math0702715/pm.tex
1: %\documentclass[11pt]{amsart}
2: \documentclass[arma]{svjour}
3: \usepackage{amsmath,amssymb,epsfig}
4: 
5: %Reintroduce lines 486-492, 496-498 in svjour.cls if publication proceeds.
6: 
7: \numberwithin{equation}{section}
8: 
9: \input{commands.tex}
10: \begin{document}
11: 
12: \title{A new well-posed nonlocal Perona-Malik type equation}
13: \author{Patrick Guidotti}
14: \titlerunning{A new Perona-Malik type equation}
15: %\date{Received: date / Revised version: date}
16: 
17: \maketitle
18: 
19: \begin{abstract}
20: A modification of the Perona-Malik equation is proposed for which the local nonlinear diffusion 
21: term is replaced by a nonlocal term of slightly reduced ``strength''. The new equation is globally 
22: well-posed (in spaces of classical regularity) and possesses desirable properties from the 
23: perspective of image processing. It admits characteristic functions as (formally linearly ``stable'') 
24: stationary solutions and can therefore be reliably employed for denoising keeping blurring in 
25: check. Its numerical implementation is stable, enhances some of the features of Perona-Malik, and 
26: avoids problems known to affect the latter.
27: \end{abstract}
28: 
29: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
30: \section{Introduction}
31: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
32: The Perona-Malik equation \cite{PM90}
33: \begin{equation}\label{pm}
34:   u_t-\dive\bigl(\frac{1}{1+|\nabla u|^2}\nabla u\bigr)=0
35: \end{equation}
36: on a domain $\Omega\overset{o}{\subset}\bbr ^n$ is usually complemented with homogeneous Neumann boundary 
37: conditions. The nonlinear diffusion coefficient
38: $$
39:  a(u)=\frac{1}{1+|\nabla u|^2}
40: $$ 
41: is sometimes replaced by $a(u)=e^{-|\nabla u|^2}$. Perona and Malik originally introduced \eqref{pm} 
42: in the context of image processing with the aim of denoising a given image $u_0$ while at the same 
43: time controlling blurring. The latter is unavoidable if a linear diffusion equation is used instead. 
44: The nonlinear diffusion determined by $a(u)$ significantly impedes diffusion in directions of steep 
45: gradients which correspond to sharp edges in the image. If one considers a large gradient across a 
46: surface, then the nonlinear diffusion coefficient leads to diffusion in tangential directions and 
47: backward diffusion in normal direction. This is the reason why models of this type are often referred 
48: to as anisotropic diffusion models. \\
49: Numerical approximations of \eqref{pm} have confirmed this prediction and have been successfully 
50: employed as a consequence. It has, however, been observed that discretizations of \eqref{pm} often lead to 
51: stair-casing in the numerical solution (see Figure \ref{eps}), or, more in general, to the preservation 
52: of gradients which do not represent any image feature but might simply be due to the presence of noise 
53: (cf. \cite{PM90,ALM92} and Figure \ref{kink_osc}). 
54: Equation \eqref{pm} poses many a challenge from the analytical point of view. It is an example of 
55: so-called forward-backward diffusion. In \cite{K97} the author shows that it is not well-posed. To 
56: circumvent this problem a variety of regularization techniques have been proposed. Spatial regularizations 
57: were the first to appear in \cite{CLMC92,ALM92}. The gain in analytical tractability is, however, offset by 
58: the introduction of blurring. Temporal and spatio-temporal regularizations have also been considered, 
59: see \cite{NS92,CEA98,CB01,B03,BC05,Ama06}. In \cite{Ama06} the author considers a purely temporal 
60: regularization which leads to time-delayed Perona-Malik equation which mimics the implicit linearization 
61: procedure common to all numerical discretizations 
62: of \eqref{pm}, that is
63: $$\begin{cases}
64:  u^{n+1}=u^n+\dive \bigl( a(u^n)\nabla u^{n+1}\bigr)\delta t&n\geq 0\, ,\\ 
65:  u^0=u_0 &n=0\, ,\end{cases}
66: $$
67: whereby the nonlinearity is always evaluated at the previous time step. The introduction of a time 
68: delay leads to a tractable analytical problem and does not seem to negatively impact the desired 
69: features of a corresponding numerical implementation. The author proves local existence of a regular 
70: solution for the delay equation.\\ 
71: Variational techniques have also been utilized to gain analytical understanding of \eqref{pm} 
72: in a one-dimensional context. The Perona-Malik equation can in fact be viewed as the gradient flow 
73: associated with the non-convex energy functional
74: $$
75:   E(u)=\int _0^1 \log\bigl( 1+|\nabla u|^2\bigr)\, ,\: u\in\oph ^1(\Omega)\, .
76: $$
77: Using concepts and techniques related to Young-measure solutions various results have been 
78: obtained. \cite{TTZ05,Z06} prove instability of certain solutions and the existence of infinitely 
79: many solutions of a certain (weak) type. \\
80: Equation \eqref{pm} was motivated by its potential applications to image processing. Whereas it 
81: poses significant and interesting mathematical challenges, its form is clearly not dictated by 
82: any physical principle. A modification is therefore proposed in this paper which is globally well-posed, 
83: and allows for special natural functions to be stationary solutions. The latter leads to a desirable 
84: dynamical behavior from the practical point of view. It also leads to stable pseudo-spectral 
85: discretizations with 
86: satisfactory properties from the perspective of image processing. It can also be viewed as a new 
87: regularization technique. It differs, however, substantially from other regularization techniques in that 
88: it actually makes piecewise constant function become stationary solutions and in that the degree of 
89: regularization can be more finely tuned. In particular it does not regularize at any given specific 
90: scale.\\
91: To simplify the discussion, analytical considerations will be restricted to the one dimensional 
92: setting. A representative numerical experiment, however, will be presented also in the natural two 
93: dimensional setting. Roughly speaking, the modification proposed consists in replacing the nonlinear term 
94: by
95: \begin{equation}\label{mpf}
96:   a_\varepsilon(u)=\frac{1}{1+[(-A)^{\frac{1-\gev}{2}}u]^2}  
97: \end{equation}
98: for $\varepsilon\in(0,\frac{1}{2})$ and where $A$ is the Neumann Laplacian. The 
99: fractional powers appearing in \eqref{mpf} can be defined in various manners depending on the 
100: choice of function spaces in which the equation is considered.
101: Since one is interested in denoising while preserving sharp edges, it is clear that $a_\gev$ should 
102: provide the very same benefits for the kind of large gradients one is interested in preserving. 
103: The associated equation 
104: \begin{equation}\label{mpm1}
105:   u_t-\bigl(\frac{1}{1+[(-A)^{1-\varepsilon}u]^2}u_x\bigr) _x=0
106: \end{equation}
107: has, however, the advantage of being quasi-linear and of avoiding backward diffusion altogether. 
108: The analysis and the numerical experiments performed in this paper therefore also show that backward 
109: diffusion is not an essential ingredient in order to preserve sharp edges. For analytical reasons 
110: the ideas just presented will be performed on an equivalent formulation of \eqref{pm} leading to 
111: an equation of the flavor of \eqref{mpf}.
112: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
113: \section{The Problem}
114: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
115: In its one dimensional formulation, the Perona-Malik equation reads
116: \begin{equation}\label{1pm}
117:  \begin{cases} u_t-\bigl(\frac{1}{1+u_x^2}u_x\bigr)_x=0 &\text{ in }(0,1)\text{ for }t>0\, ,\\
118:  u_x=0&\text{ at }x=0,1\text{ for }t>0\, ,\\
119:  u=u_0&\text{ at }t=0\text{ in }\, [0,1].
120: \end{cases}
121: \end{equation}
122: Introducing $v(x)=\int_{0}^{x}\!u(y)\,dy$ as a new unknown, it is easily seen that it satisfies
123: \begin{equation}\label{1pmi}
124:  \begin{cases}
125:   v_t-\frac{1}{1+v_{xx}^2}v_{xx}=0&\text{ in }(0,1)\text{ for }t>0\, ,\\
126:   v(t,0)=0\, ,\: v(t,1)=\int_{0}^{1}\!u_0(y)\,dy&\text{ for }t>0\, ,\\
127:   v(0,\cdot)=\int_{0}^{\cdot}\!u_0(y)\,dy&\text{ in }[0,1]\, ,
128:  \end{cases}
129: \end{equation}
130: since it is readily verified that the average of $u_0$ is preserved over time using \eqref{1pm}. 
131: By further defining 
132: $w(x)=v(x)-x\int_{0}^{1}\!u_0(y)\,dy\, ,\: x\in[0,1]\, ,$ it follows that
133: \begin{equation}\label{1pmf}
134:   \begin{cases}
135:   w_t-\frac{1}{1+w_{xx}^2}w_{xx}=0&\text{ in }(0,1)\text{ for }t>0\, ,\\
136:   w(t,0)=0=w(t,1)&\text{ for }t>0\, ,\\
137:   w(0,\cdot)=\int_{0}^{\cdot}\!u_0(y)\,dy-(\cdot)\int_{0}^{1}\!u_0(y)\,dy&\text{ in }[0,1]\, .
138:   \end{cases}
139: \end{equation}
140: Conversely, any solution of \eqref{1pmi} leads to a solution of \eqref{1pm} by setting 
141: $$
142:  u(x)=v_x(x)\, ,\: x\in[0,1]\, .
143: $$
144: The evolution equation is clearly satisfied whereas for the boundary conditions the 
145: following argument is needed. The function 
146: $$
147:  \tilde v(x)=\int_{0}^{x}\!u(y)\,dy\, ,\: x\in[0,1]\, ,
148: $$ 
149: satisfies
150: $$
151:   \tilde v_t-\frac{1}{1+\tilde v _{xx}^2}\tilde v_{xx}=-\frac{u_x(0)}{1+u_x^2(0)}\, .
152: $$
153: Since $\tilde v=v$ by $v(0)=0$ and since $v$ satisfies \eqref{1pmi}, it follows that 
154: $\frac{u_x(0)}{1+u_x^2(0)}=0$ and therefore $u_x(0)=0$. Now
155: $$
156:  \int_{0}^{1}\!u(t,y)\,dy=v(t,1)-v(t,0)=\int_{0}^{1}\!u_0(y)\,dy
157: $$
158: implies that
159: $$
160:  0=\int_{0}^{1}\!u_t(y)\,dy=\frac{u_x(1)}{1+u_x^2(1)}-\frac{u_x(0)}{1+u_x^2(0)}
161: $$
162: and therefore that $u_x(1)=0$, too. Notice that if $u$ is a piecewise constant function, then 
163: $v$ is a continuous piecewise affine function. Thus piecewise affine functions play the same 
164: role for \eqref{1pmf} as piecewise constant functions do for \eqref{1pm}.\\
165: Equation \eqref{1pmf} is fully nonlinear and clearly presents the same analytical difficulties 
166: as the original Perona-Malik equation. The following modification is proposed
167: \begin{equation}\label{mpm}
168:   \begin{cases}
169:   u_t-a_\varepsilon(u)Au=0&\text{ in }(0,1)\text{ for }t>0\, ,\\
170:   u(0,\cdot)=u_0&\text{ in }[0,1]\, ,
171:   \end{cases}  
172: \end{equation}
173: for 
174: \begin{equation}\label{mpff}
175:  a_\varepsilon(u)=\bigl(1+[(-A)^{1-\varepsilon}u]^2\bigr)^{-1}\, ,\: 
176:  \varepsilon\in(0,1/2)\, ,
177: \end{equation} 
178: and where $A$ now denotes the Dirichlet Laplacian. This is perfectly analogous to \eqref{mpf}. 
179: Notice that the intensity of the nonlinearity is barely reduced. It is therefore to be expected 
180: that solutions of \eqref{mpm} with $\varepsilon>0$ behave similarly to the solutions of \eqref{mpm} 
181: with $\varepsilon=0$ for which the original \eqref{pm} is recovered, at least at the discrete level. 
182: At the continuous level \eqref{1pm} is ill-posed and such a claim is not very meaningful.
183: It is therefore likely that \eqref{mpm} can provide a viable practical denoising tool while preventing 
184: blurring.
185: \begin{remark}\label{rem:cons}
186: It is easily verified that
187: $$
188:  \frac{d}{dt}\int_{0}^{1}\!v\,dx=0\text{ and }\frac{1}{2}\frac{d}{dt}\int_{0}^{1}\!|v|^2\,dx=
189:  -\int_{0}^{1}\!\frac{|v_x|^2}{1+[A^{1-\varepsilon}u]^2}\,dx
190: $$
191: for $v=u_x$ and a smooth solution $u$ of \eqref{mpm}.
192: \end{remark}
193: 
194: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
195: \section{Global Existence of Smooth Solutions}
196: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
197: It turns out that \eqref{mpm} can be shown to be well-posed in the classical sense as soon as 
198: $\varepsilon>0$. In order to prove this, some notation needs to be introduced. For $\alpha\in[0,1)$ 
199: let 
200: $$
201:  \opc ^\alpha _0(0,1):=\begin{cases}
202:  \{ u\in\opc ^\alpha\bigl( [0,1],\bbr\bigr)\, |\, u(j)=0\, ,\: j=0,1\} &\alpha\in(0,1)\, ,\\
203:  \opc\bigl([0,1],\bbr\bigr)& \alpha=0\, ,\end{cases}
204: $$
205: where $\opc ^\alpha\bigl( [0,1],\bbr\bigr)$ is for $\alpha>0$ the standard space of H\"older continuous 
206: functions endowed with norm
207: $$
208:  \| \cdot\| _\alpha:=\| \cdot\| _\infty +[\cdot]_\alpha\, .
209: $$
210: The semi-norm $[\cdot ]_\alpha$ is given by
211: $$
212:   [u]_\alpha:=\sup _{x\neq y}\frac{|u(x)-u(y)|}{|x-y|^\alpha}\, ,\: u\in\opc ^\alpha\bigl( [0,1],
213:   \bbr\bigr)
214: $$
215: if $\alpha>0$. Given $\alpha\in[0,1)$, the operator $A_\ga$ is given by
216: $$
217:   \begin{cases} 
218:     \operatorname{dom}{A_\alpha}=\opc ^{2+\alpha}_0(0,1):=\{ u\in\opc ^{2+\alpha}(0,1)\, |\, 
219:     u,u_{xx}\in\opc ^\alpha _0(0,1)\} &\text{ for }\alpha>0\\
220:     \operatorname{dom}{A_0}=\opc ^2_0(0,1):=\{ u\in\opc ^2(0,1)\, |\, u(j)=0\, ,\: j=0,1\}&
221:     \text{ for }\alpha=0
222:   \end{cases}
223: $$
224: and
225: $$
226:   A_\alpha u=u_{xx}\, ,\: u\in\operatorname{dom}(A_\alpha)\, .
227: $$
228: Once $\alpha\in[0,1)$ is fixed, the following simplified notation will be used
229: $$
230:   A:=A_\alpha\, ,\: E_0:=\opc ^\alpha _0(0,1)\, ,\: E_1=\opc ^{2+\alpha}_0(0,1)\, .
231: $$
232: It follows from \cite[Corollary 3.1.21, Corollary 3.1.32, Theorem 3.1.34]{Lun95} that $A$ is a sectorial 
233: operator and therefore generates an analytic semi-group $\{e^{tA}\, |\, t\geq 0\}$ on $E_0$. Then its 
234: fractional powers $(-A)^\gamma$, $\gamma\in(0,1)$, can be defined as the inverses of
235: $$
236:  (-A)^{-\gamma} :=\frac{1}{\Gamma(\gamma)}\int_{0}^{\infty}\!t^{\gamma -1}e^{tA}\,dt
237: $$
238: defined on their range, that is, by
239: $$
240:  \operatorname{dom}\bigl((-A)^\gamma\bigr)=\operatorname{R}\bigl((-A)^{-\gamma}\bigr)^{-1}\, ,\:
241:  (-A)^\gamma =\bigl((-A)^{-\gamma}\bigr)^{-1}\, .
242: $$
243: The domain of $(-A)^\gamma$ might not be an interpolation space between $E_1$ and $E_0$ but it 
244: satisfies the interpolation inequality
245: \begin{equation}\label{int-ineq}
246:  \| (-A)^\gamma x\| _{E_0}\leq c\| x\| _{E_0}^{1-\gamma}\| Ax\| _{E_0}^\gamma\, ,\: x\in
247:  \operatorname{dom}(A)\, .
248: \end{equation}
249: It can, however, be sandwiched between known interpolation spaces
250: \begin{equation}\label{sandwich}
251:  (E_0,E_1)_{\gamma,1}\hookrightarrow D\bigl((-A)^\gamma\bigr)\hookrightarrow(E_0,E_1)_{\gamma,\infty}
252: \end{equation}
253: where, for $\gamma\in(0,1)$ and $p\in[1,\infty]$, $(E_0,E_1)_{\gamma,p}$ is the standard real 
254: interpolation functor. Observe that for $\alpha=0$ it can proved that
255: $$
256:  (E_0,E_1)_{\gamma,\infty}=\opc ^{2\gamma}_0(0,1)\, ,\: \gamma\in(0,1)\setminus\{\frac{1}{2}\}\, .
257: $$
258: H\"older spaces of the time variable will also be useful. For $\beta\in(0,1)$, let
259: \begin{equation}\label{mrfs}
260:   \bbe _0:=\opc ^\beta\big([0,T],E_0\bigr)\, ,\\ \bbe _1:=\opc ^{1+\beta}\bigl([0,T],E_0\bigr)
261:   \cap\opc ^{\beta}\bigl([0,T],E_1\bigr)\, .
262: \end{equation}
263: If $u\in\bbe _1$, then
264: $$
265:  \frac{1}{1+[(-A)^{1-\varepsilon}u(t)]^2}A
266: $$
267: is sectorial for every  fixed $t$. By H\"older maximal regularity (see \cite[Proposition 6.1.3]{Lun95}) 
268: it follows that any solution of
269: $$\begin{cases}
270:  v_t-\frac{1}{1+[(-A)^{1-\varepsilon}u]^2}Av=0 &\text{ for }t>0\\
271:  v(0)=u_0&\end{cases}
272: $$
273: satisfies $u\in\bbe _1$ for any $u_0\in E_1$ such that $Au_0\in (E_0,E_1)_{\beta,\infty}$. In order 
274: to apply the mentioned theorem, one needs also to take into account the additional facts that
275: $$
276:  [t\mapsto A(t):=\frac{1}{1+[(-A)^{1-\varepsilon}u(t)]^2}A]\in\opc ^\beta\bigl([0,T],\mathcal{L}
277:  (E_1,E_0)\bigr)
278: $$
279: and 
280: $$
281:  \operatorname{dom}\bigl( A(t)\bigr)=\operatorname{dom}\bigl(A(0)\bigr)\, ,\: t\in[0,T]\, .
282: $$
283: It is now possible to prove the following result.
284: \begin{theorem}\label{thm:ex}
285: Let $u_0\in E_1$ such that $Au_0\in (E_0,E_1)_{\beta,\infty}$. Then there exists $T^+(u_0)$ and a unique 
286: $u:\bigl[0,T^+(u_0)\bigr)\to E_1$ such that $u\big |_{[0,T]}$ is a solution of \eqref{mpm} on 
287: the time interval $[0,T]$ for any $0<T<T^+(u_0)$. Furthermore
288: $$
289:  u\in\opc ^{1+\beta}\bigl( [0,T^+(u_0)),E_0\bigr)\cap\opc ^{\beta}\bigl( [0,T^+(u_0)),E_1\bigr)
290: $$
291: \end{theorem}
292: \begin{proof}
293: Let $v\in\bbe _1$ and define $\Phi(v)$ to be the solution of
294: $$
295:  u_t-\frac{1}{1+[(-A)^{1-\varepsilon}v]^2}Au=0\, ,\: u(0)=u_0\, .
296: $$
297: Define $r=2\| u_0-\Phi(u_0)\| _{\bbe _1}$ and let
298: $$
299:  B_r:=\{ v\in\bbe _1\, |\, \| v-u_0\| _{\bbe _1}\leq r\}\, .
300: $$
301: It follows that
302: \begin{multline*}
303:   \| \Phi(v)-u_0\| _{\bbe _1}\leq\| \Phi(v)-\Phi(u_0)\| _{\bbe _1}+\| \Phi(u_0)-u_0\| _{\bbe _1}=\\
304:   \big\|\bigl[ \partial _t-a_\varepsilon(v)A\bigr]^{-1}(0,u_0)-\bigl[ \partial _t-a_\varepsilon(u_0)A
305:   \bigr]^{-1}(0,u_0)\big\| _{\bbe _1}+\frac{r}{2}\\ 
306:   =\Big\|\bigl[ \partial _t-a_\varepsilon(u_0)A\bigr]^{-1}\Big\{ 
307:   1-\bigl[ \partial _t -a_\varepsilon(u_0)A\bigr]\bigl[ \partial _t-a_\varepsilon(v)A\bigr]^{-1}
308:  \Big\}(0,u_0)\Big\| _{\bbe _1}+\frac{r}{2}\\\leq r
309: \end{multline*}
310: if $T>0$ is chosen small enough since $v\in B_r$ and $v(0)=u_0$, so that $v$ is uniformly close 
311: to $u_0$. To see this, consider
312: \begin{multline}\label{eq1}
313:  \|[a_\varepsilon(v)-a_\varepsilon(u_0)]A\| _{\mathcal{L}(\bbe _1,\bbe _0)}\leq
314:  \| a_\varepsilon(v)-a_\varepsilon(u_0)\| _{\mathcal{L}(\bbe _0)}\\\leq\|(-A)^{1-\varepsilon}(v-u_0)
315:  \|_{\bbe _0} 
316: \end{multline}
317: which follows from
318: $$
319:  \| \frac{1}{1+w^2(t)}-\frac{1}{1+w^2(s)}\| _{E_0}\leq \| w(s)-w(t)\|_{E_0}
320: $$
321: by setting $w=(-A)^{1-\varepsilon}v$ since $v(0)=u_0$ and $a_\varepsilon(u_0)$ are independent of 
322: the time variable.
323: To further estimate \eqref{eq1}, one can use the interpolation inequality \eqref{int-ineq} and 
324: \begin{equation}\label{eq2}
325:  v(t)-v(s)=\int _s^t\dot v(\tau)\, d\tau\text{ and } \dot v\in\opl _\infty E_0\, 
326: \end{equation}
327: to obtain
328: \begin{multline*}
329:   \| (-A)^{1-\varepsilon}[v(t)-u_0]\| _{E_0}\leq c\|v(t)-u_0\| _{E_1}^{1-\varepsilon}\| 
330:   v(t)-u_0\| _{E_0}^\varepsilon\\\leq crt^{\beta(1-\varepsilon)+\varepsilon}\, ,\: s,t\leq T\, ,
331: \end{multline*}
332: and
333: \begin{multline*}
334:   \| (-A)^{1-\varepsilon}[v(t)-v(s)]\| _{E_0}\leq c\| v(t)-v(s)\| _{E_1}^{1-\varepsilon}
335:  \| v(t)-v(s)\| ^\varepsilon _{E_0}\\\leq cr|t-s|^{\beta(1-\varepsilon)+\varepsilon}\, ,\: s,t\leq T\, .
336: \end{multline*}
337: It follows that
338: $$
339:  \bigl[(-A)^{1-\varepsilon}(v-u_0)\bigr] _{\beta}\leq cr T^{(1-\beta)\varepsilon}\, .
340: $$
341: and finally that
342: $$
343:  \|[a_\varepsilon(v)-a_\varepsilon(u_0)]A\| _{\mathcal{L}(\bbe _1,\bbe _0)}\leq crT^{(1-\beta)
344:  \varepsilon}
345: $$
346: Thus the claimed inequality follows by a simple Neumann series argument made possible by choosing 
347: $T$ small enough. Next observe that $\Phi(v_1)-\Phi(v_2)$ solves
348: \begin{equation*}\begin{cases}
349:  [\Phi(v_1)-\Phi(v_2)] _t=a_\varepsilon(v_1)A[\Phi(v_1)-\Phi(v_2)]+\bigl( a_\varepsilon(v_1)-
350:  a_\varepsilon(v_2)\bigr)A\Phi(v_2)\\ [\Phi(v_1)-\Phi(v_2)](0)=0
351: \end{cases}
352: \end{equation*}
353: It follows that
354: $$
355:  \Phi(v_1)-\Phi(v_2)=\bigl[ \partial _t-a_\varepsilon(v_1)A\bigr]^{-1}\Big\{ 0,\bigl[ a_\varepsilon(v_1)-
356:  a_\varepsilon(v_2)\bigr]A\Phi(v_2)\Big\}\, .
357: $$
358: Thus
359: \begin{multline*}
360:   \|\Phi(v_1)-\Phi(v_2)\| _{\bbe _1}\leq c(r)\|[a_\varepsilon(v_1)-a_\varepsilon(v_2)]A\Phi(v_2)\| _{E_0}
361:   \\\leq c(r)\| A^{1-\varepsilon}[v_1-v_2]\| _{\bbe _0}\| \Phi(v_2)\| _{\bbe _1}
362: \end{multline*}
363: Now, since $v_1-v_2\in\bbe _1$ it follows that
364: $$
365:  \| A^{1-\varepsilon}[v_1-v_2]\| _{\bbe _0}\leq c\| v_1-v_2\| _{\bbe _1} T^{\varepsilon(1-\beta)}
366: $$
367: by making use of the interpolation inequality \eqref{int-ineq} and of \eqref{eq2} in a perfectly 
368: similar way to the calculations used to obtain the self-map property.
369: Thus Banach fixed-point Theorem implies the existence of a unique solution 
370: $$
371:  u\in\bbe _1
372: $$ 
373: provided the time interval is chosen short enough. It follows from \cite[Proposition 6.1.3]{Lun95} that 
374: $$
375:   \dot u(t)\in (E_0,E_1)_{\beta,\infty}\, ,\: t\in (0,T]
376: $$ 
377: which gives $Au(T)\in(E_0,E_1)_{\beta,\infty}$. Thus, by using the same argument, the solution can be 
378: extended to a larger interval of existence. Repeating this argument indefinitely, a solution is obtained 
379: on a maximal interval of existence $\bigl[ 0,T^+(u_0)\bigr)$ with the stated properties.
380: \end{proof}
381: Next classical Sobolev spaces are needed
382: $$
383:  \opw ^2_p(0,1)=\big\{ u\in\opl _p(0,1)\, \big |\, \partial^j u\in\opl _p(0,1)\text{ for }
384:  0\leq j\leq 2\big\}
385: $$
386: for $p\in(0,\infty)$ where the derivatives have of course to be understood in the distributional 
387: sense. The simple embedding
388: \begin{equation}\label{rkineq}
389:   \opw ^s_p(0,1)\hookrightarrow\opc ^{s-1/p}(0,1)\, ,\: sp>1\, ,
390: \end{equation}
391: will be very useful in the proof of the next Lemma.
392: \begin{lemma}\label{lem:glob-ex}
393: If it can be shown that $u\in\opl _\infty\bigl(\bigl[ 0,T^+(u_0)\bigr),\opw ^2_p(0,1)\bigr)$ for all 
394: $p\in(1,\infty)$, then the solution exists globally in time.
395: \end{lemma}
396: \begin{proof}
397: Observe that
398: $$
399:  \dot u=a_\varepsilon(u)\triangle u\in\opl _\infty\bigl(\bigl[ 0,T^+(u_0)\bigr),\opl _p(0,1)\bigr)
400: $$
401: follows from the assumption and the form of $a_\varepsilon$. For $s,t\in\bigl[ 0,T^+(u_0)\bigr]$, one has
402: \begin{multline*}
403:  \| A^{1-\varepsilon}\bigl( u(t)-u(s)\bigr)\| _{\opc ^{\tilde{\rho}}}\\\leq c\,
404:  \| A^{1-\varepsilon}\bigl( u(t)-u(s)\bigr)\| _{\opw ^{2\rho}_p}\leq c\,
405:  \| A^{1-\varepsilon+\rho}\bigl( u(t)-u(s)\bigr)\| _{\opl _p}\\\leq
406:  c\, \| A\bigl( u(t)-u(s)\bigr)\| _{\opl _p}^{1-\varepsilon+\rho}\|\int_{s}^{t}\!\dot u(\tau)\,d\tau
407:  \| _{\opl _p}^{\varepsilon-\rho}\leq c\, |t-s|^{\varepsilon-\rho}
408: \end{multline*}
409: as follows from \eqref{rkineq} and \eqref{int-ineq}. Hereby it needs to be assumed that
410: $$
411:  \tilde\rho =2\rho-\frac{1}{p}>0\text{ and }0<\rho<\varepsilon\, .
412: $$
413: This can always be achieved by choosing $p$ large enough and yields
414: \begin{equation}\label{reg}
415:  A^{1-\varepsilon}u\in\opc ^{\varepsilon-\rho}\bigl(\bigl[ 0,T^+(u_0)\bigr],\opc ^{\tilde{\rho}}(0,1)
416:  \bigr)\, .  
417: \end{equation}
418: Denote the function space in \eqref{reg} by $\bbe _0$. Let then $v$ be the solution of
419: $$
420:  \dot v -a_\varepsilon(u)Av=0\, ,\:v(0)=u_0
421: $$
422: on $\bigl[ 0,T^+(u_0)\bigr]$. It satisfies $v\in\bbe _1$ and one has
423: $$
424:  v\big | _{[0,T]}=u\big |_{[0,T]}\, ,\: T<T^+(u_0)
425: $$
426: by uniqueness. Since $v$ is uniformly (H\"older) continuous, so must be $u$. It can therefore be 
427: extended as a solution to $\bigl[ 0,T^+(u_0)\bigr]$ and as a consequence of Theorem \ref{thm:ex} beyond 
428: any finite time. The solution is therefore global. 
429: \end{proof}
430: \begin{lemma}\label{lem:smooth}
431: Let $u_0\in\opc^\infty\bigl([0,1]\bigr)$ satisfy compatibility conditions to all orders, that is, 
432: assume that $u_0\in\operatorname{dom}\bigl((-A)^k\bigr)$, $k\in\bbn$. Then the solution of \eqref{mpm} 
433: satisfies
434: $$
435:  u\in\opc ^{1+\beta}\Bigl(\bigl[ 0,T^+(u_0)\bigr),\opc ^\infty\bigl([0,1]\bigr)\Bigr)\, ,
436: $$
437: for $\beta<\varepsilon$.
438: \end{lemma}
439: \begin{proof}
440: Let $u$ be the solution of \eqref{mpm}. Then 
441: $$
442:  (-A)^{1-\varepsilon}u\in\opc ^\beta\Bigl(\bigl[ 0,T^+(u_0)\bigr),\opc ^{\alpha+2\beta}_0(0,1)
443: \Bigr)
444: $$ 
445: since $\beta<\varepsilon$ and $u$ is therefore also a solution of \eqref{pm} in $\bbe _1$ for 
446: $\alpha+\beta$. By repeating this boot-strapping argument indefinitely one obtains the claim.
447: \end{proof}
448: \begin{remark}\label{rem:timereg}
449: Time regularity can also be improved by similar boot-strapping arguments but it will play no role in 
450: this paper. 
451: \end{remark}
452: It is now possible to prove the following global regularity results.
453: \begin{theorem}\label{thm:mthm}
454: The maximal solution of \eqref{mpm} exists globally.
455: \end{theorem}
456: \begin{proof}
457: According to Lemma \ref{lem:glob-ex} it is enough to show 
458: $$
459:  \| Au(t)\| _{\opw ^2_p}\leq c\, ,\: t\in\bigl[ 0,T^+(u_0)\bigr)\text{ for }p\in(0,\infty)\, .
460: $$
461: By Lemma \ref{lem:smooth} it follows from \eqref{pm} that
462: $$\begin{cases}
463:  A\dot u-A\bigl( a_\varepsilon(u)Au\bigr)=0\, ,\\
464:  Au(0)=u_0\, .
465: \end{cases}
466: $$
467: Thus $v:=Au$ satisfies
468: \begin{equation}\label{Aeq}\begin{cases}
469:  \dot v-A\bigl( a_\varepsilon(u)v\bigr)=0\, ,\\
470:  v(0)=u_0\, .
471: \end{cases}
472: \end{equation}
473: Observe that $a_\gev(u(t))A\cdot$ is, for any fixed time $t$, formally adjoint to 
474: the operator $A\bigl[ a_\varepsilon\bigl(u(t)\bigr)\cdot\bigr]$. 
475: The evolution operator $U_{\mathbb{A}}(t,\tau)$ generated by $\mathbb{A}=a_\varepsilon(u)A$ on $E_0$ 
476: satisfies
477: $$
478:  \| U_{\mathbb{A}}(t,\tau)u_0\| _{\opl _p(0,1)}\leq \| u_0\| _{\opl _p(0,1)}\, .
479: $$
480: This follows from the Trotter-Kato product formula (see \cite{Pa83}) and the fact that the operator
481: $$
482:  c(x)A
483: $$
484: generates a contraction semigroup on $\opl _p(0,1)$ if ellipticity is assumed (cf. \cite{A07}).  
485: Let $T<T^+(u_0)$ and consider the family of generators $\mathbb{B}=\mathbb{A}(T-t)$, $t\in[0,T]$. 
486: Then an easy computation based on
487: $$
488:  \frac{d}{d\tau}U_{\mathbb{A}}(t,\tau)u_0=-U_{\mathbb{A}}(t,\tau)\mathbb{A}(\tau)u_0\, ,\:
489:  u_0\in\operatorname{dom}{\mathbb{A}(0)}\, ,
490: $$
491: reveals that
492: $$
493:  U^*_{\mathbb{B}}(T,T-t)v_0\, ,\: t\in[0,T]\, ,
494: $$
495: satisfies \eqref{Aeq} if $U^*_{\mathbb{B}}$ on $\opl _{p'}(0,1)$ is taken to be the evolution operator 
496: dual to $U_{\mathbb{B}}$. Now, since
497: $$
498:  \| U^*(t,\tau)\| _{\mathcal{L}(\opl _{p'})}=\| U(t,\tau)\| _{\mathcal{L}(\opl _p)}\leq 1
499: $$
500: it follows that
501: $$
502:  \| v(t)\| _{\opl _{p'}}\leq \| v_0\| _{\opl _{p'}}=\| Au_0\| _{\opl_{p'}}
503: $$
504: for any $p\in(1,\infty)$ and therefore 
505: $$
506:  \| Au(t)\| _{\opl _p}\leq c\,\| u\| _{\opw ^2_p}\leq c\,\| Au_0\| _{\opl _p}
507: $$
508: for all $p\in(1,\infty)$.
509: The proof is thus complete since $T<T^+(u_0)$ is arbitrary and the embedding constant $c$ does not 
510: depend on $T$.
511: \end{proof}
512: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
513: \section{Stationary Solutions}%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
514: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
515: Among constant functions, $u\equiv 0$ is clearly  a stationary solution of \eqref{mpm}. It 
516: corresponds to a constant solution with value $\int _0^1 u_0(x)\, dx$ of the original Perona-Malik 
517: equation with initial value $u_0$. It turns out, however, that other more desirable 
518: functions are equilibria of \eqref{mpm} as well. Even though the next result remains valid for \eqref{mpm}, 
519: it is formulated in the context of the modified Perona-Malik equation with periodic boundary conditions 
520: instead of homogeneous Dirichlet conditions. This allows the simplify a little the argument.
521: \begin{proposition}\label{sationary}
522: Let $u$ be a continuous $1$-periodic piecewise affine function. Then
523: $$
524:  \frac{1}{1+[(-A)^{1-\varepsilon}u]^2}Au\equiv 0\, ,
525: $$
526: for $A$ taken to be the realization of $\partial_{xx}$ in the space of Radon measures $M_\pi$. 
527: The subscript $\pi$ indicates that the periodic case is considered. 
528: \end{proposition}
529: \begin{proof}
530: If $u$ has the properties stated in the proposition, it follows that
531: $$
532:  \partial_{xx}u=\sum _{j=1}^nc_j\delta _{x_j}
533: $$ 
534: for some constants $c_j\in\bbr$ and locations $x_1,\dots,x_n\in[0,1)$. The claim would therefore 
535: follow if it can proved that
536: $$
537:  \frac{1}{1+[(-A)^{1-\varepsilon}u]^2} 
538: $$
539: is a continuous function vanishing at $x_j\, ,\: j=1,\dots,n$.
540: It is clear that it is enough to consider the case of a function with a single kink. Furthermore, since 
541: periodicity is assumed, Poisson's summation formula implies that the singularity in 
542: $(-A)^{1-\varepsilon}u$ in [0,1) is the same as the singularity at the same location for 
543: $(-A)^{1-\varepsilon}u$ periodically extended to real line. By making use of standard localization 
544: techniques, it is therefore sufficient to 
545: consider a function of the real line with a single kink in the origin and $A=\partial_{xx}$ on the 
546: full line. Choose the prototype $u(x)=|x|$. Since
547: $$
548:   (-A)^{1-\varepsilon}u=(-A)^{-\varepsilon}\delta
549: $$
550: it follows that $\mathcal{F}\bigl((-A)^{1-\varepsilon}u\bigr)=\frac{1}{|\xi |^{2\varepsilon}}$ and 
551: therefore that
552: $$
553:  (-A)^{1-\varepsilon}u=c_\varepsilon\frac{1}{|x|^{1-2\varepsilon}}\, .
554: $$
555: for $C_\varepsilon=\sqrt{\frac{2}{\pi}}\Gamma(1-2\varepsilon)\sin(\pi\varepsilon)$. 
556: Remember that it has been assumed initially that $\varepsilon\in(0,1/2)$ to make the nonlinearity 
557: strong enough. It follows that, in the general case, $(-A)^{1-\varepsilon}u$ has integrable 
558: singularities at $x_j$, $j=1,\dots,n$, and is otherwise smooth by hypo-ellipticity. This shows 
559: that
560: $$
561:  \frac{1}{1+[(-A)^{1-\varepsilon}u]^2}
562: $$
563: is a continuous function which vanishes exactly at $x_j$, $j=1,\dots,n$. It is therefore a well-defined 
564: multiplier for $\sum _{j=1}^nc_j\delta _{x_j}$ and one has
565: $$
566:  \frac{1}{1+[(-A)^{1-\varepsilon}u]^2}Au=0
567: $$
568: as claimed.
569: \end{proof}
570: \begin{remark}\label{rem:norealstationary}
571: It should be observed that it is not possible (to the best of our knowledge) to make sense of 
572: \eqref{pm} as generating a flow on $M_\pi$ (or on any of the standard Besov spaces containing Dirac 
573: distributions) and so the stationary solutions of the proposition are only formally such. 
574: A short computation would even show that they are formally linearly 
575: stable. This might help explain the fact that, in numerical calculations of \eqref{mpm}, it is 
576: observed that solution starting close to such a stationary solution tend to stay in its vicinity 
577: for a long time before being driven to the trivial steady state. Even more is actually true. Numerical 
578: experiments (see Figure \ref{regev}) show that smooth solutions tend to become piecewise affine at first 
579: but are eventually completely smoothed out. 
580: \end{remark}
581: \begin{remark}\label{rem:linearization}
582: Since the weakened nonlinearity \eqref{mpff} does not give rise to forward-backward diffusion for the 
583: linearization, global well-posedness can be proved. Its linearization, however, will be closer and closer 
584: to the Perona-Malik linearization as $\gev$ decreases. As a consequence more and more of its eigenvalues 
585: will become negative when linearizing in the presence of large gradients. This is supporting evidence 
586: that the numerical benefits of Perona-Malik should not go lost in the modified \eqref{mpm}.  
587: \end{remark}
588: \begin{remark}\label{rem:asym-behavior}
589: The new equation is globally well-posed for smooth initial conditions. It is, however, not immediately 
590: clear what their large time asymptotic behavior should be. Numerical experiments seem to suggest that 
591: any solution will eventually converge to a trivial steady-state. This behavior might, however, be due 
592: to numerical diffusion. In the next section a reason will be given which seems to exclude the latter 
593: possibility.
594: \end{remark}
595: \begin{remark}\label{rem:stability}
596: It is also observed that the original Perona-Malik equations do not admit piece-wise constant 
597: functions as formal stationary points unless an ad-hoc concept of generalized stationary solution 
598: is introduced (see \cite{K97}) since the multiplication of distributions is in general not well-defined. 
599: Its discretization, however, seems to exhibit ``almost stability'' of such solutions in the sense 
600: explained here. This might be an indication that the Perona-Malik equation is regularized by 
601: discretization and could explain the success of its implementations in spite of the ill-posedness of 
602: its analytical counterpart.
603: \end{remark}
604: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
605: \section{Numerical implementation and experiments}
606: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
607: Next a numerical discretization of \eqref{mpm} is derived in a periodic context and used to perform 
608: numerical experiments intended to illustrate and demonstrate the claimed improvements on the classical 
609: Perona-Malik equation. The periodic Laplacian $A$ is discretized spectrally by means of the fast Fourier 
610: transform $F_n$
611: \begin{equation}\label{plaplacian}
612:   A_n=F_n^{-1}\Lambda _nF_n
613: \end{equation}
614: where $n=2^m$ denotes the number of grid points used and
615: $$
616:  D_n=4\pi ^2\operatorname{diag}((-\frac{n}{2}+1)^2,(-\frac{n}{2}+2)^2,\dots, 
617:  0,1,\dots,\frac{n^2}{4})\, .
618: $$
619: The time variable is discretized by forward semi-implicit Euler so that
620: \begin{equation}\label{time-stepping}
621:   u^{k+1}=\bigl[\opid _n-\frac{h_t}{1+(A_n^{1-\varepsilon}u^k)^2}A_n  \bigr]^{-1}u^k
622: \end{equation}
623: where $u^k$ is the spatial $n$-vector at time $k\, h_t$ for the time step $h_t>0$. Observe 
624: that, setting $\varepsilon=0$, the classical Perona-Malik equation is recovered.\\
625: The following experiments are presented here. First the evolution of the function
626: $$
627:   u_0(x):=100x^2(1-x^2)\, ,\: x\in[0,1]\, .
628: $$
629: is considered. Figure \ref{eps} depicts the first derivative of the function $u^k$ after $100$ 
630: time steps of size $h_t=0.06$ for $m=8$. The blue curve corresponds to the Perona-Malik solutions 
631: and the stair-casing 
632: phenomenon is apparent. The red, magenta and cyan curves correspond to solution of the modified 
633: equation for $\varepsilon=0.3,0.2,0.1$, respectively. They clearly benefit from all effects of the 
634: Perona-Malik equation without leading to stair-casing. Some decrease in contrast is the only price 
635: paid. Notice that the black curve depicts the first derivative of the initial value. Comparison with 
636: the other curves reveals the de-blurring effect of the equation.
637: \begin{figure}
638: \begin{center}
639: \includegraphics[scale=0.2]{pm1.jpg}
640: \caption{The gradient of the solution for different values of $\varepsilon$.}\label{eps}
641: \end{center}
642: \end{figure}
643: It has been proved that continuous piecewise affine functions are steady-states for the modified 
644: Perona-Malik equation. There is, however, no mechanism by which a smooth solution can develop a 
645: singularity in finite time. The solution therefore feels the presence of these nontrivial steady-states 
646: but never leaves the regime where diffusion, slowly but surely, drives it to a trivial steady-state. 
647: This is exemplified in Figure \ref{regev} where various stages of the evolution of a smooth solution are 
648: with initial condition
649: $$
650:  u_0(x)=\sin(2\pi x)+2\sin(4\pi x)\, ,\: x\in[0,1]\, .
651: $$
652: is depicted. Clearly the convergence towards the trivial steady state might be due to numerical 
653: diffusion (see Remark \ref{rem:asym-behavior}). However, Remark \ref{rem:cons} provides an indirect 
654: way to test the numerics. If follows from Remark \ref{rem:cons} that
655: $$
656:  \int_{0}^{1}\!u_x^2\,dx=\int_{0}^{1}\!(u_0)_x^2\,dx-2\int_{0}^{t}\!\int_{0}^{1}\!
657:  \frac{u_{xx}^2}{1+[A^{1-\varepsilon}u]^2}\,dx\,d\tau\, .
658: $$
659: This conservation relation can be tracked for any smooth solution and, in the particular case considered, 
660: the relative numerical deviation observed between the left and right-hand-side of it amounts to a mere 
661: $0.2$\% for a fully converged solution.
662: \begin{figure}
663: \begin{center}
664: \includegraphics[scale=0.5]{regev.jpg}
665: \caption{Tendency to evolve towards or close to piece-wise affine functions.}\label{regev}
666: \end{center}
667: \end{figure}
668: Scheme \eqref{time-stepping} also preserves the piecewise affine structure of initial values for 
669: large integration times. The solution with initial condition
670: $$
671:  u_0(x)=5-|10\,x-5|\, ,\: x\in(0,1)\, ,
672: $$
673: is computed for $100$ time steps with $h_t=0.06$ and $m=8$. The solutions are plotted in Figure \ref{kink}. 
674: The color coding is as in the previous experiment. In spite of the fact that $u_0$ is a steady-state 
675: for any choice of $\varepsilon\in(0,\frac{1}{2})$, numerical dissipation is stronger for larger 
676: $\varepsilon$ as the relative strength of the non-linearity decreases. The initial value and the solution 
677: to $\varepsilon=0.1$ are indistinguishable in the plot. The differ by $0.2\%$ in the supremum norm. 
678: Continuous piecewise linear functions are not steady-states for the original Perona-Malik equation since 
679: the non-linearity is not well-defined for such functions. In spite of this its numerical counterpart 
680: delivers results comparable to those for small positive $\varepsilon$.
681: \begin{figure}
682: \begin{center}
683: \includegraphics[scale=0.2]{pm2.jpg}
684: \caption{Behavior close to formally stationary solutions.}\label{kink} 
685: \end{center}
686: \end{figure}
687: Figure \ref{kink_osc} shows clearly one of the claimed enhancements of \eqref{mpm} over the original 
688: Perona-Malik equations. The initial condition (in red)
689: $$
690:  u_0(x)=5-|10\,x-5|+0.2\sin(64\pi x)\, ,\: x\in(0,1)\, ,
691: $$
692: is evolved to time $t=2$ (in blue) with $\varepsilon=0.3$. The new equation can manifestly 
693: differentiate between high low contrast gradients and high contrast gradients, which are remarkably 
694: well preserved. Observe that the initial oscillatory condition would be left virtually unchanged by 
695: the original Perona-Malik equation for the same and longer time ranges.
696: \begin{figure}
697: \begin{center}
698: \includegraphics[scale=0.5]{kink_osc.jpg}
699: \caption{The effect of \eqref{mpm} on high frequency low contrast oscillations. The initial condition 
700: (in red) is rapidly evolved to the almost piecewise affine function (in blue).}\label{kink_osc}
701: \end{center}
702: \end{figure}
703: Even though this paper is concerned with a one-dimensional modification of Perona-Malik, equation 
704: \eqref{mpm} can be considered in a natural two dimensional setting. The qualities of the its 
705: one dimensional counterpart considered here do carry over to that case. Figure \ref{teaser} 
706: shows the evolution of a noisy test image every pixel of which has been corrupted by about 15\% noise 
707: in the gray-scale. For other tests and details of the two-dimensional implementation we refer to 
708: \cite{GL07}.
709: \begin{figure}
710: \begin{center}
711: \includegraphics[scale=0.6]{teaser.jpg}
712: \caption{The denoising effect obtained for the 2D version of \eqref{mpm} with 
713:          $\varepsilon=0.6$ and Neumann boundary conditions. The initial condition is Lenna's image 
714:          corrupted with about 15\% salt and pepper noise.}\label{teaser}
715: \end{center}
716: \end{figure}
717: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
718: \section{Conclusions}
719: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
720: A modified Perona-Malik equations has been proposed by strength reduction in the non-linearity via 
721: a non-local term. Global well-posedness of the new equation and the fact that piecewise constant 
722: functions are steady-states for it are its main analytical advantages. From the practical point 
723: of view, the new equation delivers enhanced benefits as compared to Perona-Malik and suppresses known 
724: shortcomings associated to it. It provides an easy, effective, and stable tool for image de-noising. 
725: These claims are corroborated by numerical experiments.
726: 
727: \bibliographystyle{plain}
728: \bibliography{../../lite}
729: 
730: \address{103 Multipurpose Science and Technology Building\\ Department of Mathematics\\ 
731:          University of California\\ Irvine, CA 92697-3875 USA\\ email: {gpatrick@math.uci.edu}}
732: 
733: 
734: \end{document}
735: