math0612597/cm8.tex
1: % G. Crasta, A. Malusa
2: %
3: % A variational approach to the macroscopic electrodynamics
4: % of anisotropic hard superconductors
5: %
6: % last update: dec. 20, 2006
7: %
8: \documentclass[11pt, psamsfonts]{amsart}
9: \usepackage{amssymb, amsmath, amsthm}
10: %\usepackage[notref,notcite]{showkeys}
11: \usepackage{graphicx}
12: %\usepackage[final, hypertex, backref]{hyperref}
13: \usepackage[final, hypertex]{hyperref}
14: %geometry
15: \usepackage[a4paper, centering]{geometry}
16: %per impostare la pagina normale
17: %\geometry{text={12.6cm, 21cm}}
18: %per impostare la pagina media
19: \geometry{text={15cm, 22cm}}
20: %per impostare la pagina larga
21: %\geometry{text={17cm, 23cm}}
22: 
23: % .pdf figures
24: %\newcommand{\picturename}[1]{#1.pdf}
25: % .eps figures
26: \newcommand{\picturename}[1]{#1.eps}
27: 
28: %\newif\iflargepage
29: %\largepagefalse
30: 
31: \newtheorem{theorem}{Theorem}[section]
32: \newtheorem{prop}[theorem]{Proposition}
33: \newtheorem{lemma}[theorem]{Lemma}
34: \newtheorem{cor}[theorem]{Corollary}
35: \theoremstyle{definition}
36: \newtheorem{dhef}[theorem]{Definition}
37: \theoremstyle{remark}
38: \newtheorem{rk}[theorem]{Remark}
39: \newtheorem{ehse}[theorem]{Example}
40: 
41: \newtheorem{claim}[theorem]{{\it Claim}}
42: \newenvironment{claim-b}
43:    {\begin{claim}\rm}
44:    {\end{claim}}
45: 
46: % Modify page dim
47: %\iflargepage
48: %\textwidth17cm
49: %\textheight22cm
50: %\hoffset-20mm
51: %\voffset-20mm
52: %\parindent8pt
53: %\fi
54: 
55: \long\def\elimina#1{} \long\def\TODO#1{\par\noindent\textbf{TODO:
56: } #1\par}
57: 
58: \def\R{\mathbb{R}}
59: \def\C{\mathbb{C}}
60: \def\N{\mathbb{N}}
61: \def\U{\mathcal{U}}
62: \def\Wuu{W^{1,1}_0}
63: %\def\haus{d_H}
64: \def\inrad#1{r_{#1}}
65: \def\inradius{\inrad{\Omega}}
66: \def\pscal#1#2{\left\langle#1,\,#2\right\rangle}
67: %\def\estig{\sup_{\Omega} g}}
68: %\def\estig{\|g\|_{\infty}}
69: \def\Hausn#1{\mathcal{H}^{n-1}(#1)}
70: \def\haus{\mathcal{H}^{n-1}}
71: \def\dimh{\textrm{dim}_{\mathcal{H}}}
72: \def\K{\mathcal{K}}
73: \def\convopen{\mathcal{K}^n_0}
74: %support function
75: \def\support#1{h_{#1}}
76: \def\supp{\support{K}}
77: %gauge
78: \def\gau#1{\rho_{#1}}
79: \def\gauge{\rho}
80: \def\gaugem{\rho_{-}}
81: \def\pgauge{\gauge^0}
82: %Minkowski distance from a set $S$
83: \def\dists{\delta_S}
84: %Minkowski distance from the boundary
85: \def\distb#1{d_{#1}}
86: %Minkowski distance from $\partial\Omega$
87: \def\dist{\distb{\Omega}}
88: \def\distm{\distb{\Omega}^-}
89: %signed Minkowski distance from $\partial\Omega$
90: \def\sdist{\dist^s}
91: %cut point
92: \def\cut{m}
93: %unit vector
94: \def\uv{\nu}
95: %principal curvatures
96: \def\curv{\kappa}
97: %$\rho$-principal curvatures
98: \def\curvg{\tilde{\kappa}}
99: \def\curvgm{\tilde{\kappa}^-}
100: %normal
101: \def\nor{\nu}
102: %\def\nor{\mathbf{n}}
103: %ray length
104: \def\leno{\tilde{l}}
105: %ray length extended to closure
106: \def\len{l}
107: %tubular neighborhood
108: \def\tube#1{\mathcal{T}(#1)}
109: %ridge set
110: \def\ridge{\mathcal{R}}
111: \def\xia{\hat{\xi}_t}
112: \def\Lip{\textrm{Lip}}
113: \def\Lipr{\Lip_{\rho}}
114: \def\bw{\overline{W}}
115: \def\bv{\overline{V}}
116: \def\vf{v}
117: \def\meas{\mathcal{L}^n}
118: \def\matr{\mathcal{M}}
119: %\def\uu{w}
120: %\def\vv{z}
121: %\def\vz{v}
122: \def\uz{u}
123: \def\Cb{C_b}
124: \def\dato{\overline{u}}
125: \def\dt{\delta t^n_i}
126: \def\macroy{y}
127: \def\nt{\vec{n}}
128: \def\normp#1{{\left\|#1\right\|}_p}
129: \def\Uz{\vec{U}_0}
130: \def\Usol{\vec{U}}
131: \def\cz{\vec{c}_0}
132: \def\gd{\gauge_{\Delta}}
133: \def\elle#1{L^{#1}(A,\R^3)}
134: \def\wp#1{W^{1,#1}(A,\R^3)}
135: 
136: \DeclareMathOperator{\relint}{ri} \DeclareMathOperator{\relbd}{rb}
137: \DeclareMathOperator{\inte}{int} \DeclareMathOperator{\spt}{supp}
138: \DeclareMathOperator{\conv}{co} \DeclareMathOperator{\dive}{div}
139: \DeclareMathOperator{\trace}{Tr} \DeclareMathOperator{\diag}{diag}
140: \DeclareMathOperator{\proj}{\Pi} \DeclareMathOperator{\curl}{curl}
141: 
142: 
143: \begin{document}
144: \title[Electrodynamics of hard superconductors]
145: {A variational approach to the macroscopic electrodynamics
146: of anisotropic hard superconductors}%
147: 
148: 
149: \author[G.~Crasta]{Graziano Crasta}
150: \address{Dipartimento di Matematica ``G.\ Castelnuovo'', Univ.\ di Roma I\\
151: P.le A.\ Moro 2 -- 00185 Roma (Italy)}
152: \email[Graziano Crasta]{crasta@mat.uniroma1.it}
153: 
154: \author[A.~Malusa]{Annalisa Malusa}
155: \email[Annalisa Malusa]{malusa@mat.uniroma1.it}
156: 
157: \date{December 20, 2006}
158: 
159: \keywords{Minimum problems with constraints, Euler equation, hard superconductors, Bean's model}
160: \subjclass[2000]{Primary 35C15; Secondary 49J30, 49J45, 49K20}
161: 
162: 
163: \begin{abstract}
164: We consider the Bean's critical state model for anisotropic superconductors.
165: A variational problem solved by the quasi--static evolution
166: of the internal magnetic field is obtained as the $\Gamma$-limit of functionals
167: arising from the Maxwell's equations combined with a power law for the
168: dissipation. Moreover, the quasi--static approximation of the internal
169: electric field is recovered, using a first order necessary condition.
170: 
171: If the sample is a long cylinder subjected to an axial uniform external
172: field, the macroscopic electrodynamics is explicitly determined.
173: %A class of minimum problems
174: %with a gradient constraint is considered. We determine the Euler equation solved by the optimal function
175: %coupled with its dual function.
176: %Moreover we find explicitly both the optimal function and its dual function.
177: %These results are applied to a variational model for the mixed
178: %state of hard superconductors in order to describe the macroscopic
179: %response of a sample to an applied external magnetic field.
180: \end{abstract}
181: 
182: \maketitle
183: 
184: 
185: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
186: % MSC 2000
187: %
188: %35–XX PARTIAL DIFFERENTIAL EQUATIONS
189: %35Axx General theory
190: %35A35 Theoretical approximation to solutions
191: %35Cxx Representations of solutions
192: %35C05 Solutions in closed form
193: %35C10 Series solutions, expansion theorems
194: %35C15 Integral representations of solutions of PDE
195: %35C20 Asymptotic expansions
196: %35C99 None of the above, but in this section
197: %35Jxx Partial differential equations of elliptic type
198: %35J05 Laplace equation, reduced wave equation
199: %           (Helmholtz), Poisson equation [see Also{ 31Axx, 31Bxx}
200: %35J20 Variational methods for second-order, elliptic equations
201: %35Qxx Equations of mathematical physics and other
202: %           areas of application [See also 35J05, 35J10,
203: %           35K05, 35L05]
204: %
205: %49–XX CALCULUS OF VARIATIONS AND OPTIMAL CONTROL; OPTIMIZATION
206: %49Jxx Existence theories
207: %49J10 Free problems in two or more independent variables
208: %49J30 Optimal solutions belonging to restricted classes
209: %           (Lipschitz controls, bang-bang controls, etc.)
210: %49Kxx Necessary conditions and sufficient conditions
211: %49K20 Problems involving partial differential equations
212: %49K24 Problems involving differential inclusions
213: %49K30 Optimal solutions belonging to restricted classes
214: %49Lxx Hamilton-Jacobi theories, including dynamic
215: %programming
216: %49L20 Dynamic programming method
217: %49L25 Viscosity solutions
218: %49L99 None of the above, but in this section
219: %49Qxx Manifolds
220: %49Q10 Optimization of the shape other than minimal surfaces
221: %
222: %52–XX CONVEX AND DISCRETE GEOMETRY
223: %52A++ General convexity
224: %52A20 Convex sets in $n$ dimensions (including
225: %              convex hypersurfaces) [see Also{ 53A07,  53C45}
226: %52A40 Inequalities and extremum problems
227: %
228: %53–XX DIFFERENTIAL GEOMETRY
229: %53Axx Classical differential geometry
230: %53A07 Higher-dimensional and -codimensional surfaces in Euclidean n-space
231: %
232: %74–XX MECHANICS OF DEFORMABLE SOLIDS
233: %74Bxx Elastic materials
234: %74B05 Classical linear elasticity
235: %74Kxx Thin bodies, structures
236: %74K10 Rods (beams, columns, shafts, arches, rings, etc.)
237: %
238: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
239: 
240: 
241: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
242: \section{Introduction}
243: 
244: %The aim of this paper is to determine the evolution of the electric and magnetic field inside
245: %a long cylindrical type--II  anisotropic superconductor
246: %placed into a nonstationary, uniform, and axial magnetic field $\vec{H}_s$.
247: 
248: It is well known that a superconductor is a conductor which is able to pass an electric current
249: without dissipation. The transition from the normally conducting state
250: to the superconducting one occurs at a critical temperature $T_c$, depending on the material,
251: below which the material exhibits (almost) perfect conductivity.
252: We are interested in the response of a superconducting material  to an applied external
253: magnetic field $\vec{H}_s$  under isothermal conditions below its critical temperature.
254: 
255: The superconductors can be classified in terms of a material parameter $\kappa>0$,
256: known as the Ginzburg--Landau parameter.
257: For the so called type--II superconductors (corresponding to $\kappa>1/\sqrt{2}$)
258: there exist two critical magnetic field intensities,
259: ${H}_{c_1} < {H}_{c_2}$, such
260: that for $|\vec{H}_s| < {H}_{c_1}$ the material is in the superconducting state and the
261: magnetic field is excluded from the bulk of the sample except in thin boundary layers,
262: while for $|\vec{H}_s| > {H}_{c_2}$ the material
263: behaves as a normal conductor,
264: %is in the normal conducting
265: and the magnetic field penetrates it fully.
266: For ${H}_{c_1}<|\vec{H}_s|< {H}_{c_2}$ a third
267: state exists, known as ``mixed state'' as well as ``vortex state''.
268: The mixed state is characterized by a partial penetration
269: of the magnetic field into the sample, which occurs, at a mesoscopic level, by
270: means of thin filaments of normally conducting material carrying magnetic flux
271: and circled by a vortex of superconducting current.
272: Type--I superconductors are those with ${H}_{c_1}={H}_{c_2}$, so that
273: the mixed state does not occur.
274: 
275: Although from a physical point of view the most interesting description of the mixed
276: state for type--II superconductors is given by the mesoscopic Ginzburg--Landau model,
277: in designing magnets and other large--scale applications of superconducting materials,
278: engineers use macroscopic models, involving averaged variables. One of the
279: most reliable macroscopic models is the Bean's critical state model (see \cite{Bean}).
280: We refer to \cite{Cha} for a derivation of this model
281: as a macroscopic version of the Ginzburg--Landau model
282: under suitable assumptions.
283: 
284: The basic idea of the Bean's phenomenological model for isotropic materials is that,
285: because of physical limitations imposed by the material properties, the current
286: density $|\vec{J}|$ cannot
287: exceed a critical value $J_c$ without destroying the superconducting phase.
288: 
289: Moreover it is assumed that any electromotive
290: force due to external field variations induces the maximum current density flow,
291: according to the most effective way of shielding field variations. Hence
292: the current density $|\vec{J}|$ is forced to be $J_c$ in the part of the sample
293: where the field is penetrated, while $\vec{J}=0$ in the remaining part of the sample.
294: 
295: At a mesoscopic level, the critical current density $J_c$ corresponds to the
296: balance between a repulsive vortex--vortex interaction and attractive forces
297: towards the pinning centers. At a macroscopic level, the electric field is zero
298: when $|\vec{J}|<J_c$ and abruptly rises to arbitrarily large values if $J_c$ is
299: overrun.
300: 
301: Since for isotropic materials it is well--established that the electric field and the
302: current have the same direction (at least for slowly varying external fields),
303: in the original Bean's model
304: the Ohm's law is replaced by a vertical current--voltage law
305: \[
306: \vec{E} \parallel \vec{J}\,, \quad
307: |\vec{J}|\leq J_c\,,\qquad
308: \vec{E}= 0 \ \textrm{if}\ |\vec{J}| < J_c\,,
309: \]
310: %\begin{cases}
311: %0 & \textrm{if\ }|\vec{J}|<|\vec{J}_c|\,, \\
312: %[0, +\infty) & \textrm{if\ }|\vec{J}|=|\vec{J}_c|\,,
313: %\end{cases}
314: %\]
315: which can be interpreted as the limit
316: for $p\to +\infty$ in the power--law
317: \begin{equation}\label{f:power}
318: \vec{E} \parallel \vec{J}\,, \quad |\vec{E}|=e_c \left(\frac{|\vec{J}|}{|\vec{J}_c|}\right)^p
319: \end{equation}
320: (see, e.g., \cite{Bran,BaPr,BLc}).
321: Notice that the direction of the electric field is obtained by exploiting the isotropy of the
322: material and cannot be obtained as a consequence of the power law approximation, which involves
323: only the intensities of the fields.
324: Hence this approach is not appropriate to deal with anisotropic materials
325: (see anyhow \cite{BLb} for some
326: explicit computation of the electric field in the case of infinite slab geometry,
327: and \cite{BKR} in the case of a cylindrical body with elliptic section).
328: On the other hand, many type--II superconducting materials are anisotropic. Moreover,
329: even if the material is isotropic, the presence of flux-flow Hall effects has to be described
330: in terms of an anisotropic resistivity (see, e.g., \cite{Bran}).
331: 
332: Concerning the constraints on the current, in the anisotropic case the Bean's law dictates that
333: there exists a compact set $\Delta$ containing the origin as an interior point, such that
334: $\vec{J}$ cannot lie outside $\Delta$ without destroying the superconducting state.
335: %In what follows we shall always assume that $\Delta$ is a convex compact set.
336: Moreover $\vec{J}\in \partial \Delta$ in the penetrated region, and $\vec{J}=0$ elsewhere.
337: 
338: Section \ref{s:phis} of this paper is devoted to the description of the Bean's law as
339: a limit of a power--like law fulfilled by the dissipation. This approach
340: allows us to determine \textit{a priori} the direction of the electric field in terms of
341: the direction of the current for anisotropic materials.
342: 
343: In Section \ref{s:bean} we deal with a variational statement of the critical state
344: proposed in \cite{BLd}, which takes
345: the form of a quasistatic evolution of the penetrated magnetic field, obtained combining
346: the finite--difference expression of Faraday's law with the Bean's law.
347: We shall give a mathematical justification of the variational model
348: as a limiting case of the power law model for dissipation.
349: The result is proposed in terms of
350: $\Gamma$--convergence of functionals, which is nowadays a classical tool in the mathematical methods for
351: the material science (see, e.g., \cite{Brai,DM} and the references therein).
352: 
353: In the last part of the paper we focus our attention to the special case of a long
354: cylindrical anisotropic superconductor placed into a nonstationary, uniform axial
355: magnetic field. The parallel geometry enables us to make a two dimensional reduction
356: of the problem,
357: which can then be explicitly solved.
358: %and we can compute the electrodynamics explicitly.
359: 
360: The plan of this part of the paper is the following. In Section \ref{s:bean2} we describe the
361: two dimensional reduction of the problem in the case of parallel geometry.
362: In Section \ref{s:distance} we introduce some
363: technical tool needed for the analytical description of the fields.
364: In Section \ref{s:minp} we find explicitly the solution to a general class of minimum problems
365: with a gradient constraint and we determine the Euler equation solved by the optimal function
366: coupled with its dual function. Moreover we find the explicit form of the dual function.
367: In Section \ref{s:evol} the previous results are applied to the variational model for the critical
368: state, and we find the quasistatic evolution of both the magnetic field and the dissipation
369: inside the superconductor. Finally, a passage to the limit on the time layer gives
370: the explicit form of the macroscopic electrodynamics.
371: 
372: For what concerns the magnetic field, our result
373: generalizes the one, valid for isotropic materials, obtained by Barrett and Prigozhin
374: in \cite{BaPr} with a different method
375: based on a evolutionary variational inequality.
376: Due to the fact that the magnetic field is explicitly known,
377: we can compute the full penetration time in the case of monotonic external fields
378: as well as we can depict the well known hysteresis phenomenon.
379: On the other hand, the knowledge of the electric field allows us to give a detailed
380: description of the evolution of the dissipation
381: for cylinders of anisotropic materials with a general geometry of the cross section
382: (see Figures \ref{BB3} and \ref{BB4}).
383: 
384: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
385: \section{Notation and preliminaries}
386: 
387: For $\xi\in \R^N$, $|\xi|$ will be the Euclidean norm, and
388: $\pscal{\xi}{\xi'}$ will denote the scalar product with $\xi'\in \R^N$.
389: The symbol $\times$ will be used for the cross product of vectors
390: in $\R^3$.
391: Given $a,b\in\R$, $a\vee b$ and $a\wedge b$ will denote
392: respectively the maximum and the minimum of $a$ and $b$.
393: 
394: \smallskip
395: 
396: Given $A\subset\R^N$, we shall denote by $\Lip(A)$, $C(A)$,
397: $\Cb(A)$ and $C^k(A)$, $k\in\N$ the set of functions $u\colon
398: A\to\R$ that are respectively Lipschitz continuous, continuous,
399: bounded and continuous, and $k$-times continuously differentiable
400: in $A$. Moreover, $C^{\infty}(A)$ will denote the set of functions
401: of class $C^k(A)$ for every $k\in\N$, while $C^{k,\alpha}(A)$ will
402: be the set of functions of class $C^k(A)$ with H\"older continuous
403: $k$-th partial derivatives with exponent $\alpha\in [0,1]$.
404: Finally $L^p(A)$, $W^{1,p}(A)$, $W^{1,p}_0(A)$, and $L^p(A,\R^d)$,
405: $W^{1,p}(A,\R^d)$, $W^{1,p}_0(A,\R^d)$, $d>1$, will be the usual Lebesgue
406: and Sobolev spaces of scalar or vectorial functions respectively.
407: 
408: \smallskip
409: 
410: Let $X$ be subset of $\R^N$, $N\geq 2$. We shall denote by
411: $\partial X$ its boundary, by $\inte X$ its interior, and by $\overline{X}$ its closure.
412: The characteristic function of $X$
413: will be denoted by $\chi_X$.
414: The set $X$ is said to be
415: of class $C^k$, $k\in\N$,
416: if for every point $x_0\in\partial X$
417: there exists a ball $B=B_r(x_0)$ and a one-to-one
418: mapping $\psi\colon B\to D$ such that
419: $\psi\in C^k(B)$, $\psi^{-1}\in C^k(D)$,
420: $\psi(B\cap X)\subseteq\{x\in\R^n;\ x_n > 0\}$,
421: $\psi(B\cap\partial X)\subseteq\{x\in\R^n;\ x_n = 0\}$.
422: If the maps $\psi$ and $\psi^{-1}$ are of class
423: $C^{\infty}$ or $C^{k,\alpha}$ ($k\in\N$, $\alpha\in [0,1]$),
424: then $X$ is said to be of class
425: $C^{\infty}$ or $C^{k,\alpha}$ respectively.
426: 
427: \smallskip
428: 
429: Let $D\subset\R^N$ be a compact convex set containing
430: $0$ as an interior point and with boundary of class $C^2$.
431: The gauge function $\gau{D}$ of the set $D$ is the convex, positively
432: 1--homogeneous function defined by
433: \[
434: \gau{D}(\xi) = \inf\{ t\geq 0;\ \xi\in t D\}\,,
435: \qquad\xi\in\R^N.
436: \]
437: Since $D$ is a compact set containing a neighborhood of
438: $0$, there exist two positive constants $c_1<c_2$ such that
439: \begin{equation}\label{robdd}
440: c_1|\xi| \leq \gau{D}(\xi)\leq c_2 |\xi|\,, \qquad \forall\ \xi \in \R^N\,.
441: \end{equation}
442: 
443: The indicator function of the set $D$ is defined
444: by
445: \begin{equation}\label{indk}
446: I_{D}(\xi)=
447: \begin{cases}
448: 0 & \textrm{if}\ \xi\in D \\
449: +\infty & \textrm{if}\ \xi\not\in D\,.
450: \end{cases}
451: \end{equation}
452: 
453: Since $D$ has a smooth boundary, the subgradient of the indicator function
454: can be explicitly computed, and
455: \begin{equation}\label{subindk}
456: \partial I_{D}(\xi)=
457: \begin{cases}
458: \{\alpha D\gau{D}(\xi)\colon \alpha \geq 0\} & \textrm{if}\ \xi\in \partial D \\
459: \emptyset & \textrm{if}\ \xi\not\in D \\
460: \{0\} & \textrm{if}\ \xi\in \inte D
461: \end{cases}
462: \end{equation}
463: (see e.g.\ \cite{Rock}, Section 23).
464: 
465: In what follows, a family of objects, even if parameterized by a continuous
466: parameter, will also be often called a ``sequence'' not to overburden
467: notation.
468: 
469: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
470: \section{The physical setting}\label{s:phis}
471: 
472: Since we shall deal with a macroscopic model,
473: the coarse--grained electrodynamics will be formulated
474: in terms of
475: \begin{itemize}
476: \item[(i)] the flux density $\vec{B}$ within the sample, which is the average of the
477: microscopic field intensity;
478: \item[(ii)] the magnetic field $\vec{H}$, which is assumed to be linearly connected
479: to $\vec{B}$ by $\vec{B}=\mu_0 \vec{H}$;
480: %(this requirement is always
481: %satisfied if $\vec{H}_s$ is well apart
482: %from the lower and upper critical fields);
483: \item[(iii)] the averaged current density $\vec{J}$, linked to  $\vec{H}$ by the
484: Amp\`ere's law $\vec{J}=\curl \vec{H}$.
485: \end{itemize}
486: Moreover, on neglecting finite size effects, we can assume that
487: \begin{itemize}
488: \item[(iv)] the magnetic source $\vec{H}_s$ enters
489: as a boundary condition for the flux density at the surface of the sample,
490: requiring that $\nt\times\vec{H} = \nt\times\vec{H}_s$ on that surface,
491: where $\nt$ is the outward normal vector.
492: %requiring that the magnetic field
493: %$\vec{H}$ is continuous.
494: \end{itemize}
495: %Moreover, the connection to observable quantities is done by characterizing the external field
496: %sources and the sample response.
497: %\begin{itemize}
498: %\item[(iv)] On neglecting finite size effects, the magnetic source $\vec{H}_s$ enters
499: %as a boundary condition for the flux density at the surface of the sample,
500: %requiring that the magnetic field
501: %$\vec{H}$ is continuous.
502: %\item[(v)] Then the measured magnetic moment of the sample per unit volume
503: %is $\vec{M}=\langle\vec{H}\rangle-\vec{H}_s$. Here the average concerns the whole volume of the
504: %superconductor.
505: %\end{itemize}
506: 
507: Time variations of the external field $\vec{H}_s$ induce in the sample an electric field
508: $\vec{E}$, according to Faraday's
509: law.
510: %\[
511: %\curl\vec{E}+\mu_0\frac{\partial \vec{H}}{\partial t}=0\,.
512: %\]
513: On the other hand,
514: the electric field leads
515: a current $\vec{J}$ which
516: induces an internal magnetic field $\vec{H}$,
517: according to Amp\`ere's law.
518: Finally, we recall that $\dive \vec{H}=0$ (Gauss' law).
519: 
520: Summarizing, we are considering the so-called eddy current model
521: for Maxwell equations:
522: \[
523: \begin{cases}
524: \curl\vec{E}+\mu_0\dfrac{\partial \vec{H}}{\partial t}=0
525: &\textrm{(Faraday's law)},\\
526: \vec{J}=\curl \vec{H}
527: &\textrm{(Amp\`ere's law)},\\
528: \dive\vec{H} = 0
529: &\textrm{(Gauss' law)},\\
530: \nt\times(\vec{H} -\vec{H}_s) = 0
531: &\textrm{on the surface}.
532: \end{cases}
533: \]
534: 
535: In order to complete the physical setting,
536: it remains to find an appropriate version of the constitutive law
537: $\vec{E}(\vec{J})$. Our starting point is a reading of the macroscopic behaviour
538: of the material in terms of the dissipation
539: $\dot{S}=\pscal{\vec{E}(\vec{J})}{\vec{J}}$
540: (here $S$ denotes the entropy of the system).
541: Namely, the Bean's model dictates that if $\vec{J}$ is in the interior of
542: an allowed region $\Delta$, which is assumed to be a convex compact set of $\R^3$
543: having the origin as an interior point,
544: then $\dot{S}$ vanishes as no electric field is generated in stationary
545: condition, while when $\vec{J}$ touches
546: the boundary $\partial \Delta$ of the allowed region a huge dissipation occurs,
547: destroying the superconducting phase.
548: The fact that $\dot{S}=0$ if $\vec{J}$ is an interior point of $\Delta$,
549: while $\dot{S} = +\infty$ if $\vec{J}\not\in\Delta$
550: suggests that the Bean's law can be
551: interpreted as the limit as $p\to\infty$ of a power law for the dissipation
552: $\pscal{\vec{E}_p(\vec{J})}{\vec{J}}$, that is
553: \begin{equation}\label{f:pdissip}
554: \pscal{\vec{E}_p(\vec{J})}{\vec{J}}=\frac{c}{p}\left(\gauge_\Delta(\vec{J})\right)^{p}\,.
555: \end{equation}
556: This approximation allows us to recover the direction of
557: the electric field $\vec{E}(\vec{J})$ when
558: $\vec{J}\in\partial\Delta$.
559: Namely, we claim that, under the physically consistent assumption that
560: \begin{equation}\label{f:nulel}
561: \lim_{t\to 0^+}\vec{E}_p(t\vec{J})=0\,, \quad \forall\ \vec{J}\in\R^3\,,
562: \end{equation}
563: the relation (\ref{f:pdissip}) implies, for $p>1$,
564: \begin{equation}\label{f:pelet}
565: \vec{E}_p(\vec{J})=\frac{c}{p}\left(\gauge_\Delta(\vec{J})\right)^{p-1} D \gauge_\Delta(\vec{J})\,.
566: \end{equation}
567: In order to prove (\ref{f:pelet}), we
568: differentiate (\ref{f:pdissip}), obtaining
569: \[
570: \pscal{D\vec{E}_p(\vec{J})}{\vec{J}}+ \vec{E}_p(\vec{J})=
571: c\left(\gauge_\Delta(\vec{J})\right)^{p-1} D \gauge_\Delta(\vec{J})\,.
572: \]
573: Hence, fixed $\vec{J}\neq 0$, the function $v(t)= \vec{E}_p(t\vec{J})$, $t>0$,
574: is a solution of the O.D.E.
575: \begin{equation*}
576: \begin{cases}
577: t\, v'(t)+v(t)=c\, t^{p-1}w \,, \\
578: v(1)= \vec{E}_p(\vec{J})\,,
579: \end{cases}
580: \end{equation*}
581: where $w=\left(\gauge_\Delta(\vec{J})\right)^{p-1} D \gauge_\Delta(\vec{J})$. Then
582: \[
583: v(t)=\frac{1}{t}\left[\vec{E}_p(\vec{J})+\frac{cw}{p}(t^p-1)\right]\,,
584: \]
585: and, by (\ref{f:nulel}),
586: \[
587: \vec{E}_p(\vec{J})-\frac{c}{p}\left(\gauge_\Delta(\vec{J})\right)^{p-1} D \gauge_\Delta(\vec{J}) =
588: \lim_{t\to 0^+}[\vec{E}_p(\vec{J})+\frac{cw}{p}(t^p-1)]=0\,.
589: \]
590: %Notice that formula (\ref{f:pelet}) contains both the current--voltage power law and the
591: %information that the electric field has the same direction of $D \gauge_\Delta(\vec{J})$.
592: 
593: As a consequence of (\ref{f:pelet}), the direction of $\vec{E}_p(\vec{J})$ is given
594: by $D \gauge_\Delta(\vec{J})$, and it does not depend on $p$.
595: In conclusion, recalling (\ref{subindk}),
596: the anisotropic version of the Bean's law is the following.
597: \begin{itemize}
598: \item[(B1)] There exists a convex
599: compact set $\Delta$ containing the origin as an interior point and
600: such that $\vec{J}(x,t)\in \Delta$ for every $x\in\Omega$, $t\geq 0$.
601: \item[(B2)] If we denote by $I_\Delta$ the indicator function
602: of the set $\Delta$, the constitutive law $\vec{E}(\vec{J})$ is given by
603: $\vec{E}\in \partial I_\Delta(\vec{J})$.
604: \end{itemize}
605: 
606: %\begin{rk}
607: %{}From (\ref{f:pelet}) we see that
608: %$\vec{E}(\vec{J}) = D\varphi_p(\vec{J})$,
609: %where $\varphi_p(\vec{J}) = c\, \gauge_\Delta(\vec{J})^p/p^2$
610: %converges to the indicator function of the set $\Delta$
611: %as $p\to +\infty$.
612: %Hence it is reasonable to assume that
613: %$\vec{E}$ has to be an element of the subdifferential $\partial I_\Delta(\vec{J})$
614: %(this assertion will be proved by the variational analysis of
615: %Section~\ref{s:bean-forse}).
616: %We remark that, for $\vec{J}\in\partial\Delta$,
617: %the pointwise limit of the right-hand side of (\ref{f:pelet}) as $p\to +\infty$
618: %is zero.
619: %\end{rk}
620: 
621: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
622: \section{A variational model for the mixed state}\label{s:bean}
623: 
624: Let $A\subset\R^3$ be the region occupied by the superconductor, and
625: $\partial A$ its surface.
626: We assume that $A$ is a bounded, open subset of $\R^3$, and that
627: $\partial A$ is of class $C^{1,1}$.
628: Moreover, we assume that $A$ has no enclosed cavity.
629: In our model the presence of cavities is not allowed,
630: since the boundary condition in such a cavity cannot be given in terms of
631: the external magnetic field.
632: (The magnetic field must be constant on the boundary of every enclosed cavity.)
633: In order to avoid the presence of cavities, we assume that $A$ has second Betti number $0$
634: (see e.g.\ \cite{Gold} for the definition of Betti numbers).
635: %We recall that, loosely speaking, the second Betti number of $A$
636: %is the number of enclosed cavities.
637: %In our model the presence of cavities is not allowed,
638: %since the boundary condition in such a cavity cannot be given in terms of
639: %the external magnetic field.
640: %(The magnetic field must be constant on the boundary of every enclosed cavity.)
641: 
642: 
643: Let $T>0$ be fixed. For every $n\in\N^+$, we take a partition
644: $P^n=\left\{t^n_i\right\}_{i=0}^{k(n)}$
645: %$0=t^n_0<t^n_1<\cdots<t^n_{n}=T$
646: of the interval $[0,T]$, and we set $\delta t^n_i=t^n_i-t^n_{i-1}$.
647: %Finally, let us set $\vcn_i=\vecH_s(t^n_i)$, and
648: %$h^n_i(x)$ the magnetic field intensity at time layer $i$, given by $H_s(t^n_i)+u(t^n_i,x)$.
649: The least action principle proposed in \cite{BLd} is the following: starting from an initial field profile
650: $\vec{H}^n_i(x)$ in $A$ at the time layer $i$,
651: and under a small change of the external drive, the new profile $\vec{H}^n_{i+1}(x)$
652: at the time layer $i+1$
653: is the unique solution to the minimum problem
654: \begin{equation}\label{varmod3d}
655: \min\left\{
656: \int_A  |\vec{V}-\vec{H}^n_i|^2\, dx;\
657: \vec{V}\in \vec{H}_s(t^n_{i+1})+X_2,\
658: \curl\vec{V}\in \Delta\ \textrm{a.e.}
659: \right\}\,,
660: \end{equation}
661: where, for every $p>1$,
662: \begin{equation*}%\label{f:Xp}
663: X_p = \left\{\vec{V}\in \elle{p};\
664: \curl\vec{V}\in \elle{p},\,
665: \dive\vec{V} = 0,\,
666: \nt\times\vec{V} = 0\, \textrm{on}\, \partial A
667: \right\}.
668: \end{equation*}
669: 
670: In \cite{BLd}
671: the trustworthiness of this variational principle was motivated
672: in terms of nonequilibrium thermodynamical principles
673: in analogy with the ohmic case.
674: In this section we shall give a mathematical justification, in terms of
675: $\Gamma$--convergence of functionals, of the variational model
676: as a limiting case of the power law model for dissipation.
677: %In point of fact, we are going to use the theory of $\Gamma$--convergence.
678: %However, since the model considered is very simple, we prefer to write the result in term
679: %of convergence of minima instead of $\Gamma$--convergence of functionals.
680: 
681: Our starting point is the fact that, assuming that the power law (\ref{f:pdissip})
682: holds true, the discretized version of the Faraday's law
683: \begin{equation}\label{f:farlp}
684: \curl\vec{E}^n_{i+1}+\frac{\mu_0}{c\dt}(\vec{H}^n_{i+1}-\vec{H}^n_{i})=0
685: \end{equation}
686: is the Euler equation for the functional
687: \[
688: F_p(\vec{V})=\int_A \frac{1}{p}\gd(\curl\vec{V})^p+
689: \lambda\left|\vec{V}-\vec{H}^n_i\right|^2\, dx\,,\qquad
690: \vec{V}\in \vec{H}_s(t^n_{i+1})+X_p
691: \]
692: where $\lambda := \frac{\mu_0}{2c\dt}>0$.
693: 
694: In other words, it can be easily checked that the magnetic field $\vec{H}^n_{i+1}$
695: is the unique minimum point of $F_p$ in $\vec{H}_s(t^n_{i+1})+ X_p$.
696: It is clear that, up to a translation of a constant vector,
697: it is equivalent to minimize $F_p$ on $X_p$ by
698: changing $\vec{H}^n_i$ with $\vec{H}^n_i-\vec{H}_s(t^n_{i+1})$.
699: 
700: The following properties of the spaces $X_p$ will be useful in the sequel
701: (see \cite[Thm.~2.2]{YLZ} and \cite[Thm.~3.1]{Wa}).
702: 
703: \begin{theorem}\label{t:Xp}
704: Let $A\subset\R^3$ be a bounded open set of class $C^{1,1}$.
705: Let $1<p<\infty$. Then the following hold.
706: \begin{itemize}
707: \item[(i)]
708: Every function $\vec{V}\in X_p$ belongs to $\wp{p}$.
709: Moreover, if $p>3$, then $\vec{V}\in C^{0,\alpha}(\overline{A})$
710: and there exists a constant $C_p>0$, depending only on $p$ and $A$,
711: such that
712: \[
713: {\left\|\vec{V}\right\|}_{C^{0,\alpha}(\overline{A})}\leq
714: C_p \normp{\curl\vec{V}}\,.
715: \]
716: 
717: \item[(ii)]
718: If $A$ has second Betti number $0$, then there exists a constant
719: $M_p > 0$, depending only on $p$ and $A$,
720: such that
721: \[
722: \normp{D\vec{V}} \leq M_p \normp{\curl\vec{V}}
723: \]
724: for every $\vec{V}\in X_p$.
725: \end{itemize}
726: \end{theorem}
727: 
728: \begin{rk}\label{r:Xp}
729: {}From Theorem~\ref{t:Xp} we infer that $X_p$ is a (proper)
730: subspace of $\wp{p}$.
731: Moreover, for $p>3$ the map defined by
732: $X_p\ni\vec{V}\mapsto\normp{\curl\vec{V}}$ is a norm in $X_p$ equivalent
733: to the standard $W^{1,p}$ norm.
734: \end{rk}
735: 
736: From now on we shall always assume $p>3$. In this case,
737: by Theorem \ref{t:Xp}(i), a function $\vec{V}\in X_p$ is continuous
738: in $\overline{A}$,
739: and the boundary condition $\nt\times \vec{V}$ is understood to
740: be pointwise fulfilled. On the other hand, the divergence--free requirement
741: on $\vec{V}\in X_p$ is understood in the sense of distributions, i.e.
742: \[
743: \int_A \vec{V}\cdot D\vec{\Psi} =0\,, \qquad \forall \vec{\Psi}\in C^\infty_0(A, \R^3)\,.
744: \]
745: 
746: For our subsequent considerations, we need to define all the functionals $F_p$
747: on the same Banach space.
748: For this reason,
749: naming $\vec{U}_0=\vec{H}^n_i-\vec{H}_s(t^n_{i+1})$,
750: we set
751: \begin{equation}\label{f:Fp}
752: F_p(\vec{V})=
753: \begin{cases}
754: \displaystyle
755: \int_A \frac{1}{p}\gd(\curl\vec{V})^p+
756: \lambda\left|\vec{V}-\vec{U}_0\right|^2\, dx\,,
757: &\textrm{if $\vec{V}\in X_p$},\\
758: +\infty,
759: &\textrm{otherwise in $\elle{2}$}.
760: \end{cases}
761: \end{equation}
762: We also define the functional
763: \begin{equation}\label{f:F}
764: F(\vec{V})=
765: \begin{cases}
766: \displaystyle\int_A
767: I_{\Delta}(\curl\vec{V})+
768: |\vec{V}-\Uz|^2\, dx\,,
769: &\textrm{if $\vec{V}\in X_2$},\\
770: +\infty,
771: &\textrm{otherwise in $\elle{2}$},
772: \end{cases}
773: \end{equation}
774: where $I_{\Delta}$ is the indicator function of the convex set $\Delta$, defined in (\ref{indk}).
775: 
776: \begin{lemma}\label{l:basicp}
777: The functionals $F_p$, $p>3$, and $F$ are lower semicontinuous
778: in the strong $L^2$ topology.
779: Moreover, the functionals $F_p$ are equicoercive, i.e.\
780: there exists a constant $\alpha>0$ such that
781: \[
782: F_p(\vec{V})\geq \frac{\lambda}{2} \|\vec{V}\|_2^2-\alpha
783: \qquad
784: \forall \vec{V}\in \elle{2},\ \forall p>3\,.
785: \]
786: \end{lemma}
787: 
788: \begin{proof}
789: Let $3< p<\infty$, and let $(\vec{V}_k)\subset \elle{2}$
790: be a sequence converging to $\vec{V}$ in $L^2$. We have to prove that
791: \[
792: F_p(\vec{V})\leq \liminf_{k \to \infty}F_p(\vec{V}_k)\,.
793: \]
794: Without loss of generality we can assume that
795: $\vec{V}_k\in X_p$ for every $k$ and
796: \begin{equation}\label{f:limfin}
797: \lim_{k\to\infty} F_p(\vec{V}_k) = C < +\infty\,.
798: \end{equation}
799: 
800: We claim that $\vec{V}\in X_p$.
801: Namely, the divergence-free requirement is stable under
802: strong $L^2$ convergence.
803: Moreover, from (\ref{f:limfin}) and (\ref{robdd}) we have that
804: there exists a constant $C_1>0$ such that
805: $\|\curl\vec{V}_k\|_p \leq C_1$ for every $k\in\N$.
806: {}From Theorem~\ref{t:Xp}(i) we infer that the sequence
807: $(\vec{V}_k)$ is equicontinuous and equibounded in $\overline{A}$.
808: Hence we can pass to a subsequence, that we do not relabel,
809: converging uniformly to $\vec{V}$ in $\overline{A}$,
810: so that the boundary condition $\nt\times\vec{V} = 0$ on $\partial A$
811: holds.
812: Finally, from Remark~\ref{r:Xp},
813: $(\vec{V}_k)$ converges to $\vec{V}$ weakly in $\wp{p}$, and
814: $\|\curl\vec{V}\|_p \leq \liminf_k \|\curl\vec{V}_k\|_p$,
815: proving the claim.
816: 
817: By the convexity of $\gd$ and the convergence of
818: $\curl\vec{V}_k$ to $\curl\vec{V}$ in the weak $L^p$ topology,
819: we obtain that
820: \[
821: \int_A\gd(\curl\vec{V})^p \leq \liminf_{k\to \infty} \int_A\gd(\curl\vec{V}_k)^p,
822: \]
823: which implies  the semicontinuity inequality
824: $F_p(\vec{V})\leq C$.
825: 
826: The semicontinuity of $F$ can be proved in a similar way,
827: while the equicoerciveness of $(F_p)$ easily follows
828: choosing $\alpha = \frac{\lambda}{2}\|\vec{U}_0\|_2^2$.
829: \end{proof}
830: 
831: \begin{rk}\label{r:minFp}
832: For every $p>3$, the functional $F_p$ is lower semicontinuous in $\elle{2}$,
833: coercive,
834: and strictly convex in its effective domain $X_p$.
835: Hence it admits a unique minimizer $\vec{U}_p$,
836: which belongs to $X_p$.
837: Similarly, the functional $F$ admits a unique minimizer $\vec{U}\in X_2$
838: with $\gd(\curl\vec{U})\leq 1$.
839: \end{rk}
840: 
841: The mathematical justification of the Bad\'\i a and L\'opez model is
842: based on the fact that
843: the sequence $(\vec{U}_p)$ of
844: minimizers of $(F_p)$ converges to the minimizer $\vec{U}$ of $F$.
845: As is customary in material science,
846: we are going to prove this statement
847: using the technique of $\Gamma$-convergence (see \cite{Brai} and the
848: reference therein for examples of applications of the theory to
849: different models).
850: 
851: \begin{dhef}\label{d:gamma}
852: The sequence $(F_p)$ $\Gamma$-converges to the functional
853: $F$ (with respect to the strong $L^2$ topology) if the following
854: two conditions are satisfied.
855: \begin{itemize}
856: \item[(i)]
857: For every $\vec{V}\in \elle{2}$ there exists a sequence $(\vec{V}_p)$
858: (called a \textsl{recovering sequence}), converging to $\vec{V}$
859: in $\elle{2}$, and such that
860: \[
861: \limsup_{p\to\infty} F_p(\vec{V}_p) \leq F(\vec{V})\,.
862: \]
863: \item[(ii)]
864: For every sequence $(\vec{V}_p)\subset\elle{2}$ converging strongly to
865: $\vec{V}$, we have
866: \[
867: \liminf_{p\to\infty} F_p(\vec{V}_p) \geq F(\vec{V})\,.
868: \]
869: \end{itemize}
870: \end{dhef}
871: 
872: \begin{rk}\label{r:gc}
873: Since the functionals $F_p$ are equicoercive (see Lemma \ref{l:basicp}),
874: the convergence of the minimizers $(\vec{U}_p)$ to $\vec{U}$ will follow
875: if we prove that
876: $(F_p)$ $\Gamma$-converges to $F$ and
877: that $(\vec{U}_p)$ is a precompact sequence in $\elle{2}$
878: (see e.g.\ \cite{Brai}).
879: \end{rk}
880: 
881: The main tool needed in the proof of the $\Gamma$--convergence
882: of $(F_p)$ to $F$
883: is the following lemma.
884: 
885: \begin{lemma}\label{l:rholim}
886: Let $(\vec{V}_p)\in \elle{2}$ be a sequence converging to
887: $\vec{V}$ in $\elle{2}$. If
888: \[
889: \liminf_{p\to\infty} F_p(\vec{V}_p) < +\infty
890: \]
891: then
892: $\gd(\curl\vec{V})\leq 1$ a.e.\ in $A$.
893: \end{lemma}
894: 
895: \begin{proof}
896: Without loss of generality we can assume that
897: \[
898: \begin{split}
899: \liminf_{p\to\infty} F_p(\vec{V}_p) =
900: \lim_{p\to\infty} F_p(\vec{V}_p) = L < +\infty,\\
901: \curl\vec{V}_p\in \elle{p}\,,\quad F_p(\vec{V}_p)\leq L_1\qquad \forall p>3.
902: \end{split}
903: \]
904: %hence $\curl\vec{V}_p\in \elle{p}$
905: %and $F_p(\vec{V}_p)\leq L_1$ for every $p>3$.
906: 
907: We are going to prove that the sequence $(\vec{V}_p)$
908: is bounded in $\wp{q}$ for every $q>3$.
909: In view of Theorem~\ref{t:Xp}, it is enough show that
910: for every given $q>3$
911: there exists a constant $C>0$ such that
912: ${\|\curl\vec{V}_p\|}_q \leq C$ for every $p>q$.
913: 
914: Given $3<q<p$, by (\ref{robdd}) and H\"older's inequality we have
915: \begin{equation}\label{f:gac1}
916: \begin{split}
917: \int_A |\curl\vec{V}_p|^q\, dx & \leq
918: \frac{1}{c_1^q}\int_{A} \gd(\curl\vec{V}_p)^q\, dx
919: \\ & \leq
920: \frac{1}{c_1^q}\left(\frac{p}{|A|}\right)^{q/p}|A|
921: \left(\int_A \frac{1}{p}\,\gd(\curl\vec{V}_p)^p\, dx\right)^{q/p}\,.
922: \end{split}
923: \end{equation}
924: On the other hand we get
925: \[
926: \int_A \frac{1}{p}\,\gd(\curl\vec{V}_p)^p\, dx \leq F_p(\vec{V}_p)
927: \leq L_1\,.
928: %\leq F_p(\Uz-\cz)
929: %\leq (1+\lambda\, |\cz|^2) |A|\,.
930: \]
931: Collecting all the previous estimates,
932: for every $p > q$ we obtain
933: \begin{equation}\label{f:gacn}
934: \int_A |\curl\vec{V}_p|^q\, dx \leq
935: \frac{ |A|}{c_1^q}\,
936: \left(\frac{p\, L_1}{|A|}\right)^{q/p}
937: \leq
938: \frac{|A|}{c_1^q}\,
939: \exp\left(\frac{q L_1}{e |A|}\right)\,.
940: \end{equation}
941: %where $C>0$ does not depend on $q$ and $p$.
942: Then the sequence $(\vec{V}_p)$ is bounded in $\wp{q}$,
943: and hence $(\vec{V}_p)$ converges to $\vec{V}$ in the weak topology of
944: $\wp{q}$.
945: %We remark that, since $\dive\vec{V}_p=0$ for every $p$,
946: %by weak convergence we get $\dive\Usol = 0$.
947: 
948: This fact and
949: the convexity of the function $\gd$ imply that
950: \begin{equation}\label{f:gacnpu}
951: \int_B \gd(\curl\vec{V})\, dx \leq \liminf_{p\to \infty}
952: \int_B \gd(\curl\vec{V}_p)\, dx
953: \end{equation}
954: for every open set $B\subseteq A$
955: (see e.g.\ \cite[Thm.~4.2.1]{Butt}).
956: On the other hand, by H\"older's inequality
957: \[
958: \begin{split}
959: \int_B \gd(\curl\vec{V}_p)\, dx & \leq
960: \left(\int_B \gd(\curl\vec{V}_p)^p\, dx\right)^{1/p}|B|^{(p-1)/p} \\
961: & \leq p^{1/p}\left(\frac{1}{p}\int_B \gd(\curl\vec{V}_p)^p dx\right)^{1/p}|B|^{(p-1)/p}\,.
962: \end{split}
963: \]
964: Then, using (\ref{f:gacn}) and (\ref{f:gacnpu}) we get
965: $\displaystyle\int_B \gd(\curl\vec{V})\, dx \leq |B|$.
966: Since $\gd(\curl\vec{V})\in L^1(A)$, we conclude that
967: \[
968: \gd(\curl\vec{V}(x))=\lim_{r\to 0^+} \frac{1}{|B_r(x)|}
969: \int_{B_r(x)} \gd(\curl\vec{V})\, dx \leq 1\,,
970: \]
971: for a.e.\ $x\in A$.
972: \end{proof}
973: 
974: %Now we are able to prove the $\Gamma$--convergence result.
975: 
976: \begin{theorem}\label{t:gammac}
977: The sequence $(F_p)$ $\Gamma$-converges to $F$.
978: \end{theorem}
979: 
980: \begin{proof}
981: (i)
982: Let $\vec{V}\in \elle{2}$.
983: It is not restrictive to assume that $F(\vec{V})<+\infty$,
984: so that $\vec{V}\in X_2$ and $\gd(\curl\vec{V})\leq 1$ a.e.
985: Then it is enough to choose $\vec{V}_p = \vec{V}$ as recovery sequence.
986: 
987: \noindent
988: (ii)
989: Let $\vec{V}_p \in \elle{2}$ be functions converging to
990: $\vec{V}$ in $\elle{2}$, such that
991: \[
992: \liminf\limits_{p\to\infty} F_p(\vec{V}_p)<+\infty.
993: \]
994: By Lemma~\ref{l:rholim} we have that $\gd(\curl\vec{V})\leq 1$
995: a.e.\ in $A$, and hence
996: \[
997: F(\vec{V}) = \int_A |\vec{V}-\vec{U}_0|^2\, dx
998: = \lim_{p\to\infty} \int_A |\vec{V}_p-\vec{U}_0|^2\, dx
999: \leq \liminf_{p\to\infty} F_p(\vec{V}_p)\,,
1000: \]
1001: concluding the proof.
1002: \end{proof}
1003: Notice that the $\Gamma$--convergence result is obtained under the sole
1004: assumption that $\Uz\in\elle{2}$. Actually in our model
1005: $\Uz=\vec{H}^n_i-\vec{H}_s(t^n_{i+1})$ so that $\Uz\in X_2+\vec{c}_0$,
1006: $\cz\in\R^3$, and $\gd(\curl\Uz)\leq 1$. This additional regularity
1007: is needed in order to recover the convergence of the minimizers.
1008: 
1009: \begin{theorem}
1010: Let $\Uz\in \cz + X_2$, $\cz\in\R^3$, with $\gd(\curl\Uz)\leq 1$ a.e.\ in $A$.
1011: Let $\vec{U}_p$ be the unique minimizer of $F_p$, $p>3$,
1012: and let $\Usol$ be the unique minimizer of $F$.
1013: Then $(\vec{U}_p)$ converges to $\vec{U}$ in the weak
1014: topology of $\wp{q}$ for every $q>3$.
1015: %In particular, $(\vec{U}_p)$ converges to $\Usol$
1016: %in $\elle{2}$.
1017: \end{theorem}
1018: 
1019: \begin{proof}
1020: Reasoning as in (\ref{f:gac1}), for $3<q<p$ we have
1021: \begin{equation}\label{f:gac2}
1022: \int_A |\curl\vec{U}_p|^q\, dx \leq
1023: \frac{1}{c_1^q}\left(\frac{p}{|A|}\right)^{q/p}|A|
1024: \left(\int_A \frac{1}{p}\,\gd(\curl\vec{U}_p)^p\, dx\right)^{q/p}\,.
1025: \end{equation}
1026: On the other hand, by the minimality of $\vec{U}_p$ we get
1027: \[
1028: \int_A \frac{1}{p}\,\gd(\curl\vec{U}_p)^p\, dx \leq F_p(\vec{U}_p)
1029: \leq F_p(\Uz-\cz)
1030: \leq (1+\lambda\, |\cz|^2) |A|\,,
1031: \]
1032: so that
1033: %Collecting all the previous estimates,
1034: %for every $p > q$ we obtain
1035: \[
1036: \int_A |\curl\vec{U}_p|^q\, dx \leq
1037: \frac{1}{c_1^q}\,
1038: [p(1+\lambda\, |\cz|^2)]^{q/p} |A|
1039: \leq
1040: \frac{1}{c_1^q}\,|A|\,
1041: \exp[q(1+\lambda\, |\cz|^2)/e]\,.
1042: \]
1043: Then the sequence $(\vec{U}_p)$ is bounded in
1044: $\wp{q}$.
1045: In particular,
1046: it is precompact in $\elle{2}$.
1047: Then, by Remark~\ref{r:gc} we obtain the
1048: convergence of $(\vec{U}_p)$ to $\vec{U}$ in $\elle{2}$.
1049: Finally, notice that the previous estimate implies that
1050: $(\vec{U}_p)$ converges to $\vec{U}$ in
1051: the weak topology of $W^{1,q}$ for every $q>3$.
1052: \end{proof}
1053: 
1054: 
1055: 
1056: 
1057: %The mathematical justification of the Bad\'\i a and L\'opez model is the following.
1058: %
1059: %\begin{theorem}\label{t:conv3d}
1060: %Let $\Uz\in \cz + X_2$, $\cz\in\R^3$, with $\gd(\curl\Uz)\leq 1$ a.e.\ in $A$.
1061: %Let $\vec{U}_p$ be the unique minimizer in
1062: %$X_p$ of the functional $F_p$. Then there exists
1063: %$\Usol \in X_2$, with $\gd(\curl\Usol)\leq 1$ a.e.\ in $A$,
1064: %such that $\vec{U}_p$ converge to $\Usol$
1065: %in the weak topology of $W^{1,q}(A)^3$ for every $q>3$.
1066: %Moreover
1067: %$\Usol$ is the unique minimizer of the functional
1068: %\[
1069: %F(\vec{V})=\int_A \left[
1070: %I_{\Delta}(\curl\vec{V})+
1071: %|\vec{V}-\Uz|^2\right]\, dx\,, \quad
1072: %\vec{V}\in X_2\,.
1073: %\]
1074: %\end{theorem}
1075: %
1076: %\begin{proof}
1077: %It is easy to verify that $F_p$ is lower semicontinuous in
1078: %the weak topology of $W^{1,p}(A)$.
1079: %On the other hand, from Remark~\ref{r:Xp} we infer that
1080: %$F_p$ is coercive in $X_p$, hence the existence of
1081: %a minimizer $\vec{U}_p$ follows from the direct method
1082: %of calculus of variations.
1083: %The uniqueness of the minimizer is a consequence of the
1084: %strict convexity of $F_p$.
1085: %
1086: %We are going to prove that the sequence $(\vec{U}_p)$
1087: %is bounded in $W^{1,q}(A)$ for every $q>3$.
1088: %In view of Theorem~\ref{t:Xp}, it is enough show that
1089: %there exists a constant $C>0$ such that
1090: %${\|\curl\vec{U}_p\|}_q \leq C$ for every $p>q$.
1091: %
1092: %Given $3<q<p$, by (\ref{robdd}) and H\"older's inequality we have
1093: %\begin{equation}\label{f:gac1}
1094: %\begin{split}
1095: %%\int_{A} |D\vec{U}_p|^q\, dx & \leq
1096: %\int_A |\curl\vec{U}_p|^q\, dx & \leq
1097: %\frac{1}{c_1^q}\int_{A} \gd(\curl\vec{U}_p)^q\, dx
1098: %\\ & \leq
1099: %\frac{1}{c_1^q}\left(\frac{p}{|A|}\right)^{q/p}|A|
1100: %\left(\int_A \frac{1}{p}\,\gd(\curl\vec{U}_p)^p\, dx\right)^{q/p}\,.
1101: %\end{split}
1102: %\end{equation}
1103: %On the other hand, by the minimality of $\vec{U}_p$ we get
1104: %\[
1105: %\int_A \frac{1}{p}\,\gd(\curl\vec{U}_p)^p\, dx \leq F_p(\vec{U}_p)
1106: %\leq F_p(\Uz-\cz)
1107: %\leq (1+\lambda\, |\cz|^2) |A|\,.
1108: %\]
1109: %Collecting all the previous estimates,
1110: %for every $p\geq q$ we obtain
1111: %\[
1112: %\int_A |\curl\vec{U}_p|^q\, dx \leq
1113: %\frac{1}{c_1^q}\,
1114: %[p(1+\lambda\, |\cz|^2)]^{q/p} |A|
1115: %\leq
1116: %\frac{1}{c_1^q}\,|A|\,
1117: %\exp[q(1+\lambda\, |\cz|^2)/e]\,.
1118: %\]
1119: %%where $C>0$ does not depend on $q$ and $p$.
1120: %Then the sequence $(\vec{U}_p)$ is bounded in $W^{1,q}(A)^3$, and we can extract a subsequence,
1121: %still denoted by $(\vec{U}_p)$, converging to a function $\Usol$ in the weak topology of
1122: %$W^{1,q}(A)^3$.
1123: %We remark that, since $\dive\vec{U}_p=0$ for every $p$,
1124: %by weak convergence we get $\dive\Usol = 0$.
1125: %
1126: %We claim that $\gd(\curl\Usol)\leq 1$ a.e.\ in $A$.
1127: %Namely, for every open set $B\subseteq A$
1128: %we have
1129: %\[
1130: %\int_B \gd(\curl\Usol)\, dx \leq \liminf_{p\to \infty}
1131: %\int_B \gd(\curl\vec{U}_p)\, dx
1132: %\]
1133: %due to the convexity of the function $\gd$
1134: %(see e.g.\ \cite[Thm.~4.2.1]{Butt}).
1135: %On the other hand, by H\"older's inequality
1136: %\[
1137: %\begin{split}
1138: %\int_B \gd(\curl\vec{U}_p)\, dx & \leq
1139: %\left(\int_B \gd(\curl\vec{U}_p)^p\, dx\right)^{1/p}|B|^{(p-1)/p} \\
1140: %& \leq p^{1/p}\left(\frac{1}{p}\int_B \gd(\curl\vec{U}_p)^p dx\right)^{1/p}|B|^{(p-1)/p}\,.
1141: %\end{split}
1142: %\]
1143: %Then $\displaystyle\int_B \gd(\curl\Usol)\, dx \leq |B|$.
1144: %Since $\gd(\curl\Usol)\in L^1(\Omega)$, for a.e.\ $x\in A$
1145: %\[
1146: %\gd(\curl\Usol(x))=\lim_{r\to 0^+} \frac{1}{|B_r(x)|}
1147: %\int_{B_r(x)} \gd(\curl\Usol)\, dx \leq 1\,,
1148: %\]
1149: %proving the claim.
1150: %
1151: %It remains to prove that $\Usol$ is the unique minimizer of $F$
1152: %in $X_2$.
1153: %The uniqueness follows from the strict convexity of $F$.
1154: %Observe that $F(\vec{V}) = +\infty$ for every $\vec{V}\not\in X_{\infty}$,
1155: %so it is enough to prove that
1156: %$F(\Usol)\leq F(\vec{V})$ for every $\vec{V}\in X_{\infty}$.
1157: %The fact that $\Usol$ is a minimizer
1158: %is an easy consequence of the fact that $\vec{U}_p$ is the unique minimizer of the
1159: %functional $\widetilde{F}_p(\vec{V})=F_p(\vec{V})-\frac{1}{p}|A|$.
1160: %Namely, the sequence $\widetilde{F}_p$
1161: %is nondecreasing and converges monotonically to $F$.
1162: %Hence for $q>3$ we have
1163: %\[
1164: %\widetilde{F}_{q}(\Usol) \leq \liminf_{p\to \infty}\widetilde{F}_{q}(\vec{U}_p) \leq
1165: %\liminf_{p\to \infty}\widetilde{F}_{p}(\vec{U}_p) \leq
1166: %\liminf_{p\to \infty}\widetilde{F}_{p}(\vec{V})
1167: %=F(\vec{V})\,,
1168: %\]
1169: %for all $\vec{V}\in X_{\infty}$. A passage to the limit in $q$ concludes the proof.
1170: %\end{proof}
1171: %
1172: 
1173: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1174: \section{Cylindrical superconductors}\label{s:bean2}
1175: 
1176: In what follows we shall consider a long cylindrical type--II superconductor
1177: occupying the region $\Omega\times\R$, with a simply connected
1178: cross section $\Omega\subset \R^2$, placed into a nonstationary uniform axial magnetic field
1179: $\vec{H}_s(t)=(0,0,H_s(t))$.
1180: 
1181: Notice that, due to the parallel geometry, the problem admits a dimensional reduction.
1182: Namely, we can assume that there exists a function
1183: $u$ depending only on $x=(x_1,x_2)\in \Omega$ such that $\vec{H}=(0,0,H_s(t)+u(x,t))$, so that
1184: $\curl \vec{H}=(u_{x_2}(x,t),-u_{x_1}(x,t), 0)$
1185: and $\dive\vec{H}=0$.
1186: In particular the current
1187: density $\vec{J}$ is parallel to the cross section plane and its projection
1188: on this plane is $(J_1,J_2)=(Du)^\bot$, where $(Du)^\bot$ is the
1189: rotation of $-\frac{\pi}{2}$ of $Du$.
1190: Hence, under the additional hypothesis that
1191: the allowed region $\Delta$ is symmetric with respect to the $z=0$ plane, the constitutive law becomes
1192: $Du\in K$, where $K$ is the rotation of the section $z=0$ of the set $\Delta$,
1193: and $\vec{E}= (E_1,E_2, 0)$ with $(E_1,E_2)\in \partial I_K(Du)$.
1194: 
1195: In this case, if $c^n_i=H_s(t^n_i)$, and
1196: $h^n_i(x)$ is the magnetic field intensity at time layer $i$, given by $H_s(t^n_i)+u(t^n_i,x)$,
1197: the variational formulation is
1198: \begin{equation}\label{varmod}
1199: \min\left\{
1200: \int_\Omega  (w-h^n_i)^2\, dx;\
1201: w\in c^n_{i+1}+W^{1,1}_0(\Omega),\
1202: Dw\in K\ \textrm{a.e.~in}\ \Omega
1203: \right\}\,.
1204: \end{equation}
1205: In the following sections we shall give an explicit representation
1206: of the unique minimizer $h^n_{i+1}$ and we shall show that the previous quasistatic evolution
1207: converges to a function $h(x,t)$ (also explicit), which determines the evolution of the internal
1208: magnetic field.
1209: 
1210: Further to the dimensional reduction and a rescaling, we obtain that
1211: $h^n_{i+1}-c^n_{i+1}$ is the unique minimum point of the functional
1212: \[
1213: J(v)=\int_\Omega\left[
1214: I_K(Dv)+
1215: (v-\dato)^2\right]\, dx\,,\qquad  v\in W^{1,1}_0(\Omega),
1216: \]
1217: where $\dato=h^n_{i}-c^n_{i+1}$.
1218: 
1219: Although in this case we deal with an unbounded
1220: superconductor $A=\Omega\times\R$,
1221: the above reduction allows us to
1222: prove, using the same arguments as in
1223: Section~\ref{s:bean} (see also \cite{GNP}, Section 2),
1224: a $\Gamma$--convergence result, and,
1225: consequently, the convergence of the minimizers.
1226: Even if the case $N=2$ is the
1227: sole meaningful physical situation, the following results hold for
1228: every dimension $N \geq 2$.
1229: 
1230: \begin{theorem}
1231: Let $\Omega\subset\R^N$ be a bounded open set, let $K\subset\R^N$ be a nonempty, compact, convex set
1232: with $0 \in \inte K$, and let $\gauge_K$ be the gauge function of $K$.
1233: Let $G_p$, $p\geq 1$, and $G$ be the
1234: functionals defined respectively by
1235: \[
1236: G_p(v)=
1237: \begin{cases}
1238: \displaystyle{\int_\Omega \frac{1}{p}\,\gauge_K(Dv)^p\, dx\,,} & v\in W^{1,p}_0(\Omega) \\
1239: +\infty & \textrm{otherwise in }\ L^1(\Omega)
1240: \end{cases}
1241: \]
1242: and
1243: \[
1244: G(v)=
1245: \begin{cases}
1246: 0, & v\in W^{1,\infty}_0(\Omega),\ \gauge(Dv)\in K\\
1247: +\infty & \textrm{otherwise in }\ L^1(\Omega)\,.
1248: \end{cases}
1249: \]
1250: Then $(G_p)$ $\Gamma$--converges to $G$ in $L^r(\Omega)$ for every $r\geq 1$.
1251: As a consequence, for a given
1252: $\dato \in L^2(\Omega)$
1253: the functionals
1254: \[
1255: {J}_p(v)=
1256: \begin{cases}
1257: \displaystyle{\int_\Omega \left[\frac{1}{p}\,\gauge_K(Dv)^p\, dx+
1258: \left(v-\dato\right)^2\right]\, dx}\,, & v\in W^{1,p}_0(\Omega) \\
1259: +\infty & \textrm{otherwise in }\ L^2(\Omega)
1260: \end{cases}
1261: \]
1262: $\Gamma$--converge to the functional
1263: \[
1264: {J}(v)=
1265: \begin{cases}
1266: \displaystyle{\int_\Omega \left[
1267: I_K(Dv)+
1268: (v-\dato)^2\right]\, dx}\,, & v\in W^{1,1}_0(\Omega)\\
1269: +\infty & \textrm{otherwise in }\ L^2(\Omega)\,,
1270: \end{cases}
1271: \]
1272: in $L^2(\Omega)$.
1273: If in addition  $ \dato \in c_0+W^{1,\infty}_0(\Omega)$, $c_0\in\R$, with $\gauge(D\dato)\leq 1$ a.e. in $\Omega$,
1274: then the minimizers $u_p$ of $J_p$ converge to the unique minimizer $\uz$ of the functional $J$
1275: in the weak topology of $W^{1,q}_0(\Omega)$ for every $q>1$.
1276: \end{theorem}
1277: 
1278: 
1279: 
1280: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1281: \section{The Minkowski distance function}\label{s:distance}
1282: 
1283: Here and hereafter we shall deal with
1284: \begin{equation}\label{f:Omega}
1285: \Omega\subset\R^N\
1286: \textrm{nonempty, bounded, open connected set
1287: of class $C^2$},
1288: \end{equation}
1289: and
1290: \begin{equation}\label{f:ipoK}
1291: \begin{split}
1292: & K\subset\R^N\,\ \textrm{nonempty, compact, convex set},\
1293: 0 \in \inte K, \\
1294: &\partial K\ \textrm{of class}\ C^2\ \textrm{with strictly positive principal curvatures}\,.
1295: \end{split}
1296: \end{equation}
1297: %In Section \ref{s:minp} we shall assume $n\geq 2$, while in Sections \ref{s:bean2} and
1298: %\ref{s:evol} we shall consider the case $n=2$, which is the only meaningful physical situation.
1299: 
1300: The polar set of $K$
1301: is defined by
1302: \[
1303: K^0 = \{p\in\R^N;\ \pscal{p}{x}\leq 1\ \forall x\in K\}\,.
1304: \]
1305: We recall that, if $K$ satisfies (\ref{f:ipoK}), then
1306: $K^0$ also satisfies (\ref{f:ipoK}), and $K^{00} = (K^0)^0 = K$
1307: (see \cite[Thm.~1.6.1]{Sch}).
1308: 
1309: Since $K$ will be kept fixed, from now on we shall use for the gauge functions
1310: the notation $\gauge = \gauge_K$ and $\pgauge=\gauge_{K^0}$
1311: 
1312: %The gauge function of the convex set $K$ is the convex, positively
1313: %1--homogeneous function defined by
1314: %\[
1315: %\gau{K}(\xi) = \inf\{ t\geq 0;\ \xi\in t K\}\,.
1316: %\]
1317: %Since $\partial K$ is of class $C^2$, the gauge function $\gau{K}$
1318: %turns out to be twice differentiable in $\R^N\setminus \{0\}$.
1319: %Moreover, since $K$ is a compact set containing a neighborhood of
1320: %the origin, there exist $0<c_1<c_2$ such that
1321: %\begin{equation}\label{robdd}
1322: %c_1|\xi| \leq \gau{K}(\xi)\leq c_2 |\xi|\,, \qquad \forall\ \xi \in \R^N\,.
1323: %\end{equation}
1324: It can be readily seen that
1325: the gauge function $\pgauge$ of the polar set $K^0$
1326: coincides with the support function of the set $K$.
1327: As a consequence, we have that
1328: \begin{equation}\label{f:roroz}
1329: \gauge(D\pgauge(\xi))=1\, \qquad
1330: \forall\xi\in\R^N\setminus \{0\}\,.
1331: \end{equation}
1332: 
1333: %We recall that, since $K$ has a smooth boundary, the subgradient of the indicator function
1334: %$I_K$ is
1335: %\begin{equation}\label{subindk2}
1336: %\partial I_K(\xi)=
1337: %\begin{cases}
1338: %\{\alpha D\gauge(\xi)\colon \alpha \geq 0\} & \textrm{if}\ \xi\in \partial K \\
1339: %\emptyset & \textrm{if}\ \xi\not\in K \\
1340: %\{0\} & \textrm{if}\ \xi\in \inte K\,.
1341: %\end{cases}
1342: %\end{equation}
1343: 
1344: Let $\Lipr(\Omega)$ be the set of functions defined by
1345: \[
1346: \Lipr(\Omega) := \{u \in \Lip(\overline{\Omega}) \colon\ Du\in K
1347: \ \textrm{a.e.\ in \ }\Omega\}.
1348: \]
1349: We recall that $u\in \Lipr(\Omega)$ if and only if
1350: \begin{equation}\label{f:stimrho}
1351: u(x)-u(y) \leq \pgauge(x-y)\,,\qquad
1352: \end{equation}
1353: for every $x$, $y\in\Omega$ joined by a segment contained in $\Omega$.
1354: 
1355: 
1356: 
1357: The main tools needed in the following sections are the
1358: distances from the boundary of $\Omega$ associated to the
1359: Minkowski structures induced by the gauge function of $K^0$
1360: and $-K^0$ respectively.
1361: 
1362: 
1363: \begin{dhef}\label{d:dist}
1364: \ The Min\-kow\-ski distance from the
1365: boundary of $\Omega$ is
1366: \begin{equation}\label{f:d}
1367: \dist(x) = %\delta_{\R^n\setminus\Omega}(x) =
1368: \inf_{y\in\partial\Omega} \pgauge(x-y),
1369: \qquad x\in\overline{\Omega}\,.
1370: \end{equation}
1371: Similarly, we define $\distm(x)$ as the distance from the boundary
1372: induced by the gauge of $-K^0$.
1373: \end{dhef}
1374: 
1375: Notice that the function $\distm$ coincides with $\dist$ only if
1376: $K^0$ is symmetric with respect to the origin. In the remaining
1377: part of this section we illustrate some features of the function
1378: $\dist$. The analogous for $\distm$ can be obtained upon observing
1379: that $-K^0$ is the polar set of $-K$.
1380: 
1381: Since $\partial\Omega$ is a compact subset of $\R^N$
1382: and $\pgauge$ is a continuous function,
1383: the infimum in the definition of $\dist$ is achieved.
1384: We shall denote by $\proj(x)$ the set of projections of $x$
1385: in $\partial\Omega$, that is
1386: \begin{equation}\label{f:Pi}
1387: \proj(x) = \{y\in\partial\Omega;\ \dist(x) = \pgauge(x-y)\},
1388: \qquad x\in\overline{\Omega}.
1389: \end{equation}
1390: 
1391: \begin{dhef}\label{d:sigma}
1392: We say that $x\in\Omega$ is a
1393: regular point of $\Omega$
1394: if $\proj(x)$ is a singleton.
1395: We say that $x\in\Omega$ is a
1396: singular point of $\Omega$
1397: if $x$ is not a regular point.
1398: We denote by $\Sigma\subseteq\Omega$ the set
1399: of all singular points of $\Omega$.
1400: \end{dhef}
1401: 
1402: \begin{rk}\label{r:HJ}
1403: It is well known that $\dist$ is a Lipschitz function
1404: in $\Omega$, and that $\dist$ is differentiable
1405: at $x\in\Omega$ if and only if $x$ is a regular point.
1406: In addition, if $\proj(x) = \{y\}$ then
1407: $D\dist(x)=D\pgauge(x-y)$, and hence, by (\ref{f:roroz}),
1408: $D\dist\in K$ almost everywhere in $\Omega$
1409: (see \cite{BaCD,CaSi,Li}; see also
1410: Theorem~\ref{t:cm6}(i) below).
1411: \end{rk}
1412: 
1413: \begin{rk}\label{r:maxm}
1414: We recall that $\dist$ (resp. $\distm$) is the unique viscosity solution
1415: of the Hamilton-Jacobi equation
1416: $\gauge(Du) = 1$ (resp.\ $-\gauge(Du) =-1$) in $\Omega$,
1417: with boundary condition
1418: $u=0$ on $\partial\Omega$.
1419: As a consequence of the maximality property of the viscosity
1420: solutions, we have that
1421: \[
1422: -\distm\leq u \leq \dist\,, \qquad \forall u\in\Lipr(\Omega),\ u=0\ \textrm{on}\ \partial\Omega
1423: \]
1424: (see \cite{Li}). Moreover it can be easily checked that
1425: \[
1426: \begin{aligned}
1427: u  &\geq -\distm\,, &\quad \forall u\in\Lipr(\Omega),\ u\geq 0\ \textrm{on}\ \partial \Omega\,, \\
1428: u  &\leq \dist\,,  & \quad \forall u\in\Lipr(\Omega),\ u\leq 0\ \textrm{on}\ \partial \Omega\,,
1429: \end{aligned}
1430: \]
1431: (see e.g.\ \cite{BaCD}, Theorem 5.9).
1432: \end{rk}
1433: 
1434: %We recall here some properties of the
1435: %singular set $\Sigma$.
1436: %In \cite{CMf} there were studied the regularity properties
1437: %of $\dist$ and $\overline{\Sigma}$.
1438: In the following theorem
1439: we collect
1440: all the results proved in \cite{CMf} that are
1441: relevant for the subsequent analysis.
1442: 
1443: \begin{theorem}\label{t:cm6}
1444: Assume that $\Omega$ and $K$ satisfy respectively $(\ref{f:Omega})$
1445: and $(\ref{f:ipoK})$. Then the following hold.
1446: \begin{itemize}
1447: \item[(i)] $\overline{\Sigma}\subset\Omega$,
1448: and the Lebesgue measure of
1449: $\overline{\Sigma}$ is zero.
1450: 
1451: \item[(ii)] %\label{l:pis}
1452: Let $x\in\Omega$ and $\macroy \in\proj(x)$.
1453: Then $\dist$ is differentiable at every $z$
1454: along the segment jointing $y$ to $x$ (without endpoints),
1455: and
1456: $
1457: D\dist(z) = \frac{\nor(\macroy )}{\gauge(\nor(\macroy ))}
1458: $,
1459: where $\nor(\macroy )$ is the (Euclidean) inward normal unit vector
1460: to $\partial\Omega$ at $\macroy $.
1461: 
1462: \item[(iii)] %\label{p:regd}
1463: The function $\dist$ is of class $C^2$ in
1464: $\overline{\Omega}\setminus\overline{\Sigma}$.
1465: \end{itemize}
1466: \end{theorem}
1467: 
1468: \begin{proof}
1469: See  Remark~4.16, Corollary~6.9,
1470: Lemma~4.3
1471: and Theorem~6.10 in \cite{CMf}.
1472: \end{proof}
1473: 
1474: {}At any point $\macroy \in\partial\Omega$
1475: there is a unique inward ``normal'' direction $p(\macroy )$
1476: with the properties
1477: $\proj(\macroy +t p(\macroy )) = \{\macroy \}$
1478: and $\dist(\macroy +t p(\macroy )) = t$
1479: for $t\geq 0$ small enough
1480: (see \cite[Remark~4.5]{CMf}).
1481: More precisely, these properties
1482: hold true for $p(\macroy ) = D\gauge(\nor(\macroy ))$
1483: and for every $t\in [0, \len(\macroy ))$,
1484: where $\len(\macroy )$ is defined by
1485: \begin{equation}\label{f:len}
1486: \len(\macroy ) =
1487: \min\{t\geq 0;\
1488: \macroy  + t D\gauge(D\dist(x))\in\overline{\Sigma}\}
1489: \end{equation}
1490: (see
1491: \cite[Propositions~4.4 and~4.8]{CMf} and \cite[Lemma~2.2]{LN}).
1492: %It can be proved that $p(\macroy ) = D\gauge(\nor(\macroy ))$
1493: %(see \cite[Lemma~2.2]{LN}
1494: %and \cite[Proposition~4.4]{CMf}).
1495: {}From Theorem~\ref{t:cm6}(ii) and the positive $0$-homogeneity
1496: of $D\gauge$ it is plain that
1497: $D\gauge(\nor(\macroy )) = D\gauge(D\dist(\macroy ))$.
1498: 
1499: {}From Theorem~\ref{t:cm6}(iii), the function $\dist$
1500: is of class $C^2$ on
1501: %$\overline{\Omega}\setminus\overline{\Sigma}$.
1502: $\partial\Omega$.
1503: Then we can define the function
1504: \begin{equation}\label{f:W}
1505: W(\macroy) = -D^2\gauge(D\dist(\macroy))\, D^2\dist(\macroy)\,,\qquad
1506: \macroy\in
1507: %\overline{\Omega}\setminus\overline{\Sigma}\,.
1508: \partial\Omega\,.
1509: \end{equation}
1510: For any $\macroy \in\partial\Omega$ let $T_{\macroy }$ denote
1511: the tangent space to $\partial\Omega$ at $\macroy $.
1512: %If $x\in\overline{\Omega}\setminus{\Sigma}$
1513: %and $\proj(x) = \{\macroy \}$,
1514: %we set $T_x = T_{\macroy }$.
1515: %It can be proved that for every $v\in T_x$,
1516: %one has $W(x)\, v\in T_x$.
1517: It can be proved that for every $v\in T_y$,
1518: one has $W(y)\, v\in T_y$.
1519: Hence, we can define the map
1520: \begin{equation}\label{f:bw}
1521: \bw(y)\colon T_y\to T_y,\quad
1522: \bw(y)\, w = W(y)\, w,
1523: \end{equation}
1524: that can be identified with a linear application
1525: from $\R^{N-1}$ to $\R^{N-1}$.
1526: 
1527: Although
1528: the matrix $\bw(\macroy)$ is not in general symmetric,
1529: its eigenvalues
1530: are real numbers,
1531: and so its eigenvectors are real
1532: (see \cite{CMf}, Remark~5.3).
1533: The eigenvalues of $\bw(\macroy)$
1534: have an important geometric interpretation.
1535: 
1536: \begin{dhef}\label{d:curv}
1537: Let $\macroy \in\partial\Omega$.
1538: The anisotropic curvatures of $\partial\Omega$ at $\macroy $,
1539: with respect to the Minkowski norm $\dist$,
1540: are the eigenvalues
1541: $\curvg_1(\macroy )\leq\cdots\leq\curvg_{n-1}(\macroy )$
1542: of $\bw(\macroy )$.
1543: %The corresponding eigenvectors are the
1544: %\textsl{principal $\gauge$-directions} of $\partial\Omega$
1545: %at $\macroy $.
1546: \end{dhef}
1547: 
1548: In what follows we shall denote by $\Pi^-$, $\curvg^-_j$, $j=1,\ldots,n-1$, $\Sigma^-$ and
1549: $\len^-$,
1550: respectively
1551: the projection, the anisotropic curvatures and the singular set
1552: associated to $\distm$, and the normal distance to the cut locus $\Sigma^-$.
1553: 
1554: 
1555: 
1556: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1557: \section{A minimum problem}\label{s:minp}
1558: 
1559: %In this section, $\Omega$ and $K$ are subsets of $\R^N$, $N\geq 2$, fulfilling
1560: %(\ref{f:Omega}) and (\ref{f:ipoK}) respectively.
1561: %We shall study a minimum problem for integral functionals defined in $\Omega$
1562: %which will be, in dimension $n=2$,
1563: %the basic tool for finding the quasistatic evolution of all the physical
1564: %quantities involved in the electrodynamic of superconductors.
1565: 
1566: Fixed $\dato\in \Lipr(\Omega)$, we want to minimize the functional
1567: \begin{equation}\label{jfunct}
1568: J(v)=\int_\Omega [I_K(Dv)+(v-\dato)^2]\, dx\,, \quad v\in W^{1,1}_0(\Omega)\,.
1569: \end{equation}
1570: Let us consider the partition
1571: $\Omega = \Omega^+\cup\Omega^-\cup \Omega^0$, with
1572: \[
1573: \begin{split}
1574: \Omega^0  =\{-\distm \leq \dato \leq \dist\}\,,\
1575: \Omega^+ =\{\dato > \dist\}\,, \
1576: \Omega^-=\{ \dato <- \distm\}\,,
1577: \end{split}
1578: \]
1579: and define the function
1580: \begin{equation}\label{umin}
1581: u(x)=
1582: \begin{cases}
1583: \dist(x) & x\in \Omega^+ \\
1584: -\distm(x) & x\in \Omega^- \\
1585: \dato(x) & x \in \Omega^0\,.
1586: \end{cases}
1587: \end{equation}
1588: The aim of this section is to show that $u$ is the unique minimizer
1589: of $J$ in $W^{1,1}_0(\Omega)$.
1590: 
1591: \begin{rk}\label{r:noint}
1592: If $x \in \overline{\Omega}^+ \cap \overline{\Omega}^-$, then
1593: $\dato(x)=\dist(x)=-\distm(x)$ and hence $x\in \partial \Omega$
1594: and $\dato(x)=0$. As a consequence, we have
1595: $u(x)=\min((\max(\dato(x), -\distm(x)), \dist(x)))$.
1596: This implies that
1597: $u\in \Lipr(\Omega)$, $u=0$ on $\partial\Omega$,
1598: $\gauge(Du)=1$ a.e.\ in $\Omega^+\cup\Omega^-$, and $u$ is a viscosity
1599: solution of $\gauge(Du)=1$ in $\Omega^+$ and of $-\gauge(Du)=-1$ in $\Omega^-$.
1600: \end{rk}
1601: 
1602: Let us define the following two subsets of $\partial\Omega$:
1603: \begin{equation}\label{f:gpm}
1604: \Gamma^+=\{y\in\partial\Omega \colon \dato(y)>0\},\qquad
1605: \Gamma^-=\{y\in\partial\Omega \colon \dato(y)<0\}\,,
1606: \end{equation}
1607: and the functions
1608: \begin{equation}\label{f:lpm}
1609: \begin{split}
1610: \lambda(y) & = \sup\{s\in [0,\len(y))\colon \dato(y+tD\gauge(\nor(y)))>t\
1611: \forall\ t\in [0,s)\}\,,\\
1612: \lambda^-(y) & = \sup\{s\in [0,\len^-(y))\colon \dato(y+tD\gaugem(\nor(y)))<-t\
1613: \forall\ t\in [0,s)\}\,.
1614: \end{split}
1615: \end{equation}
1616: Recall that $\len(y)$ and $\len^-(y)$
1617: are the normal distance to the cut locus,
1618: corresponding to $\dist$ and $\distm$ respectively, that is the length of the ``normal'' ray
1619: starting from $y$ and ending on $\overline{\Sigma}$ (see (\ref{f:len})).
1620: 
1621: \begin{rk}\label{r:lambda}
1622: It can be easily checked that $\lambda$ and $\lambda^-$ are lower semicontinuous function
1623: in $\partial \Omega$. In general we cannot expect the continuity of these functions, as it
1624: is shown by the following example.
1625: Let $\Omega=B_1(0)\subseteq \R^2$, $K=\overline{B}_1(0)\subseteq \R^2$, and
1626: define
1627: $\dato(r \cos \theta,r \sin \theta)=1-r \cos \theta$
1628: for $0\leq r<1$, $\theta\in [0,2\pi)$.
1629: We have
1630: $|D\dato(r \cos\theta, r\sin\theta)|^2=\cos^2 \theta + \sin^2 \theta=1$
1631: ($0<r<1$), so that $\dato\in \Lipr(\Omega)$.
1632: Moreover, if $y=(\cos\theta, \sin\theta)\in \partial \Omega$, $\theta\in [0,2\pi)$,
1633: $\lambda$ is
1634: the lower semicontinuous function given by
1635: \[
1636: \lambda(y)=
1637: \begin{cases}
1638: 0\,, & \theta=0\,, \\
1639: 1\,, & \theta\in (0,2\pi)\,.
1640: \end{cases}
1641: \]
1642: \end{rk}
1643: 
1644: \smallskip
1645: 
1646: The following simple lemma provides a
1647: characterization of $\Omega^+$ and $\Omega^-$ in terms of the rays starting from points
1648: of $\Gamma^+$ and $\Gamma^-$ respectively.
1649: 
1650: \begin{lemma}\label{l:opm}
1651: Let $y\in\Gamma^+$ be such that
1652: $\lambda(y)<\len(y)$.
1653: Then $y+t D\gauge(\nor(y))\in\Omega^0$
1654: for every $t\in [\lambda(y),\len(y)]$.
1655: Analogously,
1656: let $y\in\Gamma^-$ be such that
1657: $\lambda^-(y)<\len^-(y)$.
1658: Then
1659: $y+t D\gaugem(\nor(y))\in\Omega^0$
1660: for every $t\in [\lambda^-(y),\len^-(y)]$. As a consequence, we have
1661: \begin{equation}\label{f:splom}
1662: \begin{split}
1663: \Omega^+\setminus \overline{\Sigma}^+ & =\{y+t D\gauge(\nor(y))\colon y\in\Gamma^+,\
1664: t\in (0, \lambda(y))\},\\
1665: \Omega^-\setminus \overline{\Sigma}^- & =\{y+t D\gaugem(\nor(y))\colon y\in\Gamma^-,\
1666: t\in (0, \lambda^-(y))\}.
1667: \end{split}
1668: \end{equation}
1669: \end{lemma}
1670: 
1671: \begin{proof}
1672: Let $y\in\Gamma^+$ and assume that $\lambda(y)<\len(y)$.
1673: {}From the very definition of $\lambda(y)$ and the continuity of $\dato$, we have that
1674: $\dato(y+\lambda(y)\, D\gauge(\nor(y))) = \lambda(y)$.
1675: For every $t\in [\lambda(y), \len(y)]$ we have that
1676: \[
1677: \dist(y+tD\gauge(\nor(y)))= t = \dato(y+\lambda(y)\, D\gauge(\nor(y))) + t-\lambda(y) \geq
1678: \dato(y+t\, D\gauge(\nor(y)))\,,
1679: \]
1680: where the last inequality follows from the assumption $\dato\in\Lip_\rho(\Omega)$
1681: (see (\ref{f:stimrho})), and from the fact that $\pgauge((t-\lambda)D\gauge(\nor(y)))= t-\lambda$
1682: (see (\ref{f:roroz})).
1683: The inequality
1684: $\dato(y+t\, D\gauge(\nor(y))) \geq -\distm(y+t\, D\gauge(\nor(y)))$
1685: follows easily from the fact that
1686: $\dato(y) > 0$ and $\dato\in\Lipr(\Omega)$ (see Remark \ref{r:maxm}).
1687: The case $y\in\Gamma^-$ can be handled in a similar way.
1688: \end{proof}
1689: 
1690: %As a consequence of Lemma~\ref{l:opm}, we have
1691: %the following characterization of the sets $\Omega^+$ and $\Omega^-$:
1692: %\begin{equation}\label{f:splom}
1693: %\begin{split}
1694: %\Omega^+ & =\{y+t D\gauge(\nor(y))\colon y\in\Gamma^+,\
1695: %t\in (0, \lambda(y))\},\\
1696: %\Omega^- & =\{y+t D\gaugem(\nor(y))\colon y\in\Gamma^-,\
1697: %t\in (0, \lambda^-(y))\}.
1698: %\end{split}
1699: %\end{equation}
1700: 
1701: 
1702: The following change of variable formula can be proved as in
1703: \cite[Theorem~7.1]{CMf}.
1704: 
1705: \begin{theorem}\label{t:chvar}
1706: Let $\Phi, \Phi^-\colon\partial\Omega\times\R\to\R^N$
1707: be the maps defined respectively by
1708: \[
1709: \Phi(y,t) = y + t\, D\gauge(\nor(y)), \quad
1710: \Phi^-(y,t) = y + t\, D\gaugem(\nor(y)),
1711: \ \quad (y,t)\in\partial\Omega\times\R.
1712: \]
1713: Then
1714: \[
1715: \begin{split}
1716: \int_{\Omega^+} h(x)\, dx
1717: &=
1718: \int_{\Gamma^+} \gauge(\nor(y))\, \left[
1719: \int_{0}^{\lambda(y)}
1720: h(\Phi(y,t))\,
1721: \prod_{i=1}^{n-1} (1-t\, \curvg_i(y))
1722: \, dt\right]\,d\haus(y)
1723: \\
1724: \int_{\Omega^-} h(x)\, dx
1725: &=
1726: \int_{\Gamma^-} \gaugem(\nor(y))\, \left[
1727: \int_{0}^{\lambda^-(y)}
1728: h(\Phi^-(y,t))\,
1729: \prod_{i=1}^{n-1} (1-t\, \curvg^-_i(y))
1730: \, dt\right]\,d\haus(y)
1731: \end{split}
1732: \]
1733: for every $h\in L^1(\Omega)$.
1734: \end{theorem}
1735: 
1736: In what follows we will set
1737: \[
1738: M_{x}(t) = \prod_{i=1}^{n-1}\frac{1-t\curvg_i(x)}{1-\dist(x)\curvg_i(x)}\,, \quad
1739: M_{x}^-(t) = \prod_{i=1}^{n-1}\frac{1-t\curvgm_i(x)}{1-\distm(x)\curvgm_i(x)}\,.
1740: \]
1741: {}From Lemma~7.3 in \cite{CMf} we deduce that there exists a positive constant $M_0$
1742: such that
1743: \begin{equation}\label{f:estiM}
1744: 0\leq M_x(t)\leq M_0\quad
1745: \forall x\in\Omega\setminus\overline{\Sigma},\ \dist(x)\leq t < \len(y),
1746: \end{equation}
1747: where $\proj(x) = \{y\}$.
1748: An analogous estimate holds for $M_x^-(t)$.
1749: 
1750: Using Theorem \ref{t:chvar} and (\ref{f:splom}), we are able to
1751: show that the function $u$ defined in (\ref{umin}), coupled with
1752: an explicit function $\vf$, solves a system of PDEs of Monge--Kantorowich
1753: type. \textit{A posteriori} this system will be understood as a
1754: first order necessary condition fulfilled by the minimizer of $J$.
1755: 
1756: \begin{lemma}\label{l:eul}
1757: Let $u\in \Lipr(\Omega)$ be the function defined in (\ref{umin}),
1758: and let $\vf\colon\Omega\to\R$ be the function defined by
1759: \begin{equation}\label{f:defv}
1760: \vf(x)=
1761: \begin{cases}
1762: \displaystyle
1763: \int_{\dist(x)}^{\lambda(\macroy )} [\dato(\macroy +t\,D\gauge(D\dist(x)))-t]\,
1764: M_{x}(t) dt\,, & x\in\Omega^+\setminus \overline{\Sigma}^+,\ \Pi(x)=\{y\}\\ \\
1765: \displaystyle
1766: -\int_{\distm(x)}^{\lambda^-(\macroy )} [\dato(\macroy +t\,D\gaugem(D\dist(x)))+t]\,
1767: M_{x}^-(t)
1768: \, dt\,, & x\in\Omega^-\setminus \overline{\Sigma}^-,\ \Pi^-(x)=\{y\}\\
1769: 0 & \textrm{otherwise}\,.
1770: \end{cases}
1771: \end{equation}
1772: Then $\vf\in C_b(\Omega)$, $\vf\geq 0$, and the pair $(u,\vf)$ solves the system of PDEs
1773: \begin{equation}\label{f:MK}
1774: \begin{cases}
1775: -\dive(\vf\, D\gauge(Du)) = \dato-u
1776: &\textrm{in $\Omega$},\quad \textrm{(distributional)}\\
1777: \gauge(Du)\leq 1
1778: &\textrm{a.e.\ in $\Omega$},\\
1779: \gauge(Du) = 1
1780: &\textrm{a.e.\ in $\{\vf>0\}$},\\
1781: u=0 & \textrm{in $\partial\Omega$}\,.
1782: \end{cases}
1783: \end{equation}
1784: \end{lemma}
1785: 
1786: \begin{proof}
1787: Let us extend the functions $\lambda$ and $\lambda^-$
1788: to $\Omega^+\setminus\Sigma^+$ and $\Omega^-\setminus\Sigma^-$
1789: respectively by setting
1790: $\lambda(x):= \lambda(y)$ when $\proj(x) = \{y\}$ and
1791: $\lambda^-(x):= \lambda^-(y)$ when $\proj^-(x) = \{y\}$.
1792: Since $\lambda(x)\geq\dist(x)$ for every $x\in\Omega^+\setminus\Sigma^+$
1793: and $\lambda^-(x)\geq\distm(x)$ for every $x\in\Omega^-\setminus\Sigma^-$,
1794: it is plain that $\vf\geq 0$.
1795: 
1796: Let us prove that $\vf$ is a continuous function.
1797: It is convenient to rewrite $\vf$ in $\Omega^+\setminus \overline{\Sigma}^+$ in the following way:
1798: \begin{equation}\label{f:defvp}
1799: \vf(x)=
1800: \int_{\dist(x)}^{\len(\macroy )}
1801: [t\vee\dato(\macroy +t\,D\gauge(D\dist(x)))-t]\,
1802: M_{x}(t) dt\,, \quad x\in\Omega^+\setminus \overline{\Sigma}^+,\ \Pi(x)=\{y\}.
1803: \end{equation}
1804: Since the maps $\dist$, $D\dist$, $\dato$, $\len$, and $x\mapsto\proj(x)$ are continuous
1805: in $\Omega^+\setminus \overline{\Sigma}^+$,
1806: it follows that also $\vf$ is continuous in $\Omega^+\setminus \overline{\Sigma}^+$.
1807: 
1808: If $x_0\in\Omega^+\cap\overline{\Sigma}^+$,
1809: then $\lambda(x_0) = \len(x_0) = \dist(x_0)$,
1810: and the map $\lambda$ is continuous at $x_0$.
1811: Namely, since $x_0\in\overline{\Sigma}^+$, we have that
1812: $\len(x_0) = \dist(x_0)$.
1813: Moreover, since $x_0\in\Omega^+$, we have that
1814: $\dato(x_0) > \dist(x_0)$, so that $\dato > \dist$ on every segment
1815: $[y,x_0]\subset\overline{\Omega}$ with $y\in\partial\Omega$.
1816: Thus $\lambda(x_0)\geq\dist(x_0)$.
1817: Finally, from the lower semicontinuity of $\lambda$
1818: and the fact that $\lambda(x)\leq\len(x)$ for every $x\in\Omega$,
1819: we conclude that
1820: \[
1821: \dist(x_0)\leq\lambda(x_0)\leq\liminf_{x\to x_0}\lambda(x)\leq
1822: \limsup_{x\to x_0}\lambda(x)\leq
1823: \lim_{x\to x_0}\len(x) = \len(x_0) = \dist(x_0),
1824: \]
1825: so that
1826: $\lambda(x_0) = \dist(x_0) = \lim_{x\to x_0}\lambda(x)$.
1827: 
1828: {}From the very definition of $\vf$ and
1829: %the fact that $M_x(t)$ is bounded (see Lemma~7.3 in \cite{CMf})
1830: (\ref{f:estiM})
1831: we have that
1832: \[
1833: 0\leq \vf(x)\leq C[\lambda(x)-\dist(x)]\qquad
1834: \forall x\in \Omega^+\setminus\overline{\Sigma}^+.
1835: \]
1836: We remark that the same estimate trivially holds also on
1837: $\Omega^+\cap\overline{\Sigma}^+$ since $\vf$ vanishes.
1838: If $x_0\in\Omega^+\cap\overline{\Sigma}^+$,
1839: we have that
1840: \[
1841: \limsup_{x\to x_0}\vf(x)\leq
1842: \limsup_{x\to x_0} C[\lambda(x)-\dist(x)] = 0,
1843: \]
1844: hence $\lim_{x\to x_0}\vf(x) = 0 = \vf(x_0)$.
1845: 
1846: Up to now we have proved that $\vf$ is continuous in $\Omega^+$.
1847: A similar argument shows that $\vf$ is continuous in $\Omega^-$.
1848: Since $\vf$ vanishes on $\Omega^0$ and
1849: $\partial\Omega^+\cap\partial\Omega^-\cap\Omega=\emptyset$
1850: (see Remark~\ref{r:noint}),
1851: it remains to prove that $\vf$ is continuous on
1852: $\Omega\cap\partial\Omega^+$ and $\Omega\cap\partial\Omega^-$.
1853: 
1854: Let $x_0\in\Omega\cap\partial\Omega^+$.
1855: Since $x_0\in \Omega^0$, we have that $v(x_0)=0$.
1856: Moreover, by definition of $\Omega^+$,
1857: $\dato(x_0) = \dist(x_0)$.
1858: Since $\dato\in\Lipr(\Omega)$,
1859: by (\ref{f:stimrho}) and (\ref{robdd}),
1860: given $\epsilon > 0$ we have that
1861: \[
1862: \dato(x) \leq \dist(x) + 2 c_2 \epsilon
1863: \qquad\forall x\in\Omega\cap B_{\epsilon}(x_0).
1864: \]
1865: Moreover, if $x\in(\Omega^+\cap B_{\epsilon}(x_0))\setminus\overline{\Sigma}^+$,
1866: we have that $\dato\leq\dist + 2c_2\epsilon$
1867: along the segment joining $x$ to its cut point.
1868: As a consequence,
1869: the integrand in (\ref{f:defvp}) is bounded from above by
1870: $2 c_2\epsilon\, M_0$, so that
1871: \[
1872: 0\leq\vf(x)\leq C \epsilon
1873: \qquad\forall x\in(\Omega^+\cap B_{\epsilon}(x_0))\setminus\overline{\Sigma}^+.
1874: \]
1875: Since $\vf = 0$ in $\Omega^0\cup\overline{\Sigma}^+$,
1876: this inequality clearly implies that
1877: \[
1878: \lim_{x\to x_0}\vf(x) = 0 = \vf(x_0),
1879: \]
1880: i.e.\ $\vf$ is continuous at $x_0$.
1881: The continuity in $\Omega\cap\partial\Omega^-$
1882: can be proved in a similar way.
1883: 
1884: \smallskip
1885: Let us prove that the pair $(u,\vf)$ is a solution to (\ref{f:MK}).
1886: Upon observing that $\{\vf>0\}=\Omega^+ \cup \Omega^-$,
1887: by Remark~\ref{r:noint}
1888: we have only to
1889: prove that $(u,\vf)$ solves
1890: \[
1891: \int_\Omega \vf\pscal{D\gauge(Du)}{D\varphi}\, dx = \int_\Omega (\dato-u)\varphi\, dx\,,
1892: \]
1893: for all $\varphi\in C^{\infty}_0(\Omega)$.
1894: We have that
1895: \[
1896: \int_\Omega (\dato-u)\varphi\, dx=
1897: \int_{\Omega^+} (\dato-u)\varphi\, dx+\int_{\Omega^-} (\dato-u)\varphi\, dx\,.
1898: \]
1899: By using the change of variables stated in Theorem~\ref{t:chvar}, we get
1900: \[
1901: \begin{split}
1902: \int_{\Omega^-} (\dato-u)\varphi =
1903: \int_{\Gamma^-}\gaugem(\nor)
1904: \left[\int_0^{\lambda^-}(\dato(\Phi^-)-u(\Phi^-))\varphi(\Phi^-)
1905: \prod_{i=1}^{n-1} (1-t\, \curvg_i^-)
1906: \, dt\right]\,d\haus\,.
1907: \end{split}
1908: \]
1909: %where $\Phi(x,t)=x+t D\gaugem(\nor(x))$, $x\in\Gamma^-$, $t\in\R$.
1910: An integration by parts leads to
1911: \[
1912: \begin{split}
1913: &\int_{0}^{\lambda^-}
1914: (\dato(\Phi^-)-u(\Phi^-))\,\varphi(\Phi^-)\,
1915: \prod_{i=1}^{n-1} (1-t\, \curvg_i^-)
1916: \, dt  \\
1917: & =
1918: \int_{0}^{\lambda^-}
1919: \pscal{D\varphi(\Phi^-)}{D\gaugem(\nor)}
1920: \left[\int_{t}^{\lambda^-}
1921: (\dato(\Phi^-)-u(\Phi^-))\,
1922: \prod_{i=1}^{n-1} (1-s\, \curvg_i^-)\, ds\right]
1923: \, dt\,.
1924: \end{split}
1925: \]
1926: Since $u(\Phi^-(y,s)) = -s$ for every $s\in [0, \lambda^-(y)]$,
1927: %$\dato(\Phi(x,s))-u(\Phi(x,s))=0$ for $s\in [\lambda^-(x), \len^-(x)]$,
1928: we have that
1929: \[
1930: -\vf(\Phi^-(y,t)) =
1931: \int_{t}^{\lambda^-(y)}
1932: (\dato(\Phi^-(y,s))-u(\Phi^-(y,s)))\,
1933: \prod_{i=1}^{n-1}\frac{1-s\, \curvg_i^-(y)}%
1934: {1-t\, \curvg_i^-(y)}\, ds
1935: \]
1936: for every $y\in\Gamma^-$ and $t\in [0,\lambda^-(y))$.
1937: Then we obtain
1938: \[
1939: \begin{split}
1940: \int_{\Omega^-} & (\dato-u)\varphi\, dx
1941: =
1942: -\int_{\Gamma^-}\gaugem(\nor)\left[
1943: \int_{0}^{\lambda^-}
1944: \pscal{D\varphi(\Phi^-)}{D\gaugem(\nor)}
1945: \vf(\Phi)
1946: \prod_{i=1}^{n-1} (1-t\, \curvg_i^-)\, dt\right]\, d\haus\,.
1947: \end{split}
1948: \]
1949: Finally, again from Theorem~\ref{t:chvar} and the fact that
1950: \[
1951: D\gaugem(\nor(y))=
1952: - D \gauge (-\nor (y))=-D \gauge(-D\distm(y))=-D\gauge(Du(y))\,,
1953: \]
1954: we get
1955: \[
1956: \int_{\Omega^-} (\dato-u)\varphi\, dx =
1957: \int_{\Omega^-}
1958: \vf\, \pscal{D\gauge(Du)}{D\varphi}\, dx\,,
1959: \]
1960: for every $\varphi\in C^{\infty}_c(\Omega)$.
1961: In the same way,
1962: using the change of variables in $\Omega^+$,
1963: %parameterizing $\Omega^+$ by the map
1964: %$\Phi(x,t)=x+t D\gauge(\nor(x))$, $x\in\Gamma^+$, $t\in\R$,
1965: we get
1966: \[
1967: \int_{\Omega^+} (\dato-u)\varphi =
1968: \int_{\Omega^+}
1969: \vf\, \pscal{D\gauge(Du)}{D\varphi}\,,
1970: \]
1971: for every $\varphi\in C^{\infty}_c(\Omega)$,
1972: concluding the proof.
1973: \end{proof}
1974: 
1975: Now the absolute minimality of $u$ follows easily.
1976: \begin{theorem}\label{t:exmin}
1977: The function $u$ defined in (\ref{umin}) is the unique
1978: minimizer of $J$ in $W^{1,1}_0(\Omega)$.
1979: \end{theorem}
1980: 
1981: \begin{proof}
1982: Let $\vf$ be the function defined in (\ref{f:defv}). Since $\vf\geq 0$,
1983: for almost every $x\in \Omega^+\cup\Omega^-$
1984: the vector $2 \vf(x) D\gauge(Du(x))$ belongs to the subdifferential
1985: of $I_K$ at $Du(x)$ (see (\ref{subindk})).
1986: On the other hand, $\vf = 0$ in $\Omega^0$.
1987: Then we have
1988: \[
1989: I_K(Dw(x))-I_K(Du(x)) \geq 2 \vf(x) \pscal{D\gauge(Du(x))}{Dw(x)-Du(x)}
1990: \ \textrm{a.e.\ in\ }\Omega\,,
1991: \]
1992: for every $w\in W^{1,1}_0(\Omega)$.
1993: Hence, using Lemma \ref{l:eul} we obtain that for every
1994: $w\in W^{1,1}_0(\Omega)$
1995: \begin{equation}\label{f:minprf}
1996: \begin{split}
1997: J(w) - J(u)
1998: & \geq {}
1999: \int_\Omega 2 \vf \pscal{D\gauge(Du)}{Dw-Du} +
2000: \int_\Omega (w-\dato)^2
2001: + \int_\Omega (u-\dato)^2
2002: \\ = {} &
2003: \int_\Omega [2(w-u)(\dato-u)+(w-\dato)^2-(u-\dato)^2]
2004: \\ = {} &
2005: \int_\Omega(u-w)^2\,,
2006: \end{split}
2007: \end{equation}
2008: which implies that $u$ is the unique minimizer of $J$ in $W^{1,1}_0(\Omega)$.
2009: \end{proof}
2010: 
2011: 
2012: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2013: \section{The quasistatic evolution}\label{s:evol}
2014: Let $T>0$ be fixed. For every $n\in\N^+$, we take a partition
2015: $P^n=\left\{t^n_i\right\}_{i=0}^{k(n)}$
2016: %$0=t^n_0<t^n_1<\cdots<t^n_{n}=T$
2017: of the interval $[0,T]$, and we set $\delta t^n_i=t^n_i-t^n_{i-1}$.
2018: %Let $T>0$ be fixed. Let us take a partition $0=t^n_0<t^n_1<\ldots<t^n_{k(n)}=T$ of the
2019: %interval $[0,T]$, with $\Delta^n=t^n_{i+1}-t^n_i$, and let us set $c^n_i=H_s(t^n_i)$.
2020: %From now on we shall assume that $H_s$ is a continuous function in $[0,T]$.
2021: 
2022: The starting configuration of the superconductor at time $t=0$ is given by
2023: $h(x,0)=h_0(x)\in \Lipr(\Omega)$, with $h_0=H_s(0)$ on $\partial \Omega$. %Then the magnetic field
2024: %$h^n_{i+1}(x)$ inside the superconductor at time $t^n_{i+1}$ is the unique solution
2025: %to the minimum problem
2026: %\begin{equation}\label{varmodin}
2027: %\min \left\{\int_\Omega  (u-h^n_i)^2\, dx \colon\
2028: %u\in c^n_{i+1}+ W^{1,1}_0(\Omega),\ u\in \Lipr(\Omega)\right\}\,,
2029: %\end{equation}
2030: %where $c^n_{i+1}=H_s(t^n_{i+1})$.
2031: %Upon observing that $h^n_{i+1}$ is the minimizer for (\ref{varmodin}) if and only
2032: %if $h^n_{i+1}-c^n_{i+1}$ is the minimizer for (\ref{jfunct}) with
2033: %$\dato=h^n_i-c^n_{i+1}$, by Theorem \ref{t:exmin}
2034: Using the variational formulation (\ref{varmod}) and Theorem \ref{t:exmin}, we have that
2035: the internal magnetic field $h^n_{i+1}(x)=h(x, t^n_{i+1})$ is given by
2036: \begin{equation}\label{f:akkani}
2037: h^n_{i+1}(x)=\min[\max[h^n_i(x), c^n_{i+1}-\distm(x)], c^n_{i+1}+\dist(x)]\,.
2038: \end{equation}
2039: Since the model is valid only when the external field varies slowly in time,
2040: it is not restrictive to assume that $H_s$ is a continuous function obtained
2041: by the concatenation of a finite number of monotone functions of class $C^1$.
2042: Hence from now on we shall consider only the case
2043: \begin{equation}\label{f:HS}
2044: \textrm{$H_s$ is a monotone
2045: function of class $C^1([0,T])$.}
2046: \end{equation}
2047: In this case formula (\ref{f:akkani}) simplifies.
2048: Namely, if we assume that $H_s$ is a nondecreasing
2049: function we have that $c^n_i \leq c^n_{i+1}$.
2050: In particular, at the first
2051: time step we have
2052: \[
2053: h^n_1(x)=\min[\max[h_0(x), c^n_{1}-\distm(x)], c^n_{1}+\dist(x)]\,,
2054: \]
2055: and $h_0(x)-c^n_{1} \in \Lipr(\Omega)$ with
2056: $h_0(x)-c^n_{1}=c^n_0-c^n_{1}\leq 0$ on $\partial \Omega$.
2057: Hence $h_0(x)-c^n_{1} \leq \dist (x)$ in $\Omega$ (see Remark \ref{r:maxm}), which
2058: implies
2059: \[
2060: h^n_1(x)=\max[h_0(x), c^n_{1}-\distm(x)]\,.
2061: \]
2062: At the second time step we get
2063: \[
2064: h^n_2(x)=\min[\max[h^n_1(x), c^n_{2}-\distm(x)], c^n_{2}+\dist(x)]\,.
2065: \]
2066: Since
2067: $h^n_1(x)-c^n_{2}=\max[h_0(x)-c^n_{2}, -\distm(x)+c^n_1-c^n_2]$, with
2068: $c^n_1-c^n_2\leq 0$, we have that
2069: \[
2070: h^n_2(x)=\max[h_0(x), c^n_{2}-\distm(x)]\,.
2071: \]
2072: In general we get
2073: \begin{equation}\label{f:akkanicres}
2074: h^n_{i}(x)=\max[h_0(x), c^n_{i}-\distm(x)]\,.\quad \textrm{(nondecreasing external field)}
2075: \end{equation}
2076: With a similar argument, we obtain that if $H_s$ is nonincreasing in time, the
2077: $i$--th time step leads to
2078: \begin{equation}\label{f:akkanidecre}
2079: h^n_{i}(x)=\min[h_0(x), c^n_{i}+\dist(x)]\,.\quad \textrm{(nonincreasing external field)}
2080: \end{equation}
2081: Now we easily obtain the evolution of the internal magnetic field, taking the
2082: limit as $n$ goes to $+\infty$ of the quasistatic approximation
2083: \[
2084: h^n(x,t)=h^n_i(x)\,,\ t\in [t^n_i, t^n_{i+1})\,.
2085: \]
2086: 
2087: \begin{theorem}\label{t:limitsol}
2088: If $H_s$ is a nondecreasing function in $[0,T]$, then
2089: the sequence
2090: $(h^n(x,t))$ converges uniformly in $\overline{\Omega} \times [0,T]$ to
2091: \[
2092: h(x,t)=\max[h_0(x), -\distm(x)+H_s(t)]\,.
2093: \]
2094: If $H_s$ is a nonincreasing function in $[0,T]$, then
2095: the sequence
2096: $(h^n(x,t))$ converges uniformly in $\overline{\Omega}\times [0,T]$ to
2097: \[
2098: h(x,t)=\min[h_0(x), \dist(x)+H_s(t)]\,.
2099: \]
2100: \end{theorem}
2101: 
2102: \begin{proof}
2103: Assume that $H_s$ is nondecreasing (the other case being similar). Given $t\in [0,T]$, and $n\in\N$,
2104: we have that $t\in [t^n_i, t^n_{i+1})$, and
2105: \[
2106: 0\leq h(x,t)-h^n(x,t)\leq H_s(t)-H_s(t^n_i) \leq H_s(t^n_{i+1})-H_s(t^n_i)
2107: \leq \delta t^n \max_{t\in [0,T]}H_s'(t)\,,
2108: \]
2109: where $\delta t^n$ is the size of the partition $P^n$. Hence the uniform
2110: convergence in $\overline{\Omega}\times [0,T]$ follows.
2111: \end{proof}
2112: 
2113: %\begin{rk}
2114: %The sequential application of the rules
2115: %given in Theorem \ref{t:limitsol} provides the solution
2116: %if the external field is piecewise monotonic.
2117: %\end{rk}
2118: 
2119: \begin{rk}
2120: Theorem \ref{t:limitsol} shows that the evolution obtained
2121: by the direct optimization method at discrete times proposed by
2122: Bad\'\i a and L\'opez in \cite{BLd} coincides with the one obtained in
2123: \cite{BaPrb,BaPr} by
2124: Barrett and Prigozhin as a solution of an evolutionary
2125: variational inequality.
2126: \end{rk}
2127: 
2128: For what concerns the electric field induced in the superconductor, we already
2129: know that $\vec{E}=(E_1,E_2,0)$ with $(E_1,E_2)\in \partial I_K(Dh)$. This means that
2130: $\vec{E}=0$ whenever $Dh \in \textrm{int}K$, and there exists
2131: $w(x,t)\geq 0$ such that $(E_1,E_2)=w(x,t) D\gauge(Dh(x,t))$
2132: for almost every $x\in \Omega$ and for every $t\in [0,T]$.
2133: Our approach allows us to compute explicitly the quasistatic evolution
2134: of the dissipated power $w(x,t)$ as the dual function appearing in the necessary
2135: conditions stated in Lemma~\ref{l:eul}.
2136: 
2137: Namely, from Lemma~\ref{l:eul}, for every $n\in\N$
2138: and for every $i=1, \ldots, k(n)$, there exists $v^n_{i}\in C_b(\Omega)$
2139: such that
2140: \begin{equation}\label{f:faradisc}
2141: -\dive\left(\frac{v^n_{i}}{\dt}\,D\gauge(Dh^n_{i})\right)=
2142: \frac{h^n_{i-1}-h^n_{i}}{\dt}\,,
2143: \end{equation}
2144: in the sense of distributions in $\Omega$.
2145: For our convenience, we also set $v^n_0 = 0$.
2146: The function $v^n_i$ is unique, as stated in the following result.
2147: 
2148: \begin{prop}\label{p:vni}
2149: Let $v \in C_b(\Omega)$ be a function
2150: such that
2151: \begin{equation}\label{f:vvv}
2152: \begin{cases}
2153: -\dive\left(v\,D\gauge(Dh^n_{i})\right)=
2154: {h^n_{i-1}-h^n_{i}}\,,
2155: &\textrm{in $\Omega$ (distributional)} ,\\
2156: \gauge(D h^n_i) = 1
2157: &\textrm{a.e.\ in $\{v > 0\}$}.
2158: \end{cases}
2159: \end{equation}
2160: If $c^n_i\geq c^n_{i-1}$, then $v=v^n_i$, where
2161: \[
2162: v^n_i(x)=
2163: \begin{cases}
2164: \displaystyle\int_{\distm(x)}^{\len^-(y)}{(h^n_i-
2165: h^n_{i-1})(y+s D\gaugem(\nor(y)))}
2166: M_{x}^-(s)\, ds\,, & x\in \Omega \setminus \overline{\Sigma}^-\,,\ \proj^-(x)=\{y\},\\
2167: 0 & x\in \overline{\Sigma}^-
2168: \end{cases}
2169: \]
2170: whereas, if $c^n_i\leq c^n_{i-1}$, then $v=v^n_i$, where
2171: \[
2172: v^n_i(x)=
2173: \begin{cases}
2174: \displaystyle\int_{\dist(x)}^{\len(y)}{(h^n_{i-1}-
2175: h^n_{i})(y+s D\gauge(\nor(y)))}
2176: M_{x}(s)\, ds\,, & x\in \Omega \setminus \overline{\Sigma}\,,\ \proj(x)=\{y\},\\
2177: 0 & x\in \overline{\Sigma}\,.
2178: \end{cases}
2179: \]
2180: \end{prop}
2181: 
2182: \begin{proof}
2183: Assume that $c^n_{i-1}-c^n_i\geq 0$ (the other case being similar),
2184: and let $v\in C_b(\Omega)$ be a solution of (\ref{f:vvv}).
2185: By construction we have that the function
2186: $f := h^n_{i-1}-h^n_{i}$ is bounded, continuous and nonnegative in $\Omega$,
2187: and $h^n_i = \dist$ in $\spt(f)$.
2188: Using the same arguments of Lemma~6.3 in \cite{CMg}, we can prove that
2189: the pair $(\dist, v)$ satisfies
2190: \[
2191: -\dive\left(v\,D\gauge(D\dist)\right)=f,
2192: \]
2193: in the sense of distributions in $\Omega$.
2194: Now we can apply Propositions~6.5 and 6.7 in \cite{CMg} in order to
2195: conclude that $v = v^n_i$.
2196: \end{proof}
2197: 
2198: Formula (\ref{f:faradisc})
2199: can be understood as a discretized version of Faraday's law, so that
2200: the functions $w^n_{i}=\frac{v^n_{i}}{\dt}$ are the steps of
2201: the quasistatic evolution of the dissipated power.
2202: If we set
2203: \begin{equation}\label{f:gn}
2204: w^n(x,t)=w^n_i(x)\,, \quad
2205: g^n(x,t)= \frac{h^n_{i-1}(x)-h^n_i(x)}{\dt}\,,\qquad t\in [t^n_i, t^n_{i+1})\,
2206: \end{equation}
2207: and $w^n(t)\equiv w^n(\cdot,t)$, $h^n(t)\equiv h^n(\cdot,t)$, $g^n(t)\equiv g^n(\cdot,t)$,
2208: formula (\ref{f:faradisc}) can be rewritten as
2209: \begin{equation}\label{f:faradcon}
2210: -\dive\left(w^n(t)\,D\gauge(Dh^n(t))\right)= g^n(t).
2211: %\frac{h^n(t-\dt)-h^n(t)}{\dt}\,.
2212: \end{equation}
2213: 
2214: \begin{theorem}\label{t:limitpow}
2215: If $H_s$ is a nondecreasing function in $[0,T]$, then
2216: the sequence
2217: $(w^n(x,t))$ converges pointwise in $\overline{\Omega} \times [0,T]$ to
2218: \[
2219: w(x,t)=\frac{\partial H_s}{\partial t}(t)\,\int_{\distm(x)}^{\len^-(y)}
2220: \chi_{\left\{h_0(y+s D\gaugem(\nor(y)))\leq H_s(t)-s\right\}}
2221: M_{x}^-(s)\, ds\,.
2222: \]
2223: If $H_s$ is a nonincreasing function in $[0,T]$, then
2224: the sequence
2225: $(w^n(x,t))$ converges pointwise in $\overline{\Omega}\times [0,T]$ to
2226: \[
2227: w(x,t)=-\frac{\partial H_s}{\partial t}(t)\,\int_{\dist(x)}^{\len(y)}
2228: \chi_{\left\{h_0(y+s D\gauge(\nor(y)))\geq H_s(t)+s\right\}}
2229: M_{x}(s)\, ds\,.
2230: \]
2231: Moreover, in both cases, the sequence $(w^n)$ converges to $w$ in the strong
2232: topology of $L^p(\Omega)$, $p\geq 1$, uniformly in $[0,T]$.
2233: \end{theorem}
2234: 
2235: \begin{proof}
2236: We compute the limit of the
2237: sequence $(w^n)$ only in the case of nondecreasing external field, the other
2238: case being similar.
2239: 
2240: We can assume, without loss of generality, that for every
2241: $n\in\N$ the partition $P^{n+1}$ is a refinement of the
2242: partition $P^n$. Hence, for every $t\in [0,T]$ and for every
2243: $n\in\N$ there exists $i=i(n)\in\N$ such that $t\in [t^n_{i(n)},t^n_{i(n)+1})$.
2244: Moreover, since $H_s\in C^1([0,T])$, we obtain
2245: \begin{equation}\label{f:unifg}
2246: -\frac{\partial H_s}{\partial t}(t)=\lim_{n\to \infty}
2247: \frac{H_s(t^n_{i-1})-H_s(t^n_{i})}{\dt} \qquad
2248: \textrm{uniformly in\ }[0,T]\,.
2249: \end{equation}
2250: By Proposition \ref{p:vni} we get
2251: \[
2252:  w^n(x,t)=
2253: \int_{\distm(x)}^{\len^-(y)}g^n(y+s D\gaugem(\nor(y)),t)
2254: M_{x}^-(s)\, ds\,,
2255: \]
2256: where $g^n$ is the function defined in (\ref{f:gn}).
2257: Moreover we have
2258: \begin{equation}\label{f:unifg2}
2259: \begin{split}
2260: h^n_{i-1} - h^n_i & =(H_s(t^n_{i-1})-H_s(t^n_{i}))
2261: \chi_{\left\{h_0 \leq H_s(t^n_{i-1})-\distm\right\}}
2262: \\ & +
2263: (h_0-H_s(t^n_{i})+\distm)
2264: \chi_{\left\{H_s(t^n_{i-1})-\distm < h_0 < H_s(t^n_{i})-\distm\right\}}\,,
2265: \end{split}
2266: \end{equation}
2267: and
2268: \[
2269: \begin{split}
2270: 0 & \geq (h_0-H_s(t^n_{i})+\distm)
2271: \chi_{\left\{H_s(t^n_{i-1})-\distm < h_0 < H_s(t^n_{i})-\distm\right\}}\\
2272: & \geq \left(H_s(t^n_{i-1})-H_s(t^n_{i})\right)
2273: \chi_{\left\{H_s(t^n_{i-1})-\distm < h_0 < H_s(t^n_{i})-\distm\right\}}\,.
2274: \end{split}
2275: \]
2276: Collecting the previous information, we get
2277: \begin{equation}\label{f:plim1}
2278: \lim_{n\to +\infty}g^n(t) = g(t):=
2279: -\frac{\partial H_s}{\partial t}(t)\,\chi_{\left\{h_0\leq H_s(t)-\distm\right\}}\,,
2280: \end{equation}
2281: and $\lim_{n\to +\infty}\|g^n(t)-g(t)\|_{L^p(\Omega)}=0$ uniformly in $[0,T]$, for
2282: every $p\geq 1$. In addition, from (\ref{f:unifg}) and (\ref{f:unifg2}),
2283: for every regular point $x$ of $\distm$ and setting $\{y\}=\Pi(x)$, we get
2284: \begin{equation}\label{f:plim3}
2285: \lim_{n\to +\infty}g^n(y+sD\gaugem(\nor(y)), t)= g(y+sD\gaugem(\nor(y)), t)\qquad
2286: \textrm{a.e.\ }s\in [\distm(x), \len^-(y)]\,.
2287: \end{equation}
2288: %Concerning $w^n$, by Lemma~\ref{l:eul} we have the representation formula
2289: %\[
2290: %v^n_i(x)=-\int_{\distm(x)}^{\lambda^n_i(y)} \left[h^n_{i-1}(y+sD\gaugem(\nor(y)))
2291: %-c_i^n+s\right]M_{x}^-(s)\, ds\,,
2292: %\]
2293: %where $y$ is the projection of $x$ on $\partial \Omega$, and
2294: %\[
2295: %\lambda^n_i(y)= \sup\{\tau\in [0,\len(y))\colon h^n_{i-1}(y+s D\gaugem(\nor(y)))-c^n_i<-s\
2296: %\forall s \in [0,\tau)\}\,.
2297: %\]
2298: %We claim that $s>\lambda^n_i(\macroy )$ if and only if
2299: %$h^n_i(\macroy +s D\gaugem(\nor(\macroy ))=h_0(\macroy +s D\gaugem(\nor(\macroy ))$.
2300: %Namely, if $s>\lambda^n_i(\macroy )$,
2301: %then
2302: %\[
2303: %h^n_{i-1}(\macroy +s D\gaugem(\nor(\macroy ))>c^n_i-s\geq c^n_{i-1}-s\,,
2304: %\]
2305: %so that
2306: %\[
2307: %h^n_{i-1}(\macroy +s D\gaugem(\nor(\macroy ))=h_0(\macroy +s D\gaugem(\nor(\macroy ))\,,
2308: %\]
2309: %and
2310: %\[
2311: %h_0(\macroy +s D\gaugem(\nor(\macroy ))>c^n_i-s\,.
2312: %\]
2313: %{}From Lemma~\ref{l:opm} we have that,
2314: %if $\lambda^n_i(y) < \len(y)$, then $s\in [\lambda^n_i(y),\len(y)]$
2315: %if and only if
2316: %$h^n_i(y+s D\gaugem(\nor(y))=h_0(y+s D\gaugem(\nor(y))$.
2317: %Hence
2318: %\[
2319: %\begin{split}
2320: %h^n_i(y+s D\gaugem(\nor(y)))& =\max[h_0(y+s D\gaugem(\nor(y))), -s+c^n_i] \\
2321: %& =
2322: %h_0(y+s D\gaugem(\nor(y)))\,.
2323: %\end{split}
2324: %\]
2325: %On the other hand, if $s<\lambda^n_i(y)$, then
2326: %\[
2327: %h_0(y+s D\gaugem(\nor(y)))\leq h^n_{i-1}(y+s D\gaugem(\nor(y)))<c^n_i-s\,,
2328: %\]
2329: %and hence $h^n_i(y+s D\gaugem(\nor(y)))=c^n_i-s$.
2330: %Then we have
2331: %\[
2332: %\int_{\lambda^n_i(y)}^{\len(y)}\left[h^n_i(y+s D\gaugem(\nor(y)))-
2333: %h^n_{i-1}(y+s D\gaugem(\nor(y)))\right]
2334: %M_{y}^-(s)\, ds=0\,,
2335: %\]
2336: %and
2337: %\[
2338: %\begin{split}
2339: %& \int_{\distm(x)}^{\lambda^n_i(y)}\left[h^n_i(y+s D\gaugem(\nor(y)))-
2340: %h^n_{i-1}(y+s D\gaugem(\nor(y)))\right]
2341: %M_{y}^-(s)\, ds \\
2342: %& =
2343: %\int_{\distm(x)}^{\lambda^n_i(y)}\left[
2344: %c^n_i-s-h^n_{i-1}(y+s D\gaugem(\nor(y)))\right]
2345: %M_{y}^-(s)\, ds\,.
2346: %\end{split}
2347: %\]
2348: %\[
2349: %w^n_i(x)=\frac{v^n_i(x)}{\dt}=\int_{\distm(x)}^{\len(y)}\frac{h^n_i(y+s D\gaugem(\nor(y)))-
2350: %h^n_{i-1}(y+s D\gaugem(\nor(y)))}{\dt}
2351: %M_{x}^-(s)\, ds\,.
2352: %\]
2353: Finally,
2354: by (\ref{f:plim3}), (\ref{f:estiM}) and the Dominated Convergence Theorem, we conclude that
2355: \begin{equation}\label{f:plim2}
2356: \begin{split}
2357: &\lim_{n \to +\infty} w^n(x,t)=\\
2358: & -\frac{\partial H_s}{\partial t}(t)\,\int_{\distm(x)}^{\len(y)}
2359: \chi_{\left\{h_0(y+s D\gaugem(\nor(y)))\leq H_s(t)-s\right\}}
2360: M_{x}^-(t)\, ds\,:=w(x,t)\,.
2361: \end{split}
2362: \end{equation}
2363: The last part of the theorem follows from the fact that
2364: \begin{equation}\label{f:plimw}
2365: {\|w^n(t)-w(t)\|}_{L^p(\Omega)}\leq
2366: C_0\, {\|g^n(t)-g(t)\|}_{L^p(\Omega)}
2367: \end{equation}
2368: for some positive constant $C_0$.
2369: Namely, for a given $t\in [0,T]$ let us define
2370: \[
2371: \varphi^n(x, s) := g^n(y+sD\gaugem(\nor(y)), t) - g(y+sD\gaugem(\nor(y)), t),
2372: \]
2373: for $x\in\Omega\setminus\overline{\Sigma}$,
2374: $\proj(x)=\{y\}$, and $s\in [0,\len^-(y)]$.
2375: We remark that, if $x\in\Omega\setminus\overline{\Sigma}$ and
2376: $y$ is the projection of $x$,
2377: then $\varphi^n(x,\cdot) = \varphi^n(y,\cdot)$.
2378: 
2379: By the very definition of $w^n$, $w$ and $M_x$, H\"older's inequality and
2380: (\ref{f:estiM}) we have that
2381: \[
2382: {\|w^n(t)-w(t)\|}^p_{L^p(\Omega)}\leq
2383: C \int_{\Omega} \left(
2384: \int_{\distm(x)}^{\len^-(x)} |\varphi^n(x,s)|^p
2385: \prod_{i=1}^{n-1}\frac{1-s\curvgm_i(x)}{1-\distm(x)\curvgm_i(x)}
2386: \, ds
2387: \right)dx =: I
2388: \]
2389: where $\len^-(x):= \len^-(y)$ when $\proj^-(x) = \{y\}$ and
2390: $C := M_0^{p-1}\, (\max\distm)^{1/p'}$.
2391: Using the change of variables theorem (see \cite[Thm.~7.1]{CMf}) we
2392: have that
2393: \[
2394: \begin{split}
2395: I = {} &
2396: C \int_{\partial\Omega} \gaugem(\nor(y)) \left[\int_0^{\len^-(y)}
2397: \left(\int_{\sigma}^{\len^-(y)} |\varphi^n(y,s)|^p
2398: \prod_{i=1}^{n-1}\frac{1-s\curvgm_i(y)}{1-\sigma\curvgm_i(y)}
2399: \, ds \right) \right.\cdot
2400: \\ & {} \cdot
2401: \left.\prod_{i=1}^{n-1} (1-\sigma\, \curvg^-_i(y))\, d\sigma\right]
2402: d\haus(y)
2403: \\ = {} &
2404: C \int_{\partial\Omega} \gaugem(\nor(y)) \left[\int_0^{\len^-(y)}
2405: \left(\int_{\sigma}^{\len^-(y)} |\varphi^n(y,s)|^p
2406: \prod_{i=1}^{n-1}(1-s\curvgm_i(y))
2407: \, ds \right) \, d\sigma\right]
2408: d\haus(y)
2409: \\ \leq {} &
2410: C' \int_{\partial\Omega} \gaugem(\nor(y)) \left(\int_0^{\len^-(y)}
2411: |\varphi^n(y,s)|^p
2412: \prod_{i=1}^{n-1}(1-s\curvgm_i(y))
2413: \, ds \right)
2414: d\haus(y),
2415: \end{split}
2416: \]
2417: where $C' := C\, \max\distm$.
2418: Using again the change of variables theorem we finally get
2419: \[
2420: %I \leq C' \int_{\Omega} |\varphi^n(x,s)|^p\, dx,
2421: I \leq C' \int_{\Omega} |g^n(x,t)-g(x,t)|^p\, dx,
2422: \]
2423: hence (\ref{f:plimw}) follows.
2424: \end{proof}
2425: 
2426: \begin{rk}
2427: Notice that if $\distm(x)>\lambda^n_i(y)$, then $v^n_i(x)=0$, while
2428: $h^n_i(x)=c^n_i-\distm(x)$ if $\distm(x)\leq \lambda^n_i(y)$. Hence we have that
2429: $w^n(x,t) D\gauge(Dh^n(x,t))=w^n(x,t) D\gauge(-D\distm(x))$ a.e.\ in $\Omega$ and
2430: for every $t\in [0,T]$.
2431: A passage to the limit in (\ref{f:faradcon})
2432: %and the fact that $w(x,t) D\gauge(Dh(x,t))=w^n(x,t) D\gauge(-D\distm(x))$
2433: %a.e.\ in $\Omega$ and for every $t\in [0,T]$
2434: leads to
2435: \begin{equation}\label{f:faradfin}
2436: -\dive(w(x,t)D\gauge(Dh(x,t)))=- \frac{\partial H_s}{\partial t}
2437: \,\chi_{\left\{h_0\leq H_s(t)-\distm\right\}}\,,
2438: \end{equation}
2439: in the sense of distributions in $\Omega$ for every $t\in [0,T]$.
2440: Recalling that, by construction, $\frac{\partial H_s}{\partial t}
2441: \,\chi_{\left\{h_0\leq H_s(t)-\distm\right\}}= \frac{\partial h}{\partial t}(x,t)$, and
2442: comparing (\ref{f:faradfin}) with the Faraday's law, we conclude
2443: that the electric field induced by $h$ inside the superconductor is
2444: $\vec{E}(x,t)=w(x,t)D\gauge(Dh(x,t))$.
2445: \end{rk}
2446: 
2447: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2448: \section{Further remarks}
2449: %{Mathematical counterpart of the experimental observations}
2450: 
2451: As a consequence of the Theorem \ref{t:limitsol} we obtain some
2452: detailed information about the macroscopic behavior of the
2453: penetrating magnetic field.
2454: 
2455: \begin{center}
2456: \begin{figure}
2457: %\raisebox{2.5cm}{%
2458: \includegraphics[width=5cm]{\picturename{ECdistbw-crop}}\qquad
2459: \includegraphics[width=5cm]{\picturename{ECvbw-crop}}
2460: \includegraphics[width=5cm]{\picturename{ECdist3-crop}} \qquad
2461: \includegraphics[width=5cm]{\picturename{ECv3-crop}}
2462: \caption{Level sets and 3D plot of the full penetrated magnetic
2463: field  and of the dissipation for $\Omega$ Cassini's Egg, $K$ ellipse not centered at the origin,
2464: $h_0=0$}\label{BB3}
2465: \end{figure}
2466: \end{center}
2467: 
2468: \noindent\textit{Saturation time (Full penetration time).}
2469: The experiments show that, if the superconductor is posed in an
2470: increasing external field $H_s$, say $H_s(t) \to +\infty$ as
2471: $t \to +\infty$, after a while the internal magnetic field $h(x,t)$ differs
2472: from $H_s(t)$ for a stationary amount $u(x)$.
2473: In our model we obtain that if $\tau \geq 0$ is such that
2474: $H_s(\tau)=\sup_\Omega (h_0+\distm)$, then for $t\geq \tau$
2475: $h(x,t)-H_s(t)=-\distm(x)$.
2476: 
2477: In the simpler case when $H_s(t)=-at$, $a>0$, and
2478: $h_0=0$ we obtain that at time $\tau=\frac{\max_\Omega \dist}{a}$
2479: the magnetic field is penetrated in the whole superconductor and
2480: $h(x,t)=\dist(x)+H_s(t)$ for $t\geq \tau$.
2481: Figures \ref{BB3} and  \ref{BB4} picture
2482: the plots of the magnetic field and of the dissipation in the
2483: stationary configuration for anisotropic materials with a
2484: complicated geometry. We underline that
2485: the numerical computation based on the representation formula
2486: (\ref{f:plim2}) shows the presence of an hight dissipation near
2487: the points of the boundary with negative curvature, according to the
2488: experimental observations.
2489: 
2490: \begin{center}
2491: \begin{figure}
2492: \includegraphics[width=4.5cm]{\picturename{EFdistbw-crop}}\qquad\qquad
2493: \includegraphics[width=4.5cm]{\picturename{EFvbw-crop}}
2494: \includegraphics[width=5cm]{\picturename{EFdist3-crop}}\qquad
2495: \includegraphics[width=5cm]{\picturename{EFv3-crop}}
2496: \caption{Level sets and 3D plot of the full penetrated magnetic
2497: field and of the dissipation for $\Omega$ perturbed disk, $K$ ellipse not centered at the origin,
2498: $h_0=0$.}
2499: \label{BB4}
2500: \end{figure}
2501: \end{center}
2502: %\begin{center}
2503: %\begin{figure}
2504: %\includegraphics[height=4cm]{EFvbw-crop.pdf}\qquad \includegraphics[height=4cm]{EFv3-crop.pdf}
2505: %\caption{Level sets and 3D plot of the power in the situation of Fig. \ref{BB4}}\label{BB$a}
2506: %\end{figure}
2507: %\end{center}
2508: 
2509: \noindent\textit{Boundary of penetration profile.}
2510: If the evolution starts with $h_0=0$ (no initial magnetic
2511: field inside $\Omega$) and the external field is nondecreasing,
2512: then at time $t$ the induced magnetic
2513: field is penetrated in $\Omega$ only in the region
2514: $\{x\in \Omega \colon\ \distm(x)\leq H_s(t)\}$. This gives an
2515: exact macroscopic description of the boundary of penetration profile,
2516: which is given by the level sets of the Minkowski distance from
2517: the boundary of $\Omega$.
2518: 
2519: \begin{center}
2520: \begin{figure}
2521: \includegraphics[width=4cm]{\picturename{hyst0-crop}}\qquad
2522: \includegraphics[width=4cm]{\picturename{hyst1-crop}}
2523: \includegraphics[width=4cm]{\picturename{hyst2-crop}}\qquad
2524: \includegraphics[width=4cm]{\picturename{hyst3-crop}}
2525: \includegraphics[width=4cm]{\picturename{hyst4-crop}} \qquad
2526: \includegraphics[width=4cm]{\picturename{hyst5-crop}}
2527: \caption{$\Omega$ perturbed disk, $K$ ellipse centered at the origin. Profile of the internal magnetic field,
2528: starting from $h_0=0$, during a loop of the external field.}\label{BB5}
2529: \end{figure}
2530: \end{center}
2531: 
2532: \noindent\textit{Hysteresis phenomenon.}
2533: Assume that the evolution starts with $h_0=0$ (no initial magnetic
2534: field inside $\Omega$) and that the external field $H_s(t)$ is increasing
2535: in [0,T], then at time $T$ the induced magnetic
2536: field is $h(x,T)=(-\distm(x)+H_s(t))_+$.
2537: Starting from this configuration,
2538: in [T,2T] the superconductor is subject to the external field
2539: $H_s(2T-t)$. Then we have
2540: $h(x,2T)=\min\{(-\distm(x)+H_s(T))_+, \dist(x)+H_s(0)\}$ which in general
2541: is not zero. This corresponds with the experimental observation of
2542: an hysteresis phenomenon for these hard superconductors  after an external field loop.
2543: \begin{center}
2544: \begin{figure}
2545: \includegraphics[width=4cm]{\picturename{hyst-cycle}}
2546: \caption{Magnetic hysteresis loop corresponding to the previous process.}
2547: \end{figure}
2548: \end{center}
2549: 
2550: \noindent\textit{On the validity of the Badi\'a--L\'opez formalism for anisotropic
2551: materials.} In \cite{Fisher} Fisher doubts the validity of the variational
2552: formulation proposed by Bad\'ia and Lopez in the case of anisotropic
2553: materials. He declares that \textit{``one cannot be sure that the suggested
2554: least action principle chooses the Current--Voltage Characteristics correctly.
2555: Of course, if this principle is proved, it would be a very important tool
2556: to determine the Current--Voltage Characteristics via the known restriction
2557: region''}. Our results state that, if the power--law approximation for the dissipation
2558: is assumed to be reliable, then the Current--Voltage Characteristics
2559: induced by this approximation is the one selected by the minimization procedure.
2560: 
2561: 
2562: 
2563: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2564: % bibliography
2565: %
2566: %\bibliographystyle{/Ga/BibTeX/mybst}
2567: %\bibliography{/Ga/BibTeX/Graziano,/Ga/BibTeX/Ricerca,/Ga/BibTeX/Physics}
2568: 
2569: \begin{thebibliography}{10}
2570: 
2571: \bibitem{BLd}
2572: {A.}~Bad\'\i a and {C.} L\'opez, \emph{Vector magnetic hysteresis of hard
2573:   superconductors}, Phys.\ Rev.~B \textbf{65} (2001), 104514.
2574: 
2575: \bibitem{BLc}
2576: {A.}~Bad\'\i a and {C.} L\'opez, \emph{Horizons in Superconductivity Research},
2577:   ch.~Minimal model for the topology of the critical state in hard
2578:   superconductors, Nova Science Publishers, 2003.
2579: 
2580: \bibitem{BLb}
2581: {A.}~Bad\'\i a and {C.} L\'opez, \emph{Electric field in hard superconductors
2582:   with arbitrary cross section and general critical current law}, J.\ Appl.\
2583:   Phys. \textbf{95} (2004), 8035--8040.
2584: 
2585: \bibitem{BaCD}
2586: {M.} Bardi and {I.} Capuzzo~Dolcetta, Optimal Control and Viscosity Solutions
2587:   of {H}amilton-{J}acobi-{B}ellman Equations, Systems \& Control: Foundations
2588:   \& Applications, Birkh\"auser, Boston, 1997.
2589: 
2590: \bibitem{BaPrb}
2591: {J.W.} Barrett and {L.} Prigozhin, \emph{Sandpiles and superconductors: dual
2592:   variational formulations for critical-state problems}, preprint, 2005.
2593: 
2594: \bibitem{BaPr}
2595: {J.W.} Barrett and {L.} Prigozhin, \emph{{B}ean's critical-state model as the
2596:   $p\to\infty$ limit of an evolutionary $p$-{L}aplacian equation}, Nonlinear
2597:   Anal. \textbf{42} (2000), 977--993.
2598: 
2599: \bibitem{Bean}
2600: {C.P.} Bean, \emph{Magnetization of hard superconductors}, Phys.\ Rev.\ Letters
2601:   \textbf{8} (1962), 250--253.
2602: 
2603: \bibitem{BKR}
2604: {K.V.} Bhagwat, {D.} Karmakar, and {G.} Ravikumar, \emph{Critical state model
2605:   with anisotropic critical current density}, J.\ Phys.: Condens.\ Matter
2606:   \textbf{15} (2003), 1325--1337.
2607: 
2608: \bibitem{Brai}
2609: {A.} Braides, $\Gamma$-convergence for beginners, Oxford University Press, New
2610:   York, 2002.
2611: 
2612: \bibitem{Bran}
2613: {E.H.} Brandt, \emph{Electric field in superconductors with rectangular cross
2614:   section}, Phys.\ Rev.~B \textbf{52} (1995), 15442--15457.
2615: 
2616: \bibitem{Butt}
2617: {G.} Buttazzo, Semicontinuity, relaxation and integral representation in the
2618:   calculus of variations, Pitman Res.\ Notes Math.\ Ser., vol. 207, Longman
2619:   Scientific and Technical, Harlow, U.K., 1989.
2620: 
2621: \bibitem{CaSi}
2622: {P.} Cannarsa and {C.} Sinestrari, Semiconcave functions, {H}amilton-{J}acobi
2623:   equations and optimal control, Progress in Nonlinear Differential Equations
2624:   and their Applications, vol.~58, Birkh\"auser, Boston, 2004.
2625: 
2626: \bibitem{Cha}
2627: {S.J.} Chapman, \emph{A Hierarchy of Models for Type-{II} Superconductors},
2628:   SIAM Rev. \textbf{42} (2000), 555--598.
2629: 
2630: \bibitem{CMf}
2631: {G.} Crasta and {A.} Malusa, \emph{The distance function from the boundary in a
2632:   {M}inkowski space}, to appear in Trans.\ Amer.\ Math.\ Soc.,
2633:   oai:ar{X}iv:math.AP/0612226.
2634: 
2635: \bibitem{CMg}
2636: {G.} Crasta and {A.} Malusa, \emph{On a system of partial differential
2637:   equations of {M}onge-{K}antorovich type}, to appear in J.\ Differential
2638:   Equations, oai:ar{X}iv:math.AP/0612227.
2639: 
2640: \bibitem{DM}
2641: {G.} Dal~Maso, An introduction to $\Gamma$--convergence, Birkh\"auser, Boston,
2642:   1993.
2643: 
2644: \bibitem{Fisher}
2645: {L.M.} Fisher and {V.A.} Yampol'skii, \emph{Comment on ``Critical statetheory
2646:   for non-parallel flux line lattices in type-{II} superconductors''},
2647:   oai:ar{X}iv:cond-mat/0201286, 2002.
2648: 
2649: \bibitem{GNP}
2650: {A.} Garroni, {V.} Nesi, and {M.} Ponsiglione, \emph{Dielectric breakdown:
2651:   optimal bounds}, Proc.\ R.\ Soc.\ Lond.\ A \textbf{457} (2001), 2317--2335.
2652: 
2653: \bibitem{Gold}
2654: {S.I.} Goldberg, Curvature and homology, Academic Press, New York, 1970.
2655: 
2656: \bibitem{LN}
2657: {Y.Y.} Li and {L.} Nirenberg, \emph{The distance function to the boundary,
2658:   {F}insler geometry and the singular set of viscosity solutions of some
2659:   {H}amilton--{J}acobi equations}, Commun.\ Pure Appl.\ Math. \textbf{58}
2660:   (2005), 85--146.
2661: 
2662: \bibitem{Li}
2663: {P.L.} Lions, Generalized solutions of {H}amilton-{J}acobi equations, Pitman,
2664:   Boston, 1982.
2665: 
2666: \bibitem{Rock}
2667: {R.T.} Rockafellar, Convex Analysis, Princeton Univ.\ Press, Princeton, NJ,
2668:   1970.
2669: 
2670: \bibitem{Sch}
2671: {R.} Schneider, Convex bodies: the {B}runn--{M}inkowski theory, Cambridge
2672:   Univ.~Press, Cambridge, 1993.
2673: 
2674: \bibitem{Wa}
2675: {W.} von Wahl, \emph{Estimating $\nabla u$ by div${}\,u$ and curl${}\,u$},
2676:   Math.\ Methods Appl.\ Sci. \textbf{15} (1992), 123--143.
2677: 
2678: \bibitem{YLZ}
2679: {H.-M.} Yin, {B.Q.} Li, and {J.} Zou, \emph{A degenerate evolution system
2680:   modeling {B}ean's critical-state type-{II} superconductors}, Discrete
2681:   Contin.\ Dynam.\ Systems \textbf{8} (2002), 781--794.
2682: 
2683: \end{thebibliography}
2684: 
2685: \end{document}
2686: