1: \documentclass[12pt,twoside]{amsart}
2:
3: \vfuzz4pt % Don't report over-full v-boxes if over-edge is small
4: \hfuzz5pt % Don't report over-full h-boxes if over-edge is small
5:
6:
7: \baselineskip=12pt
8:
9: \evensidemargin 0in
10: \oddsidemargin 0in
11: \topmargin -.7cm
12: \textheight 23.5cm
13: \textwidth 16.3cm
14: \parindent=0cm
15:
16:
17: \usepackage{amsmath,comment}
18: \usepackage{amssymb}
19:
20: \numberwithin{equation}{section}
21:
22: \newtheorem{thm}{Theorem}[section]
23: \newtheorem{pro}[thm]{Proposition}
24: \newtheorem{lemma}[thm]{Lemma}
25: \newtheorem{cor}[thm]{Corollary}
26: \newtheorem{defi}[thm]{Definition}
27: \newtheorem{preremark}[thm]{Remark}
28: \newenvironment{remark}{\begin{preremark}\rm}{\end{preremark}}
29: \thispagestyle{empty}
30:
31: \def\N{{\mathbb N}}
32: \def\Z{{\mathbb Z}}
33: \def\R{{\mathbb R}}
34: \def\a{{\alpha}}
35:
36: \newcommand{\CVD}{\hfill $\rule{2.6mm}{2.6mm}$}
37: \newcommand{\PF}{\noindent{\bf Proof. }}
38:
39: \title[Quasilinear phase transitions]{
40: Rigidity results for some\\
41: boundary quasilinear phase transitions}
42:
43: \author[Y. Sire]{Yannick Sire}
44: \author[E. Valdinoci]{Enrico Valdinoci}
45:
46:
47: \begin{document}
48: \begin{abstract}
49: We consider a quasilinear equation
50: given in the half-space, i.e. a so called
51: boundary reaction problem. Our concerns are a geometric Poincar\'e inequality
52: and, as a byproduct of this inequality, a result on the symmetry of
53: low-dimensional
54: bounded stable solutions, under some suitable assumptions on the nonlinearities.
55: More precisely, we analyze the following boundary problem
56: $$
57: \left\{
58: \begin{matrix}
59: -{\rm div}\, (a(x,|\nabla u|)\nabla u)+g(x,u)=0 \qquad
60: {\mbox{ on $\R^n\times(0,+\infty)$}}
61: \\
62: -a(x,|\nabla u|)u_x = f(u)
63: \qquad{\mbox{ on $\R^n\times\{0\}$}}\end{matrix}
64: \right.$$
65: under some natural assumptions on the diffusion coefficient
66: $a(x,|\nabla u|)$ and the nonlinearities $f$ and $g$.
67:
68: Here, $u=u(y,x)$, with~$y\in\R^n$ and~$x\in(0,+\infty)$.
69: This type of PDE can be seen as a nonlocal problem on the boundary
70: $\partial \R^{n+1}_+$. The assumptions on $a(x,|\nabla u|)$ allow to treat in a unified way the $p-$laplacian and the minimal surface operators.
71: \end{abstract}
72:
73: \maketitle
74: \tableofcontents
75:
76: \bigskip\bigskip
77:
78: \noindent{\em Keywords:} Boundary reactions,
79: Allen-Cahn phase transitions,
80: $p-$laplacian, minimal surface operator, quasilinear equations,
81: Poincar\'e-type inequality.
82: \bigskip
83:
84: \noindent{\em 2000 Mathematics Subject Classification:}
85: 35J70, 35J65, 47G30, 35B45.
86: \bigskip\bigskip
87:
88: \section{Introduction}
89:
90: The purpose of this paper is to give some geometric results on the following problem:
91: \begin{equation}\label{eq1-provv}
92: \left\{
93: \begin{matrix}
94: -{\rm div}\, (a(x,|\nabla u|)\nabla u)+g(x,u)=0 \qquad
95: {\mbox{ on $\R^n\times(0,+\infty)$}}
96: \\
97: - a(x,|\nabla u|) u_x = f(u)
98: \qquad{\mbox{ on $\R^n\times\{0\}$.}}\end{matrix}
99: \right.
100: \end{equation}
101: Here,
102: $u=u(y,x)$, with~$y\in\R^n$ and~$x\in(0,+\infty)$.
103: Equation \eqref{eq1-provv} is a boundary problem. This type of system is a model for nonlocal
104: operators. For instance, when
105: $g=0$ and $a(x,|\nabla u|)=x^\alpha $ with $\alpha \in (-1,1)$, it
106: has been proved by~\cite{cafS}
107: that the Dirichlet-to-Neumann operator
108: $$\Gamma: u|_{\partial \R^{n+1}_+} \mapsto -x^\alpha u_x|_{\partial \R^{n+1}_+}$$
109: is the fractional laplacian $(-\Delta)^{\frac{1-\alpha}{2}}$.
110: In \cite{SV}, a symmetry result for bounded stable solutions of
111: semilinear equations involving this operator was given.
112:
113: Unfortunately, a theory describing the boundary operator for problem \eqref{eq1-provv} is not yet
114: available. However, in virtue of the results
115: by~\cite{cafS}, one could
116: interpret the operator on the boundary as a nonlocal quasilinear operator.
117:
118:
119: In this paper,
120: we develop
121: a geometric
122: analysis of the level sets
123: of stable solutions
124: of~\eqref{eq1-provv} and
125: we prove
126: a symmetry result
127: inspired by a conjecture
128: of De Giorgi~\cite{DeG}.
129:
130: We
131: want to give a geometric insight of the phase
132: transitions for equation~\eqref{eq1-provv}. Our goal
133: is to give a geometric
134: proof of the one-dimensional symmetry result
135: for boundary reactions in dimension~$n=2$,
136: inspired by
137: De Giorgi conjecture and
138: in the spirit of the proof of
139: Bernstein Theorem given in~\cite{giusti} and
140: applied in the case of boundary
141: reactions in \cite{SV}.
142:
143:
144: We focus on problem \eqref{eq1-provv} under the following structural assumptions (denoted $(S)$):
145: \begin{itemize}
146:
147: \item The function $a$ maps $(0,+\infty) \times
148: (0,+\infty)$ into $(0,+\infty)$ and
149: $$ \lim_{t\rightarrow 0^+} t a(.,t)=0.$$
150: \item The map $t \mapsto a(.,t)$ is $C^1(0,+\infty)$ and
151: \begin{equation}\label{BCOMEA}
152: t| a_t(x,t)| \leq C a(x,t)\end{equation}
153: for any~$x$, $t>0$,
154: for some constant $C>0$.
155: \item The map $x \mapsto a(x,.)$ is in
156: $L^1((0,r))$, for any $r>0$ and bounded over all open sets compactly
157: contained in $\R^{n+1}_+$, i.e. for all $K \Subset
158: \R^{n+1}_+$,
159: there exists $\mu_1$,
160: $\mu_2>0$, possibly depending on $K$,
161: such that $\mu_1 \leq
162: a(x,t) \leq \mu_2$, for any $x\in K$ and for $0<t\leq M$.
163:
164: Also, the function~$x \mapsto a(x,.)$ is an $A_2$-Muckenhoupt
165: weight, that is, there exists $\kappa>0$
166: such that
167: \begin{equation}\label{Muck}
168: \int_c^d a(x,t)\,dx \,
169: \int_c^d \frac{1}{a(x,t)}\,dx\,\le\,\kappa(d-c)^2
170: \end{equation}
171: for any $d\ge c\ge 0$ and for all $0<t\leq M$.
172:
173:
174: \item The map $(0,+\infty)\ni x\mapsto g(x,0)$
175: belongs to $L^\infty((0,r))$ for any $r>0$.
176: Also,
177: for any $x>0$, the map $\R\ni u\mapsto g(x,u)$
178: is locally Lipschitz, and given any $R$, $M>0$
179: there exists $C>0$, possibly depending on $R$ and $M$
180: in such a way that
181: \begin{equation}\label{8ikeoqoqoqoo78}
182: \sup_{{0<x<R}\atop{|u|<M}}|g_u(x,u)|\le C.
183: \end{equation}
184: \item The function
185: $f$ is locally
186: Lipschitz in
187: $\R$.
188: \end{itemize}
189:
190:
191: Equation \eqref{eq1-provv} may be understood
192: in the weak sense, namely supposing that $u\in
193: L^\infty_{\rm loc}(\overline{\R^{n+1}_+})$, with
194: \begin{equation}\label{hgasj7717177}
195: a(x,|\nabla u|)|\nabla u|^2 \in L^1 (B_R^+)
196: \end{equation}
197: for any $R>0$,
198: and that\footnote{Condition \eqref{hgasj7717177}
199: is assumed here to make sense of \eqref{eq1}.
200: We will see in the forthcoming Lemma \ref{Daf} that it is
201: always uniformly fulfilled when $u$ is bounded and for a weight $a$ satisfying natural structural assumptions.
202:
203: The structural
204: assumptions on $g$ may be easily
205: checked when $g(x,u)$ has the product-like
206: form of $g^{(1)}(x) g^{(2)}(u)$.}
207: \begin{equation}\label{eq1}
208: \int_{{\R^{n+1}_+}}
209: a(x,|\nabla u|) \nabla u\cdot
210: \nabla\xi+\int_{{\R^{n+1}_+}} g(x,u)\, \xi=
211: \int_{\partial {\R^{n+1}_+}}
212: f(u)\xi
213: \end{equation}
214: for any $\xi:B_R ^+\rightarrow \R$ which is bounded, locally
215: Lipschitz in the interior of
216: $\R^{n+1}_+$,
217: which
218: vanishes on $\R^{n+1}_+\setminus B_R$ and such that
219: \begin{equation}\label{hgasj7717177-bis}
220: a(x,|\nabla u|)|\nabla\xi|^2
221: \in L^1 (B_R^+).\end{equation}
222:
223: As usual,
224: we are using here the notation~$B_R^+:= B_R
225: \cap\R^{n+1}_+$.
226:
227: Consider now the
228: map $\mathcal{B}: \R^+ \times \R^{n+1} \backslash \left \{ 0 \right \}
229: \rightarrow {\rm{Mat}}
230: ((n+1)\times (n+1))$ defined by
231: \begin{equation}
232: \label{BDE}
233: \mathcal{B}(x,\eta)_{ij}:=
234: a(x,|\eta|)\delta_{ij}+\frac{a_t(x,|\eta|)}{|\eta|} \eta_i \eta_j
235: \end{equation}
236: for any $1 \leq i,j \leq n+1$,
237: where $a_t$ stands for the derivative of $a(x,t)$ with respect to
238: its second variable.
239:
240: A direct computation gives
241: \begin{equation}\label{6bis}
242: \frac{d}{d\varepsilon}
243: a(x,|\nabla u +\varepsilon
244: \nabla \varphi|)(\nabla u +\varepsilon
245: \nabla \varphi)\cdot \nabla \varphi |_{\varepsilon=0}=<\mathcal{B}
246: (x,\nabla u) \nabla \varphi,\nabla \varphi>
247: \end{equation}
248: for any smooth test function $\varphi$, any function $u$ with
249: nonvanishing gradient and where $<,>$ stands for the canonical inner product in $\R^{n+1}$.
250:
251: Inspired by~\eqref{6bis}, it is tempting to say that~$u$ is
252: stable
253: if
254: \begin{equation}\label{sta1}
255: \begin{array}{c}
256: \int_{B_R^+} <\mathcal{B}(x,\nabla u)\nabla \xi, \nabla \xi>+
257: \int_{B_R^+} g_u(x,u)\xi^2
258: -\int_{\partial B_R^+}
259: f'(u)\xi^2\,\ge\,0
260: \end{array}
261: \end{equation}
262: for any $\xi$ as above.
263: The above notion of
264: stability (sometimes
265: also called semistability because of the large inequality) condition
266: in~\eqref{sta1} appears naturally in the calculus
267: of variations setting and it
268: is usually related to minimization
269: and monotonicity properties.
270: In particular, \eqref{6bis} and~\eqref{sta1}
271: state that the (formal) second variation
272: of the energy functional associated
273: to the equation has a sign (see, e.g.,~\cite{Moss, FCS, AAC}
274: and Section~7 of~\cite{FSV} for further details).
275:
276: In our case, however, it is convenient to {\em relax} this definition
277: of stability. Namely, we say that~$u$ is stable
278: if~\eqref{sta1}
279: holds for any~$\xi$ of the form~$\xi:= |\nabla_y u|\phi$,
280: where~$\phi:\R^{n+1}\rightarrow \R$ is
281: Lipschitz and vanishes on $\R^{n+1}_+\setminus B_R$.
282:
283: This relaxation of the stability definition is convenient for
284: our setting, since it makes possible to write~\eqref{sta1}
285: when $f$ is only locally Lipschitz and not necessarily differentiable.
286:
287: Indeed, since the map~$y\mapsto u(y,x)$ will be taken to
288: be
289: locally Lipschitz (see~\eqref{LipA} below),
290: then so is the map~$y\mapsto f(u(y,x))$
291: and therefore
292: $$ f'(u)\xi^2 =\nabla_y \big(f(u)\big)\cdot \nabla_y u\,\phi^2$$
293: is well-defined almost everywhere, making sense of the last
294: term in~\eqref{sta1}.
295:
296: The regularity assumption we take on~$u$ (see, in particular,~\eqref{hgasj7717177}
297: and~\eqref{SA3}) make the first term in~\eqref{sta1} well-posed too.
298: \bigskip
299:
300: The main results we prove are a geometric formula,
301: of Poincar\'e-type, given in Theorem~\ref{POIN:TH},
302: and a symmetry result, given in Theorem~\ref{SYM:TH}.
303:
304: For our geometric result, we need to recall
305: the following notation. Fixed $x>0$ and~$c\in\R$, we
306: look at the level set
307: $$ S:= \{
308: y\in\R^n {\mbox{ s.t. }}
309: u(y,x)=c
310: \}.$$
311: We will consider the regular points of~$S$,
312: that is, we define
313: $$ L:=\{ y\in
314: S
315: {\mbox{ s.t. }}
316: \nabla_y u(y,x)\neq 0
317: \}.$$
318: Note that~$L$ depends on the~$x\in(0,+\infty)$
319: that we fixed at the beginning, though we do not keep
320: explicit track of this in the notation.
321:
322: For any point $y\in L$,
323: we let $\nabla_L$ to be the tangential gradient
324: along~$L$, that is, for any~$y_o\in L$
325: and any~$G:\R^n\rightarrow\R$ smooth in the vicinity of~$y_o$,
326: we set
327: \begin{equation}\label{GR} \nabla_L G(y_o):=
328: \nabla_y G(y_o)-\left(\nabla_y G(y_o)\cdot
329: \frac{\nabla_y u(y_o,x)}{|
330: \nabla_y u(y_o,x)|}\right)
331: \frac{\nabla_y u(y_o,x)}{|
332: \nabla_y u(y_o,x)|}.\end{equation}
333: Since~$L$ is a smooth manifold, in virtue of
334: the Implicit Function Theorem (and of the standard
335: elliptic
336: regularity of $u$
337: apart from the boundary of $\R^{n+1}_+$),
338: we can define
339: the principal curvatures on it, denoted by
340: $$\kappa_1(y,x),\dots,
341: \kappa_{n-1}(y,x),$$ for any~$y\in L$.
342: We will then define the total curvature
343: $$ {\mathcal{K}}(y,x):=\sqrt{
344: \sum_{j=1}^{n-1} \big(\kappa_j (y,x)\big)^2
345: }.$$
346:
347: We also define
348: $$ {\mathcal{R}}^{n+1}_+:=\{
349: (y,x)\in\R^n\times(0,+\infty){\mbox{ s.t. }}
350: \nabla_y u(y,x)\neq 0
351: \}.$$
352:
353: With this notation, we can state
354: our geometric formula:
355:
356: \begin{thm}\label{POIN:TH}
357: Assume that
358: $u$ is a bounded and stable weak
359: solution of~\eqref{eq1-provv} under assumptions $(S)$.
360:
361: Assume furthermore that
362: \begin{itemize}
363: \item For all $r>0$,
364: \begin{equation}\label{LipA} |\nabla_y u|\in
365: {L^\infty(\overline{B_r^+})}. \end{equation}
366: \item For every $(y,x) \in B_R^+ \bigcap \left \{ \nabla u \neq 0 \right \}$, we have
367: \begin{equation}\label{H1}
368: a(x,|\nabla u|) +\frac{a_t (x,|\nabla u|)}{|\nabla u|} u_x^2 \geq 0
369: \end{equation}
370: and
371: \begin{equation}\label{H2}
372: a(x,|\nabla u|) +\frac{a_t (x,|\nabla u|)}{|\nabla u|}
373: |\nabla_y u|^2 \geq \lambda(y,x) \ge0
374: \end{equation}
375: for some $\lambda(y,x)$.
376: \end{itemize}
377:
378: Assume also the following regularity assumptions:
379: \begin{equation} \label{SA0}
380: \begin{split}
381: &{\mbox{for almost any $x>0$, the map $\R^n
382: \ni y\mapsto
383: \nabla u(y,x)$}}\\
384: &{\mbox{
385: is in $W^{1,1}_{\rm loc}(\R^n, \R^n)$ ,
386: }}\end{split}\end{equation}
387: \begin{equation}\label{SA3-provv}
388: \begin{split}
389: &{\mbox{the map $\R^{n+1}_+
390: \ni (y,x)\mapsto a(x,|\nabla u|) \sum_{j=1}^n
391: \big(
392: |\nabla u_{y_j}|^2+|u_{y_j}|^2
393: \big)$}}\\
394: &{\mbox{is in $L^1(B_r^+)$, for any
395: $r>0$
396: }}\end{split}\end{equation}
397: and
398: \begin{equation}\label{SA3}
399: \begin{split}
400: &{\mbox{the map $\R^{n+1}_+
401: \ni (y,x)\mapsto
402: a(x,|\nabla u|) \big(
403: |\nabla|\nabla_y u||^2+|\nabla_y u|^2
404: \big)
405: $}}\\
406: &{\mbox{is in
407: $L^1(B_r^+)$,
408: for any
409: $r>0$.
410: }}\end{split}
411: \end{equation}
412: Then,
413: for any $R>0$ and any~$\phi:\R^{n+1}\rightarrow \R$ which is
414: Lipschitz and vanishes on $\R^{n+1}_+\setminus B_R$, we have that
415: \begin{equation}\label{rtyua77a7a}
416: \begin{split}
417: &\int_{ {\mathcal{R}}^{n+1}_+}
418: \,\phi^2 \left(
419: a(x,|\nabla u|) {\mathcal{K}}^2 |\nabla_y u|^2+
420: \lambda(y,x) \big|
421: \nabla_L |\nabla_y u|
422: \big| ^2
423: \right)\,\leq \\
424: &\, \int_{\mathcal{R}^{n+1}_+} |\nabla_y u|^2 <\mathcal{B}(x,\nabla u)\nabla \phi,\nabla \phi> .
425: \end{split}
426: \end{equation}
427: \end{thm}
428:
429: Assumption \eqref{LipA} is natural and it holds in particular in
430: the case $g := 0$, $a(x,t)=x^\a$ where $\a \in
431: (-1,1)$, as discussed in \cite{SV} (in many cases
432: of interest,
433: interior elliptic regularity then
434: ensures for free that $u$ is $C^1$
435: inside $\R^{n+1}_+$ and
436: $C^2$
437: as soon as the gradient does not vanish).
438: It is important to notice that
439: assumptions~\eqref{H1}
440: and~\eqref{H2} hold in the important
441: case of the $p-$laplace operator (i.e. $a(x,t)=t^{p-2}$, for~$p > 1$),
442: and in the case of the mean curvature operator (i.e. $a(x,t)=
443: \frac{1}{\sqrt{1+t^2}}$).
444:
445: The regularity assumptions
446: in~\eqref{SA0}, \eqref{SA3-provv}
447: and~\eqref{SA3}
448: are satisfied
449: in many cases of interest
450: (see, for instance Lemma~\ref{REGO} below).
451:
452: The result in
453: Theorem~\ref{POIN:TH} has been deeply inspired
454: by the work of~\cite{SZarma, SZcrelle},
455: where related geometric inequalities have been first introduced
456: for the Allen-Cahn equation.
457: Further progress has been done
458: in~\cite{FAR, FSV} for reactions in the interior and
459: in~\cite{SV} for reactions on the boundary.
460:
461: The advantage of formula~\eqref{rtyua77a7a} is that
462: one
463: bounds
464: tangential gradients and curvatures of level sets
465: of stable solutions in terms of the gradient
466: of the solution
467: itself.
468: That is, suitable
469: geometric quantities of interest
470: are controlled by an appropriate
471: energy term.
472:
473: On the other hand, since the geometric formula bounds a
474: weighted~$L^2$-norm of any
475: test function~$\phi$ by a
476: weighted~$L^2$-norm of its
477: gradient, we may
478: consider Theorem~\ref{POIN:TH}
479: as a weighted {P}oincar\'e
480: inequality. Again, the advantage of such a formula
481: is that the weights have a neat geometric interpretation.
482: See also~\cite{Ferrari} for further investigation
483: of Poincar\'e-type formulas.
484:
485: The second result we present is a symmetry result
486: in low dimension.
487:
488: \begin{thm}\label{SYM:TH}
489: Assume that $n=2$
490: and that the assumptions
491: in Theorem~\ref{POIN:TH}
492: hold. Suppose also that~$\lambda(y,x)$
493: in~\eqref{H2}
494: is strictly positive almost everywhere.
495: Suppose also
496: that one of the following conditions~\eqref{g=0}
497: or~\eqref{g=+} hold, namely assume that
498: either for any $M>0$
499: \begin{equation}\label{g=0}
500: {\mbox{the map $(0,+\infty)\ni x\mapsto
501: \displaystyle\sup_{|u|\le M} |g (x,u)|$ is in $L^1((0,
502: +\infty))$}}
503: \end{equation}
504: or that
505: \begin{equation}\label{g=+}
506: \inf_{{x\in\R^n}\atop{u\in\R}}g(x,u)\,u\,
507: \ge\,0.
508: \end{equation}
509:
510: Assume that the diffusion coefficient $a(.,.)$ has a product structure
511: given by
512: $$a(x,t)=\mu(x)\mathcal{A}(t),$$
513: where
514: \begin{itemize}
515: \item the function $\mu$ is positive and such that
516: \begin{equation}\label{1.21bis}
517: \mu(x) \sim x^\alpha\end{equation}
518: for $\alpha \in (-1,1)$.
519: \item One of the following two conditions is met:
520: either \begin{equation}\label{LB}
521: \mathcal{A} \in L^\infty (\R^+,\R^+)
522: \end{equation}
523: or
524: \begin{equation}\label{PB}
525: \mathcal{A}(t) \sim t^{p-2}\end{equation}
526: with~$p \geq 1+\alpha$.
527: \end{itemize}
528: Then, there exist~$\omega:(0,+\infty)
529: \rightarrow {\rm S}^1$
530: and~$u_o: \R\times [0,+\infty)\rightarrow\R$
531: such that
532: $$ u(y,x)=u_o(\omega(x)\cdot y,x)$$
533: for any~$(y,x)\in {\R^{3}_+}$.
534: \end{thm}
535:
536: The paper~\cite{CSM} gave the first contribution
537: to symmetry result for boundary reaction PDEs.
538: In particular,~\cite{CSM} gave a result
539: analogous to Theorem~\ref{SYM:TH} when~$\mu:=1$, $g:=0$
540: and~$f\in C^{1,\beta}$.
541: In~\cite{SV}, a result analogous to Theorem~\ref{SYM:TH}
542: was given when~$a(x,t)=a(x)$, that is when~$a$ is
543: independent on the gradient term. In this sense, Theorem~\ref{SYM:TH}
544: extends the results of~\cite{CSM, SV}
545: to {\em quasilinear, possibly degenerate or singular, equations}
546: (in fact, when~$a(x,t):=x^\a$ and~$g:=0$, then~$\omega$ in Theorem~\ref{SYM:TH}
547: is constant, see~\cite{SV}).
548:
549: We now discuss the assumptions of Theorem~\ref{SYM:TH}. First,
550: the assumptions on~$\mathcal{A}$ are realized for
551: mean curvature operators, for which~$\mathcal{A}(t)=\frac{1}{\sqrt{
552: 1+t^2}}$, which satisfies~\eqref{LB}
553: and for $p$-laplace operators, for which~$\mathcal{A}(t)=
554: t^{p-2}$,
555: when~$p\ge 1+\alpha$, which fulfills~\eqref{PB}.
556:
557: The structural assumption on $\mu(x)$
558: is natural in the light of
559: the representation formula obtained in \cite{cafS}
560: which relates boundary reactions to fractional
561: operator (see also \cite{SV}): in this sense,
562: the operator studied here may be seen as a
563: quasilinear analogue of the fractional laplacian.
564:
565: Theorem \ref{SYM:TH}
566: asserts that, for any $x > 0$, the function $\R^2
567: \ni y\mapsto
568: u(y,x)$ depends only on one variable. Thus,
569: Theorem \ref{SYM:TH}
570: may be seen
571: as the analogue of De Giorgi conjecture of~\cite{DeG}
572: in dimension $n=2$
573: for equation \eqref{eq1-provv}.
574:
575: Condition \eqref{g=0}
576: is fulfilled by $g:=0$, or, more generally,
577: by $g:=g^{(1)}(x) g^{(2)}(u)$,
578: with $g^{(1)}$ summable over~$\R^+$
579: and~$g^{(2)}$ locally Lipschitz.
580: Also, condition~\eqref{g=+} is fulfilled by~$g:=u^{2\ell+1}$,
581: with~$\ell\in\N$.
582:
583: When $u$ is not bounded,
584: the claim of Theorem \ref{SYM:TH}
585: does not, in general, hold
586: (a counterexample being~$a:=1$, $f:=0$,
587: $g:=0$ and~$
588: u(y_1,y_2,x):=y_1^2-y_2^2$).
589:
590: Theorem~\ref{aux:P} below
591: will also provide a result, slightly
592: more general than Theorem \ref{SYM:TH},
593: which will be
594: valid for $n\ge 2$ and without
595: conditions~\eqref{g=0} or~\eqref{g=+},
596: under an additional energy assumption.
597: \bigskip
598:
599: The rest of the paper is devoted to the proofs
600: of Theorems~\ref{POIN:TH} and ~\ref{SYM:TH}.
601: For this, some
602: preliminary energy estimate
603: will also be needed.
604:
605: \section{Some energy bounds}
606:
607: This section is devoted to some preliminary energy estimate,
608: which are needed for the proof of
609: Theorem \ref{SYM:TH}.
610:
611: Thus, throughout this section, the structural
612: assumptions
613: of Theorem \ref{SYM:TH} are in force.
614:
615: We recall that
616: \begin{equation}
617: \label{9bis} a(x,|\nabla u|)\,u_x^2
618: \in L^1(B_R^+)
619: \end{equation}
620: for any $R>0$,
621: due to \eqref{hgasj7717177}.
622:
623: We start with an elementary observation:
624:
625: \begin{lemma}
626: There exists $C>0$ in such a way that
627: \begin{equation}\label{sii1kkkkj1j1}
628: \int_{B_{2R}^+ \setminus B_R^+} \mu(x)
629: \le CR^{n+1+\alpha}
630: \end{equation}
631: for any $R\ge1$ and $\alpha \in (-1,1)$.
632: \end{lemma}
633:
634: \PF We have that
635: \begin{eqnarray*}
636: \int_{B_{2R}^+ \setminus B_R^+} \mu(x)
637: &\le&\int_{0}^{2R} \int_{B_{ 2R} } \mu(x)\,dy\,dx
638: \\ &\le& C_1 R^{n}
639: \int_{0}^{2R} \mu(x)\,dx\\
640: &\le& C_2 R^{n+1+\alpha},
641: \end{eqnarray*}
642: for suitable $C_1$, $C_2>0$,
643: due to~\eqref{1.21bis}.~\CVD
644: \medskip
645:
646: Though not explicitly needed here, we would
647: like to point out that the natural integrability
648: condition in \eqref{hgasj7717177}
649: holds uniformly for bounded solutions.
650: A byproduct of this gives an
651: energy estimate, which we will use in the proof of
652: Theorem \ref{SYM:TH}.
653:
654: \begin{lemma}\label{Daf}
655: For any $R>0$ there exists $C$, possibly
656: depending on $R$, in such a way that
657: \begin{equation}\label{2.2bis}
658: \| \mu(x) \mathcal{A}(|\nabla u|) |\nabla u|^2 \|_{L^1 (B_R^+)}\le C.
659: \end{equation}
660: Moreover, if
661: \begin{itemize}
662: \item $n=2$, and
663: \item either \eqref{g=0} or~\eqref{g=+} holds,
664: \end{itemize} then there exists $C_o>0$
665: such that
666: \begin{equation}\label{AL}
667: \int_{B_R^+}\Big(
668: a(x,|\nabla u|) +|a_t(x,|\nabla u|)| \, |\nabla u|
669: \Big)\,|\nabla u|^2\,\le\, C_o\, R^2
670: \end{equation}
671: for any $R\ge 1$.
672: \end{lemma}
673:
674: \PF We focus on the proof of~\eqref{AL}, since~\eqref{2.2bis}
675: is a simple byproduct of the arguments we are going to perform.
676:
677: The proof of Lemma~\ref{Daf}
678: consists in testing the weak formulation
679: in~\eqref{eq1}
680: with
681: $\xi:=u
682: \tau ^\ell$ where $\tau$ is a cutoff
683: function
684: such that $0\le\tau\in C^\infty_0 (B_{2R})$, with $\tau=1$
685: in $B_{R}$ and $|\nabla \tau|\le 8/R$, with $R\ge1$.
686: The parameter $\ell >1$ will be suitably chosen below.
687:
688: Note that such a $\xi$ is admissible,
689: since~\eqref{hgasj7717177-bis}
690: follows from~\eqref{hgasj7717177}.
691:
692: One then gets from \eqref{eq1} that
693: \begin{eqnarray}\label{BAB}&&\nonumber
694: \int_{{\R^{n+1}_+}}
695: a(x,|\nabla u|)\,
696: \big( |\nabla u |^2 \tau ^\ell +\ell \tau^{\ell-1} u \nabla u \cdot \nabla \tau
697: \big)+
698: \int_{{\R^{n+1}_+}} g(x,u) u \tau^\ell\\&&\qquad=
699: \int_{\R^n} f(u) u\tau^\ell.
700: \end{eqnarray}
701:
702: We now distinguish the case in which~\eqref{LB} holds
703: from the case in which~\eqref{PB} holds.
704:
705: If~\eqref{LB} holds,
706: we take $\ell =2$.
707: Thus, by Cauchy-Schwarz
708: inequality, we deduce from~\eqref{BAB} that
709: \begin{eqnarray*}
710: \int_{\R_+^{n+1}}\mu(x)\mathcal{A}(|\nabla u|)\,|\nabla u|^2\tau^2
711: \le \frac 12 \int_{\R_+^{n+1}}\mu(x)\mathcal{A}(|\nabla u|)\,|\nabla u|^2\tau^2
712: \\
713: +
714: C_* \Big(
715: \int_{\R_+^{n+1}}
716: \mu(x)\mathcal{A}(|\nabla u|)|\nabla \tau|^2+\int_{\R^{n}}|f(u)|\,|u|\,\tau^2
717: \Big)-
718: \int_{\R_+^{n+1}}
719: g(x,u)\,u\,\tau^2,
720: \end{eqnarray*}
721: for a suitable constant $C_*>0$.
722:
723: This, recalling \eqref{BCOMEA},
724: \eqref{g=0}, \eqref{g=+}, \eqref{LB} and~\eqref{sii1kkkkj1j1},
725: plainly gives~\eqref{AL}.
726:
727: If, on the other hand,~\eqref{PB} holds,
728: we take $\ell =p$. Therefore,we have
729: \begin{equation*}
730: \int_{\R_+^{n+1}}\mu(x)\mathcal{A}(|\nabla u|)\,|\nabla u|^2\tau^p
731: \sim \int_{\R_+^{n+1}}\mu(x)\,|\nabla u|^p\tau^p.
732: \end{equation*}
733: Recalling~\eqref{BAB} and
734: using \eqref{g=0}, \eqref{g=+}, \eqref{sii1kkkkj1j1}, one has
735: \begin{equation*}
736: \int_{\R_+^{n+1}}\mu(x)\,|\nabla u|^p\tau^p \leq C
737: \Big\{
738: \int_{\R_+^{n+1}}\mu(x)\,|\nabla u|^{p-1} |\nabla \tau | \tau^{p-1} +R^n
739: \Big\}.
740: \end{equation*}
741: Thus, by Young inequality, we conclude that
742: \begin{eqnarray*}
743: \int_{\R_+^{n+1}}\mu(x) |\nabla u|^p\tau^p
744: \le C\Big \{ \varepsilon \int_{\R_+^{n+1}}
745: \Big \{ \mu(x)^{1/q}|\nabla u|^{p-1} \tau^{p-1} \Big \} ^q +\\
746: C_\varepsilon \int_{\R_+^{n+1}}\mu(x)|\nabla \tau
747: |^p +R^n \Big \}
748: \end{eqnarray*}
749: for some $\varepsilon>0$ and $q=\frac{p}{p-1}$.
750:
751: Making use of~\eqref{sii1kkkkj1j1},
752: this leads to
753:
754: \begin{eqnarray*}
755: \int_{\R_+^{n+1}}\mu(x) |\nabla u|^p\tau^p
756: \le C\Big \{\int_{B_{2R}}\frac{\mu(x)}{R^p}+R^n \Big \}
757: \leq C(R^{n+1+\alpha-p} +R^n).
758: \end{eqnarray*}
759:
760: This gives the desired result as soon
761: as $p \geq 1+\alpha.$~\CVD\medskip
762:
763: \section{The Poincar\'e-type formula: proof of Theorem \ref{POIN:TH}}
764:
765: This section is devoted to the proof
766: of the geometric formula in Theorem \ref{POIN:TH}.
767: As we will see throughout the proof,
768: the assumptions in Theorem \ref{POIN:TH} are natural and quite general.
769:
770: Besides few technicalities, the
771: proof of Theorem \ref{POIN:TH} consists
772: in plugging the right test function in stability
773: condition~\eqref{sta1} and in using the linearization
774: of \eqref{eq1-provv} to get rid of the unpleasant terms.
775: Following are the rigorous details of the proof.
776:
777: By~\eqref{BDE}, we have that
778: \begin{equation}\label{26bis}
779: \begin{split}
780: &\int_{\mathcal{R}^{n+1}_+ } a(x,|\nabla u|)\nabla u \cdot \Psi_{y_j}=
781: \\&
782: -\int_{\mathcal{R}^{n+1}_+ } a(x,|\nabla u|) \nabla u_{y_j} \cdot \nabla \Psi + a_t
783: (x,|\nabla u|) \frac{\nabla u \cdot \nabla u_{y_j}}{|\nabla u|} \nabla u \cdot \Psi=\\
784: &
785: =-\int_{\mathcal{R}^{n+1}_+ } <\mathcal{B}(x,\nabla u) \nabla u_{y_j}, \Psi >.
786: \end{split}\end{equation}
787: for any~$j=1,\dots, n$ and any~$\Psi\in C^\infty
788: (\R^{n+1}_+, \R^n)$ supported in~$B_R$.
789:
790: Making use of~\eqref{eq1} and~\eqref{26bis} with~$\Psi:=\nabla\psi$,
791: we conclude that
792: \begin{equation}\label{a711aa}
793: \begin{split}
794: &\int_{\R^{n+1}_+} g_u(x,u) u_{y_j} \psi -
795: \int_{\R^n} f'(u) u_{y_j} \psi=
796: \\
797: &\int_{\R^{n+1}_+} (g(x,u))_{y_j} \psi - \int_{\R^n} (f(u))_{y_j}\psi=\\
798: &-\int_{\R^{n+1}_+} g(x,u)\psi_{y_j} + \int_{\R^n} f(u)\psi_{y_j}=
799: \\&-\int_{\mathcal{R}^{n+1}_+ } <\mathcal{B}(x,\nabla u) \nabla u_{y_j}, \nabla \psi >
800: \end{split}
801: \end{equation}
802: for any~$j=1,\dots, n$ and any~$\psi\in
803: C^\infty
804: (\R^{n+1}_+)$ supported in~$B_R$.
805:
806: A density argument (see, e.g.,
807: Lemma~3.4, Theorem~2.4 and~(2.9) in~\cite{CPSC})
808: via~\eqref{BCOMEA}
809: and~\eqref{SA3-provv}, implies that~\eqref{a711aa}
810: holds for~$\psi:=u_{y_j} \phi^2$,
811: where~$\phi$ is
812: as in the statement of
813: Theorem~\ref{POIN:TH}, therefore
814: \begin{equation}\label{3.2bis}
815: \begin{split}
816: & 0= \int_{ {B}_R^+}
817: <\mathcal{B}(x,\nabla u) \nabla u_{y_j},
818: \nabla u_{y_j} > \phi^2+ <\mathcal{B}(x,\nabla u) \nabla u_{y_j}, \nabla \phi^2> u_{y_j}+\\
819: & \int_{B_R^+} g_u(x,u) u_{y_j}^2 \phi^2 -\int_{\partial B_R^+} f'(u) u_{y_j}^2 \phi^2.
820: \end{split}
821: \end{equation}
822:
823: Let now~$r$, $\rho>0$ and~$P\in \R^{n+1}_+$ be such
824: that~$B_{r+\rho}(P)\subset \R^{n+1}_+$.
825: We consider~$\gamma$ to be either~$|\nabla_y u|$
826: or~$u_{y_j}$. In force of~\eqref{SA3-provv}
827: and~\eqref{SA3}, we see that~$\gamma$
828: is in~$W^{1,2}( B_r(P))$, and so in~$W^{1,1}_{\rm loc}
829: (B_r (P))$.
830:
831: Thus, by Stampacchia Theorem (see, e.g., Theorem~6.19
832: in~\cite{LOSS}), $\nabla \gamma=0$ for almost
833: any~$(y,x)\in B_r(P)$ such that~$\gamma(y)=0$.
834:
835: Hence, since~$P$, $r$ and~$\rho$ can be
836: chosen arbitrarily,
837: we have that
838: \begin{equation}\label{ASTA}{\mbox{
839: $\nabla |\nabla_y u| =0=\nabla u_{y_j}$
840: for almost every~$(y,x)$ such that~$\nabla_y u(y,x)=0$.}}\end{equation}
841:
842: By~\eqref{3.2bis} and~\eqref{ASTA}, we obtain
843: \begin{equation*}
844: \begin{split}
845: & 0= \int_{\mathcal{B}_R^+}
846: <\mathcal{B}(x,\nabla u) \nabla u_{y_j},
847: \nabla u_{y_j} > \phi^2+ <\mathcal{B}(x,\nabla u) \nabla u_{y_j}, \nabla \phi^2> u_{y_j}+\\
848: & \int_{B_R^+} g_u(x,u) u_{y_j}^2 \phi^2 -\int_{\partial B_R^+} f'(u) u_{y_j}^2 \phi^2.
849: \end{split}
850: \end{equation*}
851: where $\mathcal{B}_R^+=B_R^+ \bigcap \mathcal{R}^{n+1}_+.$
852: We now sum over $j=1,...,n$ to get (dropping, for short,
853: the dependences of $\mathcal{B}$) and
854: we obtain
855: \begin{equation} \label{78iddudududuudaa}
856: \begin{split}
857: &-\int_{\mathcal{B}_R^+} \sum_{j=1}^n <\mathcal{B}\nabla u_{y_j}, \nabla u_{y_j} >
858: \phi^2- \frac{1}{2}<\mathcal{B} \nabla |\nabla_y u|^2, \nabla \phi^2>=\\
859: & \int_{B_R^+} g_u(x,u) |\nabla_y u|^2 \phi^2 -\int_{\partial B_R^+} f'(u) |\nabla_y u|^2 \phi^2.
860: \end{split}
861: \end{equation}
862:
863: Now, we make use of~\eqref{sta1}
864: by taking~$\xi:=|\nabla_y u|\phi$
865: (this choice was also performed
866: in~\cite{SZarma, SZcrelle, FAR, FSV,SV};
867: note that \eqref{LipA}
868: and~\eqref{SA3} imply \eqref{hgasj7717177-bis}
869: and so they
870: make it possible to
871: use here such a test function). We thus obtain
872: \begin{equation*}
873: \begin{split}
874: & 0 \leq \int_{\mathcal{B}_R^+} <\mathcal{B}
875: \nabla |\nabla_y u|, \nabla |\nabla_y u|> \phi^2
876: + <\mathcal{B} \nabla \phi, \nabla \phi> |\nabla_y u|^2+\\
877: & 2 <\mathcal{B}\nabla |\nabla_y u| ,
878: \nabla \phi> |\nabla_y u| \phi
879: + g_u(x,u) |\nabla_y u|^2 \phi ^2 -\int_{\partial B_R^+} f'(u) |\nabla_y u|\phi^2,
880: \end{split}
881: \end{equation*}
882: where~\eqref{ASTA} has been used once more.
883:
884: This and~\eqref{78iddudududuudaa}
885: imply that
886: \begin{equation}\begin{split}
887: \label{s88818181}
888: &0 \leq \int_{\mathcal{B}_R^+} <\mathcal{B} \nabla |\nabla_y u|, \nabla |\nabla_y u|> \phi^2 +<\mathcal{B} \nabla \phi, \nabla \phi> |\nabla_y u|^2 \\
889: &-\sum_{j=1}^n <\mathcal{B}\nabla u_{y_j}, \nabla u_{y_j} > \phi^2.
890: \end{split}\end{equation}
891:
892: By using \eqref{BDE} and~\eqref{s88818181}, we are lead
893: to the following inequality
894: \begin{equation}\label{31jkl}
895: \begin{split}
896: & 0 \leq \int_\mathcal{B_R^+} a(x,|\nabla u|) \phi^2 \Big [|\nabla |\nabla_y u||^2
897: -\sum_{j=1}^n |\nabla u_{y_j}|^2\Big ]+\\&\quad
898: < \mathcal{B} \nabla \phi, \nabla \phi> |\nabla_y
899: u|^2 +\\
900: &\quad\quad
901: \frac{a_t(x,|\nabla u|) \phi^2}{|\nabla u|} \Big [(\nabla u \cdot \nabla |\nabla_y
902: u|)^2 -\sum_{j=1}^n (\nabla u \cdot \nabla u_{y_j})^2 \Big ]
903: .\end{split}
904: \end{equation}
905:
906: We denote
907: $$
908: \mathcal{H}_*:=
909: -(\partial_x |\nabla_y u|)^2+\sum_{j=1}^n u_{xy_j}^2 ,$$
910: $$\mathcal{H}_1:=|\nabla |\nabla_y u||^2-\sum_{j=1}^n |\nabla u_{y_j}|^2$$
911: $${\mbox{and }}\quad
912: \mathcal{H}_2=:(\nabla u \cdot \nabla |\nabla_y u |)^2-\sum_{j=1}^n (\nabla u
913: \cdot \nabla u_{y_j})^2.$$
914:
915: We have that
916: \begin{equation}
917: \label{A1}\begin{split}
918: & \mathcal{H}_2= (u_x \partial_x |\nabla_y u|)^2 -\sum_{j=1}^n (u_x u_{xy_j})^2+
919: (\nabla_y u \cdot \nabla_y |\nabla_y u|)^2-\sum_{j=1}^n (\nabla_y u \cdot \nabla_y u_{y_j})^2
920: \\ &\qquad=
921: -u_x^2 \mathcal{H}_*
922: +(\nabla_y u \cdot \nabla_y |\nabla_y u|)^2-\sum_{j=1}^n (\nabla_y u \cdot \nabla_y u_{y_j})^2
923: ,\end{split}\end{equation}
924: where we have just separated the $x$ and $y$ variables.
925:
926: Also, {f}rom~\eqref{GR},
927: \begin{equation}\label{3.5bis}
928: |\nabla_L G|^2=|\nabla_y G|^2-
929: \left(\nabla_y G \cdot \frac{\nabla_y u}{|\nabla_y u|}
930: \right)^2,\end{equation}
931: for any smooth function~$G:\R^n\rightarrow\R$.
932:
933: Hence, making use of~\eqref{3.5bis}
934: with~$G:=|\nabla_y u|$, we obtain that,
935: on~${{\mathcal{R}}^{n+1}_+}$,
936: \begin{equation}
937: \label{A2}
938: \begin{split}
939: &(\nabla_y u \cdot \nabla_y |\nabla_y u|)^2-\sum_{j=1}^n (\nabla_y u \cdot \nabla_y u_{y_j})^2=\\
940: &|\nabla_y u|^2 \Big [ \Big(
941: \frac{\nabla_y u}{|\nabla_y u|} \cdot \nabla_y |\nabla_y
942: u|\Big)^2-\sum_{j=1}^n \Big(
943: \frac{\nabla_y u}{|\nabla_y u|} \cdot \nabla_y u_{y_j}
944: \Big)^2 \Big ]=\\
945: &|\nabla_y u|^2 \Big [ |\nabla_y |\nabla_y u| |^2 -|\nabla_L |\nabla_y u ||^2 -\sum_{j=1}^n
946: \Big(\frac{\nabla_y u}{|\nabla_y u|} \cdot \nabla_y u_{y_j}
947: \Big)^2\Big ]=\\
948: &-|\nabla_y u|^2 |\nabla_L |\nabla_y u ||^2.
949: \end{split}
950: \end{equation}
951:
952: By a differential geometry formula
953: obtained in~\cite{SZarma, SZcrelle}
954: (see also
955: equation~(2.10) in~\cite{FSV}), we have, on~${{\mathcal{R}}^{n+1}_+}$,
956: \begin{equation}\label{A3}
957: \mathcal{H}_1=-\mathcal{H}_* -(\mathcal{K}^2 |\nabla_y u|^2+|\nabla_L |\nabla_y u||^2).
958: \end{equation}
959:
960: As a consequence of~\eqref{A1},
961: \eqref{A2} and~\eqref{A3}, we obtain that~\eqref{31jkl}
962: may be written in the following form:
963: \begin{eqnarray}
964: \label{yuiooaoo}
965: \nonumber
966: 0 \leq \int_{{{\mathcal{R}}^{n+1}_+}} a(x,|\nabla u|) \phi^2 \Big (-\mathcal{H}_* -(\mathcal{K}^2|\nabla_y u|^2 +|\nabla_L |\nabla_y u||^2)\Big )\\
967: +\frac{a_t(x,|\nabla u|) \phi^2}{|\nabla u|} \Big (-u_x^2 \mathcal{H}_* -|\nabla_y u|^2 |\nabla_L |\nabla_y u||^2 \Big )+\\
968: <\mathcal{B}\nabla \phi,\nabla \phi > |\nabla_y u |^2.\nonumber
969: \end{eqnarray}
970: We now note that, on~${{\mathcal{R}}^{n+1}_+}$, by Cauchy-Schwarz inequality, we have
971: $\mathcal{H}_* \geq 0.$
972:
973: This,~\eqref{yuiooaoo} and assumptions \eqref{H1}-\eqref{H2} complete the proof of
974: Theorem~\ref{POIN:TH}.~\CVD
975:
976: \section{The symmetry result: proof of Theorem \ref{SYM:TH}}
977:
978: As in~\cite{FSV, SV},
979: the strategy for proving Theorem \ref{SYM:TH}
980: is to test the geometric formula of
981: Theorem~\ref{POIN:TH} against an appropriate capacity-type
982: function to make the left
983: hand side vanish.
984: This would give that the curvature of the level sets for fixed~$x>0$
985: vanishes and so that these level sets are flat, as desired
986: (for this,
987: the vanishing of the tangential gradient term is also
988: useful to take care of the possible
989: plateaus of~$u$, where
990: the level sets are not smooth manifold).
991:
992: As described in the assumptions of Theorem \ref{SYM:TH}, we will
993: take some structure for the weight $a(x,|\nabla u|)$ (in fact,
994: such assumptions might be further weakened, paying the price
995: of additional
996: technicalities in the proofs).
997:
998: Some preparation is needed for the proof
999: of Theorem \ref{SYM:TH}.
1000: Indeed, Theorem \ref{SYM:TH} will
1001: follow from the subsequent Theorem~\ref{aux:P}, which
1002: is valid for any dimension~$n$ and without the restriction
1003: in either \eqref{g=0}
1004: or \eqref{g=+}.
1005:
1006: We will use the notation~$X:=(y,x)$ for points in~$\R^{n+1}_+$.
1007:
1008: Given~$\rho_1\le\rho_2$, we also define
1009: $${\mathcal{A}}_{\rho_1,\rho_2}:=\{
1010: X\in\R^{n+1}_+{\mbox{ s.t. }}|X|\in [\rho_1,\rho_2]
1011: \}.$$
1012:
1013: \begin{lemma}\label{tatay}
1014: Let~$R>0$
1015: and~$h:B_R^+\rightarrow\R$ be a nonnegative
1016: measurable function.
1017:
1018: For any~$\rho\in (0,R)$,
1019: let
1020: $$ \eta(\rho):=\int_{B^+_{\rho}} h.$$
1021: Then,
1022: $$\int_{{\mathcal{A}}_{\sqrt R, R}}\frac{h(X)}{|X|^2}\,dX
1023: \leq 2\int_{\sqrt R}^R t^{-3}\eta(t)\,dt+\frac{\eta(R)}{R^2}.
1024: $$
1025: \end{lemma}
1026:
1027: For the proof
1028: of Lemma~\ref{tatay}, see Lemma~10 in~\cite{SV}.
1029:
1030: \begin{thm}\label{aux:P}
1031: Let $u$ be as requested in Theorem \ref{POIN:TH}.
1032: Assume furthermore that
1033: there exists~$C_o\geq 1$ in such a way that
1034: \begin{equation}\label{en:bound}
1035: \int_{B^+_R} \Big(
1036: a(x,|\nabla u|)+|a_t(x,|\nabla u|)|\, |\nabla u|
1037: \Big)|\nabla u|^2\le C_o\,
1038: R^2\end{equation}
1039: for any~$R\ge C_o$.
1040:
1041: Then there exist~$\omega:(0,+\infty)
1042: \rightarrow {\rm S}^1$
1043: and~$u_o: \R\times [0,+\infty)\rightarrow\R$
1044: such that
1045: $$ u(y,x)=u_o(\omega(x)\cdot y,x)$$
1046: for any~$(y,x)\in\R^{n+1}_+$.
1047: \end{thm}
1048:
1049: \PF {F}rom Lemma~\ref{tatay} applied here with
1050: $$h(X):=
1051: \Big(a(x,|\nabla u(X)|)+|a_t(x,|\nabla u(X)|)|\, |\nabla u(X)|
1052: \Big) |\nabla
1053: u(X)|^2$$ and \eqref{en:bound}, we obtain
1054: \begin{equation}\label{7s77s88}
1055: \begin{split}&
1056: \int_{{\mathcal{A}}_{\sqrt R, R}}\frac{\Big(a(x,|\nabla u(X)|)+
1057: |a_t(x,|\nabla u(X)|)|\, |\nabla
1058: u(X)|\Big) |\nabla u(X)|^2
1059: }{ |X|^2}\\ &\qquad\leq C_1\log R
1060: \end{split}\end{equation}
1061: for a suitable~$C_1$,
1062: as long as~$R$ is large enough.
1063:
1064: Now we define
1065: $$ \phi_R(X):=\left\{
1066: \begin{matrix}
1067: \log R & {\mbox{ if $|X|\le \sqrt R$,}}\\
1068: 2\log\big( R/|X|\big)\Big)
1069: & {\mbox{ if $\sqrt R<|X|< R$,}}
1070: \\
1071: 0 & {\mbox{ if $|X|\ge R$}}
1072: \end{matrix}
1073: \right.$$
1074: and we observe that
1075: \begin{equation}\label{4.2bis}
1076: |\nabla\phi_R|\leq \frac{C_2\,\chi_{
1077: {\mathcal{A}}_{\sqrt R, R}
1078: }}{|X|},\end{equation}
1079: for a suitable~$C_2>0$.
1080:
1081: From~\eqref{BDE}
1082: and Cauchy-Schwarz
1083: inequality, we have that,
1084: for any~$w\in \R^{n+1}$,
1085: \begin{equation}\label{4.2tris}
1086: |<\mathcal{B}(x,\nabla u)
1087: w, w >| \leq \Big \{a(x,|\nabla u|)+|a_t(x,|\nabla u|)||\nabla u| \Big \} |
1088: w|^2.\end{equation}
1089: Thus, plugging~$\phi_R$ inside the geometric
1090: inequality of Theorem~\ref{POIN:TH}, we obtain
1091: \begin{eqnarray*}
1092: && (\log R)^2\int_{B^+_{\sqrt{R}}\bigcap
1093: {\mathcal{R}}^{n+1}_+
1094: }
1095: \left( a(x,|\nabla u|)
1096: {\mathcal{K}}^2 |\nabla_y u|^2+ \lambda(y,x)
1097: \big|
1098: \nabla_L |\nabla_y u|
1099: \big| ^2
1100: \right)\\&&\qquad\qquad\,\leq\,C_3
1101: \int_{
1102: {\mathcal{A}}_{\sqrt R, R}
1103: }\frac{
1104: \Big(
1105: a(x,|\nabla u|)+|a_t(x,|\nabla u|)||\nabla u|
1106: \Big)\,
1107: |\nabla_y u|^2}{|X|^2}
1108: \end{eqnarray*}
1109: for large~$R$, thanks to~\eqref{4.2bis} and~\eqref{4.2tris}.
1110:
1111: By dividing by~$(\log R)^2$,
1112: employing~\eqref{7s77s88}
1113: and taking~$R$ arbitrarily large, we conclude
1114: that~${\mathcal{K}}$ and~$\big|
1115: \nabla_L |\nabla_y u|
1116: \big|$ vanish identically
1117: on~${\mathcal{R}}^{n+1}_+$.
1118:
1119: Then, the desired result follows
1120: by Lemma~2.11 of~\cite{FSV} (applied to the function~$y\mapsto
1121: u(y,x)$, for any fixed~$x>0$).~\CVD
1122: \medskip
1123:
1124: We now complete the proof of
1125: Theorem~\ref{SYM:TH}.
1126: We observe that, under the assumptions of
1127: Theorem~\ref{SYM:TH},
1128: estimate \eqref{en:bound}
1129: holds, thanks to \eqref{AL}.
1130: Consequently, the hypotheses of Theorem~\ref{SYM:TH}
1131: imply the ones of Theorem~\ref{aux:P},
1132: from which the claim in Theorem~\ref{SYM:TH} follows.~\CVD
1133:
1134: \section{Further comments on assumptions
1135: \eqref{SA0},
1136: \eqref{SA3-provv} and~\eqref{SA3}}
1137:
1138: Having completed the proof of the main results,
1139: in this section we would like to remark
1140: that assumptions~\eqref{SA0},
1141: \eqref{SA3-provv} and~\eqref{SA3}
1142: are quite natural in many cases of interest.
1143:
1144: For instance, we assume in this section that
1145: the structural hypotheses on~$a(x,t)$ in Theorem~\ref{SYM:TH}
1146: and the bound in~\eqref{LipA}
1147: hold true.
1148:
1149: For simplicity,
1150: we also suppose that~$u$ is~$C^2_{\rm loc} (\R^{n+1}_+)$
1151: (this is the case, for instance,
1152: of mean curvature type operators or of~$p-$laplace
1153: operators if~$\nabla u$ does not vanish). The purpose
1154: of this section is then to show that
1155: conditions~\eqref{SA0},
1156: \eqref{SA3-provv} and~\eqref{SA3}
1157: are satisfied in this case.
1158:
1159: \begin{lemma}\label{Nuovo}
1160: We have
1161: $$\mu(x) \mathcal{A}(|\nabla u|)|\nabla u_{y_j}|^2 \in L^1(B_R^+)$$
1162: for every $R>0$.
1163: \end{lemma}
1164:
1165: \PF
1166: Given $|\eta|<1$, $\eta\ne 0$, we consider the incremental
1167: quotient
1168: $$ u_\eta(y,x):= \frac{u(y_1,\dots,y_j+\eta,\dots,y_n,x)-
1169: u(y_1,\dots,y_j,\dots,y_n,x)}{\eta}.$$
1170: Since $f$ is locally Lipschitz,
1171: \begin{equation}\label{Al1}
1172: [f(u)]_\eta \le C,
1173: \end{equation}
1174: for some $C>0$, due to \eqref{LipA}.
1175:
1176: Analogously, from \eqref{8ikeoqoqoqoo78}
1177: and \eqref{LipA},
1178: for any $R>0$ there exists $C_R>0$ such that
1179: \begin{equation}\label{Al2}
1180: [g(x,u)]_\eta \le C_R
1181: \end{equation}
1182: for any $x\in(0,R)$.
1183:
1184: Let now $\xi$ be as requested in~\eqref{eq1}.
1185: Then,~\eqref{eq1} gives that
1186: \begin{eqnarray}\label{5.2bis}
1187: \nonumber&&
1188: \int_{{\R^{n+1}_+}}\big[
1189: \mu(x) \mathcal{A}(|\nabla u|)\nabla u_{\eta}\cdot
1190: \nabla\xi+\big(g(x,u)\big)_\eta \,\xi\big]-
1191: \int_{\partial {\R^{n+1}_+}} \big[
1192: f(u)\big]_\eta\xi\\
1193: &=&
1194: -\int_{{\R^{n+1}_+}}\big[
1195: \mu(x) \mathcal{A}(|\nabla u|)\nabla u\cdot
1196: \nabla\xi_{-\eta}+g(x,u) \,\xi_{-\eta}
1197: \big]+
1198: \int_{\partial {\R^{n+1}_+}}
1199: f(u) \xi_{-\eta}\\
1200: &=&0. \nonumber
1201: \end{eqnarray}
1202:
1203: We concentrate on the case when~\eqref{LB} holds
1204: (the case in which~\eqref{PB} holds is then an
1205: easy
1206: modification, analogous to the one performed
1207: in the proof of Lemma~\ref{Daf}).
1208:
1209: We consider a smooth cutoff function
1210: $\tau$ such that $0\le\tau\in C^\infty_0 (B_{R+1})$, with $\tau=1$
1211: in $B_{R}$ and $|\nabla \tau|\le 2$.
1212: Taking $\xi:=u_{\eta}\tau^2$ in~\eqref{5.2bis}, one gets
1213: \begin{equation}\label{0a8h1hmclakk}
1214: \begin{split}
1215: & 2 \int_{{\R^{n+1}_+}}
1216: \mu(x) \mathcal{A}(|\nabla u|)\tau u_{\eta} \nabla u_{\eta
1217: }\cdot \nabla \tau \\&\quad
1218: +\int_{{\R^{n+1}_+}}
1219: \mu(x) \mathcal{A}(|\nabla u|)\tau^2 |\nabla u_{\eta}|^2
1220: +\int_{{\R^{n+1}_+}} \big(g(x,u)\big)_\eta
1221: u_{\eta}\,\tau^2\\&\quad\quad
1222: =\int_{\partial {\R^{n+1}_+}}
1223: \big(f(u)\big)_\eta\,u_{\eta} \tau^2.
1224: \end{split}\end{equation}
1225: We remark that the above choice of~$\xi$
1226: is admissible, since
1227: \eqref{hgasj7717177-bis}
1228: follows from~\eqref{LipA} and \eqref{9bis}.
1229:
1230: Now, by Cauchy-Schwarz
1231: inequality, we have
1232: \begin{equation*}
1233: \begin{split}
1234: &\int_{{\R^{n+1}_+}}
1235: \mu(x) \mathcal{A}(\nabla u|)\tau u_{\eta} \nabla u_{\eta}\cdot \nabla \tau \geq
1236: -\frac{\varepsilon}{2} \int_{{\R^{n+1}_+}}
1237: \mu(x) \mathcal{A}(|\nabla u|)\tau^2 |\nabla u_{\eta}|^2\\
1238: & -\frac{1}{2\varepsilon} \int_{{\R^{n+1}_+}}
1239: \mu(x) \mathcal{A}(|\nabla u|)u_{\eta}^2|\nabla \tau|^2
1240: \end{split}
1241: \end{equation*}
1242: for any $\varepsilon>0$.
1243:
1244: Therefore, by choosing $\varepsilon$ suitably small,~\eqref{0a8h1hmclakk}
1245: reads
1246: \begin{eqnarray*}
1247: &&
1248: \int_{{\R^{n+1}_+}}
1249: \mu(x) \mathcal{A}(|\nabla u|)\tau^2 |\nabla u_{\eta}|^2
1250: \\
1251: &\le& C\,
1252: \Big[
1253: \int_{B_{R+1}^+}
1254: \mu(x) \mathcal{A}(|\nabla u|)u_{\eta}^2+
1255: \int_{B_{R+1}^+} \big| \big(g(x,u)\big)_\eta
1256: u_{\eta}\big|
1257: \\
1258: &&\quad+
1259: \int_{\{|y|\le R\}\times\{x=0\}} \big|
1260: \big(f(u)\big)_\eta
1261: u_{\eta}
1262: \big|\Big].
1263: \end{eqnarray*}
1264: for some $C>0$.
1265:
1266: {F}rom~\eqref{LipA},
1267: \eqref{LB},
1268: \eqref{Al1} and \eqref{Al2},
1269: we thus control
1270: $$\int_{B_R^+}
1271: \mu(x) \mathcal{A}(|\nabla u|)\tau^2 |\nabla u_{\eta}|^2
1272: $$
1273: uniformly in $\eta$.
1274:
1275: By sending $\eta\rightarrow 0$
1276: and using Fatou Lemma,
1277: we obtain the desired claim.~\CVD
1278: \medskip
1279:
1280: Following is the regularity needed for
1281: some subsequent computations.
1282:
1283: \begin{lemma} \label{REGO}
1284: Conditions~\eqref{SA0},
1285: \eqref{SA3-provv} and~\eqref{SA3}
1286: are satisfied.\end{lemma}
1287:
1288: The proof is omitted, since it is analogous
1289: to the one of Lemma~7 in~\cite{SV}.
1290:
1291: \section*{Acknowledgments}
1292:
1293: YS would like to thank
1294: the hospitality of Universit\`a di
1295: Roma Tor Vergata, where part of this work has been
1296: done.
1297:
1298: EV has been partially supported by~{\em MIUR
1299: Me\-to\-di va\-ria\-zio\-na\-li ed equa\-zio\-ni
1300: dif\-fe\-ren\-zia\-li non\-li\-nea\-ri}.
1301:
1302:
1303:
1304: \bibliographystyle{alpha}
1305: \bibliography{bibliofile}
1306:
1307: \vfill
1308:
1309: {\em YS} --
1310: Universit\'e Aix-Marseille 3, Paul C\'ezanne --
1311: LATP --
1312: Marseille, France.
1313:
1314: {\tt sire@cmi.univ-mrs.fr}
1315: \medskip
1316:
1317: {\em EV} --
1318: Universit\`a di Roma Tor Vergata --
1319: Dipartimento di Matematica --
1320: I-00133 Rome, Italy.
1321:
1322: {\tt valdinoci@mat.uniroma2.it}
1323: \end{document}
1324: