence
1277: $||\varphi_x(t)|| \leq D$. By a Poincar\'e
1278: inequality (e.g.\ Proposition 5.3.5 in \cite{Brenner-Scott})
1279: \eqref{eq:infinitedimregularity1} holds.
1280: 
1281: \emph{Step 3.} Since $\lambda(T)=2K(\varphi(T)-\varphi^-)$, boundedness of
1282: $\lambda(T)$ in $H_0^1(0,1)$ follows. The same analysis for
1283: \eqref{eq:mildsols3}  as was performed in \emph{Step 2} may therefore
1284: be used. Using that $||\lambda(s)||$ is bounded for all $s$ gives
1285: \eqref{eq:infinitedimregularity3}.
1286: 
1287: \emph{Step 4.}  Operate with $A$ on \eqref{eq:mildsols1} to obtain
1288: \begin{equation}\label{eq:Ahalfhalf}
1289: A\varphi(t) = A^{1/2}S(t-t_0) A^{1/2}\varphi_0 + \int_{t_0}^t A^{1/2} S(t-s)
1290: A^{1/2} \big(-\delta^{-1}V'(\varphi(s)) -\lambda(s)\big) ds.
1291: \end{equation}
1292: Since $\varphi(s)$ and $\lambda(s)$ are bounded in $H_0^1(0,1)$ for all
1293: $s$ it holds that
1294: \begin{equation*}
1295: ||A^{1/2} \big(-\delta^{-1}V'(\varphi(s))-\lambda(s)\big)|| < D.
1296: \end{equation*}
1297: Therefore 
1298: \begin{equation*}
1299: ||A \varphi(t)|| \leq D (t-t_0)^{-1/2} + D \int_{t_0}^t (t-s)^{-1/2} ds \leq
1300:   D (t-t_0)^{-1/2},
1301: \end{equation*}
1302: as $T$ is finite. So \eqref{eq:infinitedimregularity2} holds.
1303: 
1304: \emph{Step 5.} In this last step we use the operator $A$ in equation
1305: \eqref{eq:mildsols3} in the following way:
1306: \begin{equation}\label{eq:H2lambda}
1307: A\lambda(t) = S(T-t) A\lambda(T) - \delta^{-1} \int_t^T A^{1/2} S(s-t)
1308: A^{1/2} \Big(\lambda(s) V''\big(\varphi(s)\big)\Big)ds.
1309: \end{equation}
1310: The bound for $\varphi(T)$ in $H^2(0,1)$ may be transferred to
1311: $\lambda(T)$ by \eqref{eq:mildsols4}, which yields $||\lambda(T)||
1312: \leq D(T-t_0)^{-1/2}$. As both $\varphi(t)$ and $\lambda(t)$ are bounded in $H_0^1(0,1)$
1313: for all $t_0 < t < T$ the integral in \eqref{eq:H2lambda} is bounded
1314: by 
1315: \begin{equation*}
1316: D \int_t^T (s-t)^{-1/2} ds.
1317: \end{equation*}  
1318: Since $T$ is finite
1319: \eqref{eq:infinitedimregularity4} holds.
1320: 
1321: \emph{Step 6.} Equations \eqref{eq:infinitedimregularity5} and \eqref{eq:infinitedimregularity6} can be
1322: proved similarly as in \emph{Step 4} and \emph{Step 5} by changing the
1323: first term in the right hand side of \eqref{eq:Ahalfhalf} to 
1324: $A^{2/7}S(t-t_0)A^{5/7} \varphi_0$. By this we see that $||A\varphi(t)||
1325: \leq D(t-t_0)^{-2/7}$, which implies \eqref{eq:infinitedimregularity6},
1326: just as in \emph{Step 5}.
1327: \end{proof}
1328: A regularity result for the spatially discretized case is now to be
1329: established.
1330: According to theory in e.g.\ \cite{Cannarsa-Sinestrari} the optimal
1331: control problem \eqref{eq:approxu}, \eqref{eq:approxphievol} has a
1332: minimizing control $\bar \alpha$.  The corresponding state is denoted
1333: $\bar \varphi$. The value function
1334: is differentiable along optimal paths, i.e.\ the derivative exists at $\bar u (\bar\varphi(s),s)$
1335: for all $s \in (t_0,T]$. The spatial G\^ateaux derivative of $\bar u$ at
1336: $(\bar\varphi(s),s)$ will be denoted $\bar\lambda(s)$. The optimal state,
1337: $\bar\varphi$, and the G\^ateaux derivative $\bar\lambda$ satisfy the
1338: following system:
1339: \begin{subequations}\label{eq:approxHamiltonsyst}
1340: \begin{align}
1341: (\bar\varphi_t,v)&=-\delta(\bar\varphi_{x},v_x)-(\delta^{-1}V'(\bar\varphi)+\bar\lambda,v), \quad
1342:   \text{for all } v \in V,\label{eq:approxHamiltonsyst1}\\
1343: \bar\varphi(t_0)&=\bar\varphi_0, \label{eq:approxHamiltonsyst2}\\
1344: -(\bar\lambda_t,v)&=
1345:   -\delta(\bar\lambda_x,v_x)-\delta^{-1}(\bar\lambda V''(\bar\varphi),v),
1346:   \quad \text{for all } v \in V, \label{eq:approxHamiltonsyst3}\\
1347: \bar\lambda(T) &= 2K(\bar\varphi(T)-\text{Proj}_V \varphi^-). \label{eq:approxHamiltonsyst4}
1348: \end{align}
1349: \end{subequations}
1350: Furthermore, the theory in \cite{Cannarsa-Sinestrari} reveals that the
1351: optimal control, $\bar\alpha$, satisfies $\bar\alpha=-\bar\lambda$.
1352: Some new notation is now introduced. We let the interval $[0,1]$ be
1353: divided into $M$ subintervals with $\Delta x = 1/M$, and let
1354: $\{v^i\}_{i=1}^{M-1}$ be the standard nodal basis in $V$; see Figure
1355: \ref{fig:basis}.
1356: \begin{figure}
1357: \centering
1358: \includegraphics[width=0.5\textwidth]{basis.eps}
1359: \caption{Two of the nine basis functions when $\Delta x=1/10$.}
1360: \label{fig:basis}
1361: \end{figure} 
1362: The \emph{interpolant}, $I$, takes a function in $C([0,1])$ to the
1363: element in $V$ which
1364: coincides with the original function at $i \Delta x$, $i=1, \ldots ,
1365: M-1$, so that, for instance, $I \lambda(i \Delta x) = \lambda( i
1366: \Delta x)$.
1367: The second difference quotient matrix $D^2$ and the mass matrix $B$ are
1368: introduced as
1369: \begin{align}
1370: D^2 &= \frac{1}{\Delta x^2} \begin{pmatrix}
1371: -2 & 1 & 0 & \cdots & 0 \\
1372: 1 & -2 & 1 &  & \vdots & \\
1373: 0 & \ddots &\ddots &\ddots & 0 \\
1374: \vdots & & 1 & -2 & 1 \\
1375: 0 & \cdots & 0 & 1 & -2
1376: \end{pmatrix}, \label{eq:diffquotient}\\
1377: \notag \\
1378: B &= \begin{pmatrix}
1379: 2/3 & 1/6 & 0 & \cdots & 0 \\
1380: 1/6 & 2/3 & 1/6 &  & \vdots & \\
1381: 0 & \ddots &\ddots &\ddots & 0 \\
1382: \vdots & & 1/6 & 2/3 & 1/6 \\
1383: 0 & \cdots & 0 & 1/6 & 2/3
1384: \end{pmatrix}. \label{eq:massmatrix}
1385: \end{align}
1386: If $\bar\varphi(t)$ and $\bar\lambda(t)$ are written in the basis
1387: $\{v^i\}_{i=1}^{M-1}$ as
1388: \begin{equation}\label{eq:coorexpr}
1389: \bar\varphi(t) =: \sum_{i=1}^{M-1} \zeta^i(t) v^i, \quad
1390: \bar\lambda(t) =: \sum_{i=1}^{M-1} \theta^i(t) v^i,
1391: \end{equation}
1392: then equations \eqref{eq:approxHamiltonsyst1} and
1393: \eqref{eq:approxHamiltonsyst3} may be rewritten as 
1394: \begin{subequations}\label{eq:FEMcoords}
1395: \begin{align}
1396: B \zeta' &= \delta D^2 \zeta - \frac{\delta^{-1}}{\Delta x} p - B
1397: \theta, \label{eq:FEMcoords1} \\
1398: -B \theta' &= \delta D^2 \theta - \frac{\delta^{-1}}{\Delta x}r, \label{eq:FEMcoords2}
1399: \end{align}
1400: \end{subequations}
1401: where 
1402: \begin{equation}\label{eq:vectdefs}
1403: \begin{split}
1404: \zeta= \begin{pmatrix} \zeta^1 \\ \vdots \\ \zeta^{M-1} \end{pmatrix},&
1405: \quad 
1406: \theta= \begin{pmatrix} \theta^1 \\ \vdots \\ \theta^{M-1}
1407: \end{pmatrix}, \\ 
1408: p=\begin{pmatrix}
1409: (V'(\bar\varphi),v^1) \\
1410: \vdots \\
1411: (V'(\bar\varphi),v^{M-1})
1412: \end{pmatrix},
1413: &\ \text{and} \   
1414: r=\begin{pmatrix}
1415: (\bar\lambda V''(\bar\varphi),v^1) \\
1416: \vdots \\
1417: (\bar\lambda V''(\bar\varphi),v^{M-1})
1418: \end{pmatrix}.
1419: \end{split}
1420: \end{equation}
1421: We now state a Lemma which will be used in the proofs of Theorem
1422: \ref{thm:diffcontrol} and \ref{thm:spatialconvergence}.
1423: \begin{lemma}\label{lem:H2proj}
1424: For any element $\psi \in H^2(0,1) \cap H^1_0(0,1)$ the projection
1425: $P\psi$, written in the nodal basis $\{v^i\}$ as
1426: \begin{equation*}
1427: P \psi =: \sum_{i=1}^{M-1} \xi^i v^i,
1428: \end{equation*}
1429: satisfies
1430: \begin{equation}\label{eq:H2proj}
1431: \big(\Delta x \sum_{i=1}^{M-1} (D^2 \xi)_i^2 \big)^{1/2} \leq C ||\psi_{xx}||,
1432: \end{equation}
1433: with a constant $C$ independent of $\Delta x$.
1434: \end{lemma} 
1435: \begin{proof}
1436: The vector $\xi$ is defined as
1437: \begin{equation*}
1438: \xi= \begin{pmatrix} \xi^1 \\ \vdots \\ \xi^{M-1} \end{pmatrix},
1439: \end{equation*}
1440: and is given by $\xi=\frac{1}{\Delta x} B^{-1} q$, where 
1441: \begin{equation*}
1442: q=\begin{pmatrix}
1443: (\psi,v^1) \\
1444: \vdots \\
1445: (\psi,v^{M-1})
1446: \end{pmatrix},
1447: \end{equation*}
1448: which follows from the fact that $(\psi,v^i)=(P\psi,v^i)$ for all $v^i$.
1449: The matrices $D^2$ and $B^{-1}$ commute, since the eigenvectors of
1450: $D^2$ and $B$ (and $B^{-1}$) are equal, so $D^2 \xi = \frac{1}{\Delta
1451: x} B^{-1} D^2 q.$ Every element of the vector $D^2
1452: q$, except the first and the last, is a $L^2(0,1)$ scalar product
1453: between $\psi$ and a translate of the function $v$ in Figure \ref{fig:v}.
1454: \begin{figure}
1455: \centering
1456: \includegraphics[width=0.5\textwidth]{v.eps}
1457: \caption{The function $v$ admits values in the interval $[-2/\Delta
1458:     x^2, 1/\Delta x^2]$ (here $\Delta x = 1/10$).}
1459: \label{fig:v}
1460: \end{figure}
1461: It is easy to check that there exists a primitive function $\bar v$ to $v$
1462: such that $\bar v(0)=\bar v(4\Delta x)=0$, and that there exists a primitive
1463: function $\Bar{\Bar{v}}$ to $\bar v$ such that $\Bar{\Bar{v}}(0)=\Bar{\Bar{v}}(4\Delta x)=0$. Furthermore $\max
1464: \Bar{\Bar{v}} \leq E$, for a constant $E$ which does not depend on $\Delta x$. Hence,
1465: \begin{equation*}
1466: |(\psi,v)|=|(\psi_{xx},\Bar{\Bar{v}})| \leq E \int_0^{4\Delta x} |\psi_{xx}|dx
1467: \leq 2E \sqrt{\Delta x} \big( \int_0^{4\Delta x} \psi_{xx}^2 dx \big)^{1/2}.
1468: \end{equation*}
1469: The same sort of bound may be obtained
1470: also for the first and the last elements of $D^2 q$
1471: by using a $2$-periodic, odd extension of $\psi$ outside
1472: $[0,1]$. We therefore have that
1473: \begin{align*}
1474: ||D^2 q||_2^2 & \leq 4E^2 \Delta x \big(\underbrace{\int_{0}^{\Delta x} \psi_{xx}^2 dx
1475:   + \int_0^{3\Delta x} \psi_{xx}^2 dx}_{\text{From } (D^2 q)_1} + \underbrace{\int_0^{4\Delta x} \psi_{xx}^2 dx}_{\text{From } (D^2 q)_2}
1476:   + \ldots \\
1477: &+ \underbrace{\int_{(M-3)\Delta x}^1 \psi_{xx}^2 dx + \int_{(M-1)\Delta
1478:   x}^1\psi_{xx}^2 dx}_{\text{From } (D^2 q)_{M-1}} \big) \\
1479: & \leq 16 E^2 \Delta x ||\psi_{xx}||^2,
1480: \end{align*}
1481: where $||\cdot||_2$ denotes the standard Euclidean vector norm. The
1482: eigenvalues of $B^{-1}$ lie in the interval $[1,3]$, and hence it
1483: holds that 
1484: \begin{equation*}
1485: ||D^2 \xi||_2 \leq \frac{12E}{\sqrt{\Delta x}} ||\psi_{xx}||,
1486: \end{equation*}
1487: which is equivalent to \eqref{eq:H2proj}. 
1488: \end{proof}
1489: \begin{thm}\label{thm:diffcontrol}
1490: There are constants $E$ and $F$, depending on the parameters of the
1491: optimal control problem, as in Theorem \ref{thm:boundedcontrol}, but
1492: not on $\varphi_0$, $t_0 \in [0,T]$ and the size of the
1493: spatial discretization, such that for all $t_0 \leq t \leq T$,
1494: \begin{align*}
1495: ||\bar\varphi_x(t)|| + ||\bar\lambda_x(t)|| & \leq E||(\varphi_0)_x|| + F, \\
1496: ||\bar\varphi_t(t)|| &\leq E\big( \Delta x
1497: \sum_{i=1}^{M-1}(D^2 \zeta(t_0))_i^2 \big)^{1/2} +F, \\
1498: \intertext{and}
1499: ||\bar\lambda_t(t)|| & \leq \Big(E\big( \Delta x
1500: \sum_{i=1}^{M-1}(D^2 \zeta(t_0))_i^2 \big)^{1/2} +F \Big)^2, \\
1501: \intertext{where}
1502: \varphi_0 &= \sum_{i=1}^{M-1} \zeta^i(t_0) v^i. 
1503: \end{align*}  
1504: %For each $C>0$, there is a $D>0$, independent of the size of the
1505: %spatial discretization, $\Delta x$, and starting time, $t_0$, such that for all
1506: %$||(\bar\varphi_0)_x||\leq C$ 
1507: %there exists an optimal state $\bar\varphi : [t_0,T] \rightarrow V$ for
1508: %\eqref{eq:approxu}, with an associated dual $\bar\lambda:[t_0,T]
1509: %\rightarrow V$ which satisfies \eqref{eq:approxHamiltonsyst}, such
1510: %that for all $t_0 \leq t \leq T$,
1511: %\begin{equation*}
1512: %||\bar\varphi_x(t)|| + ||\bar\lambda_x(t)|| \leq D.
1513: %\end{equation*}  
1514: %If, moreover
1515: %\begin{equation*}
1516: %\varphi_0 = \sum_{i=1}^{M-1} \zeta^i(t_0) v^i
1517: %\end{equation*}
1518: %satisfies
1519: %\begin{equation}\label{eq:initcondbound}
1520: %\big( \Delta x \sum_{i=1}^{M-1}(D^2 \zeta(t_0))_i^2 \big)^{1/2} \leq C,
1521: %\end{equation}
1522: %then also $||\bar\varphi_t(t)||+||\bar\lambda_t(t)|| \leq D$, for all $t_0 \leq t \leq T$.
1523: \end{thm}
1524: \begin{proof}
1525: The proof uses the same kind of techniques as in the proof of Theorem
1526: 5 on page 360 in \cite{Evans}.
1527: Here, however, the regularity of $\bar\varphi$ must also be conveyed to $\bar\lambda$.
1528: In the proof, we will write $E$ and $F$ for any constants that may depend on
1529: the parameters of the problem, but not on $\Delta x$,  $t_0$, and $\varphi_0$.
1530: 
1531: \emph{Step 1.} As in the infinite dimensional case, treated in Theorem
1532: \ref{thm:boundedcontrol}, it holds that the infimum in
1533: \eqref{eq:approxu}  can be changed to
1534: $\inf_{||\bar\alpha||_{L^{\infty}(t_0,T;V)}\leq E||\varphi_0||+F}$. The proof goes just as the proof for Theorem
1535:   \ref{thm:boundedcontrol}. By a Poincar\'e inequality (see e.g.\ Theorem 5.3.5 in
1536: \cite{Brenner-Scott}) it holds that $||\varphi_0|| \leq C ||(\varphi_0)_x||$
1537: and therefore the infimum can be written  
1538: $\inf_{||\bar\alpha||_{L^{\infty}(t_0,T;V)}\leq E||(\varphi_0)_x||+F}$.
1539: It is therefore possible to let $v=\bar\varphi$ in
1540: \eqref{eq:approxphievol} and use this boundedness of
1541: $\bar\alpha$ to see that 
1542: $||\bar\varphi(t)|| \leq E ||(\varphi_0)_x|| +F$, for all $t_0 \leq t \leq T$.
1543: 
1544: \emph{Step 2.} As already noted $\bar\lambda=-\bar\alpha$, and so by
1545: \emph{Step 1} the same bound on $||\bar\lambda(t)||$ also holds.
1546: 
1547: \emph{Step 3.} With
1548: $v=\bar\varphi_t$ in \eqref{eq:approxHamiltonsyst1} we have
1549: \begin{multline*}
1550: ||\bar\varphi_t ||^2 + \frac{\delta}{2} \frac{d}{dt} ||\bar\varphi_x||^2 = -
1551:   \delta^{-1} (V'(\bar\varphi),\bar\varphi_t)-(\bar\lambda,\bar\varphi_t) \\
1552: \leq
1553:   \delta^{-1} |V'|\cdot||\bar\varphi_t||+ ||\bar\lambda||\cdot
1554:   ||\bar\varphi_t|| 
1555: \leq \delta^{-2}|V'|^2+\frac{||\bar\varphi_t||^2}{4} +
1556:   \frac{||\bar\lambda||^2}{2} + \frac{||\bar\varphi_t||^2}{2}.
1557: \end{multline*}
1558: The boundedness of $(\bar\varphi_0)_x$ and $||\bar\lambda(t)||$ thus implies
1559: that 
1560: \begin{align*}
1561: ||\bar\varphi_x(t)|| &\leq E||(\varphi_0)_x||+F, \\
1562: ||\bar\varphi_t||_{L^2(t_0,T;V)} &\leq E||(\varphi_0)_x||+F.
1563: \end{align*}
1564: By this the same sort of bound holds for  $\bar\lambda_x(T)$.
1565: Letting $v=\bar\lambda_t$ in \eqref{eq:approxHamiltonsyst3} gives
1566: that 
1567: \begin{align*}
1568: ||\bar\lambda_x(t)|| &\leq E||(\varphi_0)_x||+F, \\
1569: ||\bar\lambda_t||_{L^2(t_0,T;V)} &\leq E||(\varphi_0)_x||+F,
1570: \end{align*}
1571: similarly as for $\bar\varphi$.
1572: 
1573: \emph{Step 4.}  All eigenvalues
1574: of $B$ lie in the interval $[1/3,1]$, and so $||B^{-1}||_2 \leq 3$
1575: (independently of $\Delta x$), where $||\cdot||_2$ denotes the
1576: operator 2-norm. From this and \eqref{eq:FEMcoords} it follows
1577: that 
1578: \begin{equation*}
1579: ||\bar\varphi_t(t_0)|| \leq E\big( \Delta x
1580: \sum_{i=1}^{M-1}(D^2 \zeta(t_0))_i^2 \big)^{1/2} +F.
1581: \end{equation*}
1582: We now introduce the notation $\hat\varphi \equiv
1583: \bar\varphi_t$ and $\hat \lambda \equiv \bar\lambda_t$ and differentiate
1584: equation \eqref{eq:approxHamiltonsyst1}
1585: with respect to time:
1586: \begin{equation*}
1587: (\hat\varphi_t,v)=-\delta(\hat\varphi_{x},v_x)-\delta^{-1}(V''(\bar\varphi)\hat\varphi,v)-(\hat\lambda,v),
1588:   \quad \text{for all } v \in V.
1589: \end{equation*}
1590: Let $v=\hat\varphi$ to obtain
1591: \begin{multline*}
1592: \half \frac{d}{dt} ||\hat\varphi||^2 = -\delta ||\hat
1593: \varphi_x||^2-\delta^{-1} (V''(\bar\varphi)\hat\varphi,\hat\varphi)-
1594: (\hat\lambda,\hat\varphi) \\
1595: \leq \delta^{-1} |V''| \cdot ||\hat\varphi||^2 +
1596: \frac{||\hat\lambda||^2}{2} + \frac{||\hat\varphi||^2}{2}.
1597: \end{multline*}
1598: The fact that $||\bar\varphi_t(t_0)|| \leq D$
1599: together with the result on boundedness of
1600: $||\hat\lambda||_{L^2(t_0,T;V)}$ and $||\hat\varphi||_{L^2(t_0,T;V)}$ in \emph{Step 3} implies 
1601: that 
1602: \begin{equation*}
1603: ||\hat\varphi(t)|| \leq E\big( \Delta x
1604: \sum_{i=1}^{M-1}(D^2 \zeta(t_0))_i^2 \big)^{1/2} +F, \quad \text{for }
1605: t_0 \leq t \leq T.
1606: \end{equation*}
1607: 
1608: \emph{Step 5.} In this step it will  be shown that
1609: $| \big((P\varphi_-)_x,w_x\big)| \leq D ||w||$ for all $w \in V$. Some new notation is introduced:
1610: \begin{equation*}
1611: P\varphi_- \equiv \sum_{i=1}^{M-1} \xi^i v^i, \quad w \equiv
1612: \sum_{i=1}^{M-1} \eta^i v^i,
1613: \end{equation*}  
1614: with corresponding  vectors $\xi$ and $\eta$.
1615: By means of a partial integration
1616: \begin{equation}\label{eq:phiprod}
1617: |\big( (P\varphi_-)_x,w_x\big)| \leq \Delta x |\eta^{T} D^2 \xi| \leq
1618: \Delta x ||\eta||_2\cdot||D^2 \xi||_2,
1619: \end{equation}
1620: with $||\cdot||_2$ denoting the Euclidean vector norm.
1621: Since the eigenvalues of $B$ lie in $[1/3,1]$ it holds that 
1622: \begin{equation*}
1623: ||w||^2 = \Delta x \eta^T B \eta \geq \frac{\Delta x}{3} ||\eta||^2_2,
1624: \end{equation*}
1625: and so by Lemma \ref{lem:H2proj}, \eqref{eq:phiprod} gives:
1626: \begin{equation*}
1627: |\big( (P\varphi_-)_x,w_x \big)| \leq \Delta x ||D^2 \xi||_2 \cdot
1628:  ||\eta||_2 \leq \Delta x \frac{D||(\varphi_-)_{xx}||}{\sqrt{\Delta x}}
1629:  \cdot \frac{\sqrt{3}}{\sqrt{\Delta x}} ||w|| = D||w||.
1630: \end{equation*}
1631: 
1632: \emph{Step 6.} It holds by equation \eqref{eq:approxHamiltonsyst4}
1633: that $\bar\lambda_x(T)=2K(\bar\varphi_x(T)-(P\varphi_-)_x)$. Using this in
1634: \eqref{eq:approxHamiltonsyst3} as well as
1635: \eqref{eq:approxHamiltonsyst1} gives
1636: \begin{align*}
1637: -(\hat\lambda(T),v) &=
1638:  -2K\delta(\bar\varphi_x(T),v_x)+2K\delta((P\varphi_-)_x,v_x)-\delta^{-1}(\bar\lambda(T)V''(\bar\varphi(T)),v) \\
1639: &=
1640:  2K(\hat\varphi(T),v)+2K(\delta^{-1}V'(\bar\varphi(T))+\bar\lambda(T),v) \\
1641: &\quad +2K\delta((P\varphi_-)_x,v_x) -\delta^{-1}(\bar\lambda(T)V''(\bar\varphi(T)),v).
1642: \end{align*}
1643: With $v=\hat\lambda(T)$ it follows that 
1644: \begin{equation*}
1645: ||\hat\lambda(T)|| \leq E\big( \Delta x
1646: \sum_{i=1}^{M-1}(D^2 \zeta(t_0))_i^2 \big)^{1/2} +F,
1647: \end{equation*}
1648: by the results in \emph{Step 4} and
1649: \emph{Step 5}. In order to bound 
1650: $\hat\lambda$ at all
1651: times, equation
1652: \eqref{eq:approxHamiltonsyst3} is differentiated with respect to time:
1653: \begin{equation*}
1654: -(\hat\lambda_t,v)=-\delta(\hat\lambda_x,v_x)-\delta^{-1}(\hat\lambda
1655:  V''(\bar\varphi),v)-\delta^{-1}(\bar\lambda V'''(\bar\varphi)\hat\varphi,v).
1656: \end{equation*}
1657: With $v=\hat\lambda$ in the previous equation the following bound is
1658: obtained:
1659: \begin{equation*}
1660: -\half\frac{d}{dt}||\hat\lambda||^2 \leq
1661:  \delta^{-1}|V''|\cdot||\hat\lambda||^2 + \delta^{-1}|V'''| \int_0^1
1662:  |\bar\lambda \hat\varphi \hat\lambda|dx.
1663: \end{equation*}
1664: Since $\bar\lambda$ is bounded by $E\big( \Delta x
1665: \sum_{i=1}^{M-1}(D^2 \zeta(t_0))_i^2 \big)^{1/2} +F$ in $H^1_0$ for all times 
1666: it is similarly bounded in $L^\infty$. The last integral in the previous inequality may therefore be
1667: estimated as follows:
1668: \begin{equation*}
1669: \int_0^1
1670:  |\bar\lambda \hat\varphi \hat\lambda|dx \leq |\bar\lambda| \cdot
1671:  ||\hat\varphi|| \cdot ||\hat\lambda|| \leq \Big(E\big( \Delta x
1672: \sum_{i=1}^{M-1}(D^2 \zeta(t_0))_i^2 \big)^{1/2} +F\Big)^2 ||\hat\lambda||.
1673: \end{equation*}
1674: Using $\frac{d}{dt}||\hat\lambda||^2 =
1675: 2||\hat\lambda||\frac{d}{dt}||\hat\lambda||$, Gr\"onwall's Lemma, and
1676: the boundedness of $||\hat\lambda(T)||$ we see that $||\hat\lambda(t)||$ is bounded by $\Big(E\big( \Delta x
1677: \sum_{i=1}^{M-1}(D^2 \zeta(t_0))_i^2 \big)^{1/2} +F\Big)^2$ for all times.
1678: \end{proof}
1679: With the error representation in Theorem \ref{thm:valueerror},
1680: Theorem \ref{thm:HJsatisfied} about $D^+ u_H$, and the regularity results of Theorem \ref{thm:infinitedimregularity}
1681: and \ref{thm:diffcontrol}, it is possible to prove Theorem
1682: \ref{thm:spatialconvergence} about spatial convergence.
1683: We first give the idea of the proof. When the first integral in
1684: \eqref{eq:errorrepr} is estimated the optimal path $\bar\varphi(s)$ is
1685: used. In order to obtain an element in $D^+
1686: u_H\big(\bar\varphi(s),s\big)$ the system of equations
1687: \eqref{eq:mildsols} is considered with $\big(\bar\varphi(s),s\big)$
1688: playing the role of $(\varphi_0,t_0)$. By Theorem~
1689: \ref{thm:HJsatisfied} the computed $\lambda(s)$ is the spatial part of
1690: an element in $D^+ u_H\big(\bar\varphi(s),s\big)$. Therefore, equation
1691: \eqref{eq:mildsols} must be used for every starting position
1692: $\big(\bar\varphi(s),s\big)$, $0 \leq s \leq T$, as depicted by Figure
1693: \ref{fig:manyphis}. 
1694: \begin{figure}
1695: \centering
1696: \includegraphics[width=0.5\textwidth]{manyphis.eps}
1697: \caption{Sketch showing the idea in Theorem \ref{thm:spatialconvergence}.}
1698: \label{fig:manyphis}
1699: \end{figure}
1700: Similarly, when the second integral in \eqref{eq:errorrepr} is
1701: estimated, the optimal path $\varphi(s)$ is used, and for each
1702: $\big(P \varphi(s),s\big)$ as starting positions the solution
1703: $\bar\lambda(s)$ to \eqref{eq:approxHamiltonsyst} is computed. 
1704: \begin{thm}\label{thm:spatialconvergence}
1705: For every constant $C>0$ and all starting points $\varphi_0$ with
1706: \begin{equation}\label{eq:startingposs}
1707: \varphi_0 = \sum_{i=1}^{M-1} \xi^i v^i \in V
1708: \end{equation}
1709: such that
1710: \begin{equation}\label{eq:startingposcond}
1711: \big( \Delta x \sum_{i=1}^{M-1} (D^2 \xi)_i^2 \big)^{1/2} \leq C,
1712: \end{equation}
1713: there is a constant $D>0$ such that
1714: \begin{equation}\label{eq:deltaxconv}
1715: |u(\varphi_0,0) - \bar u (\varphi_0,0)| \leq D \Delta x^2.
1716: \end{equation}
1717: \end{thm}
1718: \begin{remark}
1719: Every reasonable approximation in $V$ of $\varphi_+$, for any $\Delta x$,
1720: satisfies \eqref{eq:startingposcond}. The interpolant and the
1721: projection are possible choices.
1722: \end{remark}
1723: \begin{proof}
1724: As in the  proof of Theorem \ref{thm:diffcontrol}, whenever $D$ is
1725: written in this proof it means a constant independent of $\Delta x$, but (possibly) dependent on $C$.
1726:   
1727: \emph{Step 1.} It will be shown that condition \eqref{eq:startingposcond} implies
1728: \begin{equation}\label{eq:Afivefourths}
1729: ||A^{5/7} \varphi_0 || \leq D.
1730: \end{equation}
1731: Using \eqref{eq:Agamma} it follows that
1732: \begin{equation}\label{eq:Asum}
1733: ||A^\gamma \varphi || =\big( \sum_{n=1}^\infty k_n^{2\gamma} (\psi_n,\varphi)^2 \big)^{1/2},
1734: \end{equation}
1735: where $k_n$ and $\psi_n$ are the eigenvalues and eigenfunctions of
1736: $A$. Two partial integrations imply that
1737: \begin{equation*}
1738: (\psi_n,\varphi_0)= -\frac{1}{n^2 \pi^2} (\psi_n, (\varphi_0)_{xx}),
1739: \end{equation*}
1740: where $(\varphi_0)_{xx}$ is the distributional second derivative of
1741: $\varphi_0$, i.e.\ 
1742: \begin{equation*}
1743: (\varphi_0)_{xx} = \Delta x \sum_{i=1}^{M-1} (D^2 \xi)_i \bar\delta_{i\Delta x},
1744: \end{equation*}
1745: with $\bar\delta_{i\Delta x}$ the Dirac delta distribution in $x=i \Delta
1746: x$. Since $|\psi_n(x)| \leq \sqrt{2}$ for all $n \in \NatNum$ and all
1747: $x \in [0,1]$ it holds that
1748: \begin{align*}
1749: |(\psi_n,\varphi_0)| & \leq \frac{\sqrt{2}}{n^2 \pi^2} \Delta x
1750:  \sum_{i=1}^{M-1} |(D^2 \xi)_i| 
1751: \leq \frac{\sqrt{2}}{n^2 \pi^2} \Delta x \sqrt{M-1} \big(
1752:  \sum_{i=1}^{M-1} (D^2 \xi)_i^2 \big)^{1/2} \\ 
1753: &\leq \frac{\sqrt{2}}{n^2
1754:  \pi^2} \big( \Delta x \sum_{i=1}^{M-1} (D^2 \xi)_i^2 \big)^{1/2}
1755: \leq \frac{\sqrt{2} C}{n^2 \pi^2}.
1756: \end{align*}
1757: It thereby follows that the sum in \eqref{eq:Asum} is finite (and
1758: proportional to $C$) when $\gamma < 3/4$, e.g.\ $\gamma=5/7$.
1759: 
1760: \emph{Step 2.} In this step it is shown that there exists a solution
1761: $\lambda$ to \eqref{eq:mildsols} with $\bar\varphi(s)$ playing the role
1762: of $\varphi_0$, such that
1763: \begin{equation}\label{eq:Hdiff1}
1764: |H\big(P\lambda(s),\bar\varphi(s)\big)-H\big(\lambda(s),\bar\varphi(s)\big)| \leq  D \Delta
1765:  x^2 (T-s)^{-1/2}.
1766: \end{equation}
1767: We start by showing that the starting position $\varphi_0$ in
1768: \eqref{eq:startingposs} is bounded in $H_0^1(0,1)$:
1769: \begin{multline*}
1770: ||(\varphi_0)_x||^2 = -\big( (\varphi_0)_{xx},\varphi_0 \big) \\
1771: \leq D \big(\Delta x \sum_{i=1}^{M-1} (D^2 \xi)_i^2 \big)^{1/2}
1772: ||\varphi_0||_{L^\infty(0,1)} \leq D ||(\varphi_0)_x||, 
1773: \end{multline*}
1774: where the last inequality follows by a Sobolev inequality. Hence\linebreak $||(\varphi_0)_x||~ \leq~ D$.
1775: By Theorem \ref{thm:diffcontrol} it follows that
1776: $||\bar\varphi_x(s)|| \leq D$, for all $t_0 \leq s \leq T$. By Theorem \ref{thm:infinitedimregularity} it then
1777: follows that $||\lambda(s)||_{H^2(0,1)} \leq D (T-s)^{-1/2}$.    
1778: The Hamiltonian, $H$, consists of three parts; see
1779: \eqref{eq:Hamiltonian}. The first of these is the most difficult when
1780: \eqref{eq:Hdiff1} is to be proved, so we will focus on this one and 
1781: let the other two parts be treated by the reader. The difference
1782: between the first parts of the Hamiltonians in \eqref{eq:Hdiff1} is
1783: given by
1784: \begin{equation*}
1785: -\Big(\bar\varphi_x(s),\big(\lambda(s)-P\lambda(s)\big)_x\Big) = 
1786: \big(\bar\varphi_{xx}(s), \lambda(s)- P\lambda(s) \big),
1787: \end{equation*}
1788: where the factor $\delta$ is left out for convenience. We reuse
1789: notation and let 
1790: \begin{equation*}
1791: \bar\varphi(s) =: \sum_{i=1}^{M-1} \xi^i v^i,
1792: \end{equation*}
1793: so that
1794: \begin{equation*}
1795: \bar\varphi_{xx}(s) = \Delta x \sum_{i=1}^{M-1} (D^2 \xi)_i
1796: \bar\delta_{i\Delta x}.
1797: \end{equation*}
1798: As $||\bar\varphi_t(s)|| \leq D$ by Theorem \ref{thm:diffcontrol} it,
1799: again using \eqref{eq:FEMcoords},
1800: holds that 
1801: \begin{equation*}
1802: \big(\Delta x \sum_{i=1}^{M-1}(D^2 \xi)_i^2 \big)^{1/2} \leq D.
1803: \end{equation*} 
1804: Hence it holds that
1805: \begin{multline}\label{eq:beforeinterpolant}
1806: |\big(\bar\varphi_{xx}(s),\lambda(s)-P\lambda(s)\big)| \leq \Delta x 
1807: \sum_{i=1}^{M-1} |(D^2 \xi)_i \big(\lambda(i\Delta x)- P\lambda(i\Delta
1808: x)\big)| \\
1809: \leq \big(\Delta x \sum_{i=1}^{M-1}(D^2 \xi)_i^2 \big)^{1/2}
1810: \Big(\Delta x \sum_{i=1}^{M-1} \big(\lambda(i\Delta x) - P \lambda(i\Delta
1811: x)\big)^2 \Big)^{1/2} \\
1812: \leq D \Big(\Delta x \sum_{i=1}^{M-1} \big(\lambda(i\Delta x) - P \lambda(i\Delta
1813: x)\big)^2 \Big)^{1/2}.
1814: \end{multline}
1815: By the use of the interpolant the last parenthesis in
1816: \eqref{eq:beforeinterpolant} may be written
1817: \begin{equation*}
1818: \Big(\Delta x \sum_{i=1}^{M-1} \big(I \lambda(i\Delta x) - P \lambda(i\Delta
1819: x)\big)^2 \Big)^{1/2}.
1820: \end{equation*}
1821: This minor difference simplifies the situation as $I \lambda(s) -P
1822: \lambda(s) \in V$, which  makes comparison with the $L^2$ norm
1823: possible. For an element 
1824: \begin{equation*}
1825: \kappa=\sum_{i=1}^{M-1} \eta^i v^i \in V
1826: \end{equation*}
1827: it holds that 
1828: \begin{equation*}
1829: ||\kappa||^2 = \Delta x (\eta,B\eta)_2,
1830: \end{equation*}
1831: where $(\cdot,\cdot)_2$ is the Euclidean scalar product on $\Re^d$. All
1832: eigenvalues of $B$ lie in the interval $[1/3,1]$, and hence
1833: \begin{equation*}
1834: ||I\lambda -P\lambda||^2 \geq \third \Delta x \sum_{i=1}^{M-1} (\lambda(i\Delta x) - P \lambda(i\Delta
1835: x))^2. 
1836: \end{equation*}
1837: It also holds that
1838: \begin{equation*}
1839: ||I\lambda - P\lambda|| = ||P(I \lambda - \lambda)|| \leq ||I \lambda
1840:   - \lambda|| \leq D \Delta x^2 ||\lambda||_{H^2(0,1)},
1841: \end{equation*}
1842: where the last inequality may be found in e.g.\ \cite{Brenner-Scott}.
1843: By equation (\ref{eq:infinitedimregularity4}) we have that 
1844: \begin{equation*}
1845: ||\lambda(s)||_{H^2(0,1)} \leq D (T-s)^{-1/2},
1846: \end{equation*}
1847: and so \eqref{eq:Hdiff1} holds. By Theorem \ref{thm:HJsatisfied}, an
1848: element in $D^+ u_H\big(\bar\varphi(s),s\big)$ is given by
1849: $\Big(\lambda(s),-H\big(\lambda(s),\bar\varphi(s)\big)\Big)$, and
1850: thereby it is clear that the first
1851: integral in \eqref{eq:errorrepr}
1852: may be bounded by 
1853: \begin{equation*}
1854: D\Delta x^2\int_{t_0}^T (T-s)^{-1/2}ds \leq D\Delta x^2.
1855: \end{equation*}  
1856: 
1857: \emph{Step 3.} In this step a bound for the second integral in
1858: \eqref{eq:errorrepr} is derived. In \emph{Step 1} it was proved that 
1859: $||A^{5/7} \varphi_0|| \leq D$. Theorem \ref{thm:infinitedimregularity}
1860: then implies that $||\varphi(s)||_{H^2(0,1)} \leq
1861: D(s-t_0)^{-2/7}$. Therefore, by Lemma \ref{lem:H2proj} and  Theorem \ref{thm:diffcontrol} there
1862: exists a solution $\bar\lambda$ to \eqref{eq:approxHamiltonsyst}, with
1863: $\big(P\varphi(s),s \big)$ in the role of $(\varphi_0,t_0)$, such that
1864: \begin{equation*}
1865: ||\bar\lambda_t(s)|| \leq \big( D(s-t_0)^{-2/7} +F \big)^2 \leq D(s-t_0)^{-4/7}.
1866: \end{equation*}
1867: By \eqref{eq:FEMcoords2} it holds that
1868: \begin{equation*}
1869: \Big( \Delta x \sum_{i=1}^{M-1} \big(D^2 \theta(s)\big)_i^2
1870: \Big)^{1/2} \leq D(s-t_0)^{-4/7},
1871: \end{equation*}  
1872: where $\theta$ is given by \eqref{eq:coorexpr} and
1873: \eqref{eq:vectdefs}.
1874: In order to be able to use the above information to get a bound of the
1875: second integral in \eqref{eq:errorrepr}, we need that $\bar\lambda(s)$
1876: is the spatial part of an element in $D^+ \bar u(P\varphi(s),s)$. This
1877: follows from Lemma 3.3.16 and Theorem 7.4.17 in
1878: \cite{Cannarsa-Sinestrari}. As in \emph{Step 2} we are satisfied with
1879: considering only the first parts of the Hamiltonians. The only
1880: difference is that now the partial integration is performed so that
1881: $\bar\lambda(s)$ is distributionally differentiated twice:
1882: \begin{multline*}
1883: \Big| \Big(\bar\lambda_x(s),\big(\varphi(s)-P\varphi(s)\big)_x \Big) \Big| =
1884: \big| \big(\bar\lambda_{xx}(s),\varphi(s)-P\varphi(s)\big)\big| \\
1885: \leq \Big( \Delta x \sum_{i=1}^{M-1} \big(D^2 \theta(s)\big)_i^2
1886: \Big)^{1/2} \Big(\Delta x \sum_{i=1}^{M-1}\big(\varphi(i\Delta
1887:  x,s)-P\varphi(i\Delta x,s) \big)^2 \Big)^{1/2} \\
1888: \leq D(s-t_0)^{-4/7} \Delta x^2 ||\varphi_{xx}(s)|| \leq D \Delta x^2 (s-t_0)^{6/7},
1889: \end{multline*}
1890: similarly as in \emph{Step 2}. The second integral in
1891: \eqref{eq:errorrepr} is  therefore bounded by a term
1892: \begin{equation*}
1893: D \Delta x^2 \int_{t_0}^T (s-t_0)^{-6/7}ds = D \Delta x^2.
1894: \end{equation*} 
1895: 
1896: \emph{Step 4.} It remains to show that the difference 
1897: $g(P\varphi(T))-g(\varphi(T))$ is of the order $\Delta x^2$. Since
1898: $||\varphi(T)||$ is uniformly bounded for all starting positions in a
1899: bounded set in $L^2(0,1)$ we see by \eqref{eq:gdifference} that the difference in final
1900: costs is less than $D||P\varphi(T)-\varphi(T)||$. Since $P I \varphi=I\varphi$,
1901: where $I$ is the interpolant, introduced in \emph{Step2}, we have that
1902: \begin{multline*}
1903: ||P\varphi(T)-\varphi(T)|| \leq ||P\big(\varphi(t)-I\varphi(T)\big)|| +
1904:   ||I\varphi(T)-\varphi(T)|| \\
1905: \leq 2 ||I\varphi(T)-\varphi(T)|| \leq D \Delta x^2
1906:   ||\varphi_{xx}(T)|| \leq D \Delta x^2,
1907: \end{multline*}
1908: where the last inequality follows by  Theorem \ref{thm:infinitedimregularity}. 
1909: \end{proof}
1910: The next theorem provides an error estimate which makes comparison
1911: with the situation where $\varphi_+$ is used as initial position
1912: possible.
1913: \begin{thm}\label{thm:comparisonphi+}
1914: There exists a constant $D>0$ such that 
1915: \begin{equation}\label{eq:comparisonphi+}
1916: |u(\varphi_+,0)-u(P\varphi_+,0)| + |u(\varphi_+,0)-u(I\varphi_+,0)|
1917:  \leq D \Delta x^2.
1918: \end{equation}
1919: \end{thm}
1920: \begin{remark}
1921: The theorem shows that both the projection and the interpolant can be
1922: chosen when approximating $\varphi_+$ in $V$.
1923: \end{remark}
1924: \begin{proof}
1925: The semiconcavity of $u_H$ implies that for every bounded set $X
1926: \subset H_0^1(0,1)$ there exists a constant $D$, such that 
1927: \begin{equation}\label{eq:beginningconcave}
1928: \varphi^1, \varphi^2 \in X \implies u_H(\varphi^1,0) -u_H(\varphi^2,0)
1929: - (p, \varphi^1-\varphi^2) \leq D||\varphi^1 - \varphi^2||^2_{H_0^1(0,1)},
1930: \end{equation}
1931: where $p$ is the spatial part of any element in $D^+ u_H(\varphi^2,0)$
1932: (compare \eqref{eq:semiconmodulus}). In Theorem \ref{thm:HJsatisfied}
1933: it was proved that $\Big(
1934: \lambda(0),-H\big(\lambda(0),\varphi^1\big)\Big)$ is one such element,
1935: where $\lambda$ is a solution to \eqref{eq:mildsols} with
1936: $\varphi_0=\varphi^2$. We may therefore take $p=\lambda(0)$ in
1937: \eqref{eq:beginningconcave}. By Theorem
1938: \ref{thm:infinitedimregularity}, $||p|| \leq D$ for some constant $D$
1939: (it is even bounded in $H_0^1(0,1)$, but this is not needed here).
1940: Plugging this boundedness into \eqref{eq:beginningconcave} results in
1941: the inequality
1942: \begin{equation*}
1943: u_H(\varphi^1,0)-u_H(\varphi^2,0) \leq D\big(||\varphi^1 - \varphi^2||
1944: + ||\varphi^1-\varphi^2||^2_{H_0^1(0,1)}\big).
1945: \end{equation*}
1946: We may change places for $\varphi^1$ and $\varphi^2$ everywhere above,
1947: and thereby obtain
1948: \begin{equation}\label{eq:beginningsemiconres}
1949: |u_H(\varphi^1,0)-u_H(\varphi^2,0)| \leq D\big(||\varphi^1 - \varphi^2||
1950: + ||\varphi^1-\varphi^2||^2_{H_0^1(0,1)}\big).
1951: \end{equation}
1952: Consider now $\varphi^1 = \varphi_+$ and $\varphi^2 = I\varphi_+$ or
1953: $\varphi^2 = P \varphi_+$. For the interpolant, $I$, the following
1954: bounds hold:
1955: \begin{equation}\label{eq:interpbounds}
1956: \begin{split}
1957: ||\varphi_+ - I \varphi_+|| &\leq
1958:   D||(\varphi_+)_{xx}||_{L^\infty(0,1)} \Delta x^2, \\
1959: ||\varphi_+ - I \varphi_+||_{H_0^1(0,1)} &\leq
1960:   D||(\varphi_+)_{xx}|| \Delta x,
1961: \end{split}
1962: \end{equation}
1963: with a constant $D$ independent of $\Delta x$ and $\varphi_+$. Since
1964: $(\varphi_+)_{xx}$ is bounded in $L^\infty(0,1)$ this together with
1965: \eqref{eq:beginningsemiconres} directly shows that the interpolant
1966: part of \eqref{eq:comparisonphi+} is correct. The projection part is
1967: proved by using the result from \cite{Crouzeix-Thomee}, that the $L^2$
1968: projection is stable in $H_0^1(0,1)$, i.e.
1969: \begin{equation*}
1970: ||P \varphi||_{H_0^1(0,1)} \leq D ||\varphi||_{H_0^1(0,1)}.
1971: \end{equation*}
1972: It therefore holds that 
1973: \begin{multline}\label{eq:stableproj}
1974: ||\varphi_+ - P\varphi_+||_{H_0^1(0,1)} \leq ||\varphi_+ -
1975:   I\varphi_+||_{H_0^1(0,1)} +||I \varphi_+ - P\varphi_+||_{H_0^1(0,1)} \\
1976:   = ||\varphi_+ - I\varphi_+||_{H_0^1(0,1)} + ||P(I\varphi_+ -
1977:   \varphi_+)||_{H_0^1(0,1)} \leq (1+D)||\varphi_+ - I\varphi_+||_{H_0^1(0,1)}.
1978: \end{multline}
1979: The same technique as in \eqref{eq:stableproj} can also be used for
1980: the $L^2$ norm, now using the obvious bound $||P\varphi|| \leq
1981: ||\varphi||$, which implies that
1982: \begin{equation}\label{eq:stableproj2}
1983: ||\varphi_+ - P \varphi_+|| \leq 2 ||\varphi_+ - I\varphi_+||.
1984: \end{equation}
1985: The equations \eqref{eq:beginningsemiconres}, \eqref{eq:interpbounds},
1986: \eqref{eq:stableproj} and \eqref{eq:stableproj2} imply that also the
1987: projection part of \eqref{eq:comparisonphi+} is correct.
1988: \end{proof}
1989: Theorems \ref{thm:spatialconvergence} and \ref{thm:comparisonphi+}
1990: directly imply the following corollary.
1991: \begin{cor}\label{cor:truespaceconv}
1992: There exists a constant $D$, such that
1993: \begin{equation*}
1994: |u(\varphi_+,0) - \bar u(P\varphi_+,0)| + |u(\varphi_+,0) - \bar
1995:  u(I\varphi_+,0)| \leq D \Delta x^2.
1996: \end{equation*}
1997: \end{cor}
1998: \section{Discretization in time}\label{sec:timediscr}
1999: In \cite{Sandberg-Szepessy} the method \emph{Symplectic Pontryagin} for
2000: approximation of optimally controlled ODE:s is constructed and
2001: analyzed. It is a Symplectic Euler discretization for a Hamiltonian
2002: system, involving the state and
2003: dual variables associated with the control problem, with a regularized
2004: Hamiltonian. In the present situation, when the Hamiltonian is smooth,
2005: the Symplectic Pontryagin method reduces to ordinary Symplectic Euler,
2006: since no need for regularization exists. The theory in
2007: \cite{Sandberg-Szepessy} can be used to show that the difference
2008: between the value function for a system with only spatial
2009: discretization and the value function for a system with discretization
2010: in space \emph{and} time, is of the order $\Delta t$, where $\Delta t$
2011: is the size of the temporal discretization. 
2012: It is, however, desirable to achieve more than this. In order for the
2013: estimate on the temporal discretization to be useful the constant in
2014: front of $\Delta t$ in the error estimate needs to be really constant,
2015: i.e.\ independent of $\Delta x$.
2016: 
2017: The theorems in \cite{Sandberg-Szepessy} do not directly provide the
2018: desired result. This has to do with the fact that the second order
2019: difference quotient operator $D^2$, defined in
2020: \eqref{eq:diffquotient}, has norm proportional to $1/\Delta
2021: x^2$. Furthermore, the proof in \cite{Sandberg-Szepessy} requires a
2022: bound on the derivatives
2023: $\partial\tilde\lambda^{n+1}/\partial\tilde\varphi^n$, where
2024: $\tilde\varphi$ and $\tilde\lambda$ are obtained with the Symplectic
2025: Pontryagin method. The problem of large norm of $D^2$ can be handled
2026: using that it is a negative operator. But in addition to this we also
2027: need to bound $\partial\tilde\lambda^{n+1}/\partial\tilde\varphi^n$
2028: independently of $\Delta x$ in some proper sense.
2029: 
2030: The proof of convergence of the Symplectic Euler method given here is
2031: based on another technique. It uses that the present problem admits
2032: optimal controls which are regular by Theorem
2033: \ref{thm:diffcontrol}. It also involves an assumption about the
2034: derivative $\tilde\varphi_x$, and another similar assumption. Under
2035: these assumptions it is shown in Theorem \ref{thm:timediscrconv} that
2036: a minimum of a  forward Euler approximation of
2037: control problem \eqref{eq:approxu}, \eqref{eq:approxphievol} has an
2038: error $C\Delta t$ in the objective, where $C$ does not depend on
2039: $\Delta x$. In Theorem \ref{thm:sympleuler} it is shown that the
2040: solution to this minimization problem is equivalent to the solution of
2041: a Symplectic Euler scheme, and hence the desired property for the
2042: Symplectic Euler scheme is achieved. The main difference in the result
2043: when the present method is used compared to a result using the theory
2044: in \cite{Sandberg-Szepessy} is that the present result needs an
2045: assumption on the derivative $\tilde\varphi_x$ whereas
2046: \cite{Sandberg-Szepessy} needs control over
2047: $\partial\tilde\lambda^{n+1}/\partial \tilde \varphi^n$. The
2048: assumptions on  $\tilde\varphi_x$ seem easier to verify. The numerical
2049: tests performed in Section \ref{sec:NumRes} support that it is true. 
2050: 
2051: 
2052: 
2053: We now present the setting of the aforementioned discretized
2054: optimization problem. Consider the time-discrete state
2055: $\{\tilde\varphi^n\}_{n=0}^N$, which is a forward Euler approximation of
2056: the state $\bar\varphi$ in \eqref{eq:approxphievol} and is given by
2057: \begin{equation}\label{eq:timediscretephievol}
2058: (\tilde\varphi^{n+1},v)=(\tilde\varphi^n,v)+\Delta t \big( -\delta
2059:   (\tilde\varphi^n_x,v_x) +(-\delta^{-1}V'(\tilde\varphi^n) +
2060:   \tilde\alpha^n,v)\big), \quad \text{for all } v\in V,
2061: \end{equation} 
2062: where $\{\tilde\alpha^n\}_{n=0}^{N-1}$ is a time-discrete control. The
2063: discrete state $\tilde\varphi^n$ therefore corresponds to
2064: $\bar\varphi(t_n)$, where $t_n=n \frac{T}{N} \equiv n \Delta t$. By
2065: \eqref{eq:timediscretephievol} it is possible to define a discrete value
2066: function for all times $t_m$:
2067: \begin{equation}\label{eq:timediscretevalue}
2068: \tilde u (\tilde\varphi_0,t_m)=\min_{\{\tilde\alpha^n\}_{n=m}^{N-1}}
2069: \big( g(\tilde\varphi_N)+ \Delta t \sum_{n=m}^{N-1} h(\tilde \alpha^n)\big),
2070: \end{equation}
2071: where $\{\tilde\varphi^n\}$ solves \eqref{eq:timediscretephievol} and
2072: $\tilde\varphi^m=\tilde\varphi_0$. For the proof of Theorem
2073: \ref{thm:timediscrconv} we also introduce the 
2074: discrete state $\{\mathring\varphi^n\}_{n=0}^{N}$. It
2075: is also given by a forward Euler time stepping scheme, but its
2076: evolution is determined by an optimal control $\bar\alpha$ to the
2077: time-continuous problem \eqref{eq:approxphievol}:
2078: \begin{equation}\label{eq:timediscretephievol2}
2079: (\mathring\varphi^{n+1},v)=(\mathring\varphi^n,v)+\Delta t \big( -\delta
2080:   (\mathring\varphi^n_x,v_x) +(-\delta^{-1}V'(\mathring\varphi^n) +
2081:   \bar\alpha(t_n),v)\big), \ \text{for all } v\in V.
2082: \end{equation}
2083: We will consider starting positions $\bar\varphi_0$ in finite
2084: element spaces $V$ satisfying
2085: \begin{equation}\label{eq:startingpositions}
2086: \delta\big((\bar\varphi_0)_x,v_x\big)+\delta^{-1}\big(V'(\bar\varphi_0\big),v)=0, \quad
2087: \text{for all } v \in V.
2088: \end{equation}
2089: We are now ready for the theorem on time discretization convergence.
2090: \begin{thm}\label{thm:timediscrconv}
2091: Assume there exists a function $r:\Re^+ \rightarrow \Re^+$ such that
2092: for all $\Delta t \leq r(\Delta x)$ there are solutions
2093: $\{\tilde\varphi^n\}_{n=0}^N$ and $\{\mathring\varphi^n \}_{n=0}^N$ with 
2094: $\tilde\varphi^0=\mathring\varphi^0=\bar\varphi_0$, where $\bar\varphi_0$
2095: satisfies \eqref{eq:startingpositions}, and 
2096: \begin{equation}\label{eq:gradincrease}
2097: ||\tilde\varphi^{n+1}_x-\tilde\varphi^n_x|| +
2098:   ||\mathring\varphi^{n+1}_x-\mathring\varphi^n_x|| \leq C \Delta t 
2099: \end{equation}
2100: for all $0 \leq n < N$, where $C$ does not depend on $\Delta x$. Then
2101: \begin{equation*}
2102: |\tilde u(\bar\varphi_0,0)- \bar u(\bar\varphi_0,0) | \leq D \Delta t
2103: \end{equation*}
2104: for $\Delta t \leq r(\Delta x)$, where $D$ does not depend on $\Delta x$.
2105: \end{thm}
2106: \begin{remark}
2107: By the numerical computations performed in Section \ref{sec:NumRes} it
2108: seems plausible that \eqref{eq:gradincrease} holds.
2109: \end{remark}
2110: \begin{remark}
2111: The proof would be valid without inclusion of the function $r$. 
2112: However, since the forward Euler method is used it seems reasonable to
2113: believe that \eqref{eq:gradincrease} would not be valid for all
2114: $\Delta t$.
2115: \end{remark}
2116: \begin{proof}
2117: As for Theorem \ref{thm:valueerror} the proof is divided into two
2118: steps. We obtain in the first step a lower bound for $\tilde
2119: u(\bar\varphi_0,0)- \bar u (\bar\varphi_0,0)$, and in the second step a
2120: corresponding upper bound. The first step in this proof is similar to
2121: the first step in the proof of Theorem \ref{thm:valueerror}, while the
2122: corresponding second steps differ. We denote an optimal pair (control
2123: and state) for $\bar u$ by $\bar\alpha$ and $\bar\varphi$, and an optimal
2124: pair for $\tilde u$ by $\{\tilde\alpha^n\}$ and $\{\tilde\varphi^n \}$.
2125: 
2126: \emph{Step 1.} This part of the proof starts by an extension of
2127: the initially time-discrete state $\{\tilde\varphi^n\}$ to a piecewise
2128: linear time-continuous function $\tilde\varphi : [0,T] \rightarrow V$ as
2129: follows:
2130: \begin{equation*}
2131: \tilde\varphi(t) \equiv \frac{t_{n+1}-t}{\Delta t}\tilde\varphi^n +
2132: \frac{t-t_n}{\Delta t} \tilde\varphi^{n+1}, \quad \text{for } t_n \leq t
2133: \leq t_{n+1}.
2134: \end{equation*}
2135: As in the proof of Theorem \ref{thm:valueerror} we have 
2136: \begin{equation}\label{eq:udiff}
2137: \tilde u(\bar\varphi_0,0)-\bar u (\bar\varphi_0,0) = \int_0^T \frac{d}{ds}
2138: \bar u (\tilde\varphi(s),s)ds + \Delta t \sum_{i=0}^{N-1} h(\tilde \alpha^i).
2139: \end{equation}
2140: In order to be able to use that $\bar u$ solves a Hamilton-Jacobi
2141: equation we note that the right hand side in \eqref{eq:udiff} may be
2142: written
2143: \begin{equation*}
2144: \sum_{i=0}^{N-1} \int_{t_i}^{t_{i+1}} \big(\frac{d}{ds} \bar u
2145: (\tilde\varphi(s),s)+h(\tilde\alpha^i) \big)ds,
2146: \end{equation*}
2147: and thus we may focus our attention on one time interval
2148: $[t_n,t_{n+1}]$.
2149: We also note that equation \eqref{eq:approxphievol} defines a flow
2150: $\bar f:V\times V \rightarrow V$ which is defined by
2151: \begin{equation}\label{eq:fbar}
2152: (\bar f(\bar\varphi,\bar\alpha),v)= -\delta(\bar\varphi_x,v_x) +
2153:   (\delta^{-1} V'(\bar\varphi)+\bar\alpha,v), \quad \text{for all } v \in V.
2154: \end{equation} 
2155: %Similarly as the further development in the proof of Theorem
2156: %\ref{thm:valueerror} the time derivative in \eqref{eq:udiff} is for
2157: %$t_n < s < t_{n+1}$
2158: %\begin{align*}
2159: %\frac{d}{ds}\bar u(\tilde\varphi(s),s) &= \min_{p \in D^+ \bar u
2160: %  (\tilde\varphi(s),s)} \Big(p_t + \big( p_{\varphi},\bar
2161: %  f(\tilde\varphi^n,\tilde\alpha^n)\big)\Big) \\
2162: % & \equiv p^*_t(s) + \big( p^*_\varphi (s), \bar f(\tilde\varphi^n,\tilde\alpha^n)\big).
2163: %\end{align*}
2164: Let now $p(s)=\big(p_\varphi(s),p_t(s)\big)$ be any element in $D^+
2165: \bar u\big(\tilde\varphi(s),s\big)$.
2166: Similarly as in the proof of Theorem \ref{thm:valueerror} we have for
2167: almost every  $s \in [t_n,t_{n+1}]$
2168: \begin{equation}\label{eq:fdiff}
2169: \begin{split}
2170: &\frac{d}{ds}\bar u (\tilde\varphi(s),s) +
2171:   h(\tilde\alpha^i)
2172: \geq p_t(s) + \big(p_\varphi(s),\bar f(\tilde\varphi^n,\tilde\alpha^n)\big) 
2173: \\
2174: & =p_t(s) +
2175: \underbrace{\big( p_\varphi (s), \bar f(\tilde\varphi(s),\tilde\alpha^n)\big) +
2176: h(\tilde\alpha^n)}_{\geq H(p_{\varphi}(s),\tilde\varphi(s))} + \big(p_\varphi(s),\bar
2177: f(\tilde\varphi^n,\tilde\alpha^n)-\bar f(\tilde\varphi(s),\tilde\alpha^n)
2178: \big) \\
2179: &\geq \big(p_\varphi(s),\bar
2180: f(\tilde\varphi^n,\tilde\alpha^n)-\bar f(\tilde\varphi(s),\tilde\alpha^n)
2181: \big),
2182: \end{split}
2183: \end{equation}
2184: since $p_t(s)+H(p_\varphi(s),\tilde\varphi(s)) \geq 0$ as $\bar u $ is a
2185: Hamilton-Jacobi viscosity solution. By assumption
2186: \eqref{eq:gradincrease} it follows that $\tilde\varphi_x(s)$ is bounded
2187: for $0 \leq s \leq T$ independently of $\Delta x$. We are therefore
2188: free to use $\tilde\varphi(s)$  as
2189: $\bar\varphi_0$ in Theorem \ref{thm:diffcontrol} so that the
2190: $\bar\lambda(s)$ (corresponding to $\bar u (\tilde\varphi(s),s)$) is
2191: bounded in $H_0^1$ independently of $\Delta x$. For such a
2192: $\bar\lambda(s)$ we have
2193: \begin{align*}
2194:  & \big | \big(\bar\lambda(s),\bar f(\tilde\varphi^n,\tilde\alpha^n)-\bar
2195:  f(\tilde\varphi(s),\tilde\alpha^n)\big) \big| \\
2196: & \leq \delta \big| \big(\bar\lambda_x(s),\tilde\varphi^n_x-\tilde\varphi_x(s)\big) \big| +
2197:  \delta^{-1} \big|\big(V'(\tilde\varphi^n)-V'(\tilde\varphi(s)),\bar\lambda(s)\big)\big|
2198:  \leq C \Delta t,
2199: \end{align*} 
2200: with $C$ independent of $\Delta x$ by \eqref{eq:gradincrease}. 
2201: It is now used that $\bar\lambda(s)$ is the spatial part of an element
2202: in $D^+ \bar u(\tilde\varphi(s),s)$.
2203: It thereby   holds that the right hand side in \eqref{eq:fdiff}
2204: is less than $C\Delta t$ in magnitude.
2205: 
2206: \emph{Step 2.} We start by noting that
2207: \begin{equation}\label{eq:bound2}
2208: \bar u(\bar\varphi_0,0)-\tilde u (\bar\varphi_0,0) \geq
2209: g(\bar\varphi(T))+\int_0^T h(\bar\alpha)dt - \big(
2210: g(\mathring\varphi^N)+\Delta t \sum_{i=0}^{N-1} h(\bar\alpha(t_i)) \big).
2211: \end{equation}
2212: The difference between the running costs in \eqref{eq:bound2} is 
2213: \begin{equation*}
2214: \sum_{i=0}^{N-1} \int_{t_i}^{t_{i+1}} \big( h(\bar\alpha(t))-h(\bar\alpha(t_n))\big)dt.
2215: \end{equation*}
2216: Using that $h(\alpha)=||\alpha||^2/2$ we have that 
2217: \begin{multline*}
2218: |h(\bar\alpha(t))-h(\bar\alpha(t_n))| = \half
2219:  |(\bar\alpha(t)+\bar\alpha(t_n),\bar\alpha(t)-\bar\alpha(t_n))| \\
2220: \leq
2221:  \half
2222:  ||\bar\alpha(t)+\bar\alpha(t_n)||\cdot||\bar\alpha(t)-\bar\alpha(t_n)|| \leq C\Delta t,
2223: \end{multline*}
2224: where we have used the result in Theorem \ref{thm:diffcontrol} on the
2225: boundedness of the control and its derivative (remember that
2226: $\bar\alpha= - \bar\lambda$). 
2227: It remains to show that the difference between the terminal costs in
2228: \eqref{eq:bound2} behaves similarly. As in \emph{Step 1} we now extend
2229: the discrete state $\{\mathring \varphi^n\}$ to a continuous function:
2230: \begin{equation*}
2231: \mathring\varphi(t) \equiv \frac{t_{n+1}-t}{\Delta t}\mathring\varphi^n +
2232: \frac{t-t_n}{\Delta t} \mathring\varphi^{n+1}, \quad \text{for } t_n \leq t
2233: \leq t_{n+1}.
2234: \end{equation*}
2235: For $t_n < t < t_{n+1}$ the evolution equations for $\bar\varphi$ and
2236: $\mathring\varphi$ look as follows:
2237: \begin{align*}
2238: (\bar\varphi_t,v)&=-\delta(\bar\varphi_x,v_x)+(-\delta^{-1}V'(\bar\varphi)
2239:   +\bar\alpha,v), \\
2240: (\mathring\varphi_t,v)&=-\delta(\mathring\varphi_x^n,v_x)+(-\delta^{-1}V'(\mathring\varphi^n)
2241:   +\bar\alpha(t_n),v), 
2242: \end{align*}
2243: for all $v \in V$. Subtract these two equations and let
2244: $v=\bar\varphi-\mathring\varphi$ to get:
2245: \begin{align*}
2246: &\half\frac{d}{dt} ||\bar\varphi - \mathring\varphi||^2 \\
2247: &= -\delta
2248: (\bar\varphi_x - \mathring\varphi_x^n,\bar\varphi_x -\mathring\varphi_x) +\delta^{-1}(V'(\mathring\varphi^n)-V'(\bar\varphi),\bar\varphi -\mathring\varphi)\\
2249: &\quad\quad\quad + (\bar\alpha - \bar\alpha(t_n),\bar\varphi -\mathring\varphi) \\
2250: &=-\delta ||\bar\varphi_x-\mathring\varphi_x||^2 +
2251: \delta(\mathring\varphi_x^n-\mathring\varphi_x,\bar\varphi_x-\mathring\varphi_x)
2252: \\
2253: &\quad\quad\quad +\delta^{-1}(V'(\mathring\varphi^n)-V'(\bar\varphi),\bar\varphi -\mathring\varphi)
2254: + (\bar\alpha - \bar\alpha(t_n),\bar\varphi -\mathring\varphi) \\
2255: & \leq -\delta ||\bar\varphi_x-\mathring\varphi_x||^2 +
2256: \delta\frac{||\mathring\varphi^n_x - \mathring\varphi_x||^2}{2} +
2257: \delta\frac{||\bar\varphi_x - \mathring\varphi_x||^2}{2} \\
2258: &\quad + \delta^{-1} |V''|
2259: \cdot ||\mathring\varphi^n - \mathring\varphi + \mathring\varphi-\bar\varphi||\cdot
2260: ||\bar\varphi- \mathring\varphi|| + \frac{||\bar\alpha-
2261:   \bar\alpha(t_n)||^2}{2} + \frac{||\bar\varphi -\mathring\varphi||^2}{2} \\
2262: & \leq \delta\frac{||\mathring\varphi^n_x - \mathring\varphi_x||^2}{2} +\frac{\delta^{-1}|V''|}{2}||\mathring\varphi^n - \mathring\varphi||^2 \\
2263: &\quad\quad\quad +\frac{\delta^{-1}|V''| + 1}{2} ||\bar\varphi -\mathring\varphi||^2 + \frac{||\bar\alpha-
2264:   \bar\alpha(t_n)||^2}{2}.
2265: \end{align*}
2266: According to a Poincar\'e inequality (see e.g.\ Theorem 5.3.5 in
2267: \cite{Brenner-Scott}), using the Dirichlet conditions, we have that 
2268: $||\mathring\varphi^n - \mathring\varphi|| \leq C||\mathring\varphi^n_x -
2269: \mathring\varphi_x||$. If now Gr\"onwall's Lemma is  used together with
2270: the fact that
2271: $||\bar\alpha-\bar\alpha^n||+||\mathring\varphi^n_x - \mathring\varphi_x||
2272: \leq C \Delta t$, we have that $||\bar\varphi(T)-\mathring\varphi(T)||\leq
2273: \Delta t$. Since $g(\bar\varphi(T))$ is bounded independently of $\Delta
2274: x$ we have, similarly as in the proof of Theorem
2275: \ref{thm:boundedcontrol},  that
2276: $|g(\bar\varphi(T))-g(\mathring\varphi(T))|\leq C \Delta t$.
2277: \end{proof}
2278: As convergence of the forward Euler method has now been proved, the
2279: \emph{Symplectic Euler method}, which
2280: can be used to find the forward Euler solution, is now presented. It
2281: is given by the system
2282: \begin{equation}\label{eq:sympleuler}
2283: \begin{split}
2284: (\tilde\varphi^{n+1},v) &= (\tilde\varphi^n,v) + \Delta t H_\lambda
2285:   (\tilde\lambda^{n+1},\tilde\varphi^n;v) \\
2286: &= (\tilde\varphi^n,v)+\Delta t
2287:   \big( -\delta (\tilde\varphi^n_x,v_x)- \delta^{-1}(V'(\tilde\varphi^n),v) 
2288:   - (\tilde\lambda^{n+1},v)\big), \\
2289: \tilde\varphi^0 & =\bar\varphi_0, \\
2290: (\tilde\lambda^n,v) &= (\tilde\lambda^{n+1},v) + \Delta t
2291:   H_\varphi(\tilde\lambda^{n+1},\tilde\varphi^n;v) \\
2292: & = (\tilde\lambda^{n+1},v)
2293:   + \Delta t
2294:   \big(-\delta(\tilde\lambda^{n+1}_x,v_x)-\delta^{-1}(\tilde\lambda^{n+1}V''(\tilde\varphi^n),v)\big),
2295:   \\
2296: \tilde\lambda^N & = g'(\tilde\varphi^N)=2K(\tilde\varphi^N-P \varphi_-),
2297: \end{split}
2298: \end{equation} 
2299: where $g'$ is a G\^ateaux derivative and $H_\lambda(\cdot;v)$,
2300: $H_\varphi(\cdot;v)$ are G\^ateaux derivatives in the direction $v$. For
2301: every minimizer $\{\tilde\alpha^n\}$ in  \eqref{eq:timediscretevalue}
2302: there exists a solution to \eqref{eq:sympleuler} with
2303: $\tilde\lambda^{n+1} = -\tilde\alpha^n$ for all $n$. In order to prove
2304: this we first state a lemma.
2305: \begin{lemma}\label{lem:discretesemiconcavity}
2306: The value function $\tilde u(\cdot,t_n)$ is semiconcave for every $n$.
2307: \end{lemma}
2308: \begin{proof}
2309: Consider the starting positions $\tilde\varphi^0_1$, $\tilde\varphi^0_2$ and
2310: $\frac{\tilde\varphi^0_1+\tilde\varphi^0_2}{2}$ at time $0$. The
2311: time-discrete cost functional $\tilde v$ is introduced:
2312: \begin{equation*}
2313: \tilde v_{\tilde\varphi_0,t_m}(\{\tilde\alpha^n\}) = \big( g(\tilde\varphi_N)+ \Delta t \sum_{n=m}^{N-1} h(\tilde \alpha^n)\big),
2314: \end{equation*}
2315: where $\{\tilde\varphi^n\}$ solves \eqref{eq:timediscretephievol} and
2316: $\tilde\varphi_m=\tilde\varphi_0$.
2317: Let $\{\tilde\alpha^n\}$ be an optimal control for the starting
2318: position
2319: $(\frac{\tilde\varphi^0_1+\tilde\varphi^0_2}{2},0)$. We can thus write
2320: \begin{multline*}
2321: \tilde u(\tilde\varphi^0_1,0)+\tilde u(\tilde\varphi^0_2,0) - 2\tilde
2322: u(\frac{\tilde\varphi^0_1+\tilde\varphi^0_2}{2},0) \\ 
2323: \leq \tilde
2324: v_{\tilde\varphi^0_1,0}(\{\tilde\alpha^n\}) + \tilde
2325: v_{\tilde\varphi^0_2,0}(\{\tilde\alpha^n\}) -2 \tilde
2326: v_{\frac{\tilde\varphi^0_1+\tilde\varphi^0_2}{2},0}(\{\tilde\alpha^n\}).
2327: \end{multline*}
2328: The states starting in $\tilde\varphi^0_1$, $\tilde\varphi^0_2$ and
2329: $\frac{\tilde\varphi^0_1+\tilde\varphi^0_2}{2}$, all using the control
2330: $\{\tilde\alpha^n\}$, are called
2331: \begin{equation*}
2332: \tilde\varphi^n_1 \equiv \sum_{i=1}^{M-1}\xi^n_{1,i} v^i,\  \tilde\varphi^n_2
2333: \equiv \sum_{i=1}^{M-1}\xi^n_{2,i} v^i, \text{ and } \tilde\varphi^n_3 \equiv \sum_{i=1}^{M-1}\xi^n_{3,i} v^i.
2334: \end{equation*}
2335: Introducing the notation
2336: \begin{equation*}
2337: \xi^n_m= \begin{pmatrix} \xi^n_{m,1} \\ \vdots \\ \xi^{n}_{m,M-1}
2338: \end{pmatrix}, \quad
2339: p^n_m = \begin{pmatrix} \big(V'(\tilde\varphi_m^n),v^1\big) \\ \vdots \\
2340:   \big(V'(\tilde\varphi_m^n),v^{M-1}\big)\end{pmatrix}, \quad
2341: a^n = \begin{pmatrix} a^n_1 \\ \vdots \\ a^{n}_{M-1} \end{pmatrix},
2342: \end{equation*}
2343: where $m$ can be 1, 2, or 3 and
2344: \begin{equation*}
2345: \tilde\alpha^n \equiv \sum_{i=1}^{M-1}a^n_i v^i,
2346: \end{equation*}
2347: we can, using the mass matrix $B$ in \eqref{eq:massmatrix} and the second difference
2348: operator $D^2$ in \eqref{eq:diffquotient}, write the equation for $\tilde \varphi^n_m$,
2349: $m=1,2,3$, as follows:
2350: \begin{equation}\label{eq:xistates}
2351: B \xi^{n+1}_m = B \xi^n_m + \Delta t \big(\delta D^2 \xi^n_m -
2352: \frac{\delta^{-1}}{\Delta x}p^n_m + B a^n\big).
2353: \end{equation} 
2354: Introducing the state $z^n = \xi^n_1 + \xi^n_2 - 2\xi^n_3$ and
2355: using \eqref{eq:xistates} gives
2356: \begin{equation}\label{eq:discrzevol}
2357: B z^{n+1}=B z^n +\Delta t\big(\delta D^2 z^n -
2358: \frac{\delta^{-1}}{\Delta x}(p^n_1+p^n_2-2p^n_3)\big).
2359: \end{equation}
2360: Every element in the vector 
2361: \begin{equation*}
2362: p^n_1+p^n_2-2p^n_3 = \begin{pmatrix} \big( V'(\tilde\varphi^n_1) +
2363:   V'(\tilde\varphi^n_2) -2 V'(\tilde\varphi^n_3),v^1 \big) \\ \vdots \\ \big( V'(\tilde\varphi^n_1) +
2364:   V'(\tilde\varphi^n_2) -2 V'(\tilde\varphi^n_3),v^{M-1} \big)
2365: \end{pmatrix}
2366: \end{equation*}
2367: can be bounded in magnitude by
2368: \begin{multline*}
2369: ||V'(\tilde\varphi^n_1)+V'(\tilde\varphi^n_2)-2
2370:   V'(\tilde\varphi^n_3)|| \cdot ||v^i||\\
2371:  = \sqrt{\frac{2}{3}\Delta x} ||V'(\tilde\varphi^n_1)+V'(\tilde\varphi^n_2)-2
2372:   V'(\tilde\varphi^n_3)||\\
2373: =\sqrt{\frac{2}{3}\Delta x} ||V'(\tilde\varphi^n_1)+V'(\tilde\varphi^n_2)-2
2374:   V'(\frac{\tilde\varphi^n_1+\tilde\varphi^n_2}{2})|| \\+ 2\sqrt{\frac{2}{3}\Delta x}||
2375:   V'(\frac{\tilde\varphi^n_1+\tilde\varphi^n_2}{2}) -
2376:   V'(\tilde\varphi^n_3)|| =: I + II,
2377: \end{multline*}
2378: using the triangle inequality as in \eqref{eq:Vsplit}.  We first treat
2379: term $I$ above:
2380: \begin{multline*}
2381: ||V'(\tilde\varphi^n_1)+V'(\tilde\varphi^n_2)-2
2382:   V'(\frac{\tilde\varphi^n_1+\tilde\varphi^n_2}{2})||^2 \leq
2383:   \frac{|V'''|^2}{2^2} \int_0^1 |\tilde\varphi^n_1 -
2384:   \tilde\varphi^n_2|^4 dx \\
2385: \leq \frac{|V'''|^2}{2^2} \big( \max
2386:   |\tilde\varphi^n_1 - \tilde\varphi^n_2| \big)^4 =
2387:   \frac{|V'''|^2}{2^2} \big( \max |\xi^n_1 - \xi^n_2| \big)^4 \\
2388: \leq
2389: \frac{|V'''|^2}{2^2}  \big(\sum_{i=1}^{M-1} |\xi^n_{1,i} - \xi^n_{2,i}|^2 \big)^2 = \frac{|V'''|^2}{2^2}||
2390:   \xi^n_1 - \xi^n_2 ||^4_2,
2391: \end{multline*}
2392: where $||\cdot||_2$ denotes the Euclidean vector norm.
2393: Part $II$ may be bounded as follows:
2394: \begin{equation*}
2395: 2||
2396:   V'(\frac{\tilde\varphi^n_1+\tilde\varphi^n_2}{2}) -
2397:   V'(\tilde\varphi^n_3)|| \leq A||\tilde\varphi^n_1 + \tilde\varphi^n_2
2398:   - 2 \tilde\varphi^n_3|| \leq B ||z^n||_2.
2399: \end{equation*}
2400: These facts in \eqref{eq:discrzevol} give that 
2401: \begin{equation}\label{eq:znorm}
2402: ||z^{n+1}||_2 \leq C||z^n||_2 + D ||\xi^n_1-\xi^n_2||^2_2.
2403: \end{equation}
2404: By subtracting the equations for $m=1$
2405: and $m=2$ in \eqref{eq:xistates} we see that 
2406: \begin{equation*}
2407: ||\xi^{n+1}_1-\xi^{n+1}_2||_2 \leq E ||\xi^{n}_1-\xi^{n}_2||_2 \leq
2408:   \ldots \leq E^{n+1}||\xi^0_1-\xi^0_2||_2,
2409: \end{equation*}
2410: so that in \eqref{eq:znorm} we could really write
2411: $D||\xi^0_1-\xi^0_2||^2_2$ instead of
2412: $D||\xi^n_1-\xi^n_2||^2_2$. Thereby, since $z^0=0$, it holds that
2413: $||z^N||_2 \leq F||\xi^0_1 - \xi^0_2||_2^2$. 
2414: Note that the constants $A$ -- $F$ are allowed to depend on
2415: $\Delta x$.
2416: Similarly as in the proof
2417: of Theorem \ref{thm:semiconcave}, semiconcavity  of $\tilde u$ is a
2418: consequence of this.
2419: \end{proof}
2420: We are now ready for the promised theorem about the Symplectic Euler
2421: method.
2422: \begin{thm}\label{thm:sympleuler}
2423: For every minimizer $\{\tilde\alpha^n\}$ in \eqref{eq:timediscretevalue} there exists a
2424: solution to \eqref{eq:sympleuler} with $\tilde\lambda^{n+1}=-\tilde\alpha^n$ for $0
2425: \leq n \leq N-1$.
2426: \end{thm}
2427: \begin{proof}
2428: The proof is divided into three steps. In the first step it is shown
2429: that the value function $\tilde u$ is differentiable along the optimal
2430: path $\tilde\varphi^n$. In the second step it is proved that the dual
2431: variable $\tilde\lambda^n$ equals the G\^ateaux derivative of $\tilde u$, and in
2432: the last step it is shown that $\tilde\lambda^{n+1}=\tilde\alpha^n$.
2433: 
2434: \emph{Step 1.} In order to show that the discrete value function
2435: $\tilde u$ is differentiable at $(\tilde\varphi^n,t_n)$ for $0 < n \leq
2436: N$ the function
2437: \begin{equation}\label{eq:r}
2438: r(\tilde\alpha) \equiv \tilde u(\tilde\varphi,t_{n+1}) + \Delta t h(\tilde\alpha)
2439: \end{equation}
2440: is introduced, where $\tilde\varphi=\tilde\varphi^n + \Delta t \bar
2441: f(\tilde\varphi^n,\tilde\alpha)$ and $\bar f$ is given by \eqref{eq:fbar}. 
2442: Assume that $\tilde u$ is not differentiable at
2443: $(\tilde\varphi^{n+1},t_{n+1})$. Because $\tilde u$ is semiconcave it then
2444: follows that the superdifferential $D^+ \tilde
2445: u(\tilde\varphi^{n+1},t_{n+1})$ (which we let designate the
2446: superdifferentials in the G\^ateaux sense) contains more than one
2447: point.  For all $\tilde \alpha$ in a neighborhood of $\tilde \alpha^n$
2448: it holds that
2449: \begin{equation}\label{eq:rdiff}
2450: \begin{split}
2451: &r(\tilde\alpha)-r(\tilde\alpha^n) \\
2452:  =& \tilde u \big(\tilde\varphi^n + \Delta
2453: t \bar f(\tilde\varphi^n,\tilde\alpha)\big) - \tilde u \big(\tilde\varphi^n + \Delta
2454: t \bar f(\tilde\varphi^n,\tilde\alpha^n)\big) +\Delta t
2455: \big(h(\tilde\alpha)-h(\tilde\alpha^n)\big) \\
2456:  \leq  &\Delta t \big(p, \bar f(\tilde\varphi^n,\tilde\alpha)- \bar
2457: f(\tilde\varphi^n,\tilde\alpha^n)\big) + \Delta t
2458: \big(h'(\tilde\alpha^n),\tilde\alpha - \tilde\alpha^n\big) + K||\tilde\alpha -
2459: \tilde\alpha^n||^2 \\
2460:  = &\Delta t (\tilde p, \tilde\alpha- \tilde\alpha^n) + \Delta t
2461: \big(h'(\tilde\alpha^n),\tilde\alpha - \tilde\alpha^n\big) + K||\tilde\alpha -
2462: \tilde\alpha^n||^2,
2463: \end{split} 
2464: \end{equation}
2465: where $p$ is an element in $D^+ \tilde
2466: u(\tilde\varphi^{n+1},t_{n+1})$ and $\tilde p$ is given by a linear
2467: bijection of $s$, since $\bar f$ is linear in the $\alpha$
2468: variable. Since there are more than one element $p \in D^+ \tilde
2469: u(\tilde\varphi^{n+1},t_{n+1})$, there are also more than one possible $\tilde p$
2470: in equation \eqref{eq:rdiff}. 
2471: It is therefore possible to choose the element $\tilde p$ such that
2472: the linear term in \eqref{eq:rdiff} is non-vanishing.
2473: It follows that there exists $\tilde
2474: \alpha$ such that $r(\tilde\alpha) < r(\tilde\alpha^n)$, which is the
2475: sought contradiction. By this reasoning we see that $\tilde u$ is
2476: differentiable at $(\tilde\varphi^n,t_n)$ for $0<n \leq N$. 
2477: 
2478: \emph{Step 2.} It follows directly that
2479: $\tilde\lambda^N=g'(\tilde\varphi^N)$, i.e.\ the G\^ateaux derivative of
2480: $\tilde u(\cdot,t_N)$. Assume that $\tilde\lambda^{n+1} = \tilde
2481: u_\varphi(\tilde\varphi^{n+1},t_{n+1})$. It will follow from this that
2482: $\tilde\lambda^n =\tilde u_\varphi(\tilde\varphi^n,t_n)$. Since it is known
2483: that $\tilde u$ is differentiable at both $(\tilde\varphi^n,t_n)$ and
2484: $(\tilde\varphi^{n+1},t_{n+1})$ the G\^ateaux derivative of $\tilde u$ at
2485: $(\tilde\varphi^{n+1},t_{n+1})$ equals the G\^ateaux derivative at
2486: $\tilde \varphi^{n}$ of the function
2487: \begin{equation*}
2488: s(\tilde\varphi) \equiv \tilde  u (\tilde\varphi + \Delta t \bar
2489: f(\tilde\varphi,\tilde\alpha^n),t_n) + \Delta t h(\tilde\alpha^n),
2490: \end{equation*}
2491: where $\tilde\alpha^n$ is fixed. The G\^ateaux derivative of $s$ at $\tilde\varphi^n$ is
2492: given by 
2493: \begin{equation*}
2494: s'(\tilde\varphi^n) = \tilde u_\varphi(\tilde\varphi^{n+1},t_{n+1}) \circ (I + \Delta t
2495: \bar f'(\tilde\varphi^n)),
2496: \end{equation*}
2497: where $\tilde u_\varphi(\tilde\varphi^{n+1},t_{n+1}) =
2498: \tilde\lambda^{n+1}$ is a function from $V$ to $\Re$ and $\bar
2499: f'(\tilde\varphi^n)$ is a function from $V$ to $V$. This equation
2500: coincides with the $\tilde\lambda$ equation in \eqref{eq:sympleuler},
2501: which gives that $\tilde\lambda^n=\tilde u_\varphi(\tilde\varphi^n,t_n)$.
2502: By induction in $n$   it follows that $\tilde\lambda^n=\tilde
2503: u_\varphi(\tilde\varphi^n,t_n)$ for $0 < n \leq N$.
2504: 
2505: \emph{Step 3.} Knowing that $\tilde u$ is differentiable at $(\tilde
2506: \varphi^n,t_n)$ for $0 < n \leq N$ the function \eqref{eq:r} can be
2507: differentiated. Since $\tilde\alpha^n$ is a minimizer of $r$ the
2508: derivative at this argument must be zero:
2509: \begin{equation*}
2510: r'(\tilde\alpha^n)=\tilde\lambda^{n+1} \Delta t \circ I +\Delta t \tilde\alpha^n=0,
2511: \end{equation*}
2512: where it is used that $\tilde
2513: u_\varphi(\tilde\varphi^{n+1},t_{n+1})=\tilde\lambda^{n+1}$, $\bar f_\alpha
2514: = I$ and $h'(\tilde\alpha^n)=\tilde\alpha^n$. It follows that
2515: $\tilde\lambda^{n+1} = -\tilde\alpha^n$ for $0 \leq n \leq N-1$.
2516: \end{proof}
2517: \section{Numerical Results}\label{sec:NumRes}
2518: We here present some numerical results for the 
2519: Symplectic Euler scheme for a finite difference
2520: discretization of \eqref{eq:flow}, \eqref{eq:valuefunction}.
2521: The numerics is performed in this setting, partly because it is
2522: slightly simpler than using finite elements, partly because a finite
2523: difference discretization is used in \cite{Weinan-Ren-Vanden-Eijnden}.
2524: The system we will consider is therefore 
2525: \begin{equation}\label{eq:systxieta}
2526: \begin{split}
2527: \xi^{n+1} &= \xi^n + \Delta t \big(\delta D^2 \xi^n - \delta^{-1}
2528: V'(\xi^n) -\eta^{n+1}\big), \\
2529: \eta^n &= \eta^{n+1} + \Delta t \big(\delta D^2 \eta^{n+1} -
2530: \delta^{-1}\eta^{n+1}V'(\xi^n)\big), \\
2531: \eta^N &= 2K(\xi^N-\xi^-),
2532: \end{split}
2533: \end{equation}
2534: where $\xi^-$ is a finite difference approximation to $\varphi_-$, $D^2$ is defined in \eqref{eq:diffquotient} and $\xi^n$ and
2535: $\eta^n$ correspond to the nodal values of $\bar\varphi^n$ and
2536: $\bar\lambda^n$, respectively. The approximate value used together
2537: with this scheme is
2538: \begin{equation}\label{eq:FDvalue}
2539: K \Delta x ||\xi^N-\xi^-||_2^2 + \Delta t\Delta x \sum_{n=1}^{n=N}
2540: ||\eta^n||_2^2 /2,
2541: \end{equation}
2542: where $||\cdot ||_2$ denotes the ordinary Euclidean vector norm.
2543: As noted in \cite{Weinan-Ren-Vanden-Eijnden} there are several local
2544: minima to \eqref{eq:FDvalue}, corresponding to different ``strategies'' to
2545: overcome the potential barrier $V$. The switching between the two
2546: stable points proceeds by ``nucleation'', which involves a large
2547: control $\alpha$, followed by propagation of domain walls. In Figure
2548: \ref{fig:phi2walls1wallBETAmilliardR6percentM200N2006times}  the transition is shown for the cases propagation of one
2549: and two domain walls. 
2550: \begin{figure}
2551: \centering
2552: \includegraphics[width=0.8\textwidth,height=5cm]{phi2walls1wallBETAmilliardR6percentM200N2006times.eps}
2553: \caption{Snapshots of transitions between the two stable
2554:   configurations where $\varphi$ is shown at times 0, 0.2, 0.4, 0.6, 0.8 and 1 ($=T$).   To the right propagation of one wall and to the
2555:   left propagation of two walls. In these examples $K=10^9$ and
2556:   $\delta=0.06$ was used.}
2557: \label{fig:phi2walls1wallBETAmilliardR6percentM200N2006times}
2558: \end{figure}
2559: The $\lambda$ variable, which equals the negative control, is shown in
2560: Figure \ref{fig:lambda2wallsBETAmilliardR6percentM200N200} for the case of propagation of two walls.
2561: \begin{figure}
2562: \centering
2563: \includegraphics[width=\textwidth]{lambda2wallsBETAmilliardR6percentM200N200.eps}
2564: \caption{The dual variable $\lambda$ for the case of two propagating
2565:   walls corresponding to the left part of Figure \ref{fig:phi2walls1wallBETAmilliardR6percentM200N2006times}.}
2566: \label{fig:lambda2wallsBETAmilliardR6percentM200N200}
2567: \end{figure}
2568: 
2569: Apart from the Symplectic Forward Euler method previously mentioned,
2570: the Symplectic Backward Euler method can also be used. This method is
2571: given by
2572: \begin{align*}
2573: \xi^{n+1} &= \xi^n + \Delta t (\delta D^2 \xi^{n+1} - \delta^{-1}
2574: V'(\xi^{n+1}) -\eta^{n}), \\
2575: \eta^n &= \eta^{n+1} + \Delta t (\delta D^2 \eta^{n} -
2576: \delta^{-1}\eta^{n}V'(\xi^{n+1})), \\
2577: \eta^N &= 2K(\xi^N-\xi^-).
2578: \end{align*} 
2579: The approximate value for the Symplectic Backward Euler method is
2580: given by (see Chapter 4.4 in \cite{Sandberg-Szepessy})
2581: \begin{equation}\label{eq:FDBEvalue}
2582: K \Delta x ||\xi^N-\xi^-||_2^2 + \Delta t\Delta x \sum_{n=0}^{n=N-1}
2583: ||\eta^n||_2^2 /2.
2584: \end{equation}
2585: An advantage with the Backward Euler method is that it enables using a
2586: small $\Delta x$ even when $\Delta t$ is not small. This feature is
2587: however not as profound for the present case of control of a
2588: parabolic equation as for the uncontrolled case, as the control
2589: compensates for the instability, which makes it possible to use smaller
2590: $\Delta x$. Another good thing about the Backward Euler method is
2591: that it seems to underestimate the optimal value while it seems to be
2592: overestimated by the Forward Euler method.  Figure
2593: \ref{fig:valueconvFEBEdxequalsdt} shows the dependence on $\Delta
2594: x=\Delta t$ of the values \eqref{eq:FDvalue} and \eqref{eq:FDBEvalue}.
2595: \begin{figure}
2596: \centering
2597: \includegraphics[width=\textwidth,height=5cm]{valueconvFEBEdxequalsdt.eps}
2598: \caption{Convergence of the optimal values \eqref{eq:FDvalue} and
2599:   \eqref{eq:FDBEvalue} for the case of equal spacing in space and
2600:   time, i.e.\ $\Delta x=\Delta t$. The left figure shows the values
2601:   obtained by the Forward Euler method and the right shows the values
2602:   of the Backward
2603:   Euler method.}
2604: \label{fig:valueconvFEBEdxequalsdt}
2605: \end{figure}
2606: By extrapolating these fairly straight curves to $\Delta x=\Delta t=0$
2607: an approximate value of the optimal control problem is obtained. The
2608: extrapolated value from the Forward Euler curve is 8.517, and the
2609: approximate value from the Backward Euler curve is 8.526. 
2610: 
2611: We now indicate the dependence of the spatial discretization error on
2612: $\Delta x$. This is done by changing the spatial discretization
2613: $\Delta x$ while keeping the time discretization $\Delta t$
2614: constant. We let the value
2615: obtained for the smallest spatial discretization $\Delta x$ be the
2616: reference value which takes the role of an ``exact'' solution. A convergence plot can be found in Figure
2617: \ref{fig:dxconvergence2wallsBETAmilliardR3percentN200}. The slope of
2618: the upper part of this curve corresponds to a convergence rate of
2619: approximately $(\Delta x)^{2.37}$.
2620: \begin{figure}
2621: \centering
2622: \includegraphics[width=0.5\textwidth]{dxconvergence2wallsBETAmilliardR3percentN200.eps}
2623: \caption{Convergence of the optimal value (y-axis) with respect to
2624:   $\Delta x$ (x-axis). The case with two propagating domain walls and
2625:   $\Delta t=1/200$, $\delta=0.03$ and $K=10^9$.}
2626: \label{fig:dxconvergence2wallsBETAmilliardR3percentN200}
2627: \end{figure}
2628: 
2629: For the time discretization error we want to show that it is less than
2630: a linear function of $\Delta t$ with a constant which does not depend
2631: on $\Delta x$.  Time discretization convergence is therefore
2632: considered for two spatial discretizations, one having $\Delta x=1/30$
2633: and the other $\Delta x =1/100$. Since the Forward and Backward
2634: Euler methods in the limit $\Delta t \rightarrow 0$ shall have the
2635: same value we may extrapolate the values from diagrams similar to the
2636: ones in Figure \ref{fig:valueconvFEBEdxequalsdt}, but with the
2637: exception that $\Delta x$ is held fixed. The following  values are obtained from
2638: these extrapolations in the case of two propagating domain walls,
2639: using $\delta=0.03$ and $K=10^9$: 
2640: \begin{itemize}
2641: \item Forward Euler, $\Delta x=1/30$:\quad 8.841 
2642: \item Forward Euler, $\Delta x=1/100$:\quad 8.547
2643: \item Backward Euler, $\Delta x=1/30$:\quad 8.849
2644: \item Backward Euler, $\Delta x=1/100$:\quad 8.555   
2645: \end{itemize}
2646: The mean of the above values for Forward and Backward Euler can be
2647: taken as an ``exact'' reference value when convergence is
2648: studied. Hence for
2649: $\Delta x=1/30$ the reference value is taken to be 8.845 and for
2650: $\Delta x=1/100$ it is taken to be 8.551. The two convergence plots
2651: can be found in Figure
2652: \ref{fig:dtconvergence2wallsBETAmilliardR3percentM30M100}. Note that
2653: the inclination in the left curve, the values using $\Delta x=1/30$, is
2654: larger than the inclination in the right curve ($\Delta
2655: x=1/100$). This is in harmony with Theorem \ref{thm:timediscrconv} since
2656: it is allowed that (and good if) we have faster convergence for smaller $\Delta x$.  
2657: \begin{figure}
2658: \centering
2659: \includegraphics[width=\textwidth,height=5cm]{dtconvergence2wallsBETAmilliardR3percentM30M100.eps}
2660: \caption{Time discretization convergence for two different spatial
2661:   discretizations, $\Delta x=1/30$ (left), and $\Delta x=1/100$
2662:   (right).
2663: The case with two propagating domain walls and
2664:   $\delta=0.03$ and $K=10^9$.}
2665: \label{fig:dtconvergence2wallsBETAmilliardR3percentM30M100}
2666: \end{figure}
2667: 
2668: The system \eqref{eq:systxieta} can be (and has been) solved in two
2669: steps. The first step gives a starting position for the second step,
2670: and may be performed on a coarse grid, i.e.\ using large $\Delta x$
2671: and $\Delta t$. The method is to choose an initial guess $\xi_0$ (a
2672: vector containing all time steps) and with it compute the dual,
2673: $\eta_0$, using \eqref{eq:systxieta}. This computed $\eta_0$ is used
2674: in \eqref{eq:systxieta} to compute $\xi_{upd}$, an updated
2675: $\xi$. Using a damping $\nu$, a new state $\xi_1=\nu \xi_0 +
2676: (1-\nu)\xi_{upd}$ is computed which is used to obtain the dual
2677: $\eta_1$, which in turn is used to compute a new $\xi_{upd}$, and so
2678: on. When the difference $\xi_{upd}-\xi_n$ is sufficiently small the
2679: iterations are terminated, and a starting point $(\xi,\eta)$ is
2680: obtained for the second step.
2681: 
2682: Step two consists of Newton iterations of \eqref{eq:systxieta}. Since
2683: the sparse Jacobian can be computed explicitly this second step
2684: converges at a quadratic rate, making it computationally cheap to
2685: reach  an accurate solution. In the examples presented in this chapter the
2686: Newton iterations continued until the difference between two
2687: consecutive $\xi$:s and $\eta$:s was less than $10^{-13}$ in each
2688: space-time component. After convergence has been reached for some
2689: discretization, a space-time interpolation of $\xi$ and $\eta$ can be
2690: used as a starting position for a Newton iteration on a new grid. It
2691: is also possible to gradually change the parameters $\delta$ and $K$
2692: in the Newton iterations in order to be able to treat a favorite
2693: case. When the starting point is sufficiently good the Newton
2694: method terminates after 5-7 iteration steps, making it fast. As
2695: comparison, when in \cite{Weinan-Ren-Vanden-Eijnden} a limited memory
2696: BFGS method is used, about 550 iterations is needed to decrease the
2697: $L_2$-norm of the objective gradient to $10^{-10}$, even when a clever
2698: approximation of the initial Hessian was used.
2699: %\section{Discretization in $x$}
2700: %Let $V$ be a Hilbert space and $\bar V$ a closed linear subspace in
2701: %$V$, $f:V \times B \to V$ and $\bar f : \bar V \times \bar B \to \bar
2702: %V$ be flows that guide the evolution of states $X\in V$ and $\bar X
2703: %\in \bar V$ via the differential equations
2704: %\begin{align*}
2705: %\frac{dX}{dt} &= f(X(t),\alpha(t)),\\
2706: %\frac{d\bar X}{dt} &= \bar f(\bar X(t),\bar \alpha(t)),
2707: %\end{align*} 
2708: %where $X_0 \in \bar V$. We introduce the corresponding value functions
2709: %\begin{align*}
2710: %u(x,t) &=\inf_\alpha \big( g(X(T)) + \int_t^T h(\alpha(t))dt \big), \\
2711: %\bar u(x,t) &=\inf_{\bar\alpha} \big( \bar g(X(T)) + \int_t^T \bar h(\bar\alpha(t))dt \big). 
2712: %\end{align*}  
2713: %These value functions solve the two Hamilton-Jacobi equations
2714: %\begin{align*}
2715: %u_t+H(u_x,x) &=0, \quad u(x,T)=g(x), \\
2716: %\bar u_t+\bar H(\bar u_x,x) &=0, \quad \bar u(x,T)=\bar g(x) 
2717: %\end{align*}
2718: %where
2719: %\begin{align*}
2720: %H(\lambda,x) &\eqiv \min_\alpha \big(\lambda \cdot f(x,\alpha) +
2721: %h(\alpha) \big) \\
2722: %\bar H(\lambda,x) &\eqiv \min_{\bar\alpha} \big(\lambda \cdot \bar
2723: %f(x,\bar \alpha) +
2724: %\bar h(\bar\alpha) \big). 
2725: %\end{align*}
2726: %With this notation we may formulate a theorem:
2727: %\begin{thm}\label{thm:subspace}
2728: %Goes here..
2729: %\end{thm}
2730: %\begin{proof}
2731: %We begin with the lower bound for $\bar u(x,t)-u(x,t)$. First $\bar
2732: %u(x,t)$ is split:
2733: %\begin{multline*}
2734: %\inf_{\bar\alpha} \big( \bar g(\bar X(T))+\int_t^T\bar h(\bar X,
2735: %\bar\alpha)dt \big)\\ 
2736: %\geq \inf_{\bar\alpha} \big(\bar g(\bar X(T))-
2737: %g(\bar X(T))\big) + \inf_{\bar\alpha} \big(g(\bar X(T))+\int_t^T\bar h(\bar X,
2738: %\bar\alpha)dt \big).
2739: %\end{multline*}
2740: %Hence we have that
2741: %\begin{multline*}
2742: %\bar u(x,t)-u(x,t)   \\
2743: %\geq \inf_{\bar\alpha} \big(\bar g(\bar X(T))-
2744: %g(\bar X(T))\big) + \inf_{\bar\alpha}\big(u(\bar X(T),T)-u(x,t)+
2745: %\int_t^T \bar h(\bar\alpha)dt \big) \equiv I+II.
2746: %\end{multline*}
2747: %Part $II$ is now considered.
2748: %\begin{align*}
2749: %II &= \inf_{\bar\alpha} \big(\int_t^T (du(\bar X(s),s)+\bar
2750: %h(\bar\alpha))ds \big) \\
2751: %&=
2752: %\end{align*}
2753: %\end{proof}
2754: %
2755: %\section{Section}
2756: %We use the Symplectic backward Euler method:
2757: %\begin{align}
2758: %\varphi_{n+1} &= \varphi_n+\Delta t H^{\Delta
2759: %  x}_{;\lambda}(\lambda_n,\varphi_{n+1}) \label{eq:BEphi} \\ 
2760: %\lambda_n &= \lambda_{n+1}+\Delta t H^{\Delta x}_{;\varphi}
2761: %  (\lambda_n,\varphi_{n+1}) \label{eq:BElambda}
2762: %\end{align}
2763: %
2764: %We extend $\varphi$ to all time as a continuous piecewise linear function
2765: %\begin{equation}\label{eq:phiextension}
2766: %\varphi(t)=\frac{t_{n+1}-t}{\Delta t}\varphi_n + \frac{t-t_n}{\Delta
2767: %  t}\varphi_{n+1}, \quad \text{ for } t_n\leq t \le t_{n+1}
2768: %\end{equation}
2769: %
2770: %
2771: %%In order to be able to prove the following lemma it is necessary to
2772: %%change the potential $V$ outside the interval $[-1,1]$. We may thus
2773: %%let $\tilde V$ equal $V$ in a large interval containing $[-1,1]$
2774: %\begin{lemma}\label{Lem:noncollide}
2775: %Assume that there exists a constant $C$, such that for all time steps
2776: %$n$ and all $\varphi_n$, $\norm{\frac{\partial \lambda_n}{\partial
2777: %    \varphi_n}}_2 \leq C\Delta x$. Then, for all positive constants, $D$,
2778: %when $\Delta t= D\Delta x$ and $\Delta x$ is sufficiently small, two
2779: %different solutions do not intersect.
2780: %\end{lemma}
2781: %\begin{proof}
2782: %We start by showing that it suffices that
2783: %$\frac{\partial\varphi_{n+1}}{\partial\varphi_n}$ is positive definite
2784: %everywhere for the lemma to hold.
2785: %Consider the two points $\varphi_n^1$ and $\varphi_n^2$ at time $t_n$ and
2786: %the corresponding points $\varphi_{n+1}^1$ and $\varphi_{n+1}^2$ at time
2787: %$t_{n+1}$. Integration along the line
2788: %$\varphi_{n+1}\big(\varphi_n^1+s(\varphi_n^2-\varphi_n^1)\big)$
2789: %gives that
2790: %\begin{equation*}
2791: %\varphi_{n+1}^2-\varphi_{n+1}^1=\int_0^1 \frac{\partial \varphi_{n+1}}{\partial\varphi_n}\big(\varphi_n^1+s(\varphi_n^2-\varphi_n^1)\big)ds(\varphi_n^2-\varphi_n^1).
2792: %\end{equation*}
2793: %Hence
2794: %\begin{equation*}
2795: %(\varphi_n^2-\varphi_n^1)\cdot(\varphi_{n+1}^2-\varphi_{n+1}^1)= \int_0^1
2796: %  (\varphi_n^2-\varphi_n^1) \cdot \frac{\partial
2797: %  \varphi_{n+1}}{\partial\varphi_n}(\varphi_n^1+s(\varphi_n^2-\varphi_n^1))
2798: %  (\varphi_n^2-\varphi_n^1)ds > 0
2799: %\end{equation*}
2800: %if $\frac{\partial\varphi_{n+1}}{\partial\varphi_n}$ is positive
2801: %  definite. Using (\ref{eq:phiextension}) we see that
2802: %\begin{equation*}
2803: %\big(\varphi^2(t)-\varphi^1(t) \big)\cdot (\varphi_n^2-\varphi_n^1) =
2804: %\frac{t_{n+1}-t}{\Delta t}|\varphi_n^2-\varphi_n^1|^2+\frac{t-t_n}{\Delta
2805: %  t}(\varphi_{n+1}^2-\varphi_{n+1}^1) \cdot (\varphi_n^1-\varphi_n^2)>0,
2806: %\end{equation*}
2807: %when $\varphi_n^1 \neq \varphi_n^2$.
2808: %
2809: %We next differentiate equation (\ref{eq:BEphi}) in order to show that
2810: %$\frac{\partial\varphi_{n+1}}{\partial\varphi_n}$ is positive definite under
2811: %the assumptions in the lemma. Let us for simplicity denote $H^{\Delta
2812: %  x}_{;\lambda}$ by $P$ and $H^{\Delta x}_{;\varphi}$ by $L$:
2813: %\begin{align}
2814: %\frac{\partial\varphi_{n+1}}{\partial\varphi_n} &=I+\Delta t P_\lambda
2815: %\frac{\partial \lambda_n}{\partial\varphi_n}+ \Delta t
2816: %P_\varphi\frac{\partial\varphi_{n+1}}{\partial\varphi_n} \notag \\
2817: %\intertext{so that, using $P_\lambda=-I$}
2818: %\frac{\partial\varphi_{n+1}}{\partial\varphi_n} &=(I-\Delta t P_\varphi)^{-1}(I
2819: %- \Delta t \frac{\partial\lambda_n}{\partial\varphi_n}). \label{eq:phivar}
2820: %\end{align}
2821: %That $(I-\Delta t P_\varphi)$ is invertible when $\Delta x$ is small will
2822: %be evident from what follows.
2823: %
2824: %We proceed by calculating the largest and smallest eigenvalues for
2825: %$I-\varepsilon\Delta t D^2$, the main part of 
2826: %\begin{equation*}
2827: %I-\Delta t P_\varphi=I-\varepsilon\Delta t D^2-\frac{\Delta
2828: %  t}{\varepsilon} \text{diag}(V''(\varphi)).
2829: %\end{equation*}
2830: %Summation by parts,
2831: %\begin{equation*}
2832: %v\cdot D^2 v=D\tilde v \cdot D\tilde v,
2833: %\end{equation*}
2834: %where $D\tilde v=\frac{1}{\Delta x}(v_1-v_0, v_2-v_1, \ldots ,
2835: %v_M-v_{M-1})$ and $\tilde v=(0, v_1, \ldots , v_{M-1}, 0)$ entails 
2836: %$I-\varepsilon\Delta t D^2 \geq 1$. 
2837: %
2838: %In order to find an upper bound for $I-\varepsilon\Delta t D^2$ the
2839: %discrete Fourier transform is used. We use the definition used e.g.\ in
2840: %\cite{henrici} where the Fourier transform of $z=(z_0,\ldots,z_{N-1})
2841: %\in \Compl^N$ is $\hat z \in \Compl^N$, where
2842: %\begin{align*}
2843: %\hat z_k &= \frac{1}{\sqrt{N}}\sum_{j=0}^{N-1}z_j w^{-jk}, \\
2844: %w & \equiv exp(\frac{2\pi i}{N}),
2845: %\end{align*}
2846: %making it an isometry on $\Compl^N$.
2847: %
2848: %An approximation of $I-\varepsilon \Delta t D^2$ by its periodic
2849: %analogue is now considered in order to be able to use the Fourier
2850: %transform. Hence we let
2851: %\begin{equation*}
2852: %A = I-\varepsilon \Delta t D^2 +B,
2853: %\end{equation*}
2854: %where $B_{1,M-1}=B_{M-1,1}=-\varepsilon \frac{\Delta t}{\Delta x^2}$
2855: %and all other entries in $B$ are zero. Thereby $Av=\sqrt{M-1}a*v$,
2856: %where $a=(1+2\varepsilon\frac{\Delta t}{\Delta x^2},
2857: %-\varepsilon\frac{\Delta t}{\Delta x^2}, 0,\ldots,0,
2858: %-\varepsilon\frac{\Delta t}{\Delta x^2})$. Plancherels formula with a
2859: %real $v$ then gives
2860: %\begin{equation*}
2861: %v\cdot Av=\sqrt{M-1}\bar{\hat v} \cdot \widehat{a*v}=\sqrt{M-1}\sum
2862: %\hat a_i |\hat v_i|^2.
2863: %\end{equation*}
2864: %A calculation of $\hat a$ shows that all $\hat a_i$ are real and
2865: %satisfy $\hat a_i \in [\frac{1}{\sqrt{M-1}},
2866: %  \frac{1}{\sqrt{M-1}}(1+2\varepsilon \frac{\Delta t}{\Delta x^2})]$. So $A$:s
2867: %  eigenvalues all lie in the interval $[1, 1+2\varepsilon \frac{\Delta
2868: %    t}{\Delta x^2}]$.  
2869: %
2870: %The size of $v\cdot Bv$ is at most proportional to the largest
2871: %eigenvalue of A:
2872: %\begin{equation*}
2873: %|v\cdot Bv|=2\varepsilon\frac{\Delta t}{\Delta x^2}|v_1 v_{M-1}| \leq
2874: % \varepsilon\frac{\Delta t}{\Delta x^2}(v_1^2+v_{M-1}^2) \leq 
2875: %\varepsilon\frac{\Delta t}{\Delta x^2}|v|^2.
2876: %\end{equation*}
2877: %Taking $\Delta t=D\Delta x$ we have the largest eigenvalue of
2878: %$I-\varepsilon\Delta t D^2$ of the order $\Delta x^{-1}$.
2879: %
2880: %The last part of $I-\Delta t P_{\varphi}$, i.e.\ $\frac{\Delta
2881: %  t}{\varepsilon}diag\big(V''(\varphi)\big)$, does not change the
2882: %  spectrum of $I-\varepsilon\Delta t D^2$ much, as its eigenvalues are
2883: %  of the order $\Delta x$.
2884: %
2885: %So for the inverse $(I-\Delta t P_{\varphi})^{-1}$ all eigenvalues are
2886: %positive, with the smallest of order $\Delta x$. All eigenvalues of
2887: %$\Delta t(I-\Delta t
2888: %P_{\varphi})^{-1}\frac{\partial\lambda_n}{\partial\varphi_n}$ are
2889: %$\mathcal{O}(\Delta x^2)$, as $\norm{\frac{\partial \lambda_n}{\partial
2890: %    \varphi_n}}_2 =\mathcal{O}(\Delta x)$, so by (\ref{eq:phivar})
2891: %$\frac{\partial \varphi_{n+1}}{\partial\varphi_n}$ is positive definite.
2892: %\end{proof}
2893: \section{Acknowledgments}
2894: I would like to thank Anders Szepessy and Erik von Schwerin for
2895: proofreading this article and suggesting improvements. 
2896: \section{Appendix}\label{sec:appendix}
2897: In order to show existence and uniqueness of solutions to
2898: (\ref{eq:flow}) we introduce the notion of \emph{weak} solutions (see
2899: \cite{Evans}). We will let $\langle \cdot , \cdot \rangle$ denote the
2900: pairing between $H^{-1}$ and $H^1_0$.
2901: \begin{definition}\label{def:weak}
2902: We say a function
2903: \begin{equation*}
2904: \varphi \in L^2(0,T;H_0^1(0,1)), \text{ with } \varphi_t \in L^2(0,T;H^{-1}(0,1)),
2905: \end{equation*}
2906: is a \emph{weak solution} of (\ref{eq:flow}) with $\varphi_0\in L^2(0,1)$ provided
2907: \begin{align*}
2908: &\langle \varphi_t, v \rangle + \delta(\varphi_x, v_x) = (-\delta^{-1}V'(\varphi)+\alpha,v) \\
2909: \intertext{for each $v\in H_0^1(0,1)$ and a.e.\ time $t_0 \leq t \leq
2910:   T$, and}
2911: &\varphi(t_0)=\varphi_0.
2912: \end{align*}
2913: \end{definition} 
2914: Weak solutions are in fact more regular than is required in the
2915: definition when the initial state $\varphi_0\in H_0^1(0,1)$, which is
2916: used when proving the following theorem.
2917: \begin{thm}
2918: There exists a unique weak solution $\varphi$ to (\ref{eq:flow}) in \linebreak
2919: $C([t_0,T];H_0^1)$ when $\varphi_0 \in H^1_0(0,1)$. 
2920: %This solution
2921: %satisfies $\varphi \in L^2(0,T;H^2)$ and $\varphi_t \in L^2(0,T;L^2)$.
2922: This solution satisfies
2923: \begin{equation}\label{eq:phibound1}
2924: ||\varphi_x(t)||^2_{L^2} 
2925: \leq
2926: ||(\varphi_0)_x||^2_{L^2}+\frac{\delta^{-2}}{2} ||\varphi_0||^4_{L^4}-
2927: \delta^{-2} ||\varphi_0||^2_{L^2} + \delta^{-1} ||\alpha||^2_{L^2(0,T;L^2)} +\frac{\delta^{-1}}{2},
2928: \end{equation}
2929: for all $t \in [t_0,T]$.
2930: \end{thm} 
2931: \begin{proof}
2932: We start by proving existence and uniqueness of solutions to the equation
2933: (\ref{eq:flow}) when the potential $\tilde V$ is used; see figure
2934: \ref{fig:V}. Let $\breve\varphi \in L^\infty (t_0,T;H_0^1)$ and $\breve\varphi(t_0)=\varphi_0$ and
2935: define $\tilde \varphi$ by 
2936: \begin{equation*}
2937: \tilde\varphi_t=\delta\tilde\varphi_{xx} - \delta^{-1}\tilde V'(\breve\varphi) + \alpha, \quad \tilde\varphi(t_0)=\varphi_0.
2938: \end{equation*}
2939: The solution then satisfies $\tilde\varphi \in L^\infty (t_0,T;H_0^1)$, so
2940: we can define a map
2941: \begin{align*}
2942: A: & L^\infty (t_0,T;H_0^1) \rightarrow L^\infty (t_0,T;H_0^1) \\
2943: & \breve\varphi \mapsto \tilde\varphi
2944: \end{align*} 
2945: which is single valued (see \cite{Evans}). It is now shown that $A$ is a
2946: contraction on $L^\infty (t_0,T;H_0^1)$ if $T$ is small enough. Let
2947: $\tilde \varphi=A(\breve\varphi)$ and $\tilde \psi= A(\breve\psi)$. Subtracting the
2948: equations for $\tilde \varphi$ and $\tilde \psi$ gives
2949: \begin{equation*}
2950: (\tilde \varphi - \tilde \psi)_t=\delta(\tilde \varphi - \tilde \psi)_{xx}
2951:   -\delta^{-1}(\tilde V'(\breve\varphi) - \tilde V'(\breve\psi)),
2952: \end{equation*}
2953: which entails
2954: \begin{equation*}
2955: ||\tilde \varphi- \tilde \psi||_{L^\infty(t_0,T;H_0^1)} \leq K ||\tilde
2956:   V'(\breve\varphi)- \tilde V'(\breve\psi)||_{L^2(t_0,T;L^2)},
2957: \end{equation*}
2958: where the constant $K$ decreases when $T$ decreases (see
2959: \cite{Evans}). The right hand side in the previous inequality may be
2960: estimated as
2961: \begin{align*}
2962: & ||\tilde
2963:   V'(\breve\varphi)- \tilde V'(\breve\psi)||_{L^2(t_0,T;L^2)} \leq ||\tilde
2964:   V''||_{L^\infty} ||\breve\varphi - \breve\psi||_{L^2(t_0,T;L^2)} \\
2965: & \leq ||\tilde
2966:   V''||_{L^\infty} ||\breve\varphi - \breve\psi||_{L^2(t_0,T;H_0^1)} \leq ||\tilde
2967:   V''||_{L^\infty} \sqrt{T-t_0} ||\breve\varphi - \breve\psi||_{L^\infty(t_0,T;H_0^1)},
2968: \end{align*}
2969: so that $A$ is a contraction when $T$ is small enough. By splitting
2970: the interval $[t_0,T]$ into smaller subintervals and using the
2971: contraction property on each such interval we obtain the existence and
2972: uniqueness of solutions to (\ref{eq:flow}) when the potential $\tilde
2973: V$ is used. There exists  a continuous representative of solutions to \eqref{eq:flow} in the
2974: equivalence class in $L^\infty(t_0,T;H_0^1)$ (see \cite{Evans} again)
2975: which we call $\varphi$. Since the solution lives in one space dimension
2976: it is continuous as a function of both space and time. So for each
2977: $M>||\varphi_0||$ there is a time $T^*$ such that $||\varphi(t)||_{C(0,1)}
2978: <M$ for all $t \leq T^*$. Thus, in a certain time interval the solution
2979: $\varphi$ is only affected by the unchanged potential $V$ (it never
2980: touches the level where $V$ changes into $\tilde V$). Consider a
2981: time in this interval, and take the inner product with $\varphi_t$ in
2982: \eqref{eq:flow} to get (using $V'(\varphi)=\varphi^3-\varphi$):
2983: \begin{equation*}
2984: ||\varphi_t||^2_{L^2} + \frac{\delta}{2}\frac{d}{dt}||\varphi_x||^2_{L^2} +
2985:   \frac{\delta^{-1}}{4}\frac{d}{dt}||\varphi||^4_{L^4}-
2986:   \frac{\delta^{-1}}{2}\frac{d}{dt}||\varphi||^2_{L^2} 
2987:   \leq \half ||\alpha||^2_{L^2}+ \half ||\varphi_t||^2_{L^2}.
2988: \end{equation*}
2989: The $||\varphi_t||^2_{L^2}$ terms are dropped and the resulting
2990: inequality is integrated from $t_0$ to $T^*$:
2991: \begin{multline*}
2992: \delta
2993: ||\varphi_x(T^*)||^2_{L^2}+\frac{\delta^{-1}}{2}||\varphi(T^*)||^4_{L^4}-\delta^{-1}
2994: ||\varphi(T^*)||^2_{L^2}\\ 
2995: \leq \delta
2996: ||(\varphi_0)_x||^2_{L^2}+\frac{\delta^{-1}}{2} ||\varphi_0||^4_{L^4}-
2997: \delta^{-1} ||\varphi_0||^2_{L^2} + \int_{t_0}^{T^*} ||\alpha||^2_{L^2} dt.
2998: \end{multline*}
2999: It is now used that
3000: \begin{equation*}
3001: \frac{\delta^{-1}}{2}||\varphi(T^*)||^4_{L^4}-\delta^{-1}
3002: ||\varphi(T^*)||^2_{L^2} =
3003: \delta^{-1}\int_0^1\big(\frac{\varphi(x,T^*)^4}{2}-\varphi(x,T^*)^2\big)dx \geq -\half
3004: \end{equation*}
3005: so that the previous inequality implies
3006: \begin{equation}\label{eq:phibound2}
3007: \delta
3008: ||\varphi_x(T^*)||^2_{L^2} 
3009: \leq \delta
3010: ||(\varphi_0)_x||^2_{L^2}+\frac{\delta^{-1}}{2} ||\varphi_0||^4_{L^4}-
3011: \delta^{-1} ||\varphi_0||^2_{L^2} + \int_{t_0}^{T^*} ||\alpha||^2_{L^2} dt +\half.
3012: \end{equation}
3013: By Sobolev's inequality we thereby obtain a bound on  the continuous
3014: function $\varphi(T^*)$ in the supremum norm.
3015: Consequently, for all controls $\alpha\in L^2(t_0,T;L^2)$ it is possible
3016: to choose the border point $s$ in Figure \ref{fig:V} between $V$ and
3017: $\tilde V$ large enough so that the solution $\varphi$ is affected only
3018: by the unchanged potential $V$. Note also that it is possible to
3019: choose $T^*=T$ in \eqref{eq:phibound2}. Such a solution is a weak solution to
3020: \eqref{eq:flow} with the original potential $V$. It is unique in
3021: $C(t_0,T;H^1_0)$, for non-uniqueness would otherwise also hold for some
3022: modified potential $\tilde V$. The error bound \eqref{eq:phibound1}
3023: follows from \eqref{eq:phibound2}.
3024: \end{proof}
3025: 
3026: %Include all entries in the bibliography file.
3027: %\nocite{*}
3028: 
3029: 
3030: % Bibliography style.
3031: \bibliographystyle{plain}
3032: 
3033: % Bib-file
3034: \bibliography{references} 
3035: 
3036: %\begin{thebibliography}{99}
3037: %\bibitem{young}
3038: %L.C. Young, Lectures on the Calculus of Variations and Optimal Control Theory. 
3039: %Saunders Co., Philadelphia-London-Toronto, Ont. 1969
3040: 
3041: %\end{thebibliography}
3042: \end{document}
3043: