math0401142/mp6.tex
1: % > 538 000 bytes ?? YES : 
2: % Already : 714 000 bytes
3: \documentclass[10pt,twoside,leqno]{amsart}
4: \usepackage{amssymb,amsbsy,amsmath,amsfonts,amssymb,amscd,epsfig,times,graphics,color}
5: \usepackage[latin1]{inputenc}
6: \sloppy
7: \tolerance = 1000
8: 
9: %Format des cadres pour les figures sous xfig : 1-13 cm par 1-6 cm
10: 
11: \newcommand{\B}                 {\mathbb{B}}
12: \newcommand{\C}                 {\mathbb{C}}
13: \newcommand{\I}                 {\mathbb{I}}
14: \newcommand{\K}                 {\mathbb{K}}
15: \newcommand{\LL}                {\mathbb{L}}
16: \newcommand{\R}                 {\mathbb{R}}
17: \newcommand{\N}                 {\mathbb{N}}
18: \renewcommand{\P}               {\mathbb{P}}
19: \newcommand{\V}                 {\mathbb{V}}
20: \def\dl{{[\![}}
21: \def\dr{{]\!]}}
22: 
23: \newtheorem{theorem}{Theorem}
24: \newtheorem{lemma}{Lemma}
25: \newtheorem{proposition}{Proposition}
26: \newtheorem{corollary}{Corollary}
27: \newtheorem{definition}{Definition}
28: \newtheorem{remark}{Remark}
29: \newtheorem{example}{Example}
30: 
31: \setlength{\textwidth}{12.5cm}
32: % Impression :         12.5cm
33: % Impression :         15.5cm
34: % Ecriture   :         13.5cm
35: \setlength{\textheight}{20.5cm}
36: % Impression :          20.5cm
37: % Impression :          24.5cm
38: % Ecriture   :          21.5cm
39: \voffset                  0cm
40: % Impression :            0cm
41: % Impression :            -1.5cm
42: % Ecriture   :            -2.5cm
43: \hoffset                  -0.5cm
44: % Impression :            -0.5cm
45: % Impression :            -2cm
46: % Ecriture   :            -3.5cm
47: 
48: \begin{document}
49: 
50: %\large
51: 
52: \title[Characteristic foliations and envelopes of
53: holomorphy]{Characteristic foliations on maximally real submanifolds
54: of $\C^n$ and envelopes of holomorphy}
55: 
56: \author{Jo\"el Merker and Egmont Porten}
57: 
58: \address{
59: CNRS, Universit\'e de Provence, LATP, UMR 6632, CMI, 
60: 39 rue Joliot-Curie, F-13453 Marseille Cedex 13, France}
61: 
62: \email{merker@cmi.univ-mrs.fr} 
63: 
64: \address{Humboldt-Universit\"at zu Berlin, 
65: Mathematisch-Naturwissenschaftenschaft\-liche Fa\-kult\"at II, 
66: Institut f\"ur Mathematik, Rudower Chaussee 25, 
67: D-12489 Berlin, Germany}
68: 
69: \email{egmont@mathematik.hu-berlin.de}
70: 
71: \subjclass[2000]{Primary: 32D20. 
72: Secondary: 32A20, 32D10, 32V10, 32V25, 32V35}
73: 
74: \date{\number\year-\number\month-\number\day}
75: 
76: \begin{abstract}
77: Let $S$ be an arbitrary real $2$-surface, with or without boundary,
78: contained in a hypersurface $M$ of $\C^2$, with $S$ and $M$ of class
79: $\mathcal{ C}^{ 2, \alpha}$, where $0< \alpha <1$. If $S$ is totally
80: real except at finitely many complex tangencies which are hyperbolic
81: in the sense of E.~Bishop and if the union of separatrices is a tree
82: of curves without cycles, we show that every compact $K$ of $S$ is
83: CR-, $\mathcal{ W}$- and $L^{\sf p}$-removable (Theorem~1.3). Our
84: purely local techniques enable us to formulate substantial
85: generalizations of this statement, for the removability of closed sets
86: in totally real $1$-codimensional submanifolds contained in generic
87: submanifolds of CR dimension $1$.
88: \end{abstract}
89: 
90: \maketitle
91: 
92: \begin{center}
93: \begin{minipage}[t]{11cm}
94: \baselineskip =0.35cm
95: {\scriptsize
96: 
97: \centerline{\bf Table of contents}
98: 
99: \smallskip
100: 
101: {\bf Part~I \dotfill 1.}
102: 
103: \ \
104: {\bf 1.~Introduction \dotfill 1.}
105: 
106: \ \
107: {\bf 2.~Description of the proof of Theorem~1.2 and organization 
108: of the paper \dotfill 8.}
109: 
110: \ \
111: {\bf 3.~Strategy per absurdum for the proofs of
112: Theorems~1.2' and~1.4 \dotfill 13.} 
113: 
114: \ \
115: {\bf 4.~Construction of
116: a semi-local half-wedge \dotfill 19.}
117: 
118: \ \
119: {\bf 5.~Choice of a special point of $C_{\rm nr}$ to be removed 
120: locally \dotfill 32.}
121: 
122: {\bf Part~II \dotfill 45.}
123: 
124: \ \
125: {\bf 6.~Three preparatory lemmas on H\"older spaces \dotfill 45.}
126: 
127: \ \
128: {\bf 7.~Families of analytic discs half-attached
129: to maximally real submanifolds \dotfill 47.}
130: 
131: \ \
132: {\bf 8.~Geometric properties of families of half-attached analytic
133: discs \dotfill 55.}
134: 
135: \ \
136: {\bf 9.~End of proof of Theorem~1.2': application of 
137: the continuity principle \dotfill 63.}
138: 
139: \ \
140: {\bf 10.~Three proofs of Theorem~1.4 \dotfill 69.}
141: 
142: \ \
143: {\bf 11.~$\mathcal{ W}$-removability implies
144: $L^{\sf p}$-removability \dotfill 73.}
145: 
146: \ \
147: {\bf 12.~Proofs of Theorem~1.1 and of Theorem~1.3
148: \dotfill 78.}
149: 
150: \ \
151: {\bf 13.~Applications to the edge of the wedge theorem \dotfill 87.}
152: 
153: \ \
154: {\bf 14.~An example of a nonremovable three-dimensional torus \dotfill 
155: 91.}
156: 
157: \ \
158: {\bf 15.~References \dotfill 96.}
159: 
160: \smallskip
161: 
162: {\footnotesize\tt [With 24 figures]}
163: 
164: }\end{minipage}
165: \end{center}
166: 
167: \bigskip
168: 
169: \section*{\S1.~Introduction}
170: 
171: In the past fifteen years, remarkable progress has been made towards
172: the understanding of the holomorphic extendability properties of CR
173: functions. At the origin of this development, the most fundamental
174: achievement was the deep discovery, due to the effort of numerous
175: mathematicians, that the so-called {\sl CR orbits} are the adequate
176: underlying objects for the semi-local CR analysis on a general
177: embedded CR manifold. As an independent and now established theory in
178: several complex variables, one may find a precise correspondence
179: between such orbits and progressively attached analytic discs covering
180: a thick part of the envelope of holomorphy of CR manifolds, {\it cf.}
181: \cite{ b}, \cite{ trv}, \cite{ tr1}, \cite{ tu1}, \cite{ ber}, \cite{
182: tu2}, \cite{ m1}, \cite{ j2} and \cite{ p3} for a recent synthesis.
183: 
184: Within this framework, it became mathematically accessible to
185: endeavour the general study of removable singularities on embedded CR
186: manifolds $M\subset \C^n$ of arbitrary CR dimension and of arbitrary
187: codimension, not necessarily being the boundaries of (strictly)
188: pseudoconvex domains. With respect to their size or ``mass'', the
189: interesting singularities can be essentially ordered by
190: their codimension in $M$. For instance, provided it does not perturb
191: the fact that $M$ consists of a single CR orbit, an arbitrary closed
192: subset $C\subset M$ which is of vanishing 2-codimensional Hausdorff
193: content is {\it always}\, removable, as is shown in~\cite{ cs} in the
194: hypersurface case and in \cite{ mp3}, Theorem~1.1, in arbitrary
195: codimension. Hence one is left to study the removability of
196: singularities of codimension at most two. Since the general problem
197: of characterizing removability seems at the moment to be out of reach
198: (even for $M$ being a hypersurface), it is advisible to focus on
199: geometrically accessible singularities, namely singularities contained
200: in a CR submanifold of $M$. A complete study of the automatic
201: removability of two-codimensional singularities may be found in
202: Theorem~4 of~\cite{ mp1}. Having in mind the classical Painlev\'e
203: problem, we will mainly consider in this paper singularities which are
204: closed sets $C$ contained in a {\it codimension one}\, 
205: submanifold $M^1$ of $M$ which is {\it generic}\, in $\C^n$.
206: 
207: The known results on singularities of codimension one can be
208: subdivided into two strongly different groups according to the {\it CR
209: dimension} of $M$. If ${\rm CRdim}\, M \geq 2$, then a generic
210: hypersurface $M^1\subset M$ is itself a CR manifold of positive CR
211: dimension, and singularities $C\subset M^1$ can be understood on the
212: basis of the interplay between $C$ and the CR orbits of $M^1$. Deep
213: results in this direction were established when $M$ is a hypersurface
214: of $\C^n$ in~\cite{ j4}, \cite{ j5} and then generalized to CR
215: manifolds of arbitrary codimension in~\cite{ p1}: the 
216: geometric condition insuring automatic removability is simply
217: that $C$ does not contain any CR orbit of $M^1$.
218: 
219: On the other hand, if ${\rm CRdim}\, M=1$ the geometric situation
220: becomes highly different, as a generic hypersurface $M^1\subset M$ is
221: now (maximal) {\it totally real}. Fortunately, as a substitute for the
222: CR orbits of $M^1$, one can consider the so-called {\sl characteristic
223: foliation}\, of $M^1$, obtained by integrating the characteristic line
224: field $T^c M|_{M^1} \cap TM^1$. But removability theorems exploiting
225: this concept were only known for hypersurfaces in $\C^2$ and, until
226: very recently, only in the strictly pseudoconvex case. Furthermore,
227: a geometric condition insuring automatic removablitity
228: has not yet been clearly delineated.
229: 
230: Hence, with respect to the current state of the art, there was a
231: two-fold gap about codimension one removable singularities contained
232: in generic submanifolds $M$ of CR dimension one: firstly, to establish
233: a satisfying theory for non-pseudoconvex hypersurfaces in $\C^2$ and
234: secondly, to understand the situation in higher codimension. This
235: second main task was formulated as the first open problem in a list
236: p.~432 of \cite{j5} ({\it see}\, also the comments pp.~431--432 about
237: the relative geometric simplicity of the case ${\rm CRdim}\, M \geq
238: 2$). {\it A priori}, it is not clear at all whether the two directions
239: of research are related somehow, but in the present work, we shall
240: fill in this two-fold gap by devising a new semi-local approach which
241: applies uniformly with respect to codimension.
242: 
243: For the detailed discussion of our result we have to introduce some
244: terminology which will be used throughout the article. Let $M$ be a
245: generic submanifold of $\C^n$ and let $C$ be a closed subset of $M$.
246: Recall from~\cite{mp3} that a {\sl wedgelike domain} attached to a
247: generic submanifold $M'\subset \C^n$ is a domain containing a local
248: wedge of edge $M'$ at every point of $M'$. Our wedgelike domains will
249: always be {\it nonempty}. Let us define three basic notions of
250: removability. Firstly, we say that $C$ is {\sl CR-removable} if there
251: exists a wedgelike domain $\mathcal{ W}$ attached to $M$ to which
252: every continuous CR function $f\in\mathcal{ C}_{CR}^0(M\backslash C)$
253: extends holomorphically. Secondly, as in~\cite{mp3}, p.~486, we say
254: that $C$ is {\sl $\mathcal{ W}$-removable} if for every wedgelike
255: domain $\mathcal{ W}_1$ attached to $M\backslash C$, there is a
256: wedgelike domain $\mathcal{ W}_2$ attached to $M$ and a wedgelike
257: domain $\mathcal{ W}_3 \subset \mathcal{ W}_1 \cap \mathcal{ W}_2$
258: attached to $M\backslash C$ such that for every holomorphic function
259: $f\in\mathcal{ O}(\mathcal{ W}_1)$, there exists a holomorphic
260: function $F\in\mathcal{ O}(\mathcal{ W}_2)$ which coincides with $f$
261: in $\mathcal{ W}_3$. Thirdly, with ${\sf p} \in\R\cup \{+\infty\}$
262: satisfying ${\sf p} \geq 1$, we say that $C$ is {\it $L^{\sf
263: p}$-removable} if every locally integrable function $f\in L^{\sf
264: p}_{loc}(M)$ which is CR in the distributional sense on $M\backslash
265: C$ is in fact CR on all of $M$.
266: 
267: The first notion of removability is a generalization of the kind of
268: removability considered in most of the pioneering papers \cite{cs},
269: \cite{d}, \cite{fs}, \cite{j1}, \cite{kr}, \cite{l}, \cite{lu},
270: \cite{st} about removable singularities in boundaries of domains $D
271: \subset \subset \C^n$. We observe that a wedgelike open set attached
272: to a hypersurface $M$ is just a (global) one-sided neighborhood of
273: $M$, namely a domain $\omega$ with $\overline{\omega}\supset M$ such
274: that for every point $p\in M$, the domain $\omega$ contains the
275: intersection of a neighborhood of $p$ in $\C^n$ with one side of
276: $M$. If now a closed set $C$ contained in a $\mathcal{ C}^1$-smooth
277: bounded boundary $\partial D$ is CR-removable, then an application of
278: the Hartogs-Bochner theorem shows that CR functions on $\partial
279: D\backslash C$ can be holomorphically extended to $D$. The second
280: notion of removability is a way to isolate the part of the question
281: related to envelopes of holomorphy. The third notion of removability
282: has the advantage of being completely intrinsic with respect to $M$
283: and may be relevant in the study of non-embeddable CR manifolds.
284: 
285: To avoid confusion, we state precisely our submanifold notion: $Y$ is
286: a {\sl submanifold of $X$} if $Y$ and $X$ are equipped with a manifold
287: structure, if there exists an immersion $i$ of $Y$ into $X$ and if the
288: manifold topology of $Y$ and the topology of $i(Y)$ inherited from the
289: topology of $X$ coincide, so that one may identify the submanifold $Y$
290: with the subset $i(Y)\subset X$. Furthermore, our submanifolds will
291: always be connected.
292: 
293: Let us now enter the discussion of the case $n=2$. Here we shall
294: denote the submanifold $M^1\subset M$, which is a surface in $\C^2$,
295: by $S$. In~\cite{b}, E.~Bishop showed that a two-dimensional surface
296: in $\C^2$ of class at least $\mathcal{ C}^2$ having an isolated
297: complex tangency at one of its points $p$ may be represented by a
298: complex equation of the form $w=z\bar z+\lambda(z^2+\bar z^2)+{\rm
299: o}(\vert z\vert^2)$, in terms of local holomorphic coordinates $(z,w)$
300: vanishing at $p$, where the real parameter $\lambda\in [0,\infty)$ is
301: a biholomorphic invariant of $S$. The point $p$ is said to be {\sl
302: elliptic} if $\lambda\in [0,\frac{1}{2})$, {\sl parabolic} if $\lambda
303: = \frac{1}{2}$ and {\sl hyperbolic} if $\lambda \in (\frac{1}{2},
304: \infty)$. Recall that $M$ is called {\sl globally minimal} if it
305: consists of a single CR orbit ({\it cf.} \cite{tr1}, \cite{tr2};
306: \cite{mp1}, pp.~814--815; and~\cite{j4}, pp.~266--269). Throughout
307: this paper, we shall work in the $\mathcal{ C}^{ 2, \alpha}$-smooth
308: category, where $0 < \alpha < 1$. Our first main new result is as
309: follows.
310: 
311: \def\thetheorem{{\bf 1.1}}\begin{theorem}
312: Let $M$ be a globally minimal $\mathcal{ C}^{ 2, \alpha}$-smooth
313: hypersurface in $\C^2$ and let $D \subset M$ be a $\mathcal{ C}^{ 2,
314: \alpha}$-smooth surface which is
315: \begin{itemize}
316: \item[{\bf (a)}]
317: $\mathcal{C}^{2,\alpha}$-diffeomorphic to the unit
318: $2$-disc of $\R^2$ and 
319: \item[{\bf (b)}]
320: totally real outside a discrete subset of isolated complex tangencies
321: which are hyperbolic in the sense of E.~Bishop.
322: \end{itemize}
323: Then every compact subset $K$ of $D$ is CR-, $\mathcal{ W}$- and
324: $L^{\sf p}$-removable.
325: \end{theorem}
326: 
327: As a corollary, one obtains a corresponding result about holomorphic
328: extension from $\partial\Omega\backslash K$ for the case that $M$ is
329: the boundary of a relatively compact domain $\Omega\subset\C^2$. Note
330: that $\partial \Omega$ is automatically globally minimal
331: (\cite{ j4}, Section~2). We will first recall the historical
332: background of Theorem 1.1 and explain afterwards on this basis the
333: main ideas and techniques necessary for the proof. 
334: 
335: In 1988, applying a global
336: version of the Kontinuit\"atssatz, B.~J\"oricke~\cite{j1} established
337: a remarkable theorem: every compact subset of a totally real
338: $\mathcal{ C}^2$-smooth $2$-disc lying on the boundary of the unit
339: ball in $S^3=\partial \B_2\subset \C^2$ is CR-removable. This
340: discovery motivated the work~\cite{fs} by
341: F.~Forstneri$\check{\text{\rm c}}$ and E.L.~Stout, where it is shown
342: that every $\mathcal{ C}^2$-smooth compact $2$-disc contained in a
343: strictly pseudoconvex $\mathcal{ C}^2$-smooth boundary $\partial
344: \Omega$ contained in a $2$-dimensional Stein manifold $\mathcal{ M}$
345: which is totally real except at a finite number of hyperbolic complex
346: tangencies is removable; the proof mainly relies on a previous work by
347: E.~Bedford and W.~Klingenberg about the hulls of $2$-spheres contained
348: in such strictly pseudoconvex boundaries $\Omega \subset \mathcal{
349: M}$, which may be filled by Levi-flat $3$-spheres after a generic
350: small perturbation (\cite{bk}, Theorem~1). Indirectly, it followed
351: from~\cite{j1} and~\cite{fs} that such compact totally real $2$-discs
352: $D\subset \partial \Omega$ (possibly having finitely many hyperbolic
353: complex tangencies) are $\mathcal{ O}(\overline{\Omega})$-convex and
354: in particular polynomially convex if $D=\B_2$ and $\mathcal{ M}=\C^2$,
355: thanks to a previous work~\cite{st} by E.L.~Sout, where it is shown
356: (Theorem~II.10) that a compact subset $K$ of a $\mathcal{ C}^2$-smooth
357: strictly pseudoconvex boundary $\partial \Omega$ in a Stein manifold
358: is removable {\it if and only if}\, $K$ is $\mathcal{
359: O}(\overline{\Omega})$-convex. It is also established in~\cite{fs}
360: that a neighborhood of an isolated hyperbolic complex tangency in
361: $\C^2$ is polynomially convex. These papers have been followed by the
362: work~\cite{d}, where the question of $\mathcal{
363: O}(\overline{\Omega})$-convexity of {\it arbitrary compact surfaces}\,
364: $S$ (with or without boundary, not necessarily diffeomorphic to a
365: $2$-disc) contained in a $\mathcal{ C}^2$-smooth strictly pseudoconvex
366: domain $\Omega\subset \C^2$ is dealt with directly. Using K.~Oka's
367: characterization of the envelope of a compact, J.~Duval shows that the
368: essential hull $\widehat{ K}_{\rm ess}:= \overline{ \widehat{ K}_{
369: \mathcal{ O }( \overline{ \Omega} )} \backslash K}$ must cross
370: every leaf of the characteristic foliation on the totally real part of
371: $S$ and he deduces that a compact $2$-disc having only hyperbolic
372: complex tangencies is $\mathcal{ O }(\overline{ D })$-convex.
373: 
374: All the above proofs heavily rely on strong pseudoconvexity, in
375: contrast to the experience, familiar at least in the case ${\rm
376: CRdim}\, M \geq 2$, that removability should depend rather on the
377: structure of CR orbits than on Levi curvature. The first theorem for
378: the non-pseudoconvex situation was established by the second author in
379: \cite{p2}. He proved that every compact subset of a totally real disc
380: embedded in a globally minimal $\mathcal{ C}^\infty$-smooth
381: hypersurface in $\C^2$ is always CR-removable. We would like to point
382: out that, seeking theorems without any assumption of pseudoconvexity
383: leads to substantial open problems, because one loses almost all of
384: the strong interweavings between function-theoretic tools and
385: geometric arguments which are valid in the pseudoconvex realm, for
386: instance: Hopf Lemma, plurisubharmonic exhaustions, envelopes of
387: function spaces, local maximum modulus principle, Stein neighborhood
388: basis, {\it etc.}
389: 
390: To discuss the main elements of our approach, let us briefly explain
391: the geometric setup of the proof of Theorem 1.1. The characteristic
392: foliation has isolated singularities at the hyperbolic points, where it
393: looks like the phase diagram of a saddle point. In particular there
394: are four local separatrices accumulating orthogonally at each
395: hyperbolic point. Hence we can decompose the $2$-disc $D$ as a union
396: $D=T_D \cup D_o$, where $T_D$ consists of the union of the hyperbolic
397: points of $D$ together with the separatrices issuing from them, and
398: where $D_o:= D \backslash T_D$ is the remaining open submanifold of
399: $D$, contained in the totally real part of $D$. By
400: H.~Poincar\'e and I.~Bendixson's theory, $T_D$ is a tree of $\mathcal{
401: C}^{2,\alpha}$-smooth curves which contains no subset homeomorphic to
402: the unit circle, {\it cf.}~\cite{ d}. Accordingly, we decompose $K:=
403: K_{ T_D} \cup C_o$, where $K_{T_D}:= K \cap T_D$ is a proper closed
404: subset of the tree $T_D$ and where $C_o := K \cap D_o$ is a relatively
405: closed subset of $D_o$.
406: 
407: The hard part of the proof, which was actually the starting point of
408: the whole paper, will consist in removing the closed subset $C_o$ of
409: the $2$-dimensional surface $S:=D_o$ lying in $M \backslash
410: T_D$. Thereafter the removal of the remaining part $K_{T_D}$ will
411: be done by means of an investigation of the behaviour of
412: the CR orbits near $T_D$, close in spirit to our previous
413: methods in ~\cite{mp1} ({\it see}\, Section~12 below for the details).
414: 
415: Let us formulate the first crucial part of the above argument as an
416: independent theorem about the removal of closed subsets contained in a
417: totally real surface $S$. We point out that now $S$ may have {\it
418: arbitrary topology}.
419: 
420: \def\thetheorem{{\bf 1.2}}\begin{theorem}
421: Let $M$ be a $\mathcal{ C}^{ 2, \alpha }$-smooth globally minimal
422: hypersurface in $\C^2$, let $S \subset M$ be a $\mathcal{ C}^{2,
423: \alpha}$-smooth surface, open or closed, with or without boundary,
424: which is totally real at every point. Let $C$ be a proper closed
425: subset of $S$ and assume that the following topological condition
426: holds{\rm :}
427: 
428: \smallskip
429: \begin{itemize}
430: \item[$\mathcal{ F }_S^c \{C\}:$] For every closed subset $C' \subset
431: C$, there exists a simple $\mathcal{ C}^{2,\alpha}$-smooth curve
432: $\gamma: [-1, 1] \to S$, whose range is contained in a single leaf of
433: the characteristic foliation $\mathcal{ F }_S^c$ {\rm (}obtained
434: by integrating the characteristic line field $T^cM \vert_S \cap 
435: TS${\rm )}, with $\gamma (-1)
436: \not \in C'$, $\gamma (0) \in C'$ and $\gamma (1) \not \in C'$, such
437: that $C'$ lies completely in one closed side of $\gamma [-1, 1]$
438: with respect to the topology of $S$ in a neighborhood of
439: $\gamma[-1,1]$.
440: \end{itemize}
441: 
442: \smallskip
443: \noindent
444: Then $C$ is CR-, $\mathcal{ W}$- and $L^{\sf p}$-removable.
445: \end{theorem}
446: 
447: The condition $\mathcal{ F}_S^c \{C \}$ is a {\it common condition}\,
448: on $C$ and on the characteristic foliation $\mathcal{ F}_S^c$, namely
449: on the relative disposition of $\mathcal{ F}_S^c$ with respect to $C$,
450: not only on $S$; an illustration may be found in {\sc Figure~2} below.
451: In the strictly pseudoconvex context, this condition appeared
452: implicitly during the course of the proofs given in \cite{d}. Note
453: that the relevance of the characteristic foliation had earlier been
454: discovered in contact geometry, {\it cf.} \cite{be}, \cite{e}. It is
455: interesting to notice that it re-appears in the situation of Theorem
456: 1.2, where the underlying distribution $T^c M$ is allowed to be very
457: far from contact.
458: 
459: As is known, it follows from a subcase of H.~Poincar\'e and
460: I.~Bendixson's theory that if $S$ is diffeomorphic to a real $2$-disc
461: or if $S=D_o$ as above, then $\mathcal{ F}_S^c \{C\}$ is automatically
462: satisfied for an arbitry compact subset $C$ of $S$. On the contrary, it
463: may be not satisfied when for instance $S$ is an annulus equipped with
464: a radial foliation together with $C$ containing a continuous closed
465: curve around the hole of $S$. Crucially, it is elementary to construct
466: an example of such an annulus which is truly nonremovable. Indeed, the
467: small closed curve $C$ which consists of the transversal intersection
468: of a strictly convex boundary $\partial D$ with a complex line close
469: to a boundary point may be enlarged as a thin maximally real strip $S
470: \subset \partial D$ which is diffeomorphic to an annulus; in this
471: setting, $C$ is obviously nonremovable and {\it the characteristic
472: foliation is everywhere transversal to $C$}. Consequently, the
473: geometric condition $\mathcal{ F}_S^c\{C\}$ is the optimal one
474: insuring automatic removability for all choices of $M$, $S$ and
475: $C$. Further examples of closed subsets in surfaces with arbitrary
476: genus equipped with such foliations may be exhibited.
477: 
478: In the proof of Theorem~1.2, after some contraction $C'$ of $C$, we
479: may assume that no point of $C'$ is locally removable ({\it see}\,
480: Sections~2 and~3 below). Then the very assumption $\mathcal{
481: F}_S^c\{C\}$ yields the existence of a characteristic segment
482: $\gamma[-1,1]$, such that $C'$ lies on one side of
483: $\gamma[-1,1]$. Reasoning by contradiction, our aim is to show that
484: there exists at least one special point $p\in C'\cap \gamma(-1,1)$,
485: which is {\it locally}\, CR-, $\mathcal{ W}$- and $L^{\sf
486: p}$-removable. The choice of such a point $p$, achieved in Section~5
487: below, will be nontrivial.
488: 
489: The strategy for the local removal of $p$ is to construct an analytic
490: disc $A$ such that a segment of its boundary $\partial A$ is attached
491: to $S$ and touches $C'$ in only one point $p$. Several geometrical
492: assumptions have to be met to ensure that a sufficiently rich family
493: of deformations of $A$ have boundaries disjoint from $C'$, that
494: analytic extension along these discs is possible (i.e.~appropriate
495: moment conditions are satisfied), and that the union of these good
496: discs is large enough to give analytic extension to a one-sided
497: neighborhood of $M$: this is where the (semi)localization and the
498: choice of the special point $p\in C'$ will be key ingredients.
499: Let us explain why localization is crucial.
500: 
501: Working globallly, the second author produced in~\cite{ p2} a
502: convenient disc by applying the powerful E.~Bedford and W.~Klingenberg
503: theorem to an appropriate 2-sphere containing a neighborhood of the
504: {\it entire}\, singularity $C'$. This method requires global
505: properties of $S$ like $S$ being a totally real $2$-disc, which
506: ensures the existence of a nice Stein neighborhood basis of
507: $C'$. Already for real discs with isolated hyperbolic points, it is
508: not clear whether this argument can be generalized (however, we would
509: like to mention that recent results of M.~Slapar in~\cite{ sl}
510: indicate that this could be possible at least if the geometry near the
511: hyperbolic points satisfies some additional assumptions). In the case
512: where $M$ is an arbitrary globally minimal hypersurface, where $S$ has
513: arbitrary topology and has complex tangencies, the reduction to
514: E.~Bedford and W.~Klingenberg's theorem seems impossible, {\it cf.}
515: the example of an unknotted nonfillable $2$-sphere in $\C^2$
516: constructed by J.E.~Forn{\ae}ss and D.~Ma in~\cite{fm}. Also, to the
517: authors' knowledge, the possibility of filling by Levi-flat
518: $3$-spheres the (not necessarily generic) $2$-spheres lying on a {\it
519: nonpseudoconvex}\, hypersurface is a delicate open problem. In
520: addition, for the higher codimensional generalization of Theorem~1.2,
521: the idea of global filling seems to be irrelevant at present times,
522: because no analog of the E.~Bedford and W.~Klingenberg theorem is
523: known in dimension $n\geq 3$. As we aim to deal with surfaces $S$
524: having arbitrary topology and to generalize these results in arbitrary
525: codimension, we shall endeavour to firmly {\it localize the
526: removability arguments}, using only small analytic discs.
527: 
528: Thus, our way to overcome these obstacles is to consider {\it local}
529: discs $A$ which are only partially attached to $S$. The delicate point
530: is that we have at the same time (i) to control the geometry of
531: $\partial A$ near $p\in C'$ and (ii) to guarantee that the rest of the
532: boundary stays in the region where holomorphic extension is already
533: known. In fact, (ii) will be incorporated in our very special and
534: tricky choice of $p\in C'$. For (i), we have to sharpen known existence
535: theorems about partially attached analytic discs and to combine it
536: with a careful study of the complex/real geometry of the pair
537: $(M,S)$. Importantly, our construction of such analytic discs is
538: achieved elementarily in a self-contained way. A precise description
539: of the proof in the hypersurface case (only) may be found in Section~2
540: below. With some substantial extra work, we shall generalize this
541: purely local strategy of proof to higher codimension.
542: 
543: To conclude with the removal of surfaces, let us formulate a more
544: general version of Theorem~1.1, whitout the restricted topological
545: assumption that $S$ be diffeomorphic to a real disc. Applying
546: Theorem~1.2 for the removal of $K\cap (S\backslash T_S)$ and a slight
547: generalization of Theorem~4 (ii) in~\cite{mp1} for the removal of
548: $K\cap T_S$ (more precisions will be given in \S13 below), we shall
549: obtain the following statement, implying Theorem~1.1 as a 
550: direct corollary.
551: 
552: \def\thetheorem{1.3}\begin{theorem}
553: Let $M$ be a $\mathcal{ C}^{ 2, \alpha }$-smooth globally minimal
554: hypersurface in $\C^2$, let $S \subset M$ be a $\mathcal{ C}^{2,
555: \alpha}$-smooth totally real surface, open or closed, with or without
556: boundary, which is totally real outside a discrete subset of isolated
557: complex tangencies which are hyperbolic in the sense of E.~Bishop.
558: Let $T_S$ be the union of hyperbolic points of $S$ together with all
559: separatrices issued from hyperbolic points and assume that $T_S$ does
560: not contain any subset which is homeomorphich to the unit circle. Let
561: $K$ be a proper compact subset of $S$ and assume that $\mathcal{ F
562: }_{S\backslash T_S}^c \{K \cap (S \backslash T_S)\}$ holds.
563: 
564: Then $K$ is CR-, $\mathcal{ W}$- and $L^{\sf p}$-removable.
565: \end{theorem}
566: 
567: As was already emphasized, our main motivation for this work was to
568: {\it devise a local strategy of proof for Theorems~1.1, 1.2 and~1.3 in
569: order to generalize them to higher codimension}. In fact, we will
570: realize the program sketched above for generic submanifolds of CR
571: dimension $1$ and of arbitrary codimension. Thus, let $M$ be a
572: $\mathcal{C}^{2,\alpha}$-smooth globally minimal
573: generic submanifold of codimension $(n-1)$ in $\C^n$, hence of CR
574: dimension $1$, where $n\geq 2$. Let $M^1$ be a maximally real
575: $\mathcal{C}^{2,\alpha}$-smooth one-codimensional submanifold of $M$
576: which is generic in $\C^n$. As in the surface case, $M^1$ carries a
577: {\sl characteristic foliation} $\mathcal{ F}_{M^1}^c$, whose leaves
578: are the integral curves of the line distribution $TM^1\cap
579: T^cM\vert_{M^1}$. Of course the assumption that the singularity lies
580: on one side of some characteristic segment is no longer reasonable.
581: We will generalize it as a condition requiring (approximatively
582: speaking) that there be always a characteristic segment which is
583: accessible from the complement of $C$ in $M^1$ along one direction
584: normal to the characteristic segment.
585: 
586: The generalization of Theorem~1.2, which is our
587: principal result in this paper, is as follows.
588: 
589: \def\thetheorem{{\bf 1.2'}}\begin{theorem}
590: Let $M$, $M^1$, $\mathcal{F}_{M^1}^c$ be as above and let $C$ be a
591: proper closed subset of $M$. Assume that the following topological
592: condition, meaning that $C$ is not transversal to the characteristic
593: foliation, holds{\rm :}
594: \smallskip
595: \begin{itemize}
596: \item[$\mathcal{ F}_{ M^1 }^c \{C\}:$] For every closed subset $C'
597: \subset C$, there exists a simple $\mathcal{ C}^{2, \alpha}$-smooth
598: curve $\gamma: [-1,1] \to M^1$ whose range $\gamma[-1,1]$ is
599: contained in a single leaf of the characteristic foliation
600: $\mathcal{F}_{ M^1}^c$ with $\gamma(-1)\not\in C'$, $\gamma (0) \in
601: C'$ and $\gamma (1) \not \in C'$, there exists a local $(n-
602: 1)$-dimensional transversal $R^1 \subset M^1$ to $\gamma$ passing
603: through $\gamma(0)$ and there exists a thin elongated
604: open neighborhood $V_1$ of
605: $\gamma [-1,1]$ in $M^1$ such that if $\pi_{ \mathcal{F}_{ M^1}^c}:
606: V_1 \to R^1$ denotes the semi-local projection parallel to the leaves
607: of the characteristic foliation $\mathcal{F}_{ M^1}^c$, then
608: $\gamma(0)$ lies on the boundary, relatively to the topology of $R^1$,
609: of $\pi_{\mathcal{F}_{M^1}^c}(C' \cap V_1)$.
610: \end{itemize}
611: \smallskip
612: \noindent 
613: Then $C$ is CR-, $\mathcal{ W}$- and $L^{\sf p}$-removable.
614: \end{theorem}
615: 
616: The condition $\mathcal{ F}_{ M^1}^c \{C \}$, which is of course
617: independent of the choice of the transversal $R^1$ and of the thin
618: neighborhood $V_1$, is illustrated in {\sc Figure~8} of \S5.1 below;
619: clearly, in the case $n =2$, it means that $C' \cap V_1$ lies
620: completely in one side of $\gamma [-1, 1]$, with respect to the
621: topology of $M^1$, as written in the statement of
622: Theorem~1.2. Applying some of our previous results in this direction
623: (\cite{mp1}, \cite{mp3}), we shall provide in the end of Section~13
624: below some formulations of applications of Theorem~1.2', close to
625: being analogs of Theorem~1.3 in higher codimension.
626: 
627: Importantly, in order to let the geometric condition $\mathcal{
628: F}_{M^1}^c\{C\}$ appear less mysterious and to argue that it provides
629: the adequate generalization of Theorem~1.2 to higher codimension, in
630: the last Section~14 below, we shall describe an example of $M$, $M^1$
631: and $C$ in $\C^3$ violating the condition $\mathcal{ F}_{M^1}^c\{C\}$,
632: such that $C$ is transversal to the characteristic foliation and is
633: truly nonremovable. This example will be analogous in some sense to
634: the example of a nonremovable annulus discussed after the statement of
635: Theorem~1.2. Since there is no H.~Poincar\'e and I.~Bendixson 
636: theorem for
637: foliations of $3$-dimensional balls by curves, it will be even
638: possible to insure that $M$ and $M^1$ are diffeomorphic to real balls
639: of dimension $4$ and $3$ respectively. We may therefore conclude that
640: Theorem~1.2' provides the desirable answer to the (already cited {\em
641: supra}) Problem~2.1 raised by B.~J\"oricke in~\cite{ j5}, p.~432.
642: 
643: To pursue the presentation of our results, let us comment the
644: assumption that $M$ be of codimension $(n-1)$. Geometrically speaking,
645: the study of closed singularities $C$ lying in a one-codimensional
646: generic submanifold $M^1$ of a generic submanifold $M\subset \C^n$
647: which is of CR dimension $m\geq 2$ is more simple. Indeed, thanks to
648: the fact that $M^1$ is of CR dimension $m-1\geq 1$, there exist local
649: Bishop discs {\it completely attached to $M^1$}, and this helps much in
650: describing the envelope of holomorphy of a wedge attached to
651: $M\backslash C$. On the contrary, in the case where $M$ is of CR
652: dimension $1$, small analytic discs attached to a maximally real $M^1$
653: are (trivially) inexistent. This is why, in the proof of our main
654: Theorems~1.2 and~1.2', we shall deal only with small analytic discs
655: whose boundary is in part (only) contained in $M^1$. Such discs are
656: known to exist; we would like to mention that historically speaking,
657: the first construction of discs partially attached to maximally real
658: submanifolds was exhibited by S.~Pinchuk in~\cite{p}, who developed
659: the ideas of E.~Bishop~\cite{b}.
660: 
661: Finally we will test our main Theorem~1.2' in applications. First of
662: all, we clarify its relation to known removability results in CR
663: dimension greater than one. Here the motivation is simply that most
664: questions of CR geometry should be reducible to CR dimension $1$ by
665: slicing. It turns out that the main known theorems about removable
666: singularities, due to E.~Chirka, E.~L.~Stout, and B.~J\"oricke for
667: hypersurfaces, and by the authors in higher codimension (\cite{p1},
668: Theorem~1 about $L^{\sf p}$-removability; \cite{m2}, Theorem~3 about
669: CR- and $\mathcal{ W}$- removability) are all a rather direct
670: consequence of Theorem 1.2'. Since these results have not yet been
671: published in complete form, we take the occasion of including them in
672: the present paper, {\it as a corollary of Theorem~1.2'}, yet devising
673: a new geometric approach.
674: 
675: \def\thetheorem{{\bf 1.4}}\begin{theorem} 
676: Let $M$ be a $\mathcal{ C}^{2, \alpha}$-smooth globally minimal
677: generic submanifold of $\C^n$ of CR dimension $m\geq 2$ and of
678: codimension $d=n- m\geq 1$, let $M^1 \subset M$ be a $\mathcal{ C}^{2,
679: \alpha}$-smooth one-codimensional submanifold which is generic in
680: $\C^n$ and let $C \subset M^1$ be a proper closed subset of
681: $M$. Assume that the following
682: condition holds{\rm :}
683: \begin{itemize}
684: \item[$\mathcal{ O}_{M^1}^{CR}\{C\}:$] The closed subset
685: $C$ does not contain any CR orbit of $M^1$.
686: \end{itemize}
687: Then $C$ is CR-, $\mathcal{ W}$- and $L^{\sf p}$-removable.
688: \end{theorem}
689: 
690: Notice the difference with the case $m=1$, where the analog of CR
691: orbits would consist of characteristic curves: the condition
692: $\mathcal{ F}_{M^1}^c\{C\}$ does {\it not}\, say that $C$ should not
693: contain any maximal characteristic curve. In fact, we observe that
694: there cannot exist a uniform removability statement covering both the
695: case $m=1$ and the case $m\geq 2$, whence Theorem~1.2' is stronger
696: than Theorem~1.4. Indeed, the elementary example of a nonremovable
697: circle in an annulus contained in the boundary of a strictly convex
698: domain of $\C^2$ shows that $C$ may be truly nonremovable whereas it
699: does not contain any characteristic curve. In the strictly
700: pseudoconvex hypersurface setting, it is well known that Hopf's Lemma
701: implies that boundaries of Riemann surfaces contained in $C$ (and also
702: the track on $C$ of its essential hull, {\it cf.}~\cite{d}) should be
703: everywhere transversal to the characteristic foliation. Of course, this
704: implies conversely that $C$ cannot contain such boundaries (unless
705: they are empty) if $\mathcal{ F}_{M^1}^c\{C\}$ is satisfied. The
706: reason why $\mathcal{ F}_{M^1}^c\{C\}$ implies that $C$ is removable
707: also in the nonpseudoconvex setting and in arbitrary codimension will
708: be appearant later. Finally, we mention that the $L^{\sf
709: p}$-removability of $C$ in Theorem~1.4 holds more generally with no
710: assumption of global minimality on $M$, as already noticed in~\cite{
711: j5}, \cite{ p1}, \cite{mp1}. However, since the case where $M$ is not
712: globally minimal essentially reduces to the consideration of its CR
713: orbits, which are globally minimal by definition, we shall only
714: deal with globally minimal generic submanifolds $M$ throughout this
715: paper. 
716: 
717: As a final comment, we point out that, {\it because the previously known
718: proofs of Theorem~1.4 were of local type, it is satisfactory to bring
719: in this paper a purely local framework for the treatment of
720: Theorems~1.1, 1.2, 1.3 and~1.2'}.
721: 
722: Our second group of applications concerns the classical
723: edge of the wedge theorem. Typically one considers a maximally real
724: edge $E$ to which an open double wedge $(\mathcal{ W}_1, \mathcal{
725: W}_2)$ is attached from opposite directions. One may interprete this
726: configuration as a partial thickening of a generic CR manifold
727: $M\subset E\cup \mathcal{ W}_1\cup \mathcal{ W}_2$ containing $E$ as a
728: generic hypersurface. The classical edge of the wedge theorem states
729: that functions which are continuous on $ \mathcal{ W}_1 \cup \mathcal{
730: W}_2 \cup E$ and holomorphic in $\mathcal{ W}_1 \cup \mathcal{ W}_2$
731: extend holomorphically to a neighborhood of $E$. Theorem 1.2' implies
732: that it suffices to assume continuity outside a removable
733: singularities of $E$. This allows us to derive an edge of the wedge
734: theorem for meromorphic extension (Section~13 below).
735: 
736: This paper is divided in two parts: Part I contains the strategy per
737: absurdum for the proof of Theorem~1.2', the construction of what we
738: call a semi-local half-wedge and the choice of a special point to be
739: removed locally. Part~II contains the explicit construction of
740: families of half-attached analytic discs, the end of proof of
741: Theorem~1.2' and the proofs of the various applications. The reader
742: will find a more detailed description of the content of the paper in
743: \S2.16 below.
744: 
745: \subsection*{1.5.~Acknowledgements}
746: The authors would like to thank B.~J\"oricke for several valuable
747: scientific exchanges. They acknowledge generous support from the
748: European TMR research network ERBFMRXCT 98063 and they also thank the
749: universities of Berlin (Humboldt), of G\"oteborg (Chalmers), of
750: Marseille (Provence) and of Uppsala for invitations which provided
751: opportunities for fruitful mathematical research.
752: 
753: \section*{\S2.~Description of the proof of Theorem~1.2
754: and organization of the paper}
755: 
756: The main part of this paper is devoted to the proof of Theorem~1.2',
757: which will occupy Sections~3, 4, 5, 6, 7, 8 and~9 below. In this
758: preliminary section, we shall summarize the hypersurface version of
759: Theorem~1.2', namely Theorem~1.2. Our goal is to provide a
760: conceptional description of the basic geometric constructions, which
761: should be helpful to read the whole paper. Because precise, complete
762: and rigorous formulations will be developed in the next sections, we
763: allow here the use of a slightly informal language.
764: 
765: \subsection*{2.1.~Strategy per absurdum} Let $M$, $S$, and $C$ 
766: be as in Theorem~1.2. It is essentially known that both the CR- and
767: the $L^{\sf p}$-removability of $C$ are a (relatively mild)
768: consequence of the $\mathcal{ W}$-removability of $C$ ({\it see}\,
769: \S3.14 and Section~11 below). Thus, we shall describe in this section
770: only the $\mathcal{ W}$-removability of $C$. 
771: 
772: First of all, as $M$ is globally minimal, it may be proved that for
773: every closed subset $C' \subset C$, the complement $M\backslash C'$ is
774: also globally minimal ({\it see}\, Lemma~3.5 below). As $M$ is of
775: codimension one in $\C^2$, a wedge attached to $M\backslash C$ is
776: simply a connected one-sided neighborhood of $M\backslash C$ in
777: $\C^2$. Let us denote such a one-sided neighborhood by $\omega_1$. The
778: goal is to prove that there exists a one-sided neighborhood $\omega$
779: attached to $M$ to which holomorphic functions in $\omega_1$ extend
780: holomorphically. By the definition of $\mathcal{ W}$-removability,
781: this will show that $C$ is $\mathcal{ W}$-removable.
782: 
783: Reasoning by contradiction, we shall denote by $C_{\rm nr}$ the
784: smallest {\sl nonremovable} subpart of $C$. By this we mean that
785: holomorphic functions in $\omega_1$ extend holomorphically to a
786: one-sided neighborhood $\omega_2$ of $M\backslash C_{\rm nr}$ in
787: $\C^2$ and that $C_{\rm nr}$ is the smallest subset of $C$ such that
788: this extension property holds. If $C_{\rm nr}$ is empty, the
789: conclusion of Theorem~1.2 holds, gratuitously: nothing has to be
790: proved. If $C_{ \rm nr}$ is nonempty, to come to an absurd, it
791: suffices to show that at least one point of $C_{\rm nr}$ is {\sl
792: locally removable}. By this, we mean that there exists a local
793: one-sided neighborhood $\omega_3$ of at least one point of $C_{\rm
794: nr}$ such that holomorphic functions in $\omega_2$ extend
795: holomorphically to $\omega_3$. In fact, the choice of such a point
796: will be the most delicate and the most tricky part of the proof.
797: 
798: In order to be in position to apply the continuity principle, we now
799: deform slightly $M$ inside the one-sided neighborhood $\omega_2$,
800: keeping $C_{\rm nr}$ fixed, getting a hypersurface $M^d$ (with $d$
801: like ``{\sl d}eformed'') satisfying $M^d\backslash C_{\rm nr} \subset
802: \omega_2$. We notice that a local one-sided neighborhood of $M^d$ at
803: one point $p$ of $C_{\rm nr}$ always contains a local one-sided
804: neighborhood of $M$ at $p$ (the reader may draw a figure), so we may
805: well work on $M^d$ instead of working on $M$ (however, the analogous
806: property about wedges over deformed generic submanifolds is untrue in
807: codimension $\geq 2$, {\it see}\, \S3.16 below, where supplementary
808: arguments are needed).
809: 
810: Replacing the notation $C_{\rm nr}$ by the notation $C$, the notation
811: $M^d$ by the notation $M$ and the notation $\omega_2$ by the notation
812: $\Omega$, we see that Theorem~1.2 is reduced to the following main
813: proposition, whose formulation is essentially analogous to that
814: of Theorem~1.2, except that it
815: suffices to remove at least one special point.
816: 
817: \def\theproposition{2.2}\begin{proposition}
818: Let $M$ be a $\mathcal{ C}^{ 2, \alpha}$-smooth globally minimal
819: hypersurface in $\C^2$, let $S \subset M$ be a $\mathcal{ C}^{ 2,
820: \alpha}$-smooth surface which is totally real at every point. Let $C$
821: be a {\rm nonempty} proper closed subset of $S$ and assume that the
822: nontransversality condition $\mathcal{ F }_S^c \{C\}$ formulated in
823: Theorem~1.2 holds. Let $\Omega$ be an arbitrary neighborhood of $M
824: \backslash C$ in $\C^n$. Then there exists a special point $p_{\rm sp}
825: \in C$ and there exists a local one-sided neighborhood $\omega_{ p_{
826: \rm sp}}$ of $M$ in $\C^2$ 
827: at $p_{\rm sp}$ such that holomorphic functions in
828: $\Omega$ extend holomorphically to $\omega_{ p_{ \rm sp}}$.
829: \end{proposition}
830: 
831: \subsection*{2.3.~Holomorphic extension to a half-one-sided 
832: neighborhood of $M$} The choice of the special point $p_{\rm sp}$ will
833: be achieved in two main steps. According to the nontransversality
834: assumption $\mathcal{ F}_S^c\{C\}$, there exists a characteristic
835: segment $\gamma: [-1, 1] \to S$ with $\gamma(-1)\not\in C$, with
836: $\gamma(0)\in C$ and with $\gamma(1)\not \in C$ such that $C$ lies in
837: one (closed, semi-local) side of $\gamma$ in $S$. As $\gamma$ is a
838: Jordan arc, we may orient $S$ in $M$ along $\gamma$, hence we may
839: choose a semi-local open side $(S_\gamma)^+$ of $S$ in $M$ along
840: $\gamma$. In the first main step (to be conducted in Section~4 in the
841: context of the general codimensional case Theorem~1.2'), we shall
842: construct what we call a {\sl semi-local half-wedge $\mathcal{
843: HW}_\gamma^+$ attached to $(S_\gamma)^+$ along $\gamma$}. By this, we
844: mean the ``half part'' of a wedge attached to a neighborhood of the
845: characteristic segment $\gamma$ in $M$, which yields a wedge attached
846: to the semi-local one-sided neighborhood $(S_\gamma)^+$. For an
847: illustration, {\it see}\, {\sc Figure~6} below, in which one should
848: replace the notation $M^1$ by the notation $S$. Such a half-wedge may
849: be interpreted as a wedge attached to a neighborhood of $\gamma$ in
850: $S$ which is not arbitrary, but should satisfy a further property:
851: locally in a neighborhood of every point of $\gamma$, either the
852: half-wedge contains $(S_\gamma)^+$ or one of its two ribs contains
853: $(S_\gamma)^+$, as illustrated in {\sc Figure~6} below. Importantly
854: also, the cones of this attached half-wedge should vary continuously
855: as we move along $\gamma$, {\it cf.} again {\sc Figure~6}.
856: 
857: The way how we will construct this half-wedge $\mathcal{ HW
858: }_\gamma^+$ is as follows. As illustrated in {\sc Figure~1} just
859: below, we shall first construct a string of analytic discs $Z_{ r :s }
860: ( \zeta)$, where $r$ is the approximate radius of $Z_{r: s}( \partial
861: \Delta)$, whose boundaries are contained in $(S_\gamma )^+ \subset M$
862: and which touch the curve $\gamma$ only at the point $\gamma (s)$, for
863: every $s \in [-1, 1]$, namely $Z_{r: s} (1) = \gamma (s)$ and $Z_{r
864: :s}\left(\partial \Delta \backslash \{ 1\} \right) \subset (S_\gamma
865: )^+$.
866: 
867: \bigskip
868: \begin{center}
869: \input disc-string.pstex_t
870: \end{center}
871: 
872: From now on, we fix a small radius $r_0$. By deforming the discs
873: $Z_{r_0:s}(\zeta)$ in $\Omega$ near their opposite points
874: $Z_{r_0:s}(-1)$, which lie at a positive distance from the singularity
875: $C$, we shall construct in Section~4 a family of analytic disc
876: $Z_{r_0,t:s}(\zeta)$, where $t\in \R$ is a small parameter, so that
877: the disc boundaries $Z_{r_0,t:s}(\partial \Delta)$ are pivoting
878: tangentially to $S$ at the point $\gamma(s)\equiv Z_{r_0,t:s}(1)$,
879: which remains fixed as $t$ varies. Precisely, we mean that
880: $\frac{\partial Z_{r_0,t:s}}{ \partial \theta}(1)\in T_{\gamma(s)}S$
881: and that the mapping $t\longmapsto \frac{\partial Z_{r_0,t:s}}{
882: \partial \theta}(1)$ is of rank $1$ at $t=0$. This construction and
883: the next ones will be achieved thanks to the solvability Bishop's
884: equation. Furthermore, we may add a small translation parameter $\chi
885: \in \R$, getting a family $Z_{r_0, t,\chi :s} (\zeta)$ with the
886: property that the mapping $(\chi, s) \longmapsto Z_{ r_0,t, \chi
887: :s}(1)\in S$ is a diffeomorphism onto a neighborhood of $\gamma( [-1,
888: 1])$ in $S$, still with the property that the point $Z_{r_0,t,\chi
889: :s}(1)$ is fixed equal to the point $Z_{r_0, 0,\chi:s}(1)$ as $t$
890: varies. Finally, we may add a small translation parameter $\nu\in \R$
891: with $\nu >0$, getting a family $Z_{r_0,t,\chi,\nu:s}(\zeta)$ with
892: $Z_{r_0,t,\chi,0:s}(\zeta)\equiv Z_{r_0,t,\chi:s}(\zeta)$, such that
893: the mapping $(\chi,\nu,s)\longmapsto Z_{r_0,t,\chi,\nu:s}(1)$ is a
894: diffeomorphism onto the semi-local one-sided neighborhood
895: $(S_\gamma)^+$ of $S$ along $\gamma$ in $M$, provided $\nu >0$. Then
896: the semi-local attached half-wedge may be defined as
897: \def\theequation{2.4}\begin{equation}
898: \mathcal{ HW}_\gamma^+:=\left\{
899: Z_{r_0,t,\chi,\nu:s}(\rho): \
900: \vert t \vert < \varepsilon, \
901: \vert \chi \vert < \varepsilon, \ 
902: 0< \nu < \varepsilon, \
903: 1-\varepsilon < \rho < 1, \
904: -1\leq s \leq 1
905: \right\},
906: \end{equation}
907: for some small $\varepsilon >0$. In the first main technical step (to
908: be conducted in Section~4 below in the context of Theorem~1.2'), we
909: shall show that every holomorphic function $f \in \mathcal{ O} (
910: \Omega)$ extends holomorphically to $\mathcal{ HW }_\gamma^+$. To
911: prove Proposition~2.2, we shall find a special point $p_{\rm sp} \in
912: C$ such that there exists a local one-sided neighborhood $\omega_{
913: p_{\rm sp}}$ at $p_{\rm sp}$ such that holomorphic functions in
914: $\Omega \cup \mathcal{ HW }_\gamma^+$ extend holomorphically to
915: $\omega_{ p_{\rm sp}}$.
916: 
917: \subsection*{2.5.~Field of cones on $S$}
918: We continue the description of the proof of Theorem~1.2 with the full
919: family of analytic discs $Z_{ r_0, t, \chi, \nu:s} (\zeta)$. Thanks to
920: a technical application of the implicit function theorem, we can
921: arrange from the beginning that the vectors $\frac{ \partial Z_{r_0,
922: t, \chi,0: s} }{ \partial \theta } (1)$ are tangent to $S$ at the
923: point $Z_{ r_0, 0,\chi,0:s} (1) \in S$ when $t$ varies, for all fixed
924: $s$. Then by construction, the disc boundaries
925: $Z_{r_0, t, \chi, 0:s}( \partial \Delta)$ are pivoting tangentially to
926: $S$ at the point $Z_{ r_0, t, \chi, 0 :s} (1) \equiv Z_{r_0, 0, \chi,
927: 0:s} (1)$. It follows that when $t$ varies, the oriented half-lines
928: $\R^+ \cdot \frac{ \partial Z_{ r_0,t,\chi,0 : s} }{ \partial \theta}
929: (1)$ describe an open infinite oriented cone in the tangent space to
930: $S$ at the point $Z_{r_0, 0, \chi, 0:s}(1)$. Consequently, we may
931: define a {\sl field of cones} $p \mapsto {\sf C}_p$ as
932: \def\theequation{2.6}\begin{equation}
933: {\sf C}_p:= \left\{ \R^+ \cdot 
934: \frac{ \partial Z_{r_0,t,\chi,0: s} }{
935: \partial \theta} (1) : \ \vert
936: t \vert < \varepsilon \right\},
937: \end{equation}
938: at every point $p=Z_{ r_0, 0, \chi, 0: s} (1) \in S$ of a neighborhood
939: of $\gamma$ in $S$. The following figure provides an intuitive
940: illustration. One should think that the small cones are generated when
941: the small discs boundaries of {\sc Figure~1} pivote tangentially to
942: $S$.
943: 
944: \bigskip
945: \begin{center}
946: \input field-cones-hypersurface.pstex_t
947: \end{center}
948: 
949: After having defined this field of cones, we shall {\sl fill} the
950: cones as follows. Remind that a neighborhood of $\gamma$ in $S$ is
951: foliated by characteristic segments, which are approximatively
952: parallel to $\gamma$. In {\sc Figure~2} above, one should think that
953: the characteristic foliation is horizontal. So there exists a nowhere
954: vanishing vector field $p \mapsto X_p$ defined in a neighborhood of
955: $\gamma$ whose integral curves are characteristic segments. We define
956: the {\sl filled cone} ${\sf FC}_p$ by
957: \def\theequation{2.7}\begin{equation}
958: {\sf FC}_p:=\left\{
959: \lambda \cdot X_p + (1-\lambda) \cdot v_p : \ 
960: 0 \leq \lambda < 1, \
961: v_p\in {\sf C}_p
962: \right\}.
963: \end{equation}
964: Geometrically, we rotate every half-line $\R^+ \cdot v_p$ towards the
965: characteristic half-line $\R^+ \cdot X_p$ and we call the result the
966: {\sl filling} of ${\sf C}_p$. In {\sc Figure~2} above, all the cones
967: ${\sf C}_p$ coincide with their fillings. Thus we have constructed a
968: field of filled cones $p \longmapsto {\sf FC}_p$ over a neighborhood
969: of $\gamma$ in $S$.
970: 
971: \subsection*{2.8.~Small analytic discs half-attached to $S$}
972: The next main observation is that small analytic discs which are
973: half-attached to $S$ are essentially contained in the half-wedge
974: $\mathcal{ HW }_\gamma^+$, provided that they are approximatively
975: directed by the filled cone ${\sf FC}_p$. Let us be more precise. Let
976: $\partial^+ \Delta:= \{\zeta \in \partial \Delta: \ {\rm Re} \, \zeta
977: \geq 0 \}$ denote the {\sl positive half part} of the unit circle. We
978: say that an analytic disc $A: \overline{ \Delta } \to \C^2$ is {\sl
979: half-attached} to $S$ if $A( \partial^+ \Delta)$ is contained in
980: $S$. Here, $A$ is at least of class $\mathcal{ C}^1$ over
981: $\overline{\Delta}$ and holomorphic in $\Delta$. In addition, we shall
982: always assume that our discs $A$ are embeddings of $\overline{\Delta}$
983: into $\C^2$. We shall say that $A$ is {\sl approximatively straight}
984: (in an informal sense) if $A(\Delta)$ is close in $\mathcal{
985: C}^1$-norm to an open subset of the complex line generated by the
986: complex vector $\frac{\partial A}{ \partial \zeta}(1)$. Finally, we
987: say that $A$ is {\sl approximatively directed by the filled cone ${\sf
988: FC}_p$ at $p=A(1)$}, if the vector $\frac{\partial A}{\partial
989: \theta}(1)\in T_pS$ belongs to ${\sf FC}_p$. Although this
990: terminology will not be re-employed in the next sections, we may
991: formulate a crucial geometric observation as follows.
992: 
993: \def\thelemma{2.9}\begin{lemma}
994: A sufficiently small approximatively straight analytic disc $A:
995: \overline{ \Delta} \to \C^2$ of class at least $\mathcal{ C}^1$ which
996: is half-attached to $S$ and which is approximatively directed by the
997: filled cone ${\sf FC}_p$ at $p=A(1)\in S$, necessarily satisfies
998: \def\theequation{2.10}\begin{equation}
999: A\left(\overline{ \Delta } \backslash \partial^+ \Delta\right) 
1000: \subset \mathcal{ HW}_\gamma^+.
1001: \end{equation}
1002: \end{lemma}
1003: 
1004: In the context of the general Theorem~1.2',
1005: this property (with more precisions) will be checked in 
1006: Section~8 below. 
1007: 
1008: \subsection*{2.11.~Choice of a special point}
1009: In the second main step of the proof (to be conducted in Section~5 in
1010: the context of Theorem~1.2'), we shall choose the desired special
1011: point $p_{ \rm sp}$ of Proposition~2.2 to be removed locally as
1012: follows. Since we shall remove $p_{ \rm sp}$ by means of half-attached
1013: analytic discs (applying the continuity principle), we want to find a
1014: special point $p_{ \rm sp}\in C$ so that the following two conditions
1015: hold true:
1016: \begin{itemize}
1017: \item[{\bf (i)}]
1018: There exists a small approximatively straight analytic disc $A:
1019: \overline{\Delta} \to \C^2$ with $A(1)= p_{\rm sp}$ which is
1020: half-attached to $S$ such that $A$ is approximatively directed by the
1021: filled cone ${\sf FC}_{ p_{ \rm sp}}$ (so that the conclusion of
1022: Lemma~2.9 above holds true).
1023: \item[{\bf (ii)}]
1024: The same disc satisfies $A \left( \partial^+ \Delta \backslash \{1\}
1025: \right) \subset S \backslash C$.
1026: \end{itemize}
1027: 
1028: In particular, since $M \backslash C$ is contained in $\Omega$, it
1029: follows from these two conditions that the blunt disc boundary
1030: $A\left( \partial \Delta \backslash \{ 1 \} \right)$ is contained in
1031: the open subset $\Omega \cup \mathcal{ HW }_\gamma^+$, a property that
1032: will be appropriate for the application of the continuity principle,
1033: as we shall explain in Section~9 below.
1034: 
1035: To fulfill conditions {\bf (i)} and {\bf (ii)} above, we first
1036: construct a supporting real segment at a special point of the nonempty
1037: closed subset $C\subset S$.
1038: 
1039: \def\thelemma{2.12}\begin{lemma}
1040: There exists at least one special point $p_{ \rm sp } \in C$
1041: arbitrarily close to $\gamma$ in a neighborhood of which the following
1042: two properties hold true{\rm :}
1043: \begin{itemize}
1044: \item[{\bf (i')}]
1045: There exists a small $\mathcal{ C}^{2, \alpha}$-smooth open segment
1046: $H_{p_{\rm sp}}\subset S$ passing through $p_{ \rm sp}$ such that an
1047: oriented tangent half-line to $H_{p_{\rm sp}}$ at $p_{\rm sp}$ is
1048: contained in the filled cone ${\sf FC}_{p_{\rm sp}}$, as illustrated
1049: in {\sc Figure~3} below.
1050: \item[{\bf (ii')}]
1051: The same segment is a supporting segment in the following
1052: sense{\rm :} locally in a
1053: neighborhood of $p_{\rm sp}$, the set $C \backslash \{p_{\rm sp}\}$ is
1054: contained in one open side $(H_{ p_{\rm sp}})^-$ if $H_{p_{\rm sp}}$
1055: in $S$, as illustrated in {\sc Figure~3} just below.
1056: \end{itemize}
1057: \end{lemma}
1058: 
1059: \bigskip
1060: \begin{center}
1061: \input one-touch.pstex_t
1062: \end{center}
1063: 
1064: The way how we prove Lemma~2.12 is illustrated intuitively in {\sc
1065: Figure~2} above. For $\lambda\in \R$ with $0 \leq \lambda < 1$ very
1066: close to $1$, the vector field $p \longmapsto v_p^\lambda := \lambda
1067: \cdot X_p + (1-\lambda) \cdot v_p$ is very close to the characteristic
1068: vector field $p\mapsto X_p$. By construction, this vector field runs
1069: into the filled field of cones $p\mapsto {\sf FC}_p$. In {\sc
1070: Figure~2}, the integral curves of $p\mapsto v_p^\lambda$ are drawn as
1071: dotted lines, which are almost horizontal if $\lambda$ is very close
1072: to $1$. If we choose the first dotted integral curve from the lower
1073: part of {\sc Figure~2} which touches $C$ at one special point $p_{\rm
1074: sp}\in C$ and if we choose for $H_{p_{\rm sp}}$ a small segment of
1075: this first dotted integral curve, we may check that properties {\bf
1076: (i)} and {\bf (ii)} are satisfied, modulo some mild technicalities. A
1077: rigorous complete proof of Lemma~2.12 will be provided in Section~5
1078: below.
1079: 
1080: \subsection*{2.13.~Construction of analytic discs
1081: half-attached to $S$} Small analytic discs which are half-attached to
1082: a $\mathcal{ C}^{2,\alpha}$-smooth maximally real submanifold $M^1$ of
1083: $\C^n$ and which are approximatively straight will be constructed in
1084: Section~7 below. For this, we shall use the solution of Bishop's
1085: equation with parameters in H\"older spaces, obtained by A.~Tumanov
1086: in~\cite{ tu3} with an optimal loss of regularity. In Section~8, we
1087: shall check that it follows from the general constructions of
1088: Section~7 that there exists a small analytic disc $A$ half-attached to
1089: $S$ with $A(1)= p_{\rm sp}$ whose half boundary is tangent to
1090: $H_{p_{\rm sp}}$ at $p_{\rm sp}$ and which satisfies property {\bf
1091: (ii)} above, as drawn in {\sc Figure~3} above. Thus, the two geometric
1092: properties {\bf (i')} and {\bf (ii')} satisfied by the real segment
1093: $H_{p_{\rm sp}}$ may be realized by the half-boundary of a
1094: half-attached analytic disc.
1095: 
1096: \subsection*{2.14.~Translation of half-attached and continuity 
1097: principle} By means of the results of Section~7, we shall see that we
1098: may include the disc $A (\zeta)$ is a parametrized family $A_{ x,v}
1099: (\zeta)$ of analytic discs half-attached to $S$, where $x\in \R^2$ and
1100: $v \in \R$ are small, so that the mapping $x \mapsto A_{ x,0} (1) \in
1101: S$ is a local diffeomorphism onto a neighborhood of $p_{ \rm sp}$ in
1102: $S$ and so that the mapping $v \mapsto \frac{ \partial A_{ 0,v }}{
1103: \partial \theta} (1)$ is of rank $1$ at $v=0$. Furthermore, we
1104: introduce a new parameter $u \in \R$ in order to ``translate'' the
1105: totally real surface $S$ in $M$ by means of a family of $S_u \subset
1106: M$ with $S_0=S$ and $S_u \subset ( S_\gamma )^+$ for $u>0$. Thanks to
1107: the tools developed in Section~7, we deduce that there exists a
1108: deformed family of analytic discs $A_{ x, v, u} (\zeta)$ which are
1109: half-attached to $S_u$ and which satisfies $A_{x, v,0} (\zeta) \equiv
1110: A_{ x,v} (\zeta)$. In particular, this family covers a local
1111: one-sided neighborhood $\omega_{p_{\rm sp}}$ of $M$ at $p_{\rm sp}$
1112: defined by
1113: \def\theequation{2.15}\begin{equation}
1114: \omega_{p_{\rm sp}}:= \left\{
1115: A_{x,v,u}(\rho): \ 
1116: \vert x \vert < \varepsilon, \ 
1117: \vert v \vert < \varepsilon, \
1118: \vert u \vert < \varepsilon, \ 
1119: 1-\varepsilon < \rho < 1
1120: \right\},
1121: \end{equation}
1122: for some $\varepsilon>0$.
1123: 
1124: In the third and last main step of the proof (to be conducted in
1125: Section~9 below), we shall prove that every disc $A_{x, v,u} (\zeta)$
1126: with $u \neq 0$ is analytically isotopic to a point with the boundary
1127: of every disc of the isotopy being contained in $\Omega \cup \mathcal{
1128: HW}_\gamma^+$. Thanks to the continuity principle, we shall deduce
1129: that every holomorphic function $f \in \mathcal{ O}\left( \Omega \cup
1130: \mathcal{ HW}_\gamma^+ \right)$ extends holomorphically to $\omega_{
1131: p_{\rm sp}}$ minus a certain thin closed subset $\mathcal{ C}_{p_{\rm
1132: sp}}$ of $\omega_{ p_{\rm sp}}$. Finally, we shall conclude both the
1133: proof of Proposition~2.2 and the proof of Theorem~1.2 by checking that
1134: the thin closed set $\mathcal{ C}_{p_{\rm sp}}$ is in fact removable
1135: for holomorphic functions defined in $\omega_{ p_{\rm sp }} \backslash
1136: \mathcal{ C}_{ p_{ \rm sp}}$.
1137: 
1138: \subsection*{2.16.~Organization of the paper} 
1139: As was already announced, Sections~3, 4, 5, 6, 7, 8 and~9 below will be
1140: entirely devoted to the proof of Theorem~1.2', which will be
1141: endeavoured directly in arbitrary codimension, without any further
1142: reference to the hypersurface version. Only in Section~3 shall we also
1143: consider the beginning of the proof of Theorem~1.4. During the
1144: development of the proof of Theorem~1.2', in comparison to the quick
1145: description of the proof of Theorem~1.2 achieved just above, we shall
1146: unavoidably encounter some supplementary technical complications caused
1147: by the codimension being 
1148: $\geq 2$, namely technicalities which are absent in
1149: codimension $1$. We would like to mention that the crucial geometric
1150: argument which enables us to choose the desired special point will be
1151: conducted in the central Section~5 below.
1152: 
1153: Then Section~10 is devoted to summarize three geometrically distinct
1154: proofs of Theorem~1.4. In Section~11, we check that both the CR- and
1155: the $L^{\sf p}$-removability of $C$ are a consequence of the
1156: $\mathcal{ W}$-removability of $C$. In Section~12, we
1157: provide the proof of Theorem~1.1, of Theorem~1.3 and of further
1158: applications. This Section~12 may be read before entering the proof of
1159: Theorem~1.2'. Finally, in Section~13, we provide some applications of
1160: our removability results to the edge of the wedge theorem for
1161: meromorphic functions.
1162: 
1163: \section*{\S3.~Strategy per absurdum
1164: for the proofs of Theorems~1.2' and~1.4} 
1165: 
1166: \subsection*{3.1.~Preliminary}
1167: For the proof of Theorems~1.2', as in \cite{cs}, \cite{m2},
1168: \cite{mp1}, \cite{mp3}, \cite{p1}, we shall proceed by contradiction.
1169: This strategy possesses a considerable advantage: it will not be
1170: necessary to control the size of the local subsets of $C$ that are
1171: progressively removed, which simplifies substantially the presentation
1172: and the understandability of the reasonings. We shall explain how to
1173: reduce CR- and $L^{\sf p}$-removability of $C$ to its $\mathcal{
1174: W}$-removability. Also, it may be argued that the $\mathcal{
1175: W}$-removability of $C$ can be reduced to the simpler case where the
1176: functions which we have to extend are even holomorphic in a
1177: neighborhood of $M\backslash C$ in $\C^n$. Whereas such a strategy is
1178: essentially carried out in detail in previous references (with some
1179: variations), we shall for completeness recall the complete reasonings
1180: briefly here, in \S3.2 and in \S3.16 below.
1181: 
1182: \subsection*{3.2.~Global minimality of some complements} 
1183: For background notions about CR orbits in CR manifolds, we refer the
1184: reader to~\cite{su}, \cite{j2}, \cite{mp1}, \cite{j4}. We just recall
1185: a few standard facts: if $p$ belongs to a generic submanifold of
1186: $\C^n$ of class at least $\mathcal{ C}^2$, a point $q\in M$ belongs to
1187: the CR orbit $\mathcal{ O}_{CR}(M,p)$ if and only if there exists a
1188: piecewise smooth curve $\lambda: [0,1] \to M$ with $\lambda(0)=p$,
1189: $\lambda(1)=q$ such that $d\lambda(s)/ds\in T_{\gamma(s)}^cM
1190: \backslash \{0\}$ at every $s\in [0,1]$ at which $\lambda$ is
1191: differentiable; CR orbits make a partition of $M$; CR orbits are
1192: immersed $\mathcal{ C}^{1,\alpha}$-smooth submanifolds of $M$,
1193: according to H.J.~Sussmann's Theorem~4.1 in~\cite{su} specialized in
1194: the CR category; Every maximal $T^cM$-integral immersed submanifold of
1195: $M$ must contain the CR orbit of each of its points; and finally, a
1196: trivial, but often useful fact: if $N$ is a $T^cM$-integral
1197: submanifold of $M$, namely $T_pN \subset T_p^cM$ for every point $p\in
1198: N$, then the local flow of every $T^cM$-tangent vector field on $M$
1199: stabilizes locally $N$.
1200: 
1201: In the two geometric situations of Theorems~1.2' and~1.4, we shall
1202: apply the following two ~Lemmas~3.3 and~3.5 about the CR structure of
1203: the complement $M\backslash C'$, where $C'\subset C\subset M^1$ is an
1204: arbitrary proper closed subset of $C$.
1205: 
1206: \def\thelemma{3.3}\begin{lemma} Let $M$ be a $\mathcal{ C}^{2,
1207: \alpha}$-smooth generic submanifold of $\C^n$ $(n\geq 2)$ of
1208: codimension $d\geq 1$ and of CR dimension $m:=n-d\geq 1$, let $M^1$ be
1209: a $\mathcal{ C}^{2,\alpha}$-smooth 
1210: one-codimensional submanifold of $M$ which is generic in $\C^n$ and
1211: let $C'$ be an arbitrary proper closed subset of $M^1$. If either
1212: \begin{itemize}
1213: \item[{\bf (1)}]
1214: $M$ is of CR dimension $m=1$ and the condition 
1215: $\mathcal{ F}_{M^1}^c\{C'\}$ of Theorem~1.2' holds; or if
1216: \item[{\bf (2)}]
1217: $M$ is of CR dimension $m\geq 2$ and the condition 
1218: $\mathcal{ O}_{M^1 }^{CR }\{C'\}$ of Theorem~1.4 holds,
1219: \end{itemize}
1220: then for each point $p'\in C'$, there exists a piecewise $\mathcal{
1221: C}^{2, \alpha}$-smooth curve $\gamma: [0,1]\to M^1$ satisfying $d
1222: \gamma(s)/ds\in T_{\gamma(s) }M^1 \cap T_{\gamma(s)}^c M\backslash
1223: \{0\}$ at every $s\in [0,1]$ at which $\lambda$ is differentiable,
1224: such that $\gamma(0)=p'$ and $\gamma(1)$ does not belong to $C'$.
1225: \end{lemma}
1226: 
1227: \proof
1228: In the case $m=1$, we proceed by contradiction and we suppose that
1229: there exists a point $p'\in C'$ such that all piecewise $\mathcal{
1230: C}^{2,\alpha}$-smooth curves $\gamma: [0,1]\to M^1$ with $d\gamma(s)
1231: /ds\in T_{\gamma(s)}M^1\cap T_{\gamma(s)}^cM\backslash \{0\}$ at every
1232: $s\in [0,1]$ at which $\gamma$ is differentiable, which have origin
1233: $p'$ are entirely contained in $C'$. Notice that since the bundle
1234: $TM^1\cap T^cM\vert_{M^1}$ is of real dimension one and of class
1235: $\mathcal{ C}^{1,\alpha}$, such curves $\gamma$ are in fact $\mathcal{
1236: C}^{2,\alpha}$-smooth at every point. It follows immediately there
1237: cannot exist a curve $\gamma: [-1,1] \to M^1$ contained in a single
1238: leaf of the characteristic foliation $\mathcal{ F}_{M^1}^c$ with
1239: $\gamma(0)\in C'$ and $\gamma(-1), \gamma(1)\not \in C'$, which
1240: contradicts the condition $\mathcal{ F}_{M^1}^c\{C'\}$.
1241: 
1242: In the case $m\geq 2$, by genericity of $M^1$, the complex tangent
1243: bundle $T^cM^1$, which is of real dimension $(2m-2)$, is a
1244: one-codimensional subbundle of the $(2m-1)$-dimensional bundle
1245: $TM^1\cap T^cM\vert_{M^1}$, namely $T^cM^1\subset TM^1 \cap
1246: T^cM\vert_{M^1}$. Let $p\in C'$ be an arbitrary point. By the
1247: assumption $\mathcal{ O}_{M^1}^{CR}\{C'\}$, the CR orbit of $p'$ is
1248: not contained in $C'$. Equivalently, there exists a piecewise
1249: $\mathcal{ C}^{2,\alpha}$-smooth curve $\gamma: [0,1] \to M^1$
1250: satisfying
1251: \def\theequation{3.4}\begin{equation}
1252: d\gamma(s)/ds\in T_{\gamma(s)}^cM^1\backslash \{0\}\subset
1253: T_{\gamma(s)}M^1 \cap T_{\gamma(s)}^cM\backslash \{0\} 
1254: \end{equation} 
1255: at each $s\in [0,1]$ at which $\gamma$ is
1256: differentiable,
1257: such that $\gamma(0)=p'$ and $\gamma(1)$ does not belong to $C'$.
1258: Hence the conclusion of Lemma~3.3 is immediately satisfied.
1259: This completes the proof.
1260: \endproof
1261: 
1262: As an application, we deduce that under the respective assumptions
1263: $\mathcal{ F}_{M^1}^c\{C\}$ and $\mathcal{ O}_{M^1}^{CR}\{C\}$ of
1264: Theorems~1.2' and of Theorem~1.4, the complement $M \backslash C'$ is
1265: globally minimal, for every closed subset $C'\subset C$.
1266: 
1267: \def\thelemma{3.5}\begin{lemma} 
1268: Let $M$ be a $\mathcal{ C}^{2, \alpha}$-smooth generic submanifold of
1269: $\C^n$ $(n \geq 2)$ of codimension $d\geq 1$ and of CR dimension
1270: $m:=n-d \geq 1$, let $M^1$ be a $\mathcal{ C}^{2,\alpha}$-smooth
1271: one-codimensional submanifold of $M$ which is generic in $\C^n$ and
1272: let $C'$ be an arbitrary {\rm nonempty} proper closed subset of
1273: $M^1$. Assume that for each point $q'\in C'$, there exists a piecewise
1274: $\mathcal{ C}^{2,\alpha}$-smooth curve $\gamma: [0,1] \to M^1$ with
1275: $d\gamma(s) / ds\in T_{ \gamma(s)}M^1\cap T_{ \gamma(s)}^cM\backslash
1276: \{0\}$ at every $s\in [0,1]$ at which 
1277: $\gamma$ is differentiable, such that $\gamma(0)=q'$ and $\gamma(1)$
1278: does not belong to $C'$. Then the CR orbit in $M\backslash C'$ of
1279: every point $p\in M\backslash C'$ coincides with its CR orbit in $M$
1280: minus $C'$, namely
1281: \def\theequation{3.6}\begin{equation}
1282: \mathcal{ O}_{CR}(M\backslash C', p)= 
1283: \mathcal{ O}_{CR}(M,p)\backslash C'.
1284: \end{equation}
1285: In particular, as a corollary, if $M$ is globally minimal, then
1286: $M\backslash C'$ is also globally minimal.
1287: \end{lemma}
1288: 
1289: \proof
1290: Let us first explain the last sentence, which applies
1291: to the situations considered in both Theorems~1.2' and~1.4: 
1292: if $\mathcal{ O}_{CR}(M,p)=M$, then by~\thetag{3.6}, 
1293: $\mathcal{ O}_{CR}(M\backslash C',p)\equiv
1294: M\backslash C'$, which proves that
1295: $M\backslash C'$ is globally minimal.
1296: 
1297: To establish~\thetag{3.6}, we shall need the following crucial lemma,
1298: deserving an illustration: {\sc Figure~4} below.
1299: 
1300: \def\thelemma{3.7}\begin{lemma}
1301: Under the assumptions of Lemma~3.5, for every point $q'\in C' \subset
1302: M^1$, there exists a $\mathcal{ C}^{1,\alpha}$-smooth locally embedded
1303: submanifold $\Omega_{q'}$ of $M$ passing through $q'$ which is
1304: transverse to $M^1$ at $q'$ in $M$, namely which satisfies
1305: $T_{q'}\Omega_{q'}+T_{q'}M^1= T_{q'}M$, such that
1306: \begin{itemize}
1307: \item[{\bf (1)}]
1308: $\Omega_{q'}$ is a $T^cM$-integral submanifold, namely 
1309: $T_p^cM\subset T_p\Omega_{q'}$, for every point 
1310: $p\in \Omega_{q'}$.
1311: \item[{\bf (2)}]
1312: $\Omega_{q'}\backslash C'$ is contained in 
1313: a single CR orbit of $M$.
1314: \item[{\bf (3)}] $\Omega_{q'}$ is
1315: also contained in a single CR orbit of $M\backslash C'$.
1316: \end{itemize}
1317: \end{lemma}
1318: 
1319: \proof
1320: So, let $q'\in C'\subset M^1$. Since $M^1$ is generic in $\C^n$, there
1321: exists a $\mathcal{ C}^{1,\alpha}$-smooth vector field $Y$ defined in
1322: a neighborhood of $q'$ which is complex tangent to $M$ but locally
1323: transversal to $M^1$, {\it cf.} {\sc Figure~4} just below (for easier
1324: readability, we have erased the hatching of $C'$ in a neighborhood of
1325: $q'$).
1326: 
1327: \bigskip
1328: \begin{center}
1329: \input orbit-complement.pstex_t
1330: \end{center}
1331: 
1332: \noindent
1333: Following the integral curve of $Y$ issued from $q'$, we can define a
1334: point $q_\epsilon'$ in an $\epsilon$-neighborhood of $q'$ which does
1335: not belong to $M^1$. By assumption, there exists a piecewise smooth
1336: curve $\gamma: [0,1]\to M^1$ with $d\gamma(s) /ds \in T_{\gamma(s)}M^1
1337: \cap T_{\gamma(s)}^cM\backslash \{0\}$ at every $s\in [0,1]$ at which
1338: $\gamma$ is differentiable, such that $\gamma(0)=q'$ and $\gamma(1)$
1339: does not belong to $C'$. For simplicity, we shall assume that
1340: $\gamma$ consists of a single smooth piece, the case where $\gamma$
1341: consists of finitely many smooth pieces being treated in a completely
1342: similar way. With this assumption (which will simplify slightly the
1343: technicalities), it follows that there exists a vector field $X$
1344: defined in a neighborhood of the curve $\gamma([0,1])$ in $M$ which is
1345: complex tangent to $M$ and whose restriction to $M^1$ is a semi-local
1346: section of $TM^1\cap T^cM\vert_{M^1}$, such that $\gamma$ is an
1347: integral curve of $X$ and such that $\gamma(1)=\exp(X)(q')\in
1348: M^1\backslash C'$. In addition, we can assume that the vector field
1349: $Y$ is defined in the same neighborhood of $\gamma([0,1])$ in $M$ and
1350: everywhere transversal to $M^1$. If $\epsilon$ is sufficiently small,
1351: {\it i.e.} if $q_\epsilon'$ is sufficiently close to $q'$, the point
1352: $r_\epsilon':= \exp(X)(q_\epsilon')$ is still very close to
1353: $M^1$. Thus, we can define a new point $r'\in M^1$ to be the unique
1354: intersection of the integral of $Y$ issued from $r_\epsilon'$ with
1355: $M^1$. By choosing $\epsilon$ small enough, the point $r_\epsilon'$
1356: will be arbitrarily close to $\gamma(1)\not \in C'$, and consequently,
1357: we can assume that $r'$ also does not belong to $C'$, as drawn in {\sc
1358: Figure~4} above. Notice that the integral curve of $X$ from
1359: $q_\epsilon'$ to $r_\epsilon'$ is contained in $M\backslash M^1$,
1360: since the flow of $X$ stabilizes $M^1$, whence the two points
1361: $r_\epsilon'$ and $r'$ belong to the CR orbit $\mathcal{
1362: O}_{CR}(M\backslash C',q_\epsilon')$.
1363: 
1364: Let $\Omega_{r'}$ denote a small piece of the immersed submanifold
1365: $\mathcal{ O}_{CR}(M\backslash C',r')$ passing through $r'$. By the
1366: standard properties of CR orbits, we can assume that $\Omega_{r'}$ is
1367: an embedded $\mathcal{ C}^{1,\alpha}$-smooth submanifold of
1368: $M\backslash C'$ of the same CR dimension as $M\backslash C'$ and we
1369: can in addition assume that $\Omega_{r'}$ contains $r_\epsilon'$,
1370: provided $\epsilon$ is small enough. Since $Y$ is a vector field
1371: complex tangent to $M$, the submanifold $\Omega_{r'}$ is necessarily
1372: stretched along the flow lines of $Y$, hence it is transversal to
1373: $M^1$.
1374: 
1375: We then define the submanifold $\exp(-X)(\Omega_{r'})$, close to the
1376: point $q'$ (we shall argue in a while that it passes in fact through
1377: $q'$). Since the flow of $X$ stabilizes $M^1$, it follows that
1378: $\exp(-X)(\Omega_{r'})$ is transversal to $M^1$ and that
1379: $\exp(-X)(\Omega_{r'})$ is divided in two sides by its
1380: one-codimensional $\mathcal{ C}^{1,\alpha}$-smooth submanifold
1381: $M^1\cap \exp(-X)(\Omega_{r'})$. Furthermore, the flow of $X$
1382: stabilizes the two sides of $M^1$ in $M$, semi-locally in a
1383: neighborhood of $\gamma([0,1])$, {\it see} again {\sc Figure~4} above.
1384: Consequently, every integral curve of $X$ issued from every point in
1385: $\Omega_{r'}\backslash M^1$ stays in $M\backslash M^1$, hence in
1386: $M\backslash C'$ and it follows that the submanifold
1387: \def\theequation{3.8}\begin{equation}
1388: \exp(-X)(\Omega_{r'})\backslash M^1,
1389: \end{equation}
1390: consisting of two connected pieces, is contained in the single CR
1391: orbit $\mathcal{ O}_{CR}(M\backslash C', p')$. By the characteristic
1392: property of a CR orbit, this means that the two connected pieces of
1393: $\exp(-X)(\Omega_{r'})\backslash M^1$ are CR submanifolds of
1394: $M\backslash C'$ of the same CR dimension as $M\backslash C'$.
1395: Furthermore, since the intersection $M^1\cap \exp(-X)(\Omega_{r'})$ is
1396: one-codimensional, it follows by continuity that {\it the $\mathcal{
1397: C}^{1,\alpha}$-smooth submanifold $\exp(-X) (\Omega_{r'})$ is in fact a
1398: CR submanifold of $M$ of the same CR dimension as $M$}. This 
1399: proves property {\bf (1)}. 
1400: 
1401: Since $q_\epsilon'$ belongs to $\exp (-X) (\Omega_{r'})$ and since the
1402: flow of the complex tangent vector field $Y$ necessarily stabilizes
1403: the $T^cM$-integral submanifold $\exp( -X) (\Omega_{ r'})$, the point
1404: $q'$ which belongs to an integral curve of $Y$ issued from
1405: $q_\epsilon'$, must belong to the submanifold $\exp(-X)( \Omega_{
1406: r'})$, which we can now denote by $\Omega_{ q'}$, as in {\sc
1407: Figure~4} above.
1408: 
1409: Observe that locally in a neighborhood of $q'$, the integral curves of
1410: $Y$ are transversal to $M^1$ and meet $M^1$ only at one point.
1411: Shrinking if necessary $\Omega_{q'}$ a little bit and using positively
1412: or negatively oriented integral curves of $Y$ with origin all points
1413: in $\Omega_{q'} \backslash M^1$ lying in both sides, we deduce that
1414: $\Omega_{q'}\backslash C'$ is contained in the single CR orbit
1415: $\mathcal{ O}_{CR}(M\backslash C', r')$, which proves property {\bf
1416: (3)}. Using again $Y$ to join points of $C'\cap \Omega_{q'}$, we
1417: deduce also that $\Omega_{q'}$ is contained in the single CR orbit
1418: $\mathcal{ O}_{CR}(M,r')$, which proves property {\bf (2)}.
1419: 
1420: The proof of Lemma~3.7 is complete.
1421: \endproof
1422: 
1423: We can now prove Lemma~3.5. It suffices to establish that for every two
1424: points $p \in M \backslash C'$ and $q \in \mathcal{ O}_{ CR} (M, p)$
1425: with $q\not\in C'$, the point $q$ belongs in fact to the CR orbit of
1426: $p$ in $M\backslash C'$, namely $q$ belongs to 
1427: $\mathcal{ O}_{CR}(M\backslash C', p)$.
1428: 
1429: Since $q$ belongs to the CR orbit of $p$ in $M$, there exists a
1430: piecewise $\mathcal{ C}^{2,\alpha}$-smooth curve $\lambda: [0,1] \to
1431: M$ with $\lambda(0)=p$, $\lambda(1)=q$ and $d\lambda(s)/ds\in
1432: T_{\lambda(s)}^cM\backslash \{0\}$ at every $s\in [0,1]$ at which
1433: $\lambda$ is differentiable. For every $s$ with $0\leq s \leq 1$, we
1434: define a local $\mathcal{ C}^{1,\alpha}$-smooth submanifold
1435: $\Omega_{\lambda(s)}$ of $M$ passing through $\lambda(s)$ as follows:
1436: \begin{itemize}
1437: \item[{\bf (1)}]
1438: If $\lambda(s)$ does not belong to $C'$, choose for $\Omega_{
1439: \lambda(s)}$
1440: a piece of the CR orbit of $\lambda(s)$ in $M\backslash C'$.
1441: \item[{\bf (2)}]
1442: If $\lambda(s)$ belongs to $C'$, choose for $\Omega_{\lambda(s)}$
1443: the submanifold constructed in Lemma~3.7 above.
1444: \end{itemize}
1445: By Lemma~3.7, for each $s$, the complement
1446: $\Omega_{\lambda(s)}\backslash C'$ is contained in a single CR orbit
1447: of $M\backslash C'$. Since each $\Omega_{\lambda(s)}$ is a
1448: $T^cM$-integral submanifold, for each $s\in [0,1]$, a neighborhood of
1449: $\lambda(s)$ in the arc $\lambda([0,1])$ is necessarily contained
1450: $\Omega_{\gamma(s)}$. By the Borel-Lebesgue covering lemma, we can
1451: therefore find an integer $k\geq 1$ and real numbers
1452: \def\theequation{3.9}\begin{equation}
1453: 0=s_1<r_1<t_1<s_2<r_2<t_2<\cdots\cdots <
1454: s_{k-1} < r_{k-1} < t_{k-1} < s_k=1,
1455: \end{equation}
1456: such that
1457: $\lambda([0,1])$ is covered by $\Omega_{\lambda(0)} \cup
1458: \Omega_{\lambda(s_2)} \cup \cdots \cup \Omega_{\lambda(s_{k-1})}\cup
1459: \Omega_{\lambda(1)}$ and such that in addition, 
1460: $\lambda([r_j,t_j])\subset \Omega_{\lambda(s_j)}\cap \,
1461: \Omega_{\lambda(s_{j+1})}$ for $j=1,\dots,k-1$.
1462: 
1463: \def\thelemma{3.10}\begin{lemma}
1464: The following union minus $C'$
1465: \def\theequation{3.11}\begin{equation}
1466: \left(
1467: \Omega_{\lambda(0)} \cup \Omega_{\lambda(s_2)}\cup \cdots\cdots
1468: \cup \Omega_{\lambda(s_{k-1})}\cup
1469: \Omega_{\lambda(1)}\right)\backslash C'
1470: \end{equation}
1471: is contained in a single CR orbit of $M\backslash C'$.
1472: \end{lemma}
1473: 
1474: \proof
1475: It suffices to prove that for every $j=1,\dots,k-1$, the union $\left(
1476: \Omega_{\lambda(s_j)}\cup \Omega_{\lambda(s_{j+1})}\right) \backslash
1477: C'$ minus $C'$ is contained in a single CR orbit of $M\backslash C'$.
1478: 
1479: Two cases are to be considered. Firstly, assume that
1480: $\lambda([r_j,t_j])$ is not contained in $C'$, namely there exists
1481: $u_j$ with $r_j\leq u_j\leq t_j$ such that
1482: \def\theequation{3.12}\begin{equation}
1483: \gamma(u_j)\in \left(\Omega_{\lambda(s_j)}\cap 
1484: \Omega_{\lambda(s_{j+1})}\right)\backslash C'.
1485: \end{equation}
1486: Because $\Omega_{\lambda(s_j)}\backslash C'$ and
1487: $\Omega_{\lambda(s_{j+1})} \backslash C'$ are both contained in a
1488: single CR orbit of $M\backslash C'$, it follows from~\thetag{3.12}
1489: that they are contained in the same CR orbit of $M\backslash C'$, as
1490: desired.
1491: 
1492: Secondly, assume that $\lambda([r_j,t_j])$ is contained in
1493: $C'$. Choose $u_j$ arbitrary with $r_j\leq u_j \leq t_j$. By
1494: construction, $\lambda(u_j)$ belongs to $\Omega_{\lambda(s_j)}\cap
1495: \Omega_{\lambda(s_{j+1})}$ and both $\Omega_{\lambda(s_j)}$ and
1496: $\Omega_{\lambda(s_{j+1})}$ are $T^cM$-integral submanifolds of $M$
1497: passing through the point $\lambda(u_j)$. Let $Y$ be a local section
1498: of $T^cM$ defined in a neighborhood of $\lambda(u_j)$ which is not
1499: tangent to $M^1$ at $\lambda(u_j)$. On the integral curve of $Y$
1500: issued from $\lambda(u_j)$, we can choose a point
1501: $\lambda(u_j)_\epsilon$ arbitrarily close to $\lambda(u_j)$ which does
1502: not belong to $C'$. Since $Y$ is a section of $T^cM$, it is tangent to
1503: both $\Omega_{\lambda(s_j)}$ and $\Omega_{\lambda(s_{j+1})}$, hence
1504: we deduce that
1505: \def\theequation{3.13}\begin{equation}
1506: \gamma(u_j)_\epsilon\in \left(\Omega_{\lambda(s_j)}\cap 
1507: \Omega_{\lambda(s_{j+1})}\right)\backslash C'.
1508: \end{equation}
1509: Consequently, as in the first case, it follows that
1510: $\Omega_{\lambda(s_j)}\backslash C'$ and $\Omega_{\lambda(s_{j+1})}
1511: \backslash C'$ are both contained in the same CR orbit of $M\backslash
1512: C'$, as desired. This completes the proof of Lemma~3.10.
1513: \endproof
1514: 
1515: Since $p$ and $q$ belong to the set~\thetag{3.11}, we
1516: deduce that the points $p=\lambda(0)\in M\backslash C'$ and
1517: the point $q=\lambda(1)\in \mathcal{ O}_{CR}(M,p)\backslash C'$
1518: belong to the same CR orbit of $M\backslash C'$, as desired. 
1519: This completes the proof of Lemma~3.5.
1520: \endproof
1521: 
1522: \subsection*{3.14.~Reduction of CR- and of $L^{\sf p}$-removability 
1523: to $\mathcal{ W}$-removability} First of all, we remind that it
1524: follows from successive efforts of numerous mathematicians ({\it
1525: cf.}~\cite{a}, \cite{j2}, \cite{m1}, \cite{tr2}, \cite{tu1},
1526: \cite{tu2}, \cite{tu3}) that for every $\mathcal{
1527: C}^{2,\alpha}$-smooth globally minimal submanifold $M'$ of $\C^n$,
1528: there exists a wedge $\mathcal{ W}'$ attached to $M$ to which all
1529: continuous CR functions on $M$ extend holomorphically. It follows that
1530: the CR-removability of the closed subset $C\subset M^1$ claimed in
1531: Theorems~1.2' and~1.4 is an immediate consequence of its $\mathcal{
1532: W}$-removability. Based on the construction of analytic discs
1533: half-attached to $M^1$ which will be achieved in Section~7 below, we
1534: shall also be able to settle the reduction of $L^{\sf p}$-removability
1535: in the end of the paper, and we formulate a convenient lemma, whose
1536: proof is postponed to \S11 below.
1537: 
1538: \def\thelemma{3.15}\begin{lemma}
1539: Under the assumptions of Theorem~1.2' and of Theorem~1.4, if the closed subset
1540: $C\subset M^1$ is $\mathcal{ W}$-removable, then it is
1541: $L^{\sf p}$-removable, for all ${\sf p}$ with 
1542: $1\leq {\sf p}\leq \infty$.
1543: \end{lemma}
1544: 
1545: \subsection*{3.16.~Strategy per absurdum: removal of a single point 
1546: of the residual non-removable subset} Thus, it suffices to demonstrate
1547: that the closed subsets $C$ of Theorems~1.2' and~1.4 are $\mathcal{
1548: W}$-removable, {\it cf.} the definition given in Section~1. Let us fix
1549: a wedgelike domain $\mathcal{ W}_1$ attached to $M\backslash C$ and
1550: remind that all our wedgelike domains are assumed to be nonempty. Our
1551: precise goal is to establish that there exists a wedgelike domain
1552: $\mathcal{ W}_2$ attached to $M$ (including $C$) and a wedgelike
1553: domain $\mathcal{ W}_3\subset \mathcal{ W}_1 \cap \mathcal{ W}_2$
1554: attached to $M\backslash C$ such that for every holomorphic function
1555: $f\in \mathcal{ O}(\mathcal{ W}_1)$, there exists a holomorphic
1556: function $F\in \mathcal{ O}(\mathcal{ W}_2)$ which coincides with $f$
1557: in $\mathcal{ W}_3$. At first, we need some more definition.
1558: 
1559: Let $C'$ be an arbitrary closed subset of $C$. We shall say that
1560: $M\backslash C'$ {\sl enjoys the wedge extension property} if there
1561: exist a wedgelike domain $\mathcal{ W}_2'$ attached to $M\backslash
1562: C'$ and a wedgelike subdomain $\mathcal{ W}_3 \subset \mathcal{
1563: W}_1\cap \mathcal{ W}_2'$ attached to $M\backslash C$ such that, for
1564: every function $f\in\mathcal{ O}(\mathcal{ W}_1)$, there exists a
1565: function $F'\in \mathcal{ O}(\mathcal{ W}_2')$ which coincides with
1566: $f$ in $\mathcal{ W}_3$.
1567: 
1568: The notion of wedge removability can be localized as follows. Let
1569: again $C'\subset C$ be arbitrary. We shall say that a point $p' \in
1570: C'$ is {\sl locally $\mathcal{ W}$-removable with respect to $C'$} if
1571: for every wedgelike domain $\mathcal{ W}_1'$ attached to $M\backslash
1572: C'$, there exists a neighborhood $U'$ of $p'$ in $M$, there exists a
1573: wedgelike domain $\mathcal{ W}_2'$ attached to $(M \backslash C') \cup
1574: U'$ and there exists a wedgelike subdomain $\mathcal{ W}_3' \subset
1575: \mathcal{ W}_1' \cap \mathcal{ W}_2'$ attached to $M\backslash C'$
1576: such that for every holomorphic function $f \in \mathcal{ O}
1577: (\mathcal{ W}_1')$, there exists a holomorphic function $F'\in
1578: \mathcal{ O}(\mathcal{ W}_2')$ which coincides with $f$ in $\mathcal{
1579: W}_3'$.
1580: 
1581: Supppose now that $M\backslash C_1'$ and $M\backslash C_2'$ enjoy the
1582: wedge extension property, for some two closed subsets $C_1', C_2'
1583: \subset C$. Using Ayrapetian's version of the edge of the wedge
1584: theorem (\cite{a}, \cite{tu1}, \cite{tu2}), the two wedgelike domains
1585: attached to $M\backslash C_1'$ and to $M\backslash C_2'$ can be glued
1586: (after appropriate shrinking) to a wedgelike domain $\mathcal{ W}_1$
1587: attached to $M\backslash (C_1'\cap C_2')$ in such a way that
1588: $M\backslash (C_1'\cap C_2')$ enjoys the $\mathcal{ W}$-extension
1589: property. Also, if $M\backslash C'$ enjoys the wedge extension
1590: property and if $p'\in C'$ is locally $\mathcal{ W}$-removable with
1591: respect to $C'$, then again by means of the edge of the wedge theorem,
1592: it follows that there exists a neighborhood $U'$ of $p'$ in $M$ such
1593: that $(M\backslash C') \cup U'$ enjoys the wedge extension property.
1594: 
1595: Based on these preliminary remarks, we may define the following set of
1596: closed subsets of $C$:
1597: \def\theequation{3.17}\begin{equation}
1598: \mathcal{ C}:=\{C'\subset C \ \hbox{closed} \ ;
1599: \ M\backslash C' \ \text{\rm enjoys the 
1600: $\mathcal{ W}$-extension property} \}.
1601: \end{equation}
1602: Then the residual set 
1603: \def\theequation{3.18}\begin{equation}
1604: C_{\text{\rm nr}}:= \bigcap_{C' \in\mathcal{ C}} C'
1605: \end{equation}
1606: is a closed subset of $M^1$ contained in $C$. It follows from the
1607: above (abstract nonsense) considerations that $M\backslash C_{\rm nr}$
1608: enjoys the wedge extension property and that no point of $C_{\rm nr}$
1609: is locally $\mathcal{ W}$-removable with respect to $C_{\rm nr}$.
1610: Here, we may think that the letters ``nr'' abbreviate ``{\sl n}on-{\sl
1611: r}emova\-ble'', because by the very definition of $C_{\rm nr}$, none
1612: of its points should be locally $\mathcal{ W}$-removable. Notice also
1613: that $M\backslash C_{\rm nr}$ is globally minimal, thanks to
1614: Lemma~3.5.
1615: 
1616: Clearly, to establish Theorem~1.1, it is enough to show that
1617: $C_{\text{\rm nr}}=\emptyset$.
1618: 
1619: We shall argue indirectly (by contradiction) and assume that
1620: $C_{\text{\rm nr}}\neq \emptyset$. In order to derive a
1621: contradiction, it clearly suffices to show that there exists at least
1622: one point $p\in C_{\rm nr}$ which is in fact locally $\mathcal{
1623: W}$-removable with respect to $C_{\rm nr}$.
1624: 
1625: At this point, we notice that the main assumptions $\mathcal{
1626: F}_{M^1}^c\{C\}$ and $\mathcal{ O}_{M^1}^{CR}\{C\}$ of Theorem~1.2'
1627: and of Theorem~1.4 imply trivially that for every closed subset $C'$
1628: of $C$, then the condition $\mathcal{ F}_{M^1}^c\{C'\}$ and the
1629: condition $\mathcal{ O}_{M^1}^{CR}\{C'\}$ also hold true: these two
1630: assumptions are obviously stable by passing to closed subsets. In
1631: particular, $\mathcal{ F}_{M^1}^c\{C_{\rm nr}\}$ and $\mathcal{
1632: O}_{M^1}^{CR}\{C_{\rm nr}\}$ hold true. Consequently, by following a
1633: {\it per absurdum} strategy, we are led to prove two statements wich
1634: are totally similar to Theorem~1.2' and to Theorem~1.4, except that we
1635: now have only to establish that {\it a single point of $C_{\rm nr}$}
1636: is locally $\mathcal{ W}$-removable with respect to $C_{\rm nr}$. This
1637: preliminary logical consideration will simplify substantially the
1638: whole architecture of the two proofs. Another important advantage of
1639: this strategy which will not be appearant until the very end of the
1640: two proofs in Sections~9 and~10 below is that we are even allowed to
1641: select a special point $p_{\rm sp}$ of $C_{\rm nr}$ by requiring
1642: some nice geometric disposition of $C_{\rm nr}$ in a neighborhood of
1643: $p_{\rm sp}$ before removing it. Sections~4 and~5 below are devoted to
1644: such a selection.
1645: 
1646: So we are led to show that for every wedgelike domain $\mathcal{ W}_1$
1647: attached to $M\backslash C_{\rm nr}$, there exists a special point
1648: $p_{\rm sp}\in C_{\rm nr}$, there exists a neighborhood $U_{p_{\rm
1649: sp}}$ of $p_{\rm sp}$ in $M$, there exists a wedgelike domain
1650: $\mathcal{ W}_2$ attached to $\left(M\backslash C_{\rm nr}\right)\cup
1651: U_{p_{\rm sp}}$ and there exists a wedgelike domain $\mathcal{
1652: W}_3\subset \mathcal{ W}_1 \cap \mathcal{ W}_2$ attached to
1653: $M\backslash C_{\rm nr}$ such that for every holomorphic function
1654: $f\in \mathcal{ O}(\mathcal{ W}_1)$, there exists a function
1655: $F\in\mathcal{ O}(\mathcal{ W}_2)$ which coincides with $f$ in
1656: $\mathcal{ W}_3$.
1657: 
1658: A further convenient simplification of the task may be achieved by
1659: deforming slightly $M$ inside the wedge $\mathcal{ W}_1$ attached to
1660: $M\backslash C_{\rm nr}$. Indeed, by means of a partition of unity, we
1661: may perform arbitrarily small $\mathcal{ C}^{2,\alpha}$-smooth
1662: deformations $M^d$ of $M$ leaving $C_{\text{\rm nr}}$ fixed and moving
1663: $M\backslash C_{\text{\rm nr}}$ inside the wedgelike domain $\mathcal{
1664: W}_1$. Furthermore, we can make $M^d$ to depend on a single small
1665: real parameter $d\geq 0$ with $M^0=M$ and $M^d\backslash C_{\rm
1666: nr}\subset \mathcal{ W}_1$ for all $d>0$. Now, {\it the wedgelike
1667: domain $\mathcal{ W}_1$ becomes a neighborhood of $M^d$ in
1668: $\C^n$}. Let us denote by $\Omega$ this neighborhood. After some
1669: substantial technical work has been achieved, at the end of the proofs
1670: of Theorem~1.2' and~1.4 to be conducted in Sections~9 and~10 below, we
1671: shall construct a local wedge $\mathcal{ W}_{p_{\rm sp}}^d$ of edge
1672: $M^d$ at $p_{\rm sp}$ by means of small Bishop analytic discs glued to
1673: $M^d$, to $\Omega$ and to another subset (which we will call a {\sl
1674: half-wedge}, {\it see} Section~4 below) such that every holomorphic
1675: function $f\in \mathcal{ O}(\Omega)$ extends holomorphically to
1676: $\mathcal{ W}_{p_{\rm sp}}^d$. Using the stability of Bishop's
1677: equation under perturbation, we shall argue in \S9.27
1678: below that our constructions are stable under such small deformations,
1679: whence in the limit $d\rightarrow 0$, the wedges $\mathcal{ W}_{p_{\rm
1680: sp}}^d$ tend smoothly to a local wedge $\mathcal{ W}_{p_{\rm sp}}:=
1681: \mathcal{ W}_{p_{\rm sp}}^0$ of edge a neighborhood $U_{p_{\rm sp}}$
1682: of $p_{\rm sp}$ in $M^0 \equiv M$ (notice that in codimension $\geq
1683: 2$, a wedge of edge a deformation $M^d$ of $M$ does not contain in
1684: general a wedge of edge $M$, hence such an argument is needed). In
1685: addition, we shall derive univalent holomorphic extension to
1686: $\mathcal{ W}_{p_{\rm sp}}$. Finally, using again the
1687: edge of the wedge theorem to fill the space between $\mathcal{ W}_1$
1688: and $\mathcal{W}_{p_{\rm sp}}$, possibly after appropriate
1689: contractions of these two wedgelike domains, we may construct a
1690: wedgelike domain $\mathcal{ W}_2$ attached to $\left(M\backslash
1691: C\right) \cup U_{p_{\rm sp}}$ and a wedgelike domain $\mathcal{
1692: W}_3\subset \mathcal{ W}_1 \cap \mathcal{ W}_{p_{\rm sp}}$ attached to
1693: $M\backslash C$ such that for every holomorphic function $f\in
1694: \mathcal{ O}(\mathcal{ W}_1)$, there exists a function $F\in\mathcal{
1695: O}(\mathcal{ W}_2)$ which coincides with $f$ in $\mathcal{ W}_3$. In
1696: conclusion, we thus reach the desired contradiction to the definition
1697: of $C_{\text{\rm nr}}$.
1698: 
1699: As a summary of the above discussion, we have essentially shown that
1700: it suffices to prove Theorems~1.2' and~1.4 with the following two
1701: extra simplifying assumptions:
1702: \begin{itemize}
1703: \item[{\bf 1)}] 
1704: Instead of functions which are holomorphic in a
1705: wedgelike domain attached to $M\backslash C_{\rm nr}$, we consider
1706: functions which are holomorphic in a neighborhood
1707: $\Omega$ of $M\backslash
1708: C_{\rm nr}$ in $\C^{m+n}$. 
1709: \item[{\bf 2)}]
1710: Proceeding by contradiction, we
1711: have argued that it suffices to remove at least one point of $C_{\rm
1712: nr}$.
1713: \end{itemize}
1714:  
1715: Consequently, we may formulate the local statement that remains to
1716: prove: after replacing $C_{\rm nr}$ by $C$ and $M^d$ by $M$, we are
1717: led to establish the following main assertion, to which Theorems~1.2'
1718: and~1.4 are essentially reduced.
1719: 
1720: \def\thetheorem{3.19}\begin{theorem}
1721: Let $M$ be a $\mathcal{ C}^{ 2,\alpha}$-smooth globally minimal
1722: generic submanifold of $\C^n$ of CR dimension $m\geq 1$ and of
1723: codimension $d=n- m\geq 1$, let $M^1 \subset M$ be a $\mathcal{ C}^{2,
1724: \alpha}$-smooth one-codimensional submanifold which is generic in
1725: $\C^n$, and let $C$ be a {\rm nonempty} proper closed subset of $M^1$.
1726: \begin{itemize}
1727: \item[{\bf (i)}]
1728: If $m=1$, assume that the condition $\mathcal{ F}_{M^1}^c\{C\}$
1729: holds.
1730: \item[{\bf (ii)}]
1731: If $m\geq 2$, assume that the condition 
1732: $\mathcal{ O}_{M^1}^{CR}\{C\}$ holds.
1733: \end{itemize}
1734: Let $\Omega$ be an arbitrary neighborhood of $M \backslash C$ in
1735: $\C^n$. Then there exist a special point $p_{ \rm sp} \in C$, there
1736: exists a local wedge $\mathcal{ W}_{ p_{ \rm sp}}$ of edge $M$ at
1737: $p_{\rm sp}$ and there exists a subneighborhood $\Omega' \subset
1738: \Omega$ of $M \backslash C$ in $\C^n$ with $\mathcal{ W}_{p_{ \rm sp
1739: }} \cap \Omega'$ connected such that for every holomorphic function $f
1740: \in \mathcal{ O}( \Omega)$, there exists a holomorphic function $F\in
1741: \mathcal{ O} \left( \mathcal{ W}_{p_{ \rm sp}}\cup \Omega' \right)$
1742: which coincides with $f$ in $\mathcal{ W}_{ p_{\rm sp }}\cap \Omega'$.
1743: \end{theorem}
1744: 
1745: However, we remind the necessity of some supplementary arguments about
1746: the stability of our constructions under deformation. The proof of our
1747: main Theorem~3.19 will occupy Sections~4, 5, 6, 7, 8, 9 and~10 below
1748: and the deformation arguments will appear lastly in \S9.27.
1749: From now on, the main question is: {\it How to choose
1750: the special point $p_{\rm sp}$ to be removed locally}~?
1751: 
1752: \subsection*{3.20.~Choice of a special point $p\in C$ in the CR 
1753: dimension $m\geq 2$ case} In the case of CR dimension $m\geq 2$,
1754: essentially all points of $C$ can play the role of the special point
1755: $p_{\rm sp}$. However, since we want to devise a new proof of
1756: Theorem~1.4 which differs from the two proofs given in~\cite{ j4},
1757: ~\cite{p1} and in \cite{m2}, it will be convenient to choose a special
1758: point $p_1 \in M^1$ which has the property that locally in a
1759: neighborhood of $p_1$, the singularity $C\subset M^1$ lies behind a
1760: generic ``wall'' $H^1$ contained in $M^1$ and of codimension one in
1761: $M^1$. Notice that as $m\geq 2$, the dimension of a two-codimensional
1762: submanifold $H^1$ of $M$ is equal to $2m+d-2\geq n$, whence $M^1$ may
1763: perfectly be generic in $\C^n$.
1764: 
1765: \def\thelemma{3.21}\begin{lemma}
1766: Let $M$ be a $\mathcal{ C}^{2, \alpha}$-smooth globally minimal
1767: generic submanifold of $\C^n$ of CR dimension $m \geq 2$ and of
1768: codimension $d = n-m \geq 1$, let $M^1\subset M$ be a $\mathcal{
1769: C}^{2, \alpha}$-smooth one-codimensional submanifold which is generic
1770: in $\C^n$ and let $C\subset M^1$ be a {\rm nonempty} proper closed
1771: subset which does not contain any CR orbit of $M^1$. Then there
1772: exists a point $p_1\in C$ and a $\mathcal{ C}^{2,\alpha}$-smooth
1773: one-codimensional submanifold $H^1\subset M^1$ passing through $p_1$
1774: which is {\rm generic} in $\C^n$ such that $C\backslash \{p_1\}$ lies,
1775: in a neighborhood of $p_1$, in one open side $(H^1)^-$ of $H^1$ in
1776: $M^1$.
1777: \end{lemma}
1778: 
1779: \proof
1780: The proof is completely similar to the proof of
1781: Lemma~2.1 in~\cite{mp3}, {\it see} especially 
1782: {\sc Figure~1}, p.~490. 
1783: \endproof
1784: 
1785: With Lemma~3.21 at hand, we can now state a more precise version 
1786: of case {\bf (ii)} of Theorem~3.19, which will be the
1787: main removability proposition. 
1788: 
1789: \def\theproposition{3.22}\begin{proposition}
1790: Let $M$ be a $\mathcal{ C}^{2, \alpha}$-smooth globally minimal
1791: generic submanifold of $\C^n$ of CR dimension $m \geq 2$ and of
1792: codimension $d = n-m \geq 1$, let $M^1\subset M$ be a $\mathcal{
1793: C}^{2, \alpha}$-smooth one-codimensional submanifold which is generic
1794: in $\C^n$, let $p_1 \in M^1$, let $H^1\subset M^1$ be a $\mathcal{
1795: C}^{2, \alpha}$-smooth one-codimensional submanifold of $M^1$ passing
1796: through $p_1$ which is also generic in $\C^n$ {\rm (}this is possible,
1797: thanks to the assumption $m \geq 2${\rm )} and let $(H^1 )^-$ denote
1798: an open local one-sided neighborhood of $H^1$ in $M^1$. Suppose that
1799: $C \subset M^1$ is a {\rm nonempty} proper closed subset of $M^1$ with
1800: $p_1 \in C$ which is situated, locally in a neighborhood of $p_1$,
1801: only in one side of $H^1$, namely $C \subset (H^1)^- \cup \{
1802: p_1\}$. Let $\Omega$ be a neighborhood of $M \backslash C$ in
1803: $\C^n$. Then there exists a local wedge $\mathcal{ W}_{p_1}$ of edge
1804: $M$ at $p_1$ with $\Omega \cap \mathcal{ W}_{p_1}$ connected
1805: (shrinking $\Omega$ if necessary) such that for every holomorphic
1806: function $f\in\mathcal{ O} (\Omega)$, there exists a holomorphic
1807: function $F\in \mathcal{ O} \left( \mathcal{ W}_{p_1} \cup \Omega
1808: \right)$ with $F \vert_\Omega =f$.
1809: \end{proposition}
1810: 
1811: In the CR dimension $m=1$ case, the choice of a special point $p\in C$
1812: is much more delicate and will be done in the next two Sections~4
1813: and~5 below, where the analog of Proposition~3.22 appears as the main
1814: removability proposition~5.12. In the case $m=1$, the submanifold
1815: $M^1$ is of real dimension equal to $n$ and it is not difficult to
1816: generalize Lemma~3.21, obtaining a submanifold $H^1$ which is of real
1817: dimension $(n-1)$ and totally real (but not generic) in
1818: $\C^n$. However, in general, such a point $p_1\in C\cap H^1$ is {\it not
1819: locally $\mathcal{ W}$-removable}\, in general. For instance, in the
1820: hypersurface case $n=2$, locally in a neighborhood of $p_1$, the
1821: closed set $C\subset (H^1)^-\cup \{p_1\}$ may coincide with the
1822: intersection of $M$ with a local complex curve transverse to $M$ at
1823: $p_1$, hence $C$ is not locally removable. In this (trivial) example,
1824: the condition $\mathcal{ F}^c\{C\}$ is not satisfied and this
1825: justifies a more refined geometrical analysis to chase a suitable
1826: special point $p_{\rm sp}\in C$ to be removed locally.
1827: This is the main purpose of Sections~4 and~5 below.
1828: 
1829: \section*{\S4.~Construction of a semi-local half wedge}
1830: 
1831: \subsection*{4.1.~Preliminary}
1832: Let $M$ be a $\mathcal{ C}^{2, \alpha}$ globally minimal generic
1833: submanifold of $\C^n$ of CR dimension $m=1$ and of codimension $d=n-m
1834: \geq 1$ and let $M^1$ be a $\mathcal{ C}^{2,\alpha}$-smooth
1835: one-codimensional submanifold which is generic in $\C^n$. Let
1836: $\gamma:[-1,1] \to M^1$ be a $\mathcal{ C}^{2, \alpha}$-smooth curve,
1837: embedding the segment $[-1,1]$ into $M$. Later, in Section~5 below, we
1838: shall exploit the geometric condition formulated in Theorem~2.1' that
1839: such a characteristic curve should satisfy, but in the present
1840: Section~4, we shall not at all take account of this geometric
1841: condition. Our goal is to construct a semi-local half-wedge attached
1842: to a one-sided neighborhood of $M^1$ along $\gamma$ with the property
1843: that holomorphic functions in the neighborhood $\Omega$ of
1844: $M\backslash C$ in $\C^n$ do extend holomorphically to this
1845: half-wedge. First of all, we need to define what we understand by the
1846: term ``half-wedge''. Although all the geometric considerations of
1847: this section may be generalized to the CR dimension $m\geq 2$ case
1848: with slight modifications, we choose to endeavour the exposition in
1849: the case $m=1$, because our constructions are essentially needed only
1850: for this case (however in Section~10 below, since we aim to present a
1851: new proof of Theorem~1.4, we shall also use the notion of a local
1852: half-wedge, without any characteristic curve $\gamma$).
1853: 
1854: \subsection*{4.2.~Three equivalent definitions of attached 
1855: half-wedges} First of all, we define the notion of a local
1856: half-wedge. We shall denote by $\Delta_n (p,\delta)$ the open polydisc
1857: centered at $p \in \C^n$ of radius $\delta >0$. Let $p_1\in M^1$, and
1858: let ${\sf C}_1$ be an open infinite cone in the normal space $T_{p_1}
1859: \C^n/ T_{ p_1}M$. Classically, a {\sl local wedge of edge $M$ at
1860: $p_1$} is a set of the form: $\mathcal{ W}_{p_1}:= \{p+ {\sf c}_1: \,
1861: p \in M, \, {\sf c}_1 \in { \sf C}_1\}\cap \Delta_n (p_1, \delta_1 )$,
1862: for some $\delta_1 >0$. Sometimes, we shall use the following
1863: terminology: if $v_1$ is a nonzero vector in $T_{p_1} \C^n /T_{
1864: p_1}M$, we shall say that $\mathcal{ W}_{p_1}$ is a {\sl local wedge
1865: at $(p_1, v_1)$}. This definition seems to be misleading in the sense
1866: that different vectors $v_1$ seem to yield local wedges with different
1867: directions, however, there is a concrete geometric meaning in this
1868: definition that should be reminded: the positive half-line $\R^+ \cdot
1869: v_1$ generated by the vector $v_1$ is locally contained in the wedge
1870: $\mathcal{ W}_{p_1}$.
1871: 
1872: For us, a {\sl local half-wedge of edge $M$ at $p_1$} will be a set of
1873: the form
1874: \def\theequation{4.3}\begin{equation}
1875: \mathcal{ HW}_{p_1}^+:= \left\{p+{\sf c}_1: \,
1876: p\in U_1\cap (M^1)^+, \, {\sf c}_1 \in {\sf C}_1\right\}
1877: \cap \Delta_n(p_1,\delta_1). 
1878: \end{equation}
1879: This yields a first definition and we shall delineate two further
1880: definitions. Let $\Delta$ denote the unit disc in $\C$, let $\partial
1881: \Delta$ denote its boundary, the unit circle and let $\overline{
1882: \Delta}= \Delta \cup \partial \Delta$ denote its closure.
1883: Throughout this paper, we shall denote by $\zeta=\rho e^{i\theta}$
1884: the variable of $\overline{\Delta}$ with $0 \leq \rho \leq 1$ and
1885: with $\vert \theta \vert \leq \pi$.
1886: 
1887: In fact, our local half-wedges (to be constructed in this section)
1888: will be defined by means of a $\mathcal{ C}^{2, \alpha-0}$-smooth
1889: $\C^n$-valued (remind $\mathcal{ C}^{2,\alpha-0}\equiv \bigcap_{\beta
1890: < \alpha} \mathcal{ C}^{2,\beta}$) mapping $(t, \chi, \nu, \rho)
1891: \longmapsto \mathcal{ Z}_{t,\chi, \nu} (\rho)$, which comes from
1892: parametrized family of analytic discs of the form $\zeta \mapsto
1893: \mathcal{ Z}_{ t,\chi,\nu} (\zeta)$, where the parameters $t \in
1894: \R^{n-1}$, $\chi \in \R^n$, $\nu \in \R$ satisfy $\vert t \vert<
1895: \varepsilon$, $\vert \chi \vert < \varepsilon$, $\vert \nu \vert <
1896: \varepsilon$ for some small $\varepsilon >0$, and where $\mathcal{
1897: Z}_{ t,\chi,\nu} (\zeta)$ is holomorphic with respect to $\zeta$ in
1898: $\Delta$. This mapping will satisfy the following three properties:
1899: 
1900: \smallskip
1901: \begin{itemize}
1902: \item[{\bf (i)}]
1903: $(\chi,\nu) \mapsto \mathcal{ Z}_{0,\chi,\nu}(1)$ is a 
1904: diffeomorphism onto a neighborhood of $p_1$ in $M$,
1905: the mapping $\chi \mapsto \mathcal{ Z}_{0,\chi,0}(1)$ is a 
1906: diffeomorphism onto a neighborhood of $p_1$ in $M^1$
1907: and $(M^1)^+$ corresponds to $\nu>0$ in the diffeomorphism
1908: $(\chi,\nu) \mapsto \mathcal{ Z}_{0,\chi,\nu}(1)$.
1909: \item[{\bf (ii)}]
1910: $\mathcal{ Z}_{t,0,0}(1)=p_1$ and the half-boundary
1911: $\mathcal{ Z}_{t, \chi, \nu}\left(\left\{ e^{i\theta} : \ 
1912: \vert \theta \vert \leq \frac{\pi}{2} \right\}\right)$ is contained in
1913: $M$ for all $t$, all $\chi$ and
1914: all $\nu$.
1915: \item[{\bf (iii)}]
1916: The vector $v_1:= \frac{ \partial \mathcal{ Z}_{0, 0,0}}{ \partial
1917: \theta} (1)\in T_{p_1 }\C^n$ is nonzero and belongs to $T_{p_1}M^1$.
1918: Furthermore, the rank of the $\R^{ n-1}$-valued $\mathcal{
1919: C}^{1,\alpha-0}$-smooth mapping 
1920: \def\theequation{4.4}\begin{equation}
1921: \R^{n-1}\ni t \longmapsto 
1922: \frac{\partial \mathcal{ Z}_{t,0,0}}{\partial \theta}(1)
1923: \in T_{p_1}M^1
1924: \ {\rm mod} \ \left( T_{p_1}M^1\cap T_{p_1}^cM
1925: \right) \cong \R^{n-1}
1926: \end{equation}
1927: is maximal equal to $(n-1)$ at $t=0$.
1928: \end{itemize}
1929: 
1930: \smallskip
1931: \noindent
1932: By holomorphicity of the map $\zeta\mapsto \mathcal{
1933: Z}_{t,\chi,\nu}(\zeta)$, we have $\frac{\partial \mathcal{
1934: Z}_{t,\chi,\nu}}{ \partial \theta}(1) =J\cdot \frac{\partial \mathcal{
1935: Z}_{t,\chi,\nu}}{ \partial \rho}(1)$, where $J$ denotes the complex
1936: structure of $T\C^n$. Consequently, because $J$ induces an isomorphism
1937: from $T_{p_1}M/T_{p_1}^cM\to T_{ p_1}\C^n/T_{p_1}M$, it follows from
1938: property {\bf (iii)} above that the vectors $\frac{\partial \mathcal{
1939: Z}_{t,0,0}}{\partial \rho}(1)$ cover an open cone containing $Jv_1$ in
1940: the quotient space $T_{p_1}M/ T_{p_1}^cM$, as $v$ varies. Then a {\sl
1941: local half-wedge of edge $(M^1)^+$ at $p_1$} will be a set of the form
1942: \def\theequation{4.5}\begin{equation}
1943: \mathcal{ HW}_{p_1}^+:=\left\{
1944: \mathcal{ Z}_{t,\chi,\nu}(\rho)\in\C^n: \ 
1945: \vert t \vert < \varepsilon, \ 
1946: \vert \chi \vert < \varepsilon, \
1947: 0 < \nu <\varepsilon, \ 
1948: 1-\varepsilon < \rho < 1
1949: \right\}.
1950: \end{equation}
1951: We notice that a complete local wedge of edge $M$ at $p_1$ is also
1952: associated to such a family $\mathcal{ Z}_{t,\chi, \nu}(\zeta)$ and may
1953: be defined as $\mathcal{ W}_{p_1}:= \left\{ \mathcal{
1954: Z}_{t,\chi,\nu}(\rho): \ \vert t \vert < \varepsilon, \ \vert \chi
1955: \vert < \varepsilon, \ \vert \nu \vert < \varepsilon, \ 1-\varepsilon
1956: < \rho < 1 \right\}$.
1957: 
1958: As may be checked, this second definition of a half-wege is {\sl
1959: essentially equivalent} to the first one, in the sense that a half
1960: wedge in the sense of the first definition always contains a
1961: half-wedge in the sense of the second definition, and vice versa,
1962: after appropriate shrinkings of open neighborhoods and cones.
1963: 
1964: Furthermore, we may distinguish two cases: either the vector $v_1= 
1965: \frac{ \partial \mathcal{ Z}_{0, 0,0}}{ \partial
1966: \theta} (1)$ is
1967: not complex tangent to $M$ at $p_1$ or it is complex tangent to $M$ at
1968: $p_1$. In the first case, after possibly shrinking $\varepsilon>0$, it
1969: may be checked that a local half-wedge of edge $(M^1 )^+$ coincides
1970: with the intersection of a (full) local wedge $\mathcal{ W}_{ p_1}$ of
1971: edge $M$ at $p_1$ with a one-sided neighborhood $(N^1)^+$ of a local
1972: hypersurface $N^1$ which intersects $M$ locally transversally along
1973: $M^1$ at $p_1$, as drawn in the left hand side of the following
1974: figure, where $M$ is of codimension two.
1975: 
1976: \bigskip
1977: \begin{center}
1978: \input half-wedge.pstex_t
1979: \end{center}
1980: 
1981: In the second case, the vector $v_1= \frac{ \partial \mathcal{ Z}_{0,
1982: 0,0}}{ \partial \theta} (1)$ is complex tangent to $M$ at $p_1$, hence
1983: belongs to the characteristic direction $T_{p_1}M^1\cap T_{p_1}^cM$,
1984: so the vector $-Jv_1$ which is interiorly tangent to the disc
1985: $\mathcal{ Z}_{0,0,0}(\Delta)$, is tangent to $M$ at $p_1$, is not
1986: tangent to $M^1$ at $p_1$, but points towards $(M^1)^+$ at $p_1$. It
1987: may then be checked that a local half-wedge of edge $(M^1)^+$
1988: coincides with the intersection of a local wedge $\mathcal{
1989: W}_{p_1}^1$ of edge $M^1$ at $(p_1,-Jv_1)$ which contains the side
1990: $(M^1)^+$, as drawn in the right hand side of {\sc Figure~5} above, in
1991: which $M$ is of codimension one. This provides the third and the most
1992: intuitive definition of the notion of local half-wedge.
1993: 
1994: Finally, we may define the desired notion of a semi-local attached
1995: half-wedge. Let $\gamma: [-1,1]\to M^1$ be an embedded $\mathcal{
1996: C}^{2, \alpha}$-smooth segment in $M^1$. Since the normal bundle to
1997: $M^1$ in $M$ is trivial, we can choose a coherent family of one-sided
1998: neighborhoods $(M_\gamma^1)^+$ of $M^1$ in $M$ along $\gamma$. A {\sl
1999: half-wedge attached to a one-sided neighborhood $(M_\gamma^1)^+$ of
2000: $M^1$ along $\gamma$} is a domain $\mathcal{ HW}_\gamma^+$ which
2001: contains a local half-wedge of edge $(M^1)^+$ at $\gamma(s)$ for every
2002: $s\in [-1,1]$. Another essentially equivalent definition is to
2003: require that we have a family $\mathcal{ Z}_{t,\chi, \nu:s}(\rho)$ of
2004: mappings smoothly varying with the parameter $s$ such that at each
2005: point $\gamma(s)=\mathcal{ Z}_{t,\chi, \nu:s}(1)$, the three
2006: conditions {\bf (i)}, {\bf (ii)} and {\bf (iii)} introduced above to
2007: define a local half-wedge are satisfied. Intuitively speaking, the
2008: direction of the cones defining the local half wedge at the point
2009: $\gamma(s)$ are smoothly varying with respect to $s$.
2010: 
2011: \bigskip
2012: \begin{center}
2013: \input attached-half-wedge.pstex_t
2014: \end{center}
2015: 
2016: We can now state the main proposition of this section, which will be
2017: of crucial use for the proof of Theorem~3.19 {\bf (i)}. Forgetting
2018: for a while the complete content of the geometric condition $\mathcal{
2019: F}_{M^1}^c\{C\}$ formulated in the assumptions of Theorem~1.2', which
2020: we will analyze thoroughly in Section~5 below, we shall only assume
2021: that we are given a characteristic segment $\gamma : [-1,1] \to M^1$
2022: in the following proposition, whose proof is the main goal of this
2023: Section~4.
2024: 
2025: \def\theproposition{4.6}\begin{proposition} 
2026: Let $M$ be a $\mathcal{ C}^{2,\alpha}$ globally minimal generic
2027: submanifold of $\C^n$ of CR dimension $m=1$ and of codimension
2028: $d=n-m\geq 1$ and let $M^1$ be a $\mathcal{ C}^{2,\alpha}$-smooth
2029: one-codimensional submanifold which is generic in $\C^n$. Let $\gamma:
2030: [-1,1]\to M^1$ be an arbitrary $\mathcal{ C}^{2,\alpha}$-smooth
2031: curve. Then there exist a neighborhood $V_\gamma$ of $\gamma[-1,1]$ in
2032: $M$ and a semi-local one-sided neighborhood $(M_\gamma^1)^+$ of $M^1$
2033: in $M$ along $\gamma$ which is the intersection of $V_\gamma$ with a
2034: side $(M_\gamma^1)^+$ of $M^1$ along $\gamma$ and there exists a
2035: semi-local half-wedge $\mathcal{ HW }_\gamma^+$ attached to
2036: $(M_\gamma^1 )^+ \cap V_\gamma$ with $\Omega \cap \mathcal{
2037: HW}_\gamma^+$ connected {\rm (}shrinking $\Omega$ if necessary{\rm )}
2038: such that for every holomorphic function $f\in \mathcal{ O} (\Omega)$,
2039: there exists a holomorphic function $F\in \mathcal{ O} \left(
2040: \mathcal{ HW }_\gamma^+ \cup \Omega \right)$ with $F\vert_\Omega=f$.
2041: \end{proposition}
2042: 
2043: To build $\mathcal{ HW}_\gamma^+$, we shall construct families of
2044: analytic discs with boundaries in $(M_\gamma^1)^+$. First of all, we
2045: need to formulate a special, adapted version of the so-called {\sl
2046: approximation theorem} of~\cite{ bt}.
2047: 
2048: \subsection*{4.7.~Local approximation theorem}
2049: As noted in~\cite{m2}, ~\cite{mp1} and~\cite{mp3}, when dealing with
2050: some natural geometric assumptions on the singularity to be
2051: removed\,--\,for instance, a two-codimensional singularity $N\subset
2052: M$ with $T_pN \supset T_p^cM$ for some points $p\in N$ or metrically
2053: thin singularities $E\subset M$ with $H^{2m+d-2}(E)=0$\,--\,it is
2054: impossible to show {\it a priori}\, that continuous CR functions on
2055: $M$ minus the singularity are approximable by polynomials, which
2056: justifies the introduction of deformations and the use of the
2057: continuity principle in~\cite{m2}, \cite{mp1}, \cite{mp3}. On the
2058: contrary the genericity of the submanifold $M^1$ containing the
2059: singularity $C$ enables us to get an approximation Lemma~4.8 just
2060: below, in the spirit of~\cite{bt}. Together with the existence of
2061: Bishop discs attached to $M^1$, the validity of this approximation
2062: lemma on $M\backslash M^1$ is the second main reason for the relative
2063: simplicity of the geometric proofs of Theorem~1.4 provided
2064: in~\cite{j4}, \cite{m2}, \cite{p1}, in comparison the 
2065: proof of Theorem~1.2' to be conducted in this paper.
2066: 
2067: \def\thelemma{4.8}\begin{lemma}
2068: Let $M$ be a $\mathcal{ C}^{2, \alpha}$-smooth globally minimal
2069: generic submanifold of $\C^n$ of CR dimension $m \geq 1$ and of
2070: codimension $d = n-m \geq 1$, let $M^1 \subset M$ be a $\mathcal{
2071: C}^{2, \alpha}$-smooth one-codimensional submanifold which is generic
2072: in $\C^n$ which divides locally $M$ in two open sides $(M^1)^-$ and
2073: $(M^1)^+$ and let $p_1\in M^1$. Then there exist two neighborhoods
2074: $U_1$ and $V_1$ of $p_1$ in $M$ with $V_1 \subset \subset U_1$ such
2075: that for every continuous CR function $f \in \mathcal{ C}_{CR}^0
2076: \left(( M^1 )^+\cap U_1 \right)$, there exists a sequence of
2077: holomorphic polynomials $(P_\nu)_{\nu \in\N}$ wich converges uniformly
2078: towards $f$ on $(M^1)^+\cap V_1$. Of course, the same property holds
2079: in the side $(M^1)^-$ instead of $(M^1)^+$.
2080: \end{lemma}
2081: 
2082: \proof
2083: The proof is a slight modification of~Proposition~5B in~\cite{m2} and
2084: we summarize it, taking for granted that the reader is acquainted with
2085: the approximation theorem proved in~\cite{bt} ({\it see}
2086: also~\cite{j5}). Let $L_0^1$ be a maximally real submanifold passing
2087: through $p_1$ and contained in $M^1\cap U_1$, for a sufficiently small
2088: neighborhood $U_1$ of $p_1$ in $M^1$, possibly to be shrunk later. In
2089: coordinates $z=(z_1,\dots,z_n)=x+iy\in\C^n$ vanishing at $p_1$, we can
2090: assume that the tangent plane to $L_0^1$ at $p_1$ identifies with
2091: $\R^n=\{y=0\}$. As the codimensions of $L_0^1$ in $M^1$ and in $M$ are
2092: equal to $(d-1)$ and to $d$, we can include $L^1$ in a $d$-parameter
2093: family of submanifolds $L_t^1$, where $t\in \R^d$ is small, so that
2094: $L_t^1\cap V_1$ makes a foliation of $M\cap V^1$ by maximally real
2095: $\mathcal{ C}^{2,\alpha}$-smooth submanifolds, for some neighborhood
2096: $V_1\subset\subset U^1$ of $p^1$ in $M$, such that $L_t^1$ is
2097: contained in $M^1$ for $t=(t_1,\dots, t_{d-1},0)$, {\it i.e.} for
2098: $t_d=0$, such that $L_t^1\cap V_1$ is contained in $(M^1)^+$ for
2099: $t_d>0$ and such that $L_t^1 \cap V_1$ is contained in $(M^1)^-$ for
2100: $t< 0$. In addition, we can assume that all the $L_t^1$ coincide with
2101: $L_0^1$ in a neighborhood of $\partial U_1$.
2102: 
2103: We shall first treat the case where $f$ is of class $\mathcal{
2104: C}^1$. Thus, let $f$ be a $\mathcal{ C}^1$-smooth CR function on
2105: $\left( M \backslash M^1\right) \cap U_1$, let $\tau \in \R$ with
2106: $\tau > 0$, fix $t_0\in \R^d$ small with $t_{d;0} >0$, whence
2107: $L_{t_0}^1\cap V^1$ is contained in $(M^1)^+$, let $\widehat{z} \in (
2108: M^1)^+ \cap V_1$ be an arbitrary point and consider the following
2109: integral which consists of the convolution of $f$ with the Gauss
2110: kernel:
2111: \def\theequation{4.9}\begin{equation}
2112: G_\tau f(\widehat{z}):=
2113: \left(\frac{\tau}{\pi}\right)^{n/2}
2114: \int_{U_1\cap L_{t_0}} \, 
2115: e^{-\tau(z-\widehat{z})^2}\, 
2116: f(z) \, dz,
2117: \end{equation}
2118: where $(z- \widehat{z})^2:= (z_1- \widehat{ z}_1)^2 + \cdots+
2119: (z_n-\widehat{ z}_n)^2$ and $dz := dz_1 \wedge \cdots \wedge d z_n$.
2120: The point $\widehat{ z}$ belongs to some maximally real submanifold
2121: $L_{\widehat{ t}}$ with $\widehat{ t}_d>0$. We now claim that the
2122: value of $G_\tau f( \widehat{ z})$ is the same if we replace the
2123: integration on the fixed submanifold $U_1 \cap L_{ t_0}$ in the
2124: integral~\thetag{4.9} by an integration over $U_1 \cap L_{ \widehat{
2125: t}}$. This key argument will follows from Stoke's theorem, from the
2126: fact that $f$ is CR on $M\cap U_1$ and from the important fact that
2127: between $L_{ t_0}$ and $L_{\widehat{ t}}$, we can construct a $(n+
2128: 1)$-dimensional submanifold $\Sigma$ with boundary $\partial \Sigma =
2129: L_{t_0} - L_{ \widehat{ t}}$ {\it which is entirely contained in
2130: $(M^1)^+\cap U_1$}, thanks to the fact that $M^1$ is generic and of
2131: codimension $1$; indeed, we may compute
2132: \def\theequation{4.10}\begin{equation}
2133: \left\{
2134: \aligned
2135: G_\tau f(\widehat{ z})=
2136: & \
2137: \left(\frac{\tau}{\pi}\right)^{n/2}
2138: \int_{U_1\cap L_{\widehat{ t}}}
2139: e^{-\tau(z-\widehat{ z})^2}\, 
2140: f(z) \, dz+\left(\frac{\tau}{\pi}\right)^{n/2}
2141: \int_{\Sigma}\, 
2142: d\left(
2143: e^{-\tau (z-\widehat{ z})^2}\, f(z) \, dz
2144: \right) \\
2145: =
2146: & \
2147: \left(\frac{\tau}{\pi}\right)^{n/2}
2148: \int_{U_1\cap L_{\widehat{ t}}}
2149: e^{-\tau(z-\widehat{ z})^2}\, 
2150: f(z) \, dz,
2151: \endaligned\right.
2152: \end{equation}
2153: noticing that the second integral in the right hand side of the first
2154: line vanishing, because $f$ and $e^{-(z-\widehat{ z})^2}$ being CR and
2155: of class at least $\mathcal{ C}^1$, one has $d\left( e^{-\tau (z-
2156: \widehat{ z})^2}\, f(z) \, dz \right)=0$. This proves the claim.
2157: 
2158: By analyzing the real and the imaginary part of the phase function
2159: $-\tau(z-\widehat{ z})^2$ on $L_{\widehat{ t}}$, one can show by means
2160: of a standard argument (convolution with Gauss' kernel is an
2161: approximation of the Dirac mass) that the integral over $U_1 \cap
2162: L_{\widehat{ t}}$ tends towards $f(\widehat{ z})$ as $\tau$ tends to
2163: $\infty$, provided that the submanifold $U_1\cap L_{\widehat{ t}}$ is
2164: sufficiently close to the real plane $\R^n$ in $\mathcal{ C}^1$
2165: norm. Finally, by developing in power series and truncating the
2166: exponential in the first expression~\thetag{4.9} which defines $G_\tau
2167: f(\widehat{ z})$ and by integrating termwise, one constructs the
2168: desired sequence of polynomials $(P_\nu (z))_{\nu\in\N}$.
2169: 
2170: The case where $f$ is only continuous follows from standard arguments
2171: from the theory of distributions. This completes the proof of
2172: Lemma~4.8.
2173: \endproof
2174: 
2175: \subsection*{4.11.~A family of straightenings} 
2176: Our main goal in the remainder of this section is to construct a
2177: semi-local half-wedge attached to a one-sided neighborhood
2178: $(M_\gamma^1)^+$ of $M^1$ in $M$ along $\gamma$ which consists of
2179: analytic discs attached to $(M_\gamma^1)^+$. First of all, we need a
2180: convenient family of normalizations of the local geometries of $M$ and
2181: of $M^1$ along the points $\gamma(s)$ of our characteristic curve
2182: $\gamma$, for all $s$ with $-1\leq s \leq 1$.
2183: 
2184: Let $\Omega$ be a thin neighborhood of $\gamma( [-1, 1])$ in $\C^n$,
2185: say a union of polydiscs of fixed radius centered at the points
2186: $\gamma (s)$. Then there exists $n$ real valued $\mathcal{ C}^{2,
2187: \alpha}$-smooth functions $r_1 (z, \bar z), \dots, r_n (z,\bar z)$
2188: defined in $\Omega$ such that $M \cap \Omega$ is given by the $(n-1)$
2189: Cartesian equations $r_2 (z,\bar z)= \cdots= r_n (z,\bar z)=0$ and
2190: such that moreover, $M^1\cap \Omega$ is given by the $n$ Cartesian
2191: equations $r_1(z,\bar z)=r_2(z,\bar z)=\cdots=r_n(z,\bar z)=0$. We
2192: first center the coordinates at $\gamma(s)$ by setting $z':=
2193: z-\gamma(s)$. Then the defining functions centered at $z'=0$ become
2194: \def\theequation{4.12}\begin{equation}
2195: r_j\left(z'+
2196: \gamma(s), \bar z'+\overline{\gamma(s)}\right)-r_j
2197: \left(\gamma(s),\overline{\gamma(s)}\right)=:
2198: r_j'(z',\bar z':s),
2199: \end{equation}
2200: for $j=1,\dots,n$, and they are parametrized by $s\in [-1,1]$. Now,
2201: we drop the primes on coordinates and we denote by $r_j(z,\bar z:s)$,
2202: $j=1,\dots,n$, the defining equations for the new 
2203: $M_s$ and $M_s^1$, which
2204: correspond to the old $M$ and $M^1$ locally in a neighborhood of
2205: $\gamma(s)$. Next, we straighten the tangent planes by using the
2206: linear change of coordinates $z'=A_s\cdot z$, where the $n\times n$
2207: matrix $A_s$ is defined by $A_s:= 2i \, \left(\frac{\partial
2208: r_j}{\partial z_k}(0,0:s) \right)_{1\leq j,k\leq n}$. Then the
2209: defining equations for the two transformed 
2210: $M_s'$ and for $M_s^{1'}$ are given by
2211: \def\theequation{4.13}\begin{equation}
2212: r_j'(z',\bar z':s):= r_j\left(A_s^{-1} \cdot z', \overline{A}_s^{-1}
2213: \cdot \overline{ z}': s\right),
2214: \end{equation}
2215: and we check immediately that the matrix $\left( \frac{ \partial
2216: r_j'}{ \partial z_k} (0,0:s) \right)_{1 \leq j,k \leq n}$ is equal to
2217: $2i$ times the $n\times n$ identity matrix, whence $T_0 M_s'= \{y_2'=
2218: \cdots = y_n' =0\}$ and $T_0{ M_s' }^1=\{y_1'=y_2'= \cdots =
2219: y_n'=0\}$. It is important to notice that the matrix $A_s$ only
2220: depends $\mathcal{ C}^{1,\alpha}$-smoothly with respect to $s$.
2221: Consequently, if we now drop the primes on coordinates, the defining
2222: equations for $M_s$ and for $M_s^1$ are of class $\mathcal{ C}^{2,
2223: \alpha}$ with respect to $(z,\bar z)$ and only of class $\mathcal{
2224: C}^{1, \alpha}$ with respect to $s$.
2225: 
2226: Applying then the $\mathcal{ C}^{2,\alpha}$-smooth
2227: implicit, we deduce that there exist $(n-1)$ functions
2228: $\varphi_j(x,y_1:s)$, $j=2,\dots,n$, which are all of class $\mathcal{
2229: C}^{2,\alpha}$ with respect to $(x,y_1)$ in a real cube
2230: $\I_{n+1}(2\rho_1):=\{(x,y_1)\in \R^n \times \R: \ \vert x \vert <
2231: 2\rho_1, \ \vert y_1 \vert < 2\rho_1\}$, for some $\rho_1>0$, which
2232: are uniformly bounded in $\mathcal{ C}^{2,\alpha}$-norm as the
2233: parameter $s$ varies in $[-1,1]$, which are of class $\mathcal{
2234: C}^{1,\alpha}$ with respect to $s$, such that $M_s$ may be represented
2235: in the polydisc $\Delta_n (\rho_1)$ by the $(n-1)$ graphed equations
2236: \def\theequation{4.14}\begin{equation}
2237: y_2=\varphi_2(x,y_1: s), \dots\dots, \, 
2238: y_n=\varphi_n(x,y_1:s),
2239: \end{equation}
2240: or more concisely $y'=\varphi'(x,y_1:s)$, if we denote the coordinates
2241: $(z_2,\dots,z_n)$ simply by $z'=x'+iy'$. Here, by construction, we
2242: have the normalization conditions $\varphi_j(0:s)=\partial_{x_k}
2243: \varphi_{j}(0:s)=\partial_{y_1} \varphi_{j}(0:s)=0$, for $j=2,\dots,n$
2244: and $k=1,\dots,n$. Sometimes in the sequel, we shall use the notation
2245: $\varphi_j(z_1,x':s)$ instead of $\varphi_j(x,y_1:s)$. Similarly,
2246: again by means of the implicit function theorem, we obtain $n$
2247: functions $h_k(x:s)$, for $k=1,\dots,n$ , which 
2248: are of class $\mathcal{ C}^{2,\alpha}$
2249: in the cube $\I_n(2\rho_1)$ (after possibly shrinking $\rho_1$)
2250: enjoying the same regularity property with respect 
2251: to $s$, such that $M_s^1$ is
2252: represented in the polydisc $\Delta_n(\rho_1)$ by the $n$ graphed
2253: equations
2254: \def\theequation{4.15}\begin{equation}
2255: y_1=h_1(x:s), \
2256: y_2=h_2(x:s), \
2257: \dots\dots, \ y_n=h_n(x:s).
2258: \end{equation} 
2259: In addition, we can assume that
2260: \def\theequation{4.16}\begin{equation}
2261: h_j(x:s)\equiv
2262: \varphi_j\left(x,h_1(x:s):s\right), \ \ \ \ \ \ \
2263: j=2,\dots,n.
2264: \end{equation}
2265: Here, by construction, we have the normalization conditions
2266: $h_k(0:s)=\partial_{x_l}
2267: h_k(0:s)=0$ for $k,l=1,\dots,n$.
2268: 
2269: In the sequel, we shall denote by $\widehat{ z}=\Phi_s(z)$ the
2270: final change of coordinates which is centered at $\gamma(s)$ and
2271: which straightens simultaneously the tangent planes to 
2272: $M$ at $\gamma(s)$ and to $M^1$ at $\gamma(s)$ and we shall
2273: denote by $M_s$ and by $M_s^1$ the transformations of
2274: $M$ and of $M^1$.
2275: 
2276: Also, we shall remind that the following regularity
2277: properties hold for the functions $\varphi_j (x,y_1:s)$ and
2278: $h_k(x:s)$:
2279: \begin{itemize}
2280: \item[{\bf (a)}]
2281: For fixed $s$, they are of class $\mathcal{ C}^{2,\alpha}$ with
2282: respect to their principal variables, namely excluding the parameter
2283: $s$.
2284: \item[{\bf (b)}]
2285: They are of class $\mathcal{ C}^{1,\alpha}$ with respect to all their
2286: variables, including the parameter $s$.
2287: \item[{\bf (c)}]
2288: Each of their first order partial derivative with respect to one of
2289: their principal variables is of class $\mathcal{ C}^{1,\alpha}$ with
2290: respect to all their variables, including the parameter $s$.
2291: \end{itemize}
2292: Indeed, these properties are clearly satisfied for the
2293: functions~\thetag{4.13} and they are inherited after the two
2294: applications of the implicit function theorem which yielded the
2295: functions $\varphi_j(x,y_1:s)$ and $h_k(x:s)$.
2296: 
2297: \subsection*{4.17.~Contact of a small ``round'' analytic
2298: disc with $M^1$} Let $r\in \R$ with $0\leq r \leq r_1$, where $r_1$ is
2299: small in comparison with $\rho_1$. Then the ``round'' analytic disc
2300: $\overline{\Delta} \ni \zeta \to \widehat{ Z}_{1;r}(\zeta):=
2301: ir(1-\zeta)\in \C$ with values in the complex plane equipped with the
2302: coordinate $z_1=x_1+iy_1$ is centered at the point $ir$ of the
2303: $y_1$-axis, is of radius $r$ and is contained in the open upper half
2304: plane $\{z_1\in \C: y_1> 0\}$, except its boundary point $\widehat{
2305: Z}_{1;r}(1)=0$. In addition, the tangent direction
2306: $\frac{\partial}{\partial \theta} \widehat{ Z}_{1;r}(1)=r$ is directed
2307: along the positive $x_1$-axis, {\it see} in advance {\sc Figure~7}
2308: below.
2309: 
2310: As in~\cite{tu2}, \cite{tu3}, \cite{mp1}, \cite{mp3}, we denote by
2311: $T_1$ the Hilbert transform (harmonic conjugate operator) on $\partial
2312: \Delta$ vanishing at $1$, namely $(T_1 X) (1)=0$, whence
2313: $T_1(T_1(X))=-X+X(1)$.
2314: 
2315: By lifting this disc contained in the complex tangent space
2316: $T_{p_1}^cM\equiv \C_{z_1}\times \{0\}$, we may define an analytic
2317: disc parametrized by $r$ and $s$ which is attached to $M$ 
2318: of the form
2319: \def\theequation{4.18}\begin{equation}
2320: \widehat{ Z}_{r:s}(\zeta)=\left(
2321: ir(1-\zeta), \widehat{ Z}_{r:s}'(\zeta)
2322: \right)\in \C\times \C^{n-1}
2323: \end{equation}
2324: where the real part $\widehat{ X}_{r:s}'(\zeta)$ of $\widehat{
2325: Z}_{r:s}'(\zeta)$ satisfies the following Bishop type equation on
2326: $\partial \Delta$
2327: \def\theequation{4.19}\begin{equation}
2328: \widehat{ X}_{r:s}'(\zeta)= -\left[T_1
2329: \varphi'\left(\widehat{ Z}_{1;r}(\cdot), 
2330: \widehat{ X}_{r:s}'(\cdot):s\right)\right](\zeta), \ \ \ \ \ \
2331: \zeta\in\partial\Delta.
2332: \end{equation}
2333: By~\cite{tu1}, \cite{tu3}, if $r_1$ is sufficiently small, there
2334: exists a solution which is $\mathcal{ C}^{2,\alpha-0}$-smooth with
2335: respect to $(r,\zeta)$, but only $\mathcal{ C}^{1, \alpha-0}$-smooth
2336: with respect to $(r,\zeta,s)$. Notice that for $r=0$, the disc
2337: $\widehat{ Z}_{1;0}(e^{i\theta })$ is constant equal to $0$ and by
2338: uniqueness of the solution of~\thetag{4.19}, it follows that
2339: $\widehat{ Z}_{0:s}'\left(e^{i\theta}\right) \equiv 0$. It follows
2340: trivially that $\partial_\theta \widehat{ X}_{0:s}\left(e^{i\theta}
2341: \right)\equiv 0$ and that $\partial_\theta\partial_\theta \widehat{
2342: X}_{0:s} \left(e^{i\theta}\right)\equiv 0$, which 
2343: will be used in a while. Notice also that
2344: $\widehat{ X}_{r:s}(1)=0$ for all $r$ and all $s$.
2345: 
2346: On the other hand, since by assumption, we have $h_1(0:s)=0$ and
2347: $\partial_{x_k}h_1(0:s)=0$ for $k=1,\dots,n$, it follows from the
2348: chain rule that if we set
2349: \def\theequation{4.20}\begin{equation}
2350: F(r,\theta:s):= h_1\left(
2351: \widehat{ X}_{r:s}\left(e^{i\theta}\right):s\right)
2352: \end{equation}
2353: where $\theta$ satisfies $0\leq \vert \theta \vert \leq \pi$,
2354: then
2355: the following four equations hold
2356: \def\theequation{4.21}\begin{equation}
2357: F(0,\theta:s)\equiv 0, \ \ \ \ \
2358: F(r,0:s)\equiv 0, \ \ \ \ \
2359: \partial_\theta F(0,\theta:s)\equiv 0, \ \ \ \ \ 
2360: \partial_\theta F(r,0:s)\equiv 0.
2361: \end{equation}
2362: We claim that there exists a constant $C>0$ such that
2363: the following five inequalities hold
2364: for $0\leq \vert \theta \vert \leq \pi$, for
2365: $0\leq r \leq r_1$, for $s\in [-1,1]$ and
2366: for $\vert x \vert \leq \rho_1$:
2367: \def\theequation{4.22}\begin{equation}
2368: \left\{
2369: \aligned
2370: {}
2371: & \
2372: \left\vert
2373: \widehat{ X}_{r:s}\left(e^{i\theta}\right) 
2374: \right\vert\leq C\cdot r, \\
2375: & \
2376: \left\vert 
2377: \partial_\theta \widehat{ X}_{r:s}\left(e^{i\theta}\right)
2378: \right\vert \leq C\cdot r, \\
2379: & \
2380: \left\vert
2381: \partial_\theta \partial_\theta \widehat{ X}_{r:s}\left(e^{i\theta}
2382: \right) 
2383: \right\vert \leq C\cdot r^{\frac{\alpha}{2}}, \\
2384: & \
2385: \sum_{k=1}^n 
2386: \left\vert
2387: \partial_{x_k} h_1(x) 
2388: \right\vert\leq C\cdot \vert x \vert, \\
2389: & \
2390: \sum_{k_1,k_2=1}^n\, 
2391: \left\vert
2392: \partial_{x_{k_1}}\partial_{x_{k_2}}
2393: h_1(x)
2394: \right\vert \leq C.
2395: \endaligned\right.
2396: \end{equation}
2397: The best constants for each inequality are {\it a priori} distinct,
2398: but we simply take for $C$ the largest one. Indeed, the first, the
2399: second and the third inequalities are elementary consequences of the
2400: (uniform with respect to $s$) $\mathcal{ C}^{2,
2401: \frac{\alpha}{2}}$-smoothness of $\widehat{
2402: X}_{r:s}\left(e^{i\theta}\right)$ with respect to $(r,\theta)$, and of
2403: the normalization conditions~\thetag{4.21}. The fourth and the fifth
2404: inequalities are consequences of the $\mathcal{
2405: C}^2$-smoothness of $h_1$ and of its first order normalizations
2406: (complete argument may be easily be provided by imitating the
2407: reasonings of the elementary Section~6 below). 
2408: 
2409: Computing now the second derivative
2410: of $F(r,\theta:s)$ with respect to $\theta$, we obtain
2411: \def\theequation{4.23}\begin{equation}
2412: \small
2413: \left\{
2414: \aligned
2415: \partial_\theta\partial_\theta
2416: F(r,\theta:s)
2417: & 
2418: =
2419: \sum_{k=1}^n\, 
2420: \partial_{x_k} h_1 \left(
2421: \widehat{ X}_{r:s}\left(e^{i\theta}\right):s
2422: \right)\cdot
2423: \partial_\theta \partial_\theta
2424: \widehat{ X}_{k;r:s}\left(e^{i\theta}\right) + \\
2425: &
2426: +
2427: \sum_{k_1,k_2=1}^n\, 
2428: \partial_{x_{k_1}}\partial_{x_{k_2}} 
2429: h_1\left(
2430: \widehat{ X}_{r:s}\left(e^{i\theta}\right)\right)\cdot 
2431: \partial_\theta \widehat{ X}_{k_1;r,s}\left(e^{i\theta}\right)\cdot 
2432: \partial_\theta \widehat{ X}_{k_2;r,s}\left(e^{i\theta}\right),
2433: \endaligned\right.
2434: \end{equation}
2435: and we may apply the majorations~\thetag{4.22} to get
2436: \def\theequation{4.24}\begin{equation}
2437: \left\{
2438: \aligned
2439: \left\vert
2440: \partial_\theta \partial_\theta
2441: F(r,\theta:s)
2442: \right\vert\leq
2443: & \
2444: C\cdot \left\vert
2445: \widehat{ X}_{r:s}(e^{i\theta}) \right\vert
2446: \cdot C \cdot r^{\frac{\alpha}{2}}
2447: +C\cdot (C\cdot r)^2 \\
2448: \leq
2449: & \
2450: r\cdot C^3\left[
2451: r^{\frac{\alpha}{2}}+r^2
2452: \right].
2453: \endaligned\right.
2454: \end{equation}
2455: 
2456: We can now state and prove a lemma which shows that the disc
2457: boundaries $\widehat{ Z}_{r:s}(\partial\Delta)$ touches $M^1$ only at
2458: $p_1$ and lies in $(M^1)^+\cup \{p_1\}$.
2459: 
2460: \def\thelemma{4.25}\begin{lemma}
2461: If $r_1 \leq \min\left( 1, \ \left( \frac{1}{4C^3 \pi^2}
2462: \right)^{\frac{2}{\alpha}}\right)$, then $\widehat{ Z}_{r:s}(\partial
2463: \Delta \backslash \{1\})$ is contained in $(M_s^1)^+$ for all $r$ with
2464: $0 < r \leq r_1$ and all $s$ with $-1\leq s \leq 1$.
2465: \end{lemma}
2466: 
2467: \proof
2468: In the polydisc $\Delta_n(\rho_1)$, the positive half-side $(M_s^1)^+$
2469: in $M$ is represented by the single equation $y_1>h_1(x:s)$, hence we
2470: have to check that $\widehat{ Y}_{1;r}\left(e^{i\theta}\right) >
2471: \left\vert h_1\left(\widehat{ X}_{r:s}\left(e^{i\theta}\right):s\right)
2472: \right\vert$, for all $\theta$ with $0< \vert \theta \vert \leq \pi$.
2473: According to~\thetag{4.18}, the $y_1$-component $\widehat{
2474: Y}_{1;r}\left(e^{i\theta}\right)$ of $\widehat{
2475: Z}_{r:s}\left(e^{i\theta}\right)$ is given by
2476: $r\left(1-\cos \theta\right)$.
2477: 
2478: On the first hand, we observe the elementary minoration $r(1-\cos
2479: \theta) \geq r\cdot \theta^2\cdot \frac {1}{\pi^2}$, valuable for
2480: $0\leq \vert \theta \vert \leq \pi$.
2481: 
2482: On the second hand, taking account of the second and fourth
2483: relations~\thetag{4.21}, Taylor's integral formula now yields
2484: \def\theequation{4.26}\begin{equation} 
2485: F(r,\theta:s)=\int_0^\theta \,
2486: (\theta-\theta') \cdot \partial_\theta 
2487: \partial_\theta F\left(r,\theta':s\right) \cdot
2488: d\theta'.
2489: \end{equation}
2490: Observing that $r^2\leq r^{\frac{ \alpha}{2}}$, since $0<r \leq r_1\leq
2491: 1$, and using the majoration~\thetag{4.24}, we may estimate, taking
2492: account of the assumption on $r_1$ written in the statement of the
2493: lemma:
2494: \def\theequation{4.27}\begin{equation}
2495: \left\vert
2496: F(r,\theta:s) 
2497: \right\vert\leq r\cdot \frac{\theta^2}{2}
2498: \cdot C^3 [2r^{\frac{\alpha}{2}}]
2499: \leq r\cdot \theta^2 
2500: \cdot \frac{1}{4\pi^2}.
2501: \end{equation}
2502: The desired inequality $r(1-\cos\theta) > \left\vert
2503: F(r,\theta:s)
2504: \right\vert$ for all $0< \vert \theta \vert \leq \pi$ is
2505: proved.
2506: \endproof
2507: 
2508: We now fix once for all a radius $r_0$ with $0< r_0 \leq r_1$. In the
2509: remainder of the present Section~4, we shall deform the disc
2510: $\widehat{ Z}_{r_0:s}(\zeta)$ by adding many more parameters. We
2511: notice that for all $\theta$ with $0\leq \vert \theta \vert \leq
2512: \frac{\pi}{4}$, we have the trivial minoration $\partial_\theta
2513: \partial_\theta \widehat{ Y}_{1;r_0}\left(e^{i\theta}\right) = r_0
2514: \cos \theta\geq \frac{ r_0}{\sqrt{2}}$. Also, by~\thetag{ 4.24} and
2515: by the inequality on $r_1$ written in the statement of Lemma~4.25, we
2516: deduce $\left\vert \partial_\theta \partial_\theta h_1\left( \widehat{
2517: X}_{r_0:s}\left(e^{i\theta}\right) \right) \right\vert\leq \frac{
2518: r_0}{2\pi^2}$ for all $\theta$ with $0\leq \vert \theta \vert \leq
2519: \pi$. Since we shall need a generalization of Lemma~4.25 in
2520: Lemma~4.51 below, let us remember these two interesting inequalities,
2521: valid for $0\leq \vert \theta \vert \leq \frac{\pi}{4}$:
2522: \def\theequation{4.28}\begin{equation}
2523: \left\{
2524: \aligned
2525: \partial_\theta\partial_\theta
2526: \widehat{ Y}_{1;r_0}\left(e^{i\theta}\right) \geq 
2527: & \
2528: \frac{r_0}{\sqrt{2}}, \\
2529: \left\vert
2530: \partial_\theta \partial_\theta h_1\left(
2531: \widehat{ X}_{r_0:s}\left(e^{i\theta}\right)
2532: \right)
2533: \right\vert\leq 
2534: & \
2535: \frac{r_0}{2\pi^2},
2536: \endaligned\right.
2537: \end{equation}
2538: noticing of course that $\frac{r_0}{2\pi^2}< \frac{r_0}{\sqrt{2}}$.
2539: 
2540: \subsection*{4.29.~Normal deformations of the disc 
2541: $\widehat{ Z}_{r:s}(\zeta)$} So, we fix $r_0$ small with $0 < r_0 \leq
2542: r_1$ and we consider the disc $\widehat{ Z}_{r_0:s}(\zeta)$ for
2543: $\zeta\in\overline{\Delta}$. Then the point $\widehat{
2544: Z}_{r_0:s}(-1)$ belongs to $(M_s^1)^+$ for each $s$ and stays at a
2545: positive distance from $M_s^1$ as $s$ varies in $[-1,1]$. It follows
2546: that we can choose a subneighborhood $\omega_s$ of $\widehat{
2547: Z}_{r_0:s}(-1)$ in $\C^n$ which is contained in $\Omega$ and whose
2548: diameter is uniformly positive with respect to $s$.
2549: 
2550: \bigskip
2551: \begin{center}
2552: \input normal-deformations.pstex_t
2553: \end{center}
2554: 
2555: Following~\cite{tu2} and~\cite{mp1}, we shall introduce {\sl normal
2556: deformations} of the analytic discs $\widehat{ Z}_{r_0 :s}(\zeta)$
2557: parametrized by $s$ as follows. Let $\kappa: \R^{ n-1} \to \R^{n-1}$
2558: be a $\mathcal{ C}^{2, \alpha}$-smooth mapping fixing the origin and
2559: satisfying $\partial_{ x_k} \kappa_{j}(0)=\delta_j^k$ for $j,k=1,
2560: \dots, n-1$, where $\delta_j^k$ denotes Kronecker's symbol. For
2561: $j=2,\dots,n$, let $\eta_j = \eta_j( z_1, x':s)$ be a real-valued
2562: $\mathcal{ C}^{2, \alpha}$-smooth function compactly supported in a
2563: neighborhood of the point of $\R^{n+1}$ with coordinates $\left(
2564: \widehat{ Z}_{ 1;r_0:s }(-1), \widehat{ X}_{ r_0:s}'(-1) \right)$ and
2565: equal to $1$ at this point. We then define the $\mathcal{ C}^{2,
2566: \alpha}$-smooth deformed generic submanifold $M_{ s,t}$ of equations
2567: \def\theequation{4.30}\begin{equation}
2568: \left\{
2569: \aligned
2570: y'
2571: & \
2572: =\varphi'(z_1,x': s) +\kappa(t) \cdot 
2573: \eta'(z_1,x': s) \\
2574: & \
2575: =:\Phi'(z_1,x',t:s).
2576: \endaligned\right.
2577: \end{equation}
2578: Notice that $M_{ s,0} \equiv M_s$ and that $M_{ s,t}$ coincides with
2579: $M_s$ in a small neighborhood of the origin, for all $t$. If $\mu=
2580: \mu \left( e^{i\theta}: s\right)$ is a real-valued nonnegative
2581: $\mathcal{ C}^{2, \alpha}$-smooth function defined for $e^{i \theta}
2582: \in \partial \Delta$ and for $s \in [-1,1]$ whose support is
2583: concentrated near the segment $\{-1\} \times [-1,1]$, then applying
2584: the existence Theorem~1.2 of~\cite{tu3}, for each fixed $s\in [-1,1]$,
2585: we deduce the existence of a $\mathcal{ C}^{2,\alpha-0}$-smooth
2586: solution of the Bishop type equation
2587: \def\theequation{4.31}\begin{equation}
2588: \widehat{ X}_{r_0,r:s}'\left(e^{i\theta}\right)=
2589: -\left[T_1
2590: \Phi'\left(
2591: \widehat{ Z}_{1;r_0:s}(\cdot), 
2592: \widehat{ X}_{r_0,t:s}'(\cdot), 
2593: t\mu(\cdot : s):s\right)
2594: \right]\left(e^{i\theta}\right),
2595: \end{equation}
2596: which enable us to construct a deformed family of analytic disc
2597: \def\theequation{4.32}\begin{equation}
2598: \widehat{ Z}_{r_0,t:s}\left(e^{i\theta}\right):= 
2599: \left(
2600: \widehat{ Z}_{1;r_0:s}(e^{i\theta}), 
2601: \widehat{ X}_{r_0,t:s}'(e^{i\theta})+i
2602: T_1\left[\widehat{ X}_{r_0,t:s}'(\cdot)\right]
2603: \left(e^{i\theta}\right)
2604: \right)
2605: \end{equation}
2606: whose boundaries are contained in $M\cup \omega_s$, by construction.
2607: By an inspection of Theorem~1.2 in~\cite{tu3}, taking account of the
2608: regularity properties {\bf (a)}, {\bf (b)} and {\bf (c)} stated
2609: after~\thetag{4.16}, one
2610: can show that
2611: the general solution $\widehat{ Z}_{r_0,t:s}(\zeta)$ enjoys regularity
2612: properties which are completely similar:
2613: \begin{itemize}
2614: \item[{\bf (a)}]
2615: For fixed $s$, it is of class $\mathcal{ C}^{2,\alpha-0}$ with respect
2616: to $(t,\zeta)$.
2617: \item[{\bf (b)}]
2618: It is of class $\mathcal{ C}^{1,\alpha-0}$ with respect to 
2619: all the variables $(t,\zeta,s)$.
2620: \item[{\bf (c)}]
2621: Each of its first order partial derivative
2622: with respect to the principal variables
2623: $(t,\zeta)$ is of class $\mathcal{ C}^{1,\alpha-0}$
2624: with respect to all the variables $(t,\zeta,s)$.
2625: \end{itemize}
2626: 
2627: Since the solution is $\mathcal{ C}^{1,\alpha-0}$-smooth
2628: with respect to $s$, it crucially follows that 
2629: the vector
2630: \def\theequation{4.33}\begin{equation}
2631: v_{1:s}:= - \frac{\partial \widehat{ Z}_{r_0,t:s}}{\partial \rho}(1),
2632: \end{equation}
2633: which points inside the analytic disc, varies continuously with
2634: respect to $s$. The following key proposition may be established (up
2635: to a change of notations) just by reproducing the proof of Lemma~2.7
2636: in~\cite{mp1}, taking account of the uniformity of all
2637: differentiations with respect to the curve parameter $s$.
2638: 
2639: \def\thelemma{4.34}\begin{lemma}
2640: There exists a real-valued nonnegative $\mathcal{
2641: C}^{2,\alpha}$-smooth function $\mu=\mu(e^{i\theta}:s)$ defined for
2642: $e^{i\theta}\in\partial \Delta$ and $s\in [-1,1]$ of support
2643: concentrated near $\{-1\}\times [-1,1]$ such that the
2644: mapping
2645: \def\theequation{4.35}\begin{equation}
2646: \R^{n-1}\ni t \longmapsto 
2647: \left.\frac{\partial \widehat{ X}_{r_0,t:s}'}{\partial 
2648: \theta}\left(e^{i\theta}\right)\right\vert_{
2649: \theta=0}\in \R^{n-1}
2650: \end{equation}
2651: is maximal equal to $(n-1)$ at $t=0$.
2652: \end{lemma}
2653: 
2654: Geometrically speaking, since the vector $\left. \frac{ \partial
2655: \widehat{ X}_{1;r_0:s }}{ \partial \theta} \left(e^{i\theta} \right)
2656: \right\vert_{ \theta=0}$ is nonzero, it follows that
2657: when the parameter $t$ varies, then the set of lines
2658: generated by the vectors $\left.\frac{ \partial \widehat{ X}_{ r_0,
2659: t:s}}{\partial \theta}\left(e^{i \theta}\right) \right \vert_{
2660: \theta=0}$ covers an open cone in the space $T_{ p_1}M^1\equiv \R^n$
2661: equipped with coordinates $(x_1,x' )$, {\it see} again {\sc Figure~7}
2662: above for an illustration.
2663: 
2664: \subsection*{4.36.~Adding pivoting and translation parameters}
2665: Let $\chi=(\chi_1,\chi')\in \R\times \R^{n-1}$ 
2666: and $\nu \in \R$ satisfying 
2667: $\vert \chi \vert < \varepsilon$ and $\vert \nu \vert < \varepsilon$
2668: for some small $\varepsilon>0$. Then the mapping
2669: \def\theequation{4.37}\begin{equation}
2670: \left\{
2671: \aligned
2672: \R^{n+1}\ni
2673: (\chi_1,\chi',\nu)\longmapsto 
2674: & \
2675: \left( \,
2676: \chi_1+i[h_1(\chi:s)+\nu], \chi'+i\varphi'
2677: (\chi, h_1(\chi:s)+\nu:s) \,
2678: \right) \\
2679: & \ 
2680: =: \widehat{p}(\chi,\nu:s)\in M_s
2681: \endaligned\right.
2682: \end{equation}
2683: is a $\mathcal{ C}^{2,\alpha}$-smooth diffeomorphism
2684: onto a neighborhood of the origin in $M_s$ with the
2685: property that 
2686: \begin{itemize}
2687: \item[{\bf (a)}]
2688: $\nu >0$ if and only if $\widehat{ p}(\chi,\nu:s)\in (M_s^1)^+$.
2689: \item[{\bf (b)}]
2690: $\nu=0$ if and only if $\widehat{ p}(\chi,\nu:s)\in M_s^1$.
2691: \item[{\bf (c)}]
2692: $\nu<0$ if and only if $\widehat{ p}(\chi,\nu:s)\in (M_s^1)^-$.
2693: \end{itemize}
2694: If $\tau\in \R$ with $\vert \tau \vert < \varepsilon$ is a
2695: supplementary parameter, we may now define a crucial 
2696: deformation of the first
2697: component $\widehat{ Z}_{1; r_0:s}\left(e^{i \theta} \right)$ by setting
2698: \def\theequation{4.38}\begin{equation}
2699: \widehat{ Z}_{1; r_0,\tau,\chi,\nu:s}\left(e^{i\theta}\right):=
2700: ir_0\left(1-e^{i\theta}\right)[1+i\tau]+\chi_1+
2701: i[h_1(\chi:s)+\nu].
2702: \end{equation}
2703: Of course, we have $\widehat{ Z}_{ 1; r_0, 0,0,0:s} \left( e^{i
2704: \theta} \right) \equiv \widehat{ Z}_{ 1;r_0 :s} \left( e^{i \theta }
2705: \right)$. Geometrically speaking, this perturbation corresponds to
2706: add firstly a small ``rotation parameter'' $\tau$ which rotates (and
2707: slightly dilates) the disc $ir_0\left(1-e^{i\theta}\right)$ passing
2708: through the origin in $\C_{ z_1}$, to add secondly a small
2709: ``translation parameter $(\chi_1, \chi')$ which will enable to cover a
2710: neighborhood of the origin in $M_s^1$ and to add thirdly a small
2711: translation parameter $\nu$ along the $y_1$-axis. Consequently, with
2712: this first component $\widehat{ Z}_{1; r_0, \tau, \chi, \nu :s} \left(
2713: e^{i \theta}\right)$, we can construct a $\C^n$-valued analytic disc
2714: $\widehat{ Z}_{r_0, t, \tau, \chi, \nu: s} (\zeta)$ satisfying the
2715: important property
2716: \def\theequation{4.39}\begin{equation}
2717: \widehat{ Z}_{r_0,t,\tau,\chi,\nu:s}(1)=\widehat{ p}(\chi,\nu:s),
2718: \end{equation}
2719: simply by solving the perturbed Bishop type equation
2720: which extends~\thetag{ 4.31}
2721: \def\theequation{4.40}\begin{equation}
2722: \widehat{ X}_{r_0,t,\tau,\chi,\nu:s}'
2723: \left(e^{i\theta}\right)=-
2724: \left[
2725: T_1\left(
2726: \Phi'\left(\widehat{ Z}_{1;r_0,\tau,\chi,\nu:s}(\cdot), 
2727: \widehat{ X}_{r_0,t,
2728: \tau,\chi,\nu:s}'(\cdot), 
2729: t\mu(\cdot:s): s\right)
2730: \right)
2731: \right]\left(e^{i\theta}\right).
2732: \end{equation}
2733: Of course, thanks to the sympathetic stability of Bishop's equation
2734: under perturbation, the solution exists and satisfies smoothness
2735: properties entirely similar to the ones stated after~\thetag{4.32}.
2736: We can summarize the description of our final family of analytic discs
2737: \def\theequation{4.41}\begin{equation}
2738: \widehat{ Z}_{r_0,t,\tau,\chi,\nu:s}(\zeta):
2739: \left\{
2740: \aligned
2741: r_0 = 
2742: & \ 
2743: \text{\rm approximate radius}. \\
2744: t =
2745: & \
2746: \text{\rm normal deformation parameter}. \\
2747: \tau =
2748: & \
2749: \text{\rm pivoting parameter}. \\
2750: \chi =
2751: & \
2752: \text{\rm parameter of translation along} \ M^1. \\
2753: \nu =
2754: & \
2755: \text{\rm parameter of translation in} \ M \ 
2756: \text{\rm transversally to} \ M^1. \\
2757: s =
2758: & \
2759: \text{\rm parameter of the characteristic curve} \ \gamma. \\
2760: \zeta = 
2761: & \ 
2762: \text{\rm unit disc variable}.
2763: \endaligned\right.
2764: \end{equation}
2765: 
2766: For every $t$ and every $\chi$, we now want to adjust the pivoting
2767: parameter $\tau$ in order that the disc boundary $\widehat{ Z}_{r_0,
2768: t, \tau, \chi, 0:s} \left( e^{i\theta}\right)$ for $\nu=0$ is tangent
2769: to $M_s^1$. This tangency condition will be useful in order to derive
2770: the crucial Lemma~4.51 below.
2771: 
2772: \def\thelemma{4.42}\begin{lemma}
2773: Shrinking $\varepsilon$ if necessary, there exists a unique $\mathcal{
2774: C}^{1,\alpha-0}$-smooth map $(t,\chi,s) \mapsto \tau(t, \chi:s)$
2775: defined for $\vert t \vert <\varepsilon$, for $\vert \chi \vert <
2776: \varepsilon$ and for $s \in [-1,1]$ satisfying $\tau(0,
2777: 0:s)=\partial_{ t_j} \tau( 0,0:s)= \partial_{ \chi_k}\tau (0,0:s)=0$
2778: for $j=1, \dots, n-1$ and $k=1, \dots,n$, such that the vector
2779: \def\theequation{4.43}\begin{equation}
2780: \left.
2781: \frac{\partial}{\partial \theta}\right\vert_{\theta=0}
2782: \widehat{ Z}_{r_0,t,\tau(t,\chi:s),\chi,0:s}\left(e^{i\theta}\right)
2783: \end{equation}
2784: is tangent to $M_s^1$ at the point $\widehat{ Z}_{ r_0, t, \tau(t,
2785: \chi:s), \chi, 0:s}(1)= \widehat{ p} ( \chi, 0:s) \in M_s^1$.
2786: \end{lemma}
2787: 
2788: \proof
2789: We remind that $M_s$ is represented by the $(n-1)$ scalar equations
2790: $y'= \varphi'( x, y_1: s)$ and that $M_s^1$ is represented by the $n$
2791: equations $y_1=h_1(x:s)$ and $y'=\varphi '(x, h_1 (x: s):s)\equiv
2792: h'(x':s)$. We can therefore compute the Cartesian equations of the
2793: tangent plane to $M_s^1$ at the point $\widehat{p}(\chi,0:s)=
2794: \chi+ih(\chi:s))$:
2795: \def\theequation{4.44}\begin{equation}
2796: \left\{
2797: \aligned
2798: {\sf Y}_1 -h_1(\chi:s)=
2799: & \
2800: \sum_{k=1}^n\, 
2801: \partial_{x_k}h_{1}(\chi:s) \, \left[
2802: {\sf X}_x-\chi_k\right], \\
2803: {\sf Y}'-\varphi'(\chi,h_1(\chi:s):s) =
2804: & \
2805: \sum_{k=1}^n\, 
2806: \left(
2807: \partial_{x_k}\varphi'+
2808: \partial_{y_1}\varphi'\cdot
2809: \partial_{x_k}h_{1}
2810: \right)
2811: \left[
2812: {\sf X}_k-\chi_k
2813: \right].
2814: \endaligned\right.
2815: \end{equation}
2816: On the other hand, we observe that the tangent vector
2817: \def\theequation{4.45}\begin{equation}
2818: \left. \frac{\partial }{\partial \theta}\right\vert_{\theta=0}
2819: \widehat{ Z}_{r_0,t,\tau,\chi,0:s}\left(e^{i\theta}\right)= \left(
2820: r_0[1+i\tau], \ \left. \frac{\partial }{\partial
2821: \theta}\right\vert_{\theta=0} \widehat{ Z}_{r_0,t,\tau,\chi,0:s}'
2822: \left(e^{i\theta}\right) \right)
2823: \end{equation}
2824: is already tangent to $M_s$ at the point $\widehat{ p}(\chi,0:s)$, 
2825: because $M_{s,t}\equiv M_s$ in a neighborhood of the origin.
2826: More precisely, since $\Phi'\equiv \varphi'$
2827: in a neighborhood of the origin, 
2828: we may differentiate with respect to 
2829: $\theta$ at $\theta=0$ the relation
2830: \def\theequation{4.46}\begin{equation}
2831: \widehat{ Y}_{r_0,t,\tau,\chi,0:s}'\left(e^{i\theta}\right)\equiv 
2832: \varphi'\left(\widehat{ X}_{r_0,\tau,\chi,0:s}\left(e^{i\theta}\right), 
2833: \widehat{ Y}_{1;r_0,\tau,\chi,0:s}\left(e^{i\theta}\right),
2834: \right)
2835: \end{equation}
2836: which is valid for $e^{i\theta}$ close to
2837: $1$ in $\partial \Delta$, noticing in advance that it
2838: follows immediately from~\thetag{4.38} that
2839: \def\theequation{4.47}\begin{equation}
2840: \left.
2841: \frac{\partial }{\partial 
2842: \theta} \right\vert_{\theta=0}
2843: \widehat{ X}_{1;r_0,\tau,\chi,0:s}
2844: \left(e^{i\theta}\right)=r_0 \ \ \ \ \ 
2845: {\rm and} \ \ \ \ \ 
2846: \left.
2847: \frac{\partial }{\partial 
2848: \theta} \right\vert_{\theta=0}
2849: \widehat{ Y}_{1;r_0,\tau,\chi,0:s}
2850: \left(e^{i\theta}\right)=r_0\tau,
2851: \end{equation}
2852: hence
2853: we obtain by a direct application of the chain rule
2854: \def\theequation{4.48}\begin{equation}
2855: \left\{
2856: \aligned
2857: \left.
2858: \frac{\partial }{\partial \theta} 
2859: \right\vert_{\theta=0}\widehat{ Y}_{r_0,t,
2860: \tau,\chi,0:s}'\left(e^{i\theta}\right)=
2861: & \
2862: \partial_{y_1}\varphi'\cdot r_0\tau+ \\
2863: & \
2864: +
2865: \sum_{k=1}^n\,
2866: \partial_{x_k}\varphi'\cdot \left(
2867: \left.
2868: \frac{\partial }{\partial \theta} 
2869: \right\vert_{\theta=0}
2870: \widehat{ X}_{k;r_0,t,\tau,\chi,0:s}
2871: \left(e^{i\theta}\right)
2872: \right).
2873: \endaligned\right.
2874: \end{equation}
2875: On the other hand, the vector~\thetag{4.45} belongs to the tangent
2876: plane to $M_s^1$ at $\widehat{ p}(\chi,0:s)$ whose equations are
2877: computed in~\thetag{4.44} if and only if the following two conditions
2878: are satisfied
2879: \def\theequation{4.49}\begin{equation}
2880: \left\{
2881: \aligned
2882: {}
2883: &
2884: r_0\tau=
2885: \sum_{k=1}^n\, 
2886: \partial_{x_k}h_1(\chi:s)\left[
2887: \left.
2888: \frac{\partial }{\partial \theta} 
2889: \right\vert_{\theta=0}
2890: \widehat{ X}_{k;r_0,t,\tau,
2891: \chi,0:s}\left(e^{i\theta}\right)
2892: \right], \\
2893: &
2894: \left.
2895: \frac{\partial }{\partial \theta} 
2896: \right\vert_{\theta=0}
2897: \widehat{ Y}_{r_0,t,\tau,\chi,0:s}'
2898: \left(e^{i\theta}\right)= \\
2899: &
2900: \ \ \ \ \ \ \ \ \ \ \ \ 
2901: =
2902: \sum_{k=1}^n\, 
2903: \left(
2904: \partial_{x_k}\varphi'+\partial_{y_1}
2905: \varphi'\cdot \partial_{x_k}
2906: h_1
2907: \right)\cdot \left[
2908: \left.
2909: \frac{\partial }{\partial \theta} 
2910: \right\vert_{\theta=0}
2911: \widehat{ X}_{k;r_0,t,\tau,\chi,0:s}
2912: \left(e^{i\theta}\right) \right].
2913: \endaligned\right.
2914: \end{equation}
2915: We observe that the first line of~\thetag{4.49} together with the
2916: relation~\thetag{4.48} already obtained implies the second line
2917: of~\thetag{4.49} by an obvious linear combination. Consequently, the
2918: vector~\thetag{4.45} belongs to the tangent plane to $M_s^1$ at
2919: $\widehat{ p}( \chi, 0:s)$ if and only if the first line
2920: of~\thetag{4.49} is satisfied. As $r_0$ is nonzero, as the first
2921: order derivative $\partial_{x_k}h_1 (\chi :s)$ are of class $\mathcal{
2922: C}^{1, \alpha}$ and vanish at $x=0$ and as $\left. \frac{\partial
2923: }{\partial \theta} \right \vert_{ \theta=0} \widehat{ X}_{k; r_0, t,
2924: \tau, \chi,0:s} \left(e^{ i \theta } \right)$ is of class $\mathcal{
2925: C}^{1, \alpha-0}$ with respect to all variables $(t, \tau, \chi,s)$,
2926: it follows from the implicit function theorem that there exists a
2927: unique solution $\tau= \tau(t, \chi: s)$ of the first line
2928: of~\thetag{4.49} which satisfies in addition the normalization
2929: conditions $\tau( 0,0: s)= \partial_{ t_j}\tau ( 0, 0:s)= \partial_{
2930: \chi_k}\tau (0, 0:s)=0$ for $j=1, \dots, n-1$ and $k=1, \dots, n$.
2931: This completes the proof of Lemma~4.42.
2932: \endproof
2933: 
2934: We now define the analytic disc 
2935: \def\theequation{4.50}\begin{equation}
2936: \widehat{\mathcal{ Z}}_{t,\chi,\nu:s}(\zeta):=
2937: \widehat{ Z}_{r_0,t,\tau(t,\chi:s),\chi,\nu:s}(\zeta).
2938: \end{equation}
2939: 
2940: \def\thelemma{4.51}\begin{lemma}
2941: Shrinking $\varepsilon$ if necessary, 
2942: the following two properties are satisfied{\rm :}
2943: \begin{itemize}
2944: \item[{\bf (1)}]
2945: $\widehat{\mathcal{ Z}}_{
2946: t,\chi,0:s}(\partial \Delta \backslash \{1\}) \subset (M_s^1)^+$
2947: for all $t$, $\chi$, $\nu$ and $s$ with $\vert t \vert< \varepsilon$,
2948: with
2949: $\vert \chi \vert < \varepsilon$, with
2950: $\vert \nu \vert < \varepsilon$ and with
2951: $-1\leq s \leq 1$.
2952: \item[{\bf (2)}]
2953: If $\nu$
2954: satisfies $0< \nu < \varepsilon$, then 
2955: $\widehat{\mathcal{ Z}}_{t,\chi,\nu:s}(\partial \Delta)\subset
2956: (M_s^1)^+$ for all $t$, $\chi$ and $s$ with 
2957: $\vert t \vert <\varepsilon$, with $\vert \chi \vert
2958: < \varepsilon$ and with $-1\leq s \leq 1$.
2959: \end{itemize}
2960: \end{lemma}
2961: 
2962: \proof
2963: To establish property {\bf (1)}, we first observe that the disc
2964: $\widehat{\mathcal{ Z}}_{0, 0,0 :s} \left(e^{i \theta}\right)$
2965: identifies with the disc $\widehat{ Z}_{ r_0 :s}(e^{i \theta})$
2966: defined in \S4.29. According to Lemma~4.25, we know that
2967: $\widehat{\mathcal{ Z}}_{0,0,0:s}(\partial\Delta \backslash\{1\})$ is
2968: contained in $(M_s^1)^+$. By continuity, if $\varepsilon$ is
2969: sufficiently small, we can assume that for all $t$ with $\vert t\vert
2970: < \varepsilon$, for all $\chi$ with $\vert \chi \vert < \varepsilon$
2971: and for all $\theta$ with $\frac{\pi}{4}\leq \vert \theta \vert \leq
2972: \pi$, the point $\widehat{\mathcal{ Z}}_{ t,\chi,0:s}\left(
2973: e^{i\theta}\right)$ is contained in $(M_s^1)^+$. It remains to
2974: control the part of $\partial \Delta$ which corresponds to $\vert
2975: \theta \vert \leq \frac{\pi}{4}$.
2976: 
2977: Since the disc $\widehat{\mathcal{ Z}}_{
2978: t, \chi, \nu:s}\left(e^{i \theta}\right)$ is of
2979: class $\mathcal{ C}^2$ with respect to all its principal variables
2980: $\left(t, \chi, \nu,e^{i\theta}\right)$, 
2981: if $\vert t \vert < \varepsilon$, if
2982: $\vert \chi \vert < \varepsilon$ and if $0 \leq \vert \theta \vert
2983: \leq \frac{\pi }{4}$, for sufficiently small $\varepsilon$, then the
2984: inequalities~\thetag{4.28} are just perturbed a little bit, so we can
2985: assume that
2986: \def\theequation{4.52}\begin{equation}
2987: \left\{
2988: \aligned
2989: \partial_\theta\partial_\theta
2990: \widehat{\mathcal{ Y}}_{1; t,\chi,0:s}\left(e^{i\theta}\right) \geq
2991: & \
2992: r_0, \\
2993: \left\vert
2994: \partial_\theta \partial_\theta
2995: h_1\left(
2996: \widehat{\mathcal{ X}}_{t,\chi,0:s}\left(e^{i\theta}\right)
2997: \right)\right\vert \leq 
2998: & \
2999: \frac{r_0}{2}.
3000: \endaligned\right.
3001: \end{equation}
3002: We claim that the inequality 
3003: \def\theequation{4.53}\begin{equation}
3004: \widehat{\mathcal{ Y}}_{1;t,\chi,0:s}\left(e^{i\theta}\right) >
3005: \left\vert
3006: h_1\left(\widehat{\mathcal{ X}}_{t,\chi,0:s}\left(e^{i\theta}\right)
3007: \right)
3008: \right\vert
3009: \end{equation}
3010: holds for all $0<\vert \theta \vert \leq \frac{\pi}{4}$, which 
3011: will complete the proof of property {\bf (1)}. 
3012: 
3013: Indeed, we first remind that the tangency to $M_s^1$ of the vector
3014: $\left. \frac{ \partial }{\partial \theta} \right \vert_{ \theta=0}
3015: \widehat{\mathcal{ Z}}_{ t,\chi,0:s}\left(e^{ i \theta}\right)$ at the
3016: point $\widehat{ p} ( \chi,0:s)$ is equivalent to the first
3017: relation~\thetag{4.49}, which may be rewritten in terms of the
3018: components of the disc $\widehat{\mathcal{ Z}}_{ t,\chi,0:s}\left(
3019: e^{i\theta}\right)$ as follows
3020: \def\theequation{4.54}\begin{equation}
3021: \partial_\theta \widehat{\mathcal{ Y}}_{1;t,\chi,0:s}(1)=
3022: \sum_{k=1}^n\, [\partial_{x_k} \, 
3023: h_1]\left(\widehat{\mathcal{ X}}_{t,\chi,0:s}(1)\right)\cdot
3024: \partial_\theta \widehat{\mathcal{ X}}_{k;t,\chi,0:s}(1).
3025: \end{equation}
3026: Substracting this relation from~\thetag{4.53} and
3027: substracting also the relation $\widehat{ Y}_{1;t,\chi,0:s}(1)=
3028: h_1\left(\mathcal{ X}_{t,\chi,0:s}(1)
3029: \right)$, we see that it suffices to establish that
3030: for all $\theta$ with $0< \vert \theta \vert \leq \frac{\pi}{4}$, 
3031: we have the strict inequality
3032: \def\theequation{4.55}\begin{equation}
3033: \left\{
3034: \aligned
3035: {}
3036: &
3037: \widehat{\mathcal{ Y}}_{1;t,\chi,0:s}\left(e^{i\theta}\right)-
3038: \widehat{\mathcal{ Y}}_{1;t,\chi,0:s}(1)-
3039: \theta \cdot \partial_\theta 
3040: \widehat{\mathcal{ Y}}_{1;t,\chi,0:s}(1)
3041: >\\
3042: & \
3043: >
3044: \left\vert
3045: h_1\left(
3046: \widehat{\mathcal{ X}}_{t,\chi,0:s}\left(e^{i\theta}\right)
3047: \right)
3048: -h_1\left(\widehat{\mathcal{ X}}_{t,\chi,0:s}(1)
3049: \right) - \right. \\
3050: & 
3051: \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \
3052: \left. 
3053: -\sum_{k=1}^n\, [\partial_{x_k} \, 
3054: h_1]\left(\widehat{\mathcal{ X}}_{t,\chi,0:s}(1)\right)\cdot
3055: \partial_\theta \widehat{\mathcal{ X}}_{k;t,\chi,0:s}(1)
3056: \right\vert
3057: \endaligned\right.
3058: \end{equation}
3059: However, by means of Taylor's integral
3060: formula, this last inequality may be rewritten 
3061: as
3062: \def\theequation{4.56}\begin{equation}
3063: \small
3064: \int_0^\theta \, 
3065: (\theta -\theta')\cdot
3066: \partial_\theta\partial_\theta
3067: \widehat{\mathcal{ Y}}_{1;t,\chi,0:s}
3068: \left(e^{i\theta'}\right) \cdot d\theta' >
3069: \left\vert
3070: \int_0^\theta \, 
3071: (\theta-\theta') \cdot
3072: \partial_\theta \partial_\theta
3073: \left[
3074: h_1\left(
3075: \widehat{\mathcal{ X}}_{t,\chi,0:s}\left(e^{i\theta'}\right)
3076: \right)
3077: \right]\cdot d\theta'
3078: \right\vert
3079: \end{equation}
3080: and it follows immediately by means of~\thetag{4.52}.
3081: 
3082: Secondly, to check property {\bf (2)}, we observe that by
3083: the definition~\thetag{4.37}, the parameter $\nu$ corresponds to a
3084: translation of the $z_1$-component of the disc boundary $\widehat{
3085: \mathcal{ Z}}_{ t, \chi, 0:s} (\partial \Delta)$ along the $y_1$ axis.
3086: More precisely, we have
3087: \def\theequation{4.57}\begin{equation}
3088: \frac{\partial }{\partial \nu}\, 
3089: \widehat{ \mathcal{ Y}}_{1; t,\chi,\nu:s}(\zeta) \equiv 1, 
3090: \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ 
3091: \frac{\partial }{\partial \nu}
3092: \widehat{ \mathcal{ X}}_{1; t,\chi,\nu:s}(\zeta)\equiv 0.
3093: \end{equation}
3094: On the other hand, differentiating Bishop's equation~\thetag{4.40},
3095: and using the smallness of the function $\Phi'$, it may be checked
3096: that
3097: \def\theequation{4.58}\begin{equation}
3098: \left\vert 
3099: \frac{\partial }{\partial \nu}
3100: \widehat{ Z}_{r_0,t,\tau,\chi,\nu:s}'(e^{i\theta})
3101: \right\vert < < 1, 
3102: \end{equation}
3103: if $r_0$ and $\varepsilon$ are sufficiently small. It follows that
3104: the disc boundary $\widehat{ \mathcal{ Z}}_{t, \chi, \nu:s}( \partial
3105: \Delta)$ is globally moved in the direction of the $y_1$-axis as
3106: $\nu>0$ increases, hence is contained in $(M_s^1 )^+$.
3107: 
3108: The proof of Lemma~4.51 is complete.
3109: \endproof
3110: 
3111: \subsection*{4.59.~Local half-wedges}
3112: As a consequence of Lemma~4.34, of~\thetag{4.39} and of property {\bf
3113: (2)} of Lemma~4.51, we conclude that for every $s\in [-1,1]$, our
3114: discs $\widehat{\mathcal{ Z}}_{ t,\chi,\nu:s}(\zeta)$ satisfy all the
3115: requirements {\bf (i)}, {\bf (ii)} and {\bf (iii)} of \S4.2 insuring
3116: that the set defined by
3117: \def\theequation{4.60}\begin{equation}
3118: \mathcal{HW}_s^+:=\left\{
3119: \widehat{\mathcal{ Z}}_{t,\chi,\nu:s}(\rho):\, 
3120: \vert t\vert < \varepsilon, \ 
3121: \vert \chi \vert < \varepsilon, \
3122: 0 < \nu < \varepsilon, \
3123: 1-\varepsilon < \rho < 1 
3124: \right\}
3125: \end{equation}
3126: is a local half-wedge of edge $(M_s^1)^+$ at the origin
3127: in the $\widehat{ z}$-coordinates, which corresponds 
3128: to the point $\gamma(s)$ in the $z$-coordinates. Coming back to 
3129: the $z$ coordinates, we may define the family of analytic discs
3130: \def\theequation{4.61}\begin{equation}
3131: \mathcal{ Z}_{t,\chi,\nu:s}(\zeta):= 
3132: \Phi_s^{-1}\left(\widehat{ \mathcal{ Z}}_{t,\chi,\nu:s}(\zeta)\right)
3133: \end{equation}
3134: and we shall construct the desired semi-local attached
3135: half-wedge $\mathcal{ HW}_\gamma^+$ of Proposition~4.6.
3136: 
3137: \subsection*{4.62.~Holomorphic extension to an 
3138: attached half-wedge} Indeed, we can now complete the proof of
3139: Proposition~4.6. Let us denote by $\widehat{z}=\Phi_s(z)$ the
3140: parametrized change of coordinates defined in \S4.11, where the point
3141: $\gamma(s)$ in $z$-coordinates corresponds to the origin in
3142: $\widehat{z}$-coordinates. Given an arbitrary holomorphic function
3143: $f\in \mathcal{ O}(\Omega)$ as in Proposition~4.6, by the change of
3144: coordinates $\widehat{ z}=\Phi_s(z)$ and by restriction to
3145: $(M_s^1)^+$, we get a CR function $\widehat{f}_s\in \mathcal{
3146: C}_{CR}^0\left((M_s^1)^+\cap U_1\right)$, for some small neighborhood
3147: $U_1$ of the origin in $\C^n$, whose size is uniform with respect to
3148: $s$. Thanks to an obvious generalization of the approximation
3149: Lemma~4.8 with a supplementary parameter $s\in [-1,1]$, we know that
3150: there exists a second uniform neighborhood $V_1\subset \subset U_1$ of
3151: the origin in $\C^n$ such that every continuous CR function in
3152: $\mathcal{ C}_{CR}^0\left( (M_s^1)^+ \cap U_1 \right)$ is uniformly
3153: approximable by polynomials on $(M_s^1)^+ \cap V_1$. In particular,
3154: this property holds for the CR function $\widehat{f}_s$. Furthermore,
3155: choosing $r_0$ and $\varepsilon$ sufficiently small, we can insure
3156: that all the discs $\widehat{\mathcal{ Z}}_{t, 
3157: \chi,\nu:s}(\zeta)$ are attached
3158: to $(M_s^1)^+ \cap V_1$. It then follows from the maximum principle
3159: applied to the approximating sequence of polynomials that for each
3160: $s\in [-1,1]$, the function $\widehat{f}_s$ extends holomorphically to
3161: the half-wedge defined by~\thetag{4.60}. Finally, we deduce that the
3162: holomorphic function $f\in \mathcal{ O}(\Omega)$ extends
3163: holomorphically to the half-wedge attached to the one-sided
3164: neighborhood $(M_\gamma^1)^+$ defined by
3165: \def\theequation{4.63}\begin{equation}
3166: \left\{
3167: \aligned
3168: \mathcal{HW}_\gamma^+:=\left\{
3169: \mathcal{ Z}_{t,\chi,\nu:s}(\rho):\, 
3170: \vert t\vert < \varepsilon, \ 
3171: \vert \chi \vert < \varepsilon, \ 
3172: 0< \nu < \varepsilon, \right. \\
3173: \left. 
3174: 1-\varepsilon < \rho < 1, \ 
3175: -1\leq s \leq s
3176: \right\}.
3177: \endaligned\right.
3178: \end{equation}
3179: Without shrinking $\Omega$ near the points $\mathcal{ Z}_{t, \chi,
3180: \nu:s}(-1)$ (otherwise, the crucial rank property of Lemma~4.34 would
3181: degenerate), we can shrink the open set $\Omega$ in a very thin
3182: neighborhood of the characteristic segment $\gamma$ in $M$ and we can
3183: shrink $\varepsilon >0$ if necessary in order that the intersection
3184: $\Omega \cap \mathcal{ HW}_\gamma^+$ is connected. By the principle of
3185: analytic continuation, this implies that there exists a well-defined
3186: holomorphic function $F\in \mathcal{ O}\left( \Omega \cup \mathcal{ HW
3187: }_\gamma^+ \right )$ with $F \vert_\Omega = f$.
3188: 
3189: The proof of Proposition~4.6 is complete.
3190: \qed
3191: 
3192: \subsection*{4.64.~Local half-wedge in CR dimension $m\geq 2$}
3193: Repeating the above constructions in the simpler case where the
3194: curve $\gamma$ degenerates to the point $p_1$ of Theorem~3.22 and
3195: adding some supplementary parameters along the $(m-1)$ remaining
3196: complex tangent directions of $M$ at $p_1$ for the constructions of
3197: analytic discs, we can show that under the assumptions of
3198: Theorem~3.22, there exists a local half-wedge $\mathcal{ HW}_{p_1}^+$ 
3199: at $p_1$ to which (shrinking $\Omega$ if
3200: necessary) every holomorphic function $f\in \mathcal{ O}(\Omega)$
3201: extends holomorphically. We shall not write down the details. 
3202: 
3203: \subsection*{4.65.~Transition} In the
3204: case $m=2$, using such a local half-wedge and applying the continuity
3205: principle along analytic discs of whose one boundary part is contained
3206: in $M$ and whose second boundary part is contained in the local
3207: half-wedge $\mathcal{ HW}_{p_1}^+$, we shall establish that $p_1$ is
3208: $\mathcal{ W}$-removable in Section~10 below. At present, in the more
3209: delicate case $m=1$, we shall pursue in Section~5 below our geometric
3210: constructions for the choice of a special point $p_{\rm sp}\in C$
3211: which satisfies the conclusion of Theorem~3.19 {\bf (i)}.
3212: 
3213: \section*{\S5.~Choice of a special point of $C_{\rm nr}$ 
3214: to be removed locally}
3215: 
3216: \subsection*{5.1.~Choice of a first supporting hypersurface} 
3217: Continuing with the proof of Theorem~3.19 {\bf (i)}, we shall now
3218: analyze and use the important geometric 
3219: condition $\mathcal{ F}_{M^1}^c\{C\}$
3220: defined in Theorem~1.2'. We first delineate a convenient geometric
3221: situation.
3222: 
3223: \def\thelemma{5.2}\begin{lemma} 
3224: Under the assumptions of Theorem~3.19 {\bf (i)}, there exists a
3225: $\mathcal{ C}^{2, \alpha}$-smooth segment $\gamma: [-1,1]\to M^1$
3226: running in characteristic directions, namely satisfying $d \gamma(s)/
3227: ds\in T_{ \gamma(s)}^c M\cap T_{\gamma (s)}M^1 \backslash \{0\}$ such
3228: that $\gamma (-1) \not \in C$, $\gamma(0) \in C$, $\gamma (1) \not \in
3229: C$, and there exists a $\mathcal{ C}^{ 1,\alpha}$-smooth hypersurface
3230: $H^1$ of $M^1$ with $\gamma\subset H^1$ which is foliated by
3231: characteristic segments close to $\gamma$, such that locally in a
3232: neighborhood of $H^1$, the closed subset $C$ is contained in
3233: $\gamma\cup (H^1)^-$, where $(H^1)^-$ denotes an open 
3234: one-sided neighborhood of $H^1$ in $M^1$.
3235: \end{lemma}
3236: 
3237: \proof
3238: By the assumption $\mathcal{ F}_{M^1}^c\{C\}$, there exists a
3239: characteristic curve $\widetilde{\gamma}: [-1,1]\to M^1$ with
3240: $\widetilde{\gamma}(-1)\not\in C$, $\widetilde{\gamma}(0)\in C$ and
3241: $\widetilde{\gamma}(1)\not\in C$, there exists a neighborhood
3242: $V_{\widetilde{\gamma}}^1$ of $\widetilde{\gamma}$ in $M^1$ and there
3243: exists a local $(n-1)$-dimensional submanifold $R^1$ passing through
3244: $\widetilde{\gamma}(0)$ which is transversal to $\widetilde{\gamma}$
3245: such that the semi-local projection $\pi_{\mathcal{ F}_{M^1}^c}:
3246: V_{\widetilde{\gamma}}^1\to R^1$ parallel to the characteristic curves
3247: maps $C$ onto the closed subset $\pi_{\mathcal{ F}_{M^1}^c}(C)$ with
3248: the property that $\pi_{\mathcal{ F}_{M^1}^c}(\widetilde{\gamma})$
3249: lies on the {\it boundary of $\pi_{\mathcal{ F}_{M^1}^c}(C)$ with
3250: respect to the topology of $R^1$}. This property is illustrated in the
3251: right hand side of the following figure.
3252: 
3253: \bigskip
3254: \begin{center}
3255: \input tube-ellipse.pstex_t
3256: \end{center}
3257: 
3258: However, we want in addition a foliated supporting hypersurface $H^1$,
3259: which does not necessarily exist in a neighborhood of
3260: $\widetilde{\gamma}$. To construct $H^1$, let us first straighten the
3261: characteristic lines in a neighborhood of $\widetilde{ \gamma}$,
3262: getting a product $[-1,1]\times [-\delta_1,\delta_1]^{n-1}$, for some
3263: $\delta_1>0$, equipped with coordinates
3264: $(s,\chi)=(s,\chi_2,\dots,\chi_n)\in \R\times \R^{n-1}$, so that the
3265: lines $\{\chi={\rm ct.}\}$ correspond to characteristic lines. Such a
3266: straightening is only of class $\mathcal{ C}^{1,\alpha}$, because the
3267: line distribution $T^cM\vert_{M^1} \cap TM^1$ is only of class
3268: $\mathcal{ C}^{1,\alpha}$. Clearly, we may assume that $\delta_1$ is
3269: so small that there exists $s_1$ with $0< s_1 <1$ such that the two
3270: cubes $[-1,-s_1]\times [-\delta_1,\delta_1]^{n-1}$ and $[s_1,1]\times
3271: [-\delta_1,\delta_1]^{n-1}$ do not meet the singularity $C$, as drawn
3272: in {\sc Figure~8} above.
3273: 
3274: We may identify the transversal $R^1$ with $[- \delta_1,
3275: \delta_1]^{n-1}$; then the projection of $\widetilde{\gamma}$ 
3276: is the origin of
3277: $R^1$. By assumption, $\pi_{ \mathcal{ F}_{M^1}^c}(C)$ is a proper
3278: closed subset of $R^1$ with the origin lying on its boundary. We can
3279: therefore choose a point $\chi_0$ in the interior of $R^1$ lying
3280: outside $\pi_{ \mathcal{ F}_{M^1 }^c }(C)$. Also, we can choose a
3281: small open $(n- 1)$-dimensional ball $Q_0$ centered at this point
3282: which is contained in the complement $R^1\backslash \pi_{ \mathcal{
3283: F}_{M^1 }^c}(C)$. Furtermore, we can include this ball in a one
3284: parameter family of $\mathcal{ C}^{1, \alpha}$-smooth domains $Q_\tau
3285: \subset R^1$, for $\tau \geq 0$, which are parts of ellipsoids
3286: stretched along the segment which joins the point $\chi_0$ with the
3287: origin of $R^1$.
3288: 
3289: We then consider the tube domains $[-1,1] \times Q_\tau$ in $[-1, 1]
3290: \times [- \delta_1, \delta_1]^{ n-1}$. Clearly, there exists the
3291: smallest $\tau_1>0$ such that the tube $[-1,1]\times Q_{\tau_1}$ meets
3292: the singularity $C$ on its boundary $[-1,1]\times \partial
3293: Q_{\tau_1}$. In particular, there exists a point $\chi_1\in \partial
3294: Q_{\tau_1}$ such that the characteristic segment $[-1,1]\times
3295: \{\chi_1\}$ intersects $C$. Increasing a little bit the curvature of
3296: $\partial Q_{\tau_1}$ in a neighborhood of $\chi_1$ if necessary, we
3297: can assume that $\pi_{\mathcal{ F}_{M^1}^c}(C)\cap
3298: \overline{Q_{\tau_1}}=\{\chi_1\}$ in a neighborhood of $\chi_1$.
3299: Moreover, since by construction the two segments $[-1,-s_1]\times
3300: \{\chi_1\}\cup [s_1,1]\times \{\chi_1\}$ do not meet $C$, we can
3301: reparametrize the characteristic segment $[-1,1]\times \{\chi_1\}$ as
3302: $\gamma: [-1,1]\to M^1$ with $\gamma(-1)\not\in C$, $\gamma(0)\in C$
3303: and $\gamma(1)\not\in C$. Since all characteristic lines are
3304: $\mathcal{ C}^{2,\alpha}$-smooth, we can choose the parametrization to
3305: be of class $\mathcal{ C}^{2,\alpha}$. For the supporting hypersurface
3306: $H^1$, it suffices to choose a piece of $[-1,1]\times \partial
3307: Q_{\tau_1}$ near $[-1,1]\times \{\chi_1\}$. By construction, this
3308: supporting hypersurface is only of class $\mathcal{ C}^{1,\alpha}$ and
3309: we have that $C$ is contained in $\gamma \cup (H^1)^-$ semi-locally in
3310: a neighborhood of $\gamma$, as desired. This completes the proof of
3311: Lemma~5.2.
3312: \endproof
3313: 
3314: \subsection*{5.3.~Field of cones on $M^1$}
3315: With the characteristic segment $\gamma$ constructed in Lemma~5.2, by
3316: an application of Proposition~4.6, we deduce that there exists a
3317: semi-local half-wedge $\mathcal{ HW}_\gamma^+$ attached to
3318: $(M_\gamma^1)^+\cap V_\gamma$, for some neighborhood $V_\gamma$ of
3319: $\gamma$ in $M$, to which $\mathcal{ O}(\Omega)$ extends
3320: holomorphically.
3321: 
3322: Then, we remind that by~\thetag{4.37}, \thetag{4.39}
3323: and~\thetag{4.50}, for all $t$ with $\vert t \vert < \varepsilon$, the
3324: point $\widehat{ \mathcal{ Z}}_{ t,\chi,0:s}(1)$ identifies with the
3325: point $\widehat{ p}( \chi,0:s)\in M_s^1$ defined in~\thetag{4.37}
3326: (which is independent of $t$) and the mapping $\chi \mapsto \widehat{
3327: \mathcal{ Z}}_{t,\chi,0:s}(1)\in M_s^1$ is a local diffeomorphism.
3328: 
3329: Sometimes in the sequel, we shall denote the disc $\mathcal{ Z}_{t,
3330: \chi, \nu:s} (\zeta) \equiv \Phi_s^{-1} \left( \widehat{ \mathcal{
3331: Z}}_{t,\chi, \nu:s}(\zeta) \right)$ defined in~\thetag{4.61} by
3332: $\mathcal{ Z}_{t, \chi_1, \chi',\nu:s}(\zeta)$, where $\chi'= (
3333: \chi_2, \dots, \chi_n) \in \R^{n-1}$. Since the characteristic curve
3334: is directed along the $x_1$-axis, which is transversal in $T_0M_s^1$
3335: to the space $\{(0,\chi')\}$, it follows that the mapping $(s, \chi')
3336: \longmapsto \mathcal{ Z}_{
3337: t,0,\chi',0:s}(1)=\Phi_s^{-1}\left(\widehat{ p}(0,\chi',0:s)\right)$
3338: is, independently of $t$, a diffeomorphism onto its image for $s\in
3339: [-1,1]$ and for $\chi'$ close to the origin in $\R^{n-1}$. To fix
3340: ideas, we shall let $\chi'$ vary in the {\it closed}\, cube
3341: $[-\varepsilon,\varepsilon]^{n-1}$ (analogously to the fact that $s$
3342: runs in the {\it closed}\, interval $[-1,1]$) and we shall denote by
3343: $V_\gamma^1$ the closed subset of $M^1$ which is the image of this
3344: diffeomorphism.
3345: 
3346: At every point $p:=\mathcal{ Z}_{t,0,\chi',0:s}(1)= \mathcal{
3347: Z}_{0,0,\chi',0:s}(1)$ of this neighborhood $V_\gamma^1$, we define an
3348: open infinite oriented cone contained in the $n$-dimensional linear
3349: space $T_pM^1$ by
3350: \def\theequation{5.4}\begin{equation}
3351: {\sf C}_p:=\R^+\cdot 
3352: \left\{
3353: \frac{\partial \mathcal{ Z}_{
3354: t,0,\chi',0:s}}{\partial \theta} 
3355: (1): \ \vert t \vert < \varepsilon
3356: \right\}.
3357: \end{equation}
3358: The fact that ${\sf C}_p$ is indeed an open cone follows
3359: from Lemma~4.34, from~\thetag{4.61} and
3360: from the fact that $\Phi_s^{-1}$ is a biholomorphism. 
3361: This cone contains in its interior the nonzero vector 
3362: \def\theequation{5.5}\begin{equation}
3363: v_p^0:= \frac{\partial \mathcal{ Z}_{0,0,\chi',0:s}}{\partial \theta}
3364: (1)\in {\sf C}_p\subset T_pM^1 \backslash \{0\}.
3365: \end{equation}
3366: We shall say that the cone ${\sf C}_p$ is the {\sl cone created at $p$
3367: by the semi-local attached half-wedge $\mathcal{ HW}_\gamma^+$}
3368: (more precisely, by the family of analytic discs which covers
3369: this semi-local half-wedge).
3370: 
3371: As $p$ varies, the mapping $p\mapsto {\sf C}_p$ constitutes what we
3372: shall call a {\sl field of cones over $V_\gamma^1$}, {\it see} {\sc
3373: Figure~2} in Section~2 above and {\sc Figure~9} just below for
3374: illustrations.
3375: 
3376: \bigskip
3377: \begin{center} 
3378: \input field-cones.pstex_t
3379: \end{center}
3380: 
3381: The mapping $p\mapsto v_p^0$ defines a $\mathcal{
3382: C}^{1,\alpha-0}$-smooth vector field tangent to $M^1$. This vector
3383: field is contained in the field of cones $p\mapsto {\sf C}_p$. Over
3384: $V_\gamma^1$, we can also consider a nowhere zero characteristic
3385: vector field $X$ whose direction agrees with the orientation of
3386: $\gamma$ and which satisfyies $\exp(sX)(\gamma(0))=\gamma(s)$ for all
3387: $s\in [-1,1]$. Furthermore, for every $p\in V_\gamma^1$, we define the
3388: {\sl filled cone}
3389: \def\theequation{5.6}\begin{equation}
3390: {\sf FC}_p:=\R^+\cdot \left\{
3391: \lambda \cdot X_p+
3392: (1-\lambda)\cdot v_p: \ 
3393: 0\leq \lambda < 1, \ 
3394: v_p\in {\sf C}_p
3395: \right\}.
3396: \end{equation}
3397: In the right hand side of {\sc Figure~10} just below, in the tangent
3398: space $T_pM^1$ equipped with linear coordinates $(x_1,\dots,x_n)$ such
3399: that the characteristic direction $T_p^cM\cap T_pM^1$ is directed
3400: along the $x_1$-axis, we draw ${\sf C}_p$, its filling ${\sf FC}_p$
3401: and its projection $\pi'({\sf C}_p)$ onto the $(x_2,,\dots,x_n)$-space
3402: parallel to the $x_1$-axis.
3403: 
3404: \bigskip
3405: \begin{center}
3406: \input filled-cone.pstex_t
3407: \end{center}
3408: 
3409: Given an arbitrary nonzero vector $v_p \in {\sf C}_p$, where $p\in
3410: V_\gamma^1$, it may be checked that a small neighborhood of the origin
3411: in the positive half-line $\R^+\cdot Jv_p$ generated by $Jv_p$, where
3412: $J$ denotes the complex structure of $T\C^n$, is contained in the
3413: attached half-wedge $\mathcal{ HW}_\gamma^+$. More generally, the same
3414: property holds for every nonzero vector $v_p$ which belongs to the
3415: filled cone ${\sf FC}_p$. In fact, we shall establish that analytic
3416: discs which are half-attached to $M^1$ at $p$, namely send
3417: $\partial^+\Delta:=\{\zeta\in \partial \Delta: \ {\rm Re}\, \zeta \geq
3418: 0\}$ to $M^1$, which are closed to the complex line $v_p+Jv_p$ in
3419: $\mathcal{ C}^1$ norm and which are sufficiently small are contained
3420: in the attached half-wedge $\mathcal{ HW}_\gamma^+$. The interest of
3421: this property and the reason why we have defined fields of cones and
3422: their filling will be more appearant in Sections~8 and~9 below, where
3423: we apply the continuity principle to achieve the proof of Theorem~3.19
3424: {\bf (i)}. A precise statement, involving families of analytic discs
3425: $A_c(\zeta)$ which will be constructed in Section~7 below, is as
3426: follows. For the proof of a more precise statement involving families
3427: of analytic discs $A_{x,v:c}^1(\zeta)$, we refer to Section~7 and
3428: especially to Lemma~8.3 {\bf ($\mathbf{9_1}$)} below.
3429: 
3430: \def\thelemma{5.7}\begin{lemma}
3431: Fix a point $p\in V_\gamma^1$ and a vector $v_p$ in the cone ${\sf
3432: C}_p$ created by the semi-local attached half-wedge $\mathcal{
3433: HW}_\gamma^+$ at $p$. Suppose that there exist two constants $c_1>0$
3434: and $\Lambda_1 >1$ such that for every $c$ with $0< c \leq c_1$, there
3435: exists a $\mathcal{ C}^{2,\alpha-0}$-smooth analytic disc $A_c(\zeta)$
3436: with $A_c (\partial^+\Delta) \subset M^1$,
3437: such that
3438: \begin{itemize}
3439: \item[{\bf (i)}]
3440: The positive half-line generated by the boundary of $A_c$ at $\zeta=1$
3441: coincides with the positive half-line generated by $v_p$, namely
3442: $\R^+\cdot \frac{\partial A_c}{\partial \theta}(1) \equiv \R^+ \cdot
3443: v_p$.
3444: \item[{\bf (ii)}]
3445: $\left\vert A_c (\zeta) \right\vert \leq c^ 2\cdot \Lambda_1$ for all
3446: $\zeta \in \overline{ \Delta}$ and $c \cdot \frac{ 1}{ \Lambda_1} \leq
3447: \left\vert \frac{ \partial A_c}{\partial \theta}(1) \right\vert \leq
3448: c\cdot \Lambda_1$.
3449: \item[{\bf (iii)}]
3450: $\left\vert \frac{\partial A_c}{\partial \theta}\left( \rho
3451: e^{i\theta}\right) - \frac{\partial A_c}{\partial \theta}(1)
3452: \right\vert\leq c^2 \cdot \Lambda_1$ for all $\zeta=\rho e^{i\theta} \in
3453: \overline{\Delta}$.
3454: \end{itemize}
3455: If $c_1$ is sufficiently small, then for every $c$ with $0< c \leq
3456: c_1$, the set $A_c\left(\overline{\Delta} \backslash \partial^+\Delta
3457: \right)$ is contained in the semi-local half-wedge $\mathcal{
3458: HW}_\gamma^+$. Furthermore, the same conclusion holds if the nonzero
3459: vector $v_p$ belongs to the filled cone ${\sf FC}_p$.
3460: \end{lemma}
3461: 
3462: \subsection*{5.8.~Choice of the special point $p_{\rm sp}$ in 
3463: the CR dimension $m=1$ case} We can now answer the question raised
3464: after the statement of Theorem~3.19 {\bf (i)}, which was the main
3465: purpose of the present Section~5: {\it How to choose the special point
3466: $p_{\rm sp}$ to be removed locally}? In the following statement,
3467: property {\bf (ii)} will be really crucial for the removal of $p_{\rm
3468: sp}$, {\it see}\, in advance Proposition~5.12 below.
3469: 
3470: \def\thelemma{5.9}\begin{lemma}
3471: Let $\gamma$ be the characteristic segment constructed in Lemma~5.2
3472: above. Let $\mathcal{ HW }_\gamma^+$ be the semi-local attached
3473: half-wedge of edge $(M_\gamma^1 )^+\cap V_\gamma$ constructed in
3474: Proposition~4.6 above, and let $p\mapsto {\sf FC }_p$ be the filled
3475: field of cones created in a closed neighborhood $V_\gamma^1$ of
3476: $\gamma$ in $M^1$ by this semi-local
3477: attached half-wedge $\mathcal{ HW }_\gamma^+$. Then
3478: there exists a special point $p_{\rm sp}\in V_\gamma^1$ such that the
3479: following two geometric properties are fulfilled{\rm :}
3480: \begin{itemize}
3481: \item[{\bf (i)}]
3482: There exists a $\mathcal{ C}^{2, \alpha}$-smooth local supporting
3483: hypersurface $H_{ \rm sp}$ of $M^1$ passing through $p_{\rm sp}$ such
3484: that, locally in a neighborhood of $p_{ \rm sp}$, the closed subset
3485: $C$ is contained in $(H_{ \rm sp})^-\cup \{p_{ \rm sp}\}$, where
3486: $(H_{\rm sp})^-$ denotes an open one-sided neighborhood of
3487: $H_{\rm sp}$ in $M^1$.
3488: \item[{\bf (ii)}]
3489: There exists a nonzero vector $v_{\rm sp}\in T_{p_{\rm sp}}H_{\rm sp}$
3490: which belongs to the filled cone ${\sf FC}_{p_{\rm sp}}$.
3491: \end{itemize}
3492: \end{lemma}
3493: 
3494: \proof
3495: According to Lemma~5.2, the singularity $C$ is contained
3496: in $\gamma\cup (H^1)^-$, where $H^1$ is a $\mathcal{ C}^{
3497: 1,\alpha}$-smooth hypersurface containing $\gamma$ which is foliated
3498: by characteristic segments.
3499: If $\lambda\in [0,1)$ is very close to $1$, the vector field
3500: over $V_\gamma^1$ defined by
3501: \def\theequation{5.10}\begin{equation}
3502: p\longmapsto 
3503: v_p^\lambda:= \lambda\cdot X_p +(1-\lambda)\cdot v_p
3504: \in T_pM^1
3505: \end{equation}
3506: is very close to the characteristic vector field $X_p$, so the
3507: integral curves of $p\mapsto v_p^\lambda$ are very close to the
3508: integral curves of $p\mapsto X_p$, which are the characteristic
3509: segments. If $\lambda$ is sufficiently close to $1$, we can choose a
3510: subneighborhood $V_\gamma^\lambda\subset V_\gamma^1$ of $\gamma$ which
3511: is foliated by integral curves of $p\mapsto v_p^\lambda$. As in
3512: Lemma~5.2, let us fix an $(n-1)$-dimensional submanifold $R^1$
3513: transversal to $\gamma$ and passing through $\gamma(0)$. Since the
3514: vector field $p\mapsto v_p^\lambda$ is very close to the
3515: characteristic vector field, it follows that after projection onto
3516: $R^1$ parallelly to the integral curves of $p\mapsto v_p^\lambda$, the
3517: closed set $C\cap V_\gamma^\lambda$ is again a proper closed subset of
3518: $R^1$. We notice that, by its very definition, the vector
3519: $v_p^\lambda$ belongs to the filled cone ${\sf FC}_p$ for all $p
3520: \in V_\gamma^\lambda$.
3521: 
3522: We can proceed exactly as in the
3523: proof of Lemma~5.2 with the foliation of $V_\gamma^\lambda$ induced by
3524: the integral curves of the vector field $p\mapsto v_p^\lambda$,
3525: instead of the characteristic foliation, except that we want a
3526: supporting hypersurface $H_{\rm sp}$ which is of class $\mathcal{
3527: C}^{2,\alpha}$. Consequently, we first approximate the vector field
3528: $p\mapsto v_p^\lambda$ by a new vector field $p\mapsto \widetilde{
3529: v}_p^\lambda$ whose coefficients are of class $\mathcal{ C}^{2,
3530: \alpha}$ (with respect to every local graphing function of $M^1$) and
3531: which is very close to the vector field $p\mapsto v_p^\lambda$ in
3532: $\mathcal{ C}^1$-norm. Again, we get a subneighborhood $\widetilde{
3533: V}_\gamma^\lambda \subset V_\gamma^\lambda$ of $\gamma$ which is
3534: foliated by integral curves of $p\mapsto \widetilde{ v}_p^\lambda$ and
3535: a projection of $C\cap \widetilde{ V}_\gamma^\lambda$ which is a
3536: proper closed subset of $R^1$. Moreover, if the approximation is
3537: sufficiently sharp, we still have $\widetilde{ v}_p^\lambda \in {\sf
3538: FC}_p$ for all $p \in \widetilde{ V}_\gamma^\lambda$. Then by
3539: repeating the reasoning which yielded Lemma~5.2, using the foliation
3540: of $\widetilde{ V}_\gamma^\lambda$ induced by $p \mapsto \widetilde{ v
3541: }_p^\lambda$, we deduce that there exists an integral curve
3542: $\widetilde{ \gamma}$ of the vector field $p\mapsto \widetilde{
3543: v}_p^\lambda$ satisfying (after reparametrization) $\widetilde{\gamma}
3544: (-1) \not \in C$, $\widetilde{\gamma}(0)\in C$ and
3545: $\widetilde{\gamma}(1) \not\in C$, together with a $\mathcal{ C}^{
3546: 2,\alpha}$-smooth supporting hypersurface $\widetilde{ H}$ of
3547: $\widetilde{ V}_\gamma^\lambda$ which contains $\widetilde{ \gamma}$
3548: such that $C$ is contained in $\widetilde{ \gamma}\cup (\widetilde{
3549: H})^-$. The fact that $\widetilde{ H}$ is of class $\mathcal{
3550: C}^{2,\alpha}$ is due to the $\mathcal{ C}^{2,\alpha}$-smoothness of
3551: the foliation on $\widetilde{ V}_\gamma^\lambda$ induced by the vector
3552: field $p\mapsto \widetilde{ v }_\gamma^\lambda$. In {\sc Figure~11}
3553: just below, suited for the case where $M^1$ is two-dimensional, we
3554: have drawn as a dotted line the limiting integral curve $\widetilde{
3555: \gamma}$ of $p\mapsto \widetilde{ v}_p^\lambda$ having the property
3556: that $C$ lies in one closed side of $\widetilde{ \gamma}$ in
3557: $\widetilde{ V}_\gamma^\lambda$.
3558: 
3559: \bigskip
3560: \begin{center}
3561: \input perturbed-vector-field.pstex_t
3562: \end{center}
3563: 
3564: To conclude the proof of Lemma~5.9, for the desired special point $p_{
3565: \rm sp}$, it suffices to choose $\widetilde{ \gamma}(0)$. For the
3566: desired local supporting hypersurface $H_{p_{\rm sp}}$, we cannot
3567: choose directly a piece of $\widetilde{ H}$ passing through $p_{\rm
3568: sp}$, because an open interval contained in $C\cap \widetilde{
3569: \gamma}$ may well be contained in $\widetilde{ H}$ by the construction
3570: in the proof of Lemma~5.2 that we have just reapplied. Fortunately,
3571: since we know that locally in a neighborhood of $p_{\rm sp}$, the
3572: closed subset $C$ is contained in $(\widetilde{ H})^-\cup \widetilde{
3573: \gamma}$, it suffices to choose for the desired supporting
3574: hypersurface $H_{p_{\rm sp}}\subset M^1$ a piece of a $\mathcal{
3575: C}^{2,\alpha}$-smooth hypersurface passing through $p_1$, tangent to
3576: $\widetilde{ H}$ at $p_1$ and satisfying $H_{p_{\rm sp}}\backslash
3577: \{p_{\rm sp}\}\subset (\widetilde{ H})^+$ in a neighborhood of $p_{\rm
3578: sp}$. Finally, for the nonzero vector $v_{\rm sp}$, it suffices to
3579: choose any positive multiple of the vector $\widetilde{ v}_{ p_{\rm
3580: sp}}^\lambda$. This completes the proof of Lemma~5.9.
3581: \endproof
3582: 
3583: In Section~8 below, property {\bf (ii)} of Lemma~5.9 together with the
3584: observation made in Lemma~5.7 will be crucial for the local $\mathcal{
3585: W}$-removability of the special point $p_{\rm sp}$.
3586: 
3587: \subsection*{5.11.~Main removability proposition in the 
3588: CR dimension $m=1$ case} We can now formulate the main removability
3589: proposition to which Theorem~3.19 {\bf (i)} is now reduced. From now
3590: on, we localize the situation at $p_{\rm sp}$, we denote this point
3591: simply by $p_1$, we denote its supporting hypersurface simply by $H^1$
3592: and we denote its associated vector simply by $v_1$. Furthermore, we
3593: localize at $p_1$ the family of analytic discs considered in Section~4
3594: for the construction of the semi-local attached half-wedge $\mathcal{
3595: HW}_\gamma^+$, hence we get a family of analytic discs $\mathcal{
3596: Z}_{t, \chi, \nu }( \zeta)$ which satisfy properties {\bf (i)}, {\bf
3597: (ii)} and {\bf (iii)} of \S4.2 and which generate a local half-wedge
3598: $\mathcal{ HW}_{p_1}^+ \subset \mathcal{ HW}_\gamma^+$ as defined
3599: in~\thetag{4.5}. At present, we deduce from our constructions achieved
3600: in Section~4 and in the beginning of Section~5 that Theorem~3.19 {\bf
3601: (i)} is now a consequence of the following main Proposition~5.12 just
3602: below, to the proof of which Sections~6, 7, 8 and~9 below are
3603: devoted. From now on, all our geometric considerations will be
3604: localized at the special point $p_1$.
3605: 
3606: \def\theproposition{5.12}\begin{proposition} 
3607: Let $M$ be a $\mathcal{ C}^{2, \alpha}$-smooth generic submanifold of
3608: $\C^n$ of CR dimension $m =1$ and of codimension $d = n-1 \geq 1$, let
3609: $M^1\subset M$ be a $\mathcal{ C}^{2, \alpha }$-smooth
3610: one-codimensional submanifold which is generic in $\C^n$, let $p_1 \in
3611: M^1$, let $H^1\subset M^1$ be a $\mathcal{ C}^{2, \alpha}$-smooth
3612: one-codimensional submanifold of $M^1$ passing through $p_1$ and let
3613: $(H^1 )^-$ denote an open local one-sided neighborhood of $H^1$ in
3614: $M^1$. Let $C \subset M^1$ be a {\rm nonempty} proper closed subset of
3615: $M^1$ with $p_1 \in C$ which is situated, locally in a neighborhood of
3616: $p_1$, only in one side of $H^1$, namely $C \subset (H^1)^- \cup \{
3617: p_1\}$. Let $\Omega$ be a neighborhood of $M \backslash C$ in $\C^n$,
3618: let $\mathcal{ HW}_{ p_1}^+$ be a local half-wedge of edge $(M^1)^+$
3619: at $p_1$ generated by a family of analytic discs $\mathcal{
3620: Z}_{t,\chi,\nu}(\zeta)$ satisfying the properties {\bf (i)}, {\bf
3621: (ii)} and {\bf (iii)} of \S4.2, let ${\sf C}_{p_1}\subset T_{p_1}M^1$
3622: be the cone created by $\mathcal{ HW}_{p_1}^+$ at $p_1$ and let ${\sf
3623: FC}_{p_1}$ be its filling. {\rm As a main assumption}, suppose that
3624: there exists a nonzero vector $v_1\in T_{p_1} H^1$ which belongs to
3625: the filled cone ${\sf FC}_{p_1}$.
3626: \begin{itemize}
3627: \item[{\bf (I)}]
3628: If $v_1$ does not belong to $T_{p_1}^cM$, then there exists a local
3629: wedge $\mathcal{ W}_{ p_1}$ of edge $M$ at $(p_1,Jv_1)$ with
3630: $\mathcal{ W}_{ p_1} \cap \left[\Omega \cup \mathcal{
3631: HW}_{p_1}^+\right]$ connected {\rm (}shrinking $\Omega\cup \mathcal{
3632: HW}_{p_1}^+$ if necessary{\rm )} such that for every holomorphic
3633: function $f\in \mathcal{ O} \left( \Omega\cup \mathcal{ HW}_{p_1}^+
3634: \right)$, there exists a holomorphic function $F \in \mathcal{ O}
3635: \left( \mathcal{ W}_1 \cup \left[\Omega \cup \mathcal{ HW}_1^+
3636: \right]\right)$ with $F\vert_{\Omega\cup \mathcal{ HW}_{p_1}^+} =f$.
3637: \item[{\bf (II)}]
3638: If $v_1$ belongs to $T_{p_1}^cM$, then there exists a neighborhood
3639: $\omega_{p_1}$ of $p_1$ in $\C^n$ with $\omega_{p_1} \cap \left[\Omega
3640: \cup \mathcal{ HW}_{p_1}^+\right]$ connected {\rm (}shrinking
3641: $\Omega\cup \mathcal{ HW}_{p_1}^+$ if necessary{\rm )} such that for
3642: every holomorphic function $f\in \mathcal{ O} \left( \Omega\cup
3643: \mathcal{ HW}_{p_1}^+ \right)$, there exists a holomorphic function $F
3644: \in \mathcal{ O} \left( \omega_{p_1} \cup \left[\Omega \cup \mathcal{
3645: HW}_1^+ \right]\right)$ with $F\vert_{\Omega\cup \mathcal{
3646: HW}_{p_1}^+} =f$.
3647: \end{itemize}
3648: \end{proposition}
3649: 
3650: In the CR dimension $m\geq 2$ case, we observe that an
3651: analogous main removability proposition may be formulated simply by
3652: adding to the assumptions of Proposition~3.22 a local half-wedge
3653: $\mathcal{ HW}_{p_1}^+$, whose existence was established in \S4.64
3654: above. The remainder of Section~5, and then Section~6, Section~7,
3655: Section~8 and Section~9 below will be entirely devoted to the proof of
3656: Proposition~5.12.
3657: 
3658: \subsection*{5.13.~A dichotomy}
3659: Under the assumptions of Proposition~5.12, we shall indeed 
3660: distinguish two cases: 
3661: \begin{itemize}
3662: \item[{\bf (I)}]
3663: The nonzero vector $v_1$ does not belong to the characteristic
3664: direction $T_{p_1}M^1\cap T_{p_1}^cM$.
3665: \item[{\bf (II)}]
3666: The nonzero vector $v_1$ belongs to the characteristic direction
3667: $T_{p_1}M^1\cap T_{p_1}^cM$.
3668: \end{itemize}
3669: 
3670: We must clarify the main assumption that $v_1$ belongs to the filling
3671: ${\sf FC}_{p_1}$
3672: of the cone ${\sf C }_{ p_1}\subset T_{p_1} M^1$ created by the local
3673: half-wedge $\mathcal{ HW}_{ p_1}^+$. As we have observed in \S4.2, in
3674: the (generic) situation of Case {\bf (I)}, a local half-wedge may be
3675: represented geometrically as the intersection of a (complete) local
3676: wedge of edge $M$ at $p_1$, with a local one-sided neighborhood $(N^1
3677: )^+$ of a hypersurface $N^1$ passing through $p_1$, which is
3678: transversal to $M$ and which satisfies $N^1 \cap M \equiv M^1$ in a
3679: neighborhood of $p_1$. The slope of the tangent space $T_{p_1} N^1$ to
3680: $N^1$ at $p_1$ with respect to the tangent space $T_{p_1} M$ to $M$ at
3681: $p_1$ may be understood in terms of the cone ${\sf C}_{p_1}$, as we
3682: will now explain. Afterwards, we shall consider Case {\bf (II)}
3683: separately.
3684: 
3685: \subsection*{5.14.~Cones, filled cones, subcones and 
3686: local description of half-wedges in Case (I)} As in some of the
3687: assumptions of Proposition~5.12, let $M$ be a $\mathcal{ C}^{2,
3688: \alpha}$-smooth generic submanifold of $\C^n$ of CR dimension $m=1$
3689: and of codimension $d=n-1 \geq 1$, let $p_1 \in M$, let $M^1$ be a
3690: $\mathcal{ C}^{2, \alpha}$-smooth one-codimensional submanifold of $M$
3691: passing through $p_1$. For the sake of concreteness, it will be
3692: convenient to work in a holomorphic coordinate system $z=(z_1, \dots,
3693: z_n)= (x_1+i y_1, \dots, x_n+ iy_n)$ centered at $p_1$ in which $T_{
3694: p_1}M=\{y_2= \cdots =y_n =0\}$ and $T_{ p_1}M^1=\{y_1= y_2= \cdots=
3695: y_n=0\}$ (the existence of such a coordinate system which straightens
3696: both $T_{p_1}M$ and $T_{p_1}M^1$ is a direct consequence of the
3697: considerations of \S4.11). Let $\pi': T_{p_1}M^1\to T_{p_1}M^1/
3698: (T_{p_1}M^1\cap T_{p_1}^cM)$ denote the canonical projection, namey
3699: $\pi'(x_1,x_2,\dots,x_n)=(x_2,\dots,x_n)$. Sometimes, we shall denote
3700: the coordinates by $(z_1,z')= (x_1+iy_1,x'+iy')\in \C\times
3701: \C^{n-1}$. In these coordinates, the characteristic direction is given
3702: by the $x_1$-axis and we may assume that the tangent plane at $p_1$ of
3703: the one-sided neighborhood $(M^1)^+$ is given by
3704: $T_{p_1}(M^1)^+=\{y'=0, \ y_1>0\}$.
3705: 
3706: Let ${\sf C}_{ p_1} \subset T_{p_1} M^1$ be the infinite open cone
3707: created by $\mathcal{ HW}_{ p_1}^+$ at $p_1$ and let ${\sf FC}_{ p_1}
3708: \subset T_{ p_1} M^1$ be its filling. Let ${\sf C}_{ p_1}':= \pi'
3709: \left( {\sf C}_{ p_1} \right)$ be its projection onto the $x'$-space,
3710: which yields an $(n-1)$-dimensional infinite cone in the $x'$-space,
3711: open with respect to its topology. Notice that, by the
3712: definition~\thetag{ 5.6} of the filling (along the characteristic
3713: direction) of a cone, the two projections $\pi'({\sf C}_{p_1})$ and
3714: $\pi'({\sf FC}_{p_1})$ are identical. We now need to explain how
3715: these three cones ${\sf C}_{p_1}$, ${\sf FC}_{p_1}$, ${\sf C}_{p_1}'$
3716: and the nonzero vector $v_1\in {\sf FC}_{p_1}$ are disposed,
3717: geometrically, {\it see} {\sc Figure~12} just below.
3718: 
3719: \bigskip
3720: \begin{center}
3721: \input cones-half-wedge.pstex_t
3722: \end{center}
3723: 
3724: Because the disc $\mathcal{ Z}_{ t, \chi, \nu}$ of Proposition~5.12
3725: (which is a localization in a neighborhood of the special point of the
3726: discs constructed in Section~4) is small, the tangent vector
3727: $\frac{\partial \mathcal{ Z}_{0, 0,0 }}{ \partial \theta} (1)$ is
3728: necessarily close to the complex tangent plane $T_{p_1 }^cM$: this may
3729: be checked directly by differentiating Bishop's equation~\thetag{4.40}
3730: with respect to $\theta$, using the fact that the $\mathcal{
3731: C}^1$-norm of $\Phi'$ is small. Moreover, since this vector $\frac{
3732: \partial \mathcal{ Z}_{ 0,0,0 }}{ \partial \theta }(1)$ also belongs
3733: to $T_{p_1} M^1$, it is in fact close to the positive
3734: $x_1$-axis. Furthermore, since the vector $v_1$ belongs to the filling
3735: of the open cone ${\sf C}_{ p_1}$ which contains the vector $\frac{
3736: \partial \mathcal{ Z}_{ 0,0,0 }}{ \partial \theta }(1)$, and since in
3737: the proof of Lemma~5.9 above we have chosen the special point, the
3738: supporting hypersurface and the vector $v_1$ with a parameter
3739: $\lambda$ very close to $1$, so that the vector field $p \mapsto
3740: \widetilde{ v}_p^\lambda$ was very close to the characteristic vector
3741: field $p \mapsto X_p$, it follows that the vector $v_1\equiv
3742: \widetilde{ v}_{ p_{ \rm sp}}^\lambda$ is even closer to the positive
3743: $x_1$-axis. However, we suppose in Case {\bf (I)} that $v_1$ is {\it
3744: not}\, directed along the $x_1$-axis, so $v_1$ has coordinates
3745: $(v_{1;1},v_{ 2;1}, \dots,v_{ n;1}) \in \R^n$ with $v_{1;1}>0$, with
3746: $\vert v_{j;1}\vert < < v_{ 1;1}$ for $j=2, \dots,n$ and with at least
3747: one $v_{j;1}$ being nonzero.
3748: 
3749: We need some general terminology. Let ${\sf C}$ be an open infinite
3750: cone in a real linear subspace $E$ of dimension $q\geq 1$. We say
3751: that ${\sf C}'$ is a {\sl proper subcone} and we write ${\sf
3752: C}'\subset \subset {\sf C}$ ({\it see} the left hand side of {\sc
3753: Figure~10} above for an illustration) if the intersection of ${\sf
3754: C}'$ with the unit sphere of $E$ is a relatively compact subset of the
3755: intersection of ${\sf C}$ with the unit sphere of $E$, this property
3756: being independent of the choice of a norm on $E$. We say that ${\sf
3757: C}$ is a {\sl linear cone} if it may be defined by ${\sf C}=\{x\in E:
3758: \ \ell_1(x)>0,\dots,\ell_q(x)>0\}$ for some $q$ linearly independent
3759: real linear forms $\ell_1,\dots,\ell_q$ on $E$.
3760: 
3761: In the $(x_2,\dots,x_n)$-space, we now choose an open infinite
3762: strictly convex linear proper subcone ${\sf C}_1'\subset \subset {\sf
3763: C }_{ p_1}'$ with the property that $v_1$ belongs to its filling ${\sf
3764: FC}_1'$, {\it cf.} {\sc Figure~12} above. Here, we may assume that
3765: ${\sf C}_1'$ is described by $(n-1)$ strict inequalities $\ell_1'(x')
3766: >0, \dots, \ell_{n-1}'(x')>0$, where the $\ell_j'( x')$ are linearly
3767: independent linear forms. It then follows that there exists a linear
3768: form $\sigma(x_1,x')$ of the form $\sigma(x_1, x')= x_1+a_2x_2+ \cdots
3769: +a_nx_n$ such that the original filled cone ${\sf FC}_{p_1}$ is
3770: contained in the linear cone
3771: \def\theequation{5.15}\begin{equation}
3772: {\sf C}_1:= \left\{
3773: (x_1,x')\in \R^n : \
3774: \ell_1'(x')>0,\dots,
3775: \ell_{n-1}'(x')>0, \ 
3776: \sigma(x_1,x')>0
3777: \right\},
3778: \end{equation}
3779: which contains the vector $v_1$. This cone is automatically filled,
3780: namely ${\sf C}_1\equiv {\sf FC}_1$.
3781: 
3782: We remind that by genericity of $M$, the complex structure $J$ of
3783: $T\C^n$ induces an isomorphim $T_{p_1}M/T_{p_1}^cM \to
3784: T_{p_1}\C^n/T_{p_1}M$. Hence $J{\sf C}_{p_1}'$ and $J{\sf C}_1'$ are
3785: open infinite strictly convex linear proper cones in
3786: $T_{p_1}\C^n/T_{p_1}M\cong \{(0,y')\in \C^n\}$. Since $J{\sf C}_1'$
3787: is a proper subcone of $J{\sf C}_{p_1}'$ and since in the classical
3788: definition of a wedge, only the projection of the cone on the quotient
3789: space $T_{p_1}M/T_{p_1}^cM$ has a contribution to the wedge, it then
3790: follows that the complete wedge $\mathcal{ W}_{p_1}$ associated to the
3791: family $\mathcal{ Z}_{t,\chi,\nu}(\zeta)$ ({\it cf.} the paragraph
3792: after~\thetag{4.5}) contains a wedge of the form
3793: \def\theequation{5.16}\begin{equation}
3794: \mathcal{ W}_1:=\left\{
3795: p+{\sf c}_1': \ p\in M, \ 
3796: {\sf c}_1' \in J{\sf C}_1'
3797: \right\}\cap \Delta_n(p_1,\delta_1),
3798: \end{equation}
3799: for some $\delta_1$ with $0 < \delta_1 < \varepsilon$, where
3800: $\varepsilon$ is as in \S4.2. Furthermore, as observed in \S4.2,
3801: there exists a $\mathcal{ C}^{2, \alpha}$-smooth hypersurface $N^1$ of
3802: $\C^n$ passing through $p_1$ with the property that $N^1 \cap M \equiv
3803: M^1$ locally in a neighborhood of $p_1$ such that, shrinking $\delta_1
3804: >0$ if necessary, the local half-wedge $\mathcal{ HW}_{ p_1}^+$
3805: contains a local half-wedge $\mathcal{ HW}_1^+$ of edge $(M^1)^+$ at
3806: $p_1$ which is described as the geometric intersection of the complete
3807: wedge $\mathcal{ W}_{p_1}$ with a one-sided neighborhood $(N^1)^+$,
3808: namely
3809: \def\theequation{5.17}\begin{equation}
3810: \mathcal{ HW}_1^+:=
3811: \mathcal{ W}_1\cap (N^1)^+.
3812: \end{equation}
3813: An illustration for the case $n =2$ where $M \subset \C^2$ is a
3814: hypesurface is provided in the left hand side of {\sc Figure~12}. In
3815: addition, it follows from the definition of $\mathcal{ HW}_{p_1}^+$ by
3816: means of the segments $\mathcal{ Z}_{t, \chi, \nu} \left ((1-
3817: \varepsilon, 1) \right)$ that we can assume that
3818: \def\theequation{5.18}\begin{equation}
3819: T_{p_1}(N^1)^+=T_{p_1}M \oplus J(\Sigma^1)^+,
3820: \end{equation}
3821: where $(\Sigma_1 )^+$ is the hyperplane one-sided neighborhood $\{(
3822: x_1, x'): \ \sigma (x_1, x')>0\} \subset T_{p_1} M^1$. Equivalently,
3823: $T_{p_1}(N^1)^+$ is represented by the inequality
3824: $y_1+a_2y_2+\cdots+a_ny_n>0$. Consequently, there exists a $\mathcal{
3825: C}^{2,\alpha}$-smooth function $\psi(x,y')$ with
3826: $\psi(0)=\partial_{x_k}\psi(0)= \partial_{y_j}\psi(0)=0$ for
3827: $k=1,\dots,n$ and $j=2,\dots,n$ such that $N^1$ is represented by the
3828: equation $y_1+a_2y_2+\cdots+a_ny_n=\psi(x,y')$ and $(N^1)^+$ by the
3829: inequation $y_1+a_2y_2+\cdots+a_ny_n>\psi(x,y')$.
3830: 
3831: \subsection*{5.19.~Cones, filled cones, subcones and 
3832: local description of half-wedges in Case (II)} Secondly, we assume
3833: that the nonzero vector $v_1$ of Proposition~5.12 belongs to the
3834: characteristic direction $T_{p_1 }M^1 \cap T_{p_1 }^cM$. In this
3835: case, as observed in \S4.2, the half-wedge $\mathcal{ HW}_{p_1}^+$
3836: coincides with a local wedge of edge $M^1$ at $(p_1,Jv_1)$. After a
3837: real dilatation of the $z_1$-axis, we can assume that
3838: $v_1=(1,0,\dots,0)$. Choosing an open infinite strictly convex linear
3839: proper subcone ${\sf C}_2\subset \subset {\sf C}_{p_1}\subset
3840: T_{p_1}M^1=\R_x^n$ defined by $n$ strict inequalities
3841: $\ell_1(x)>0,\dots, \ell_n(x)>0$, where the $\ell_j(x)$ are linearly
3842: independent real linear forms\,--\,of course with ${\sf C}_2$
3843: containing the vector $v_1$\,--\,it follows that there exists
3844: $\delta_1>0$ such that the half-wedge $\mathcal{ HW}_{p_1}^+$ contains
3845: the following local wedge of edge $M^1$ at $p_1$:
3846: \def\theequation{5.20}\begin{equation}
3847: \mathcal{ W}_2:=\left\{
3848: p+{\sf c}_2: \ 
3849: p\in M^1, \
3850: {\sf c}_2 \in J{\sf C}_2
3851: \right\}\cap \Delta_n
3852: (p_1,\delta_1).
3853: \end{equation}
3854: We remind that it was observed in \S4.2 ({\it cf.} especially the
3855: right hand side of {\sc Figure}~5) that $\mathcal{ W}_2$ contains
3856: $(M^1)^+$ locally in a neighborhood of $p_1$. In \S5.22 below, we
3857: shall provide a more concrete representation of $\mathcal{ W}_2$
3858: in an appropriate system of coordinates.
3859: 
3860: \subsection*{5.21.~A trichotomy}
3861: Let us pursue this discussion more concretely by introducing further
3862: normalizations. Our goal will now be to construct appropriate
3863: normalized coordinate systems. Analyzing further the dichotomy 
3864: introduced in \S5.13 by taking account of the 
3865: presence of the one-codimensional 
3866: submanifold $H^1\subset M^1$,
3867: we shall distinguish three cases by dividing Case {\bf (I)} in two
3868: subcases {\bf ($\mathbf{ I_1}$)} and {\bf ($\mathbf{ I_2}$)} as
3869: follows:
3870: \begin{itemize}
3871: \item[{\bf ($\mathbf{I_1}$)}]
3872: The nonzero vector $v_1$ does not belong to the characteristic
3873: direction $T_{p_1}M^1\cap T_{p_1}^cM$ and
3874: $\dim_\R \left(T_{p_1}H^1 \cap T_{p_1}^cM \right)=0$.
3875: \item[{\bf ($\mathbf{I_2}$)}]
3876: The nonzero vector 
3877: $v_1$ does not belong to the characteristic direction 
3878: $T_{p_1}M^1\cap T_{p_1}^cM$ and
3879: $\dim_\R \left(T_{p_1}H^1 \cap T_{p_1}^cM \right)=1$ 
3880: (this possibility can only occur when $n\geq 3$).
3881: \item[{\bf (II)}]
3882: The nonzero vector $v_1$ belongs to the
3883: characteristic direction $T_{p_1}M^1\cap 
3884: T_{p_1}^cM$.
3885: \end{itemize}
3886: 
3887: In case {\bf ($\mathbf{I_1}$)}, we notice that the
3888: assumption $T_{p_1}H^1\cap T_{p_1}^cM=\{0\}$ implies that
3889: $v_1$ does not belong to the characteristic direction, 
3890: because $v_1\in T_{p_1}H^1$. Also, 
3891: in case {\bf (II)}, we notice that $\dim_\R \left(T_{ p_1} H^1 \cap
3892: T_{ p_1}^c M \right)=1$ because $v_1 \in T_{ p_1} H^1$, because
3893: $T_{p_1}H^1 \subset T_{ p_1}M^1$ and because the characteristic
3894: direction $T_{p_1}M^1\cap T_{p_1}^cM$ is one-dimensional.
3895: 
3896: In each of the above three cases, it will be convenient in
3897: Section~8 below to work with simultaneously normalized defining
3898: (in)equations for $M$, for $M^1$, for $(M^1)^+$, for $H^1$, for $(H^1
3899: )^+$, for ${\sf C }_1'$, for $v_1$, for ${\sf C }_1$, for $(N^1)^+$
3900: and for $\mathcal{ HW}_{ p_1}^+$, in a single coordinate system
3901: centered at $p_1$. In the next paragraphs, we shall set up further
3902: elementary normalization lemmas in a {\it common system of
3903: coordinates}, firstly for Case {\bf ($\mathbf{I_1}$)}, secondly for
3904: Case {\bf ($\mathbf{I_2}$)} and thirdly for Case {\bf (II)}.
3905: 
3906: First of all, in the above coordinate system $(z_1,z')$ with $T_{p_1
3907: }M=\{ y_2= \cdots= y_n =0\}$ and with $T_{p_1 }M^1=\{y_1 = y_2=
3908: \cdots= y_n=0\}$, by means of the implicit function theorem, we can
3909: represent locally $M$ by $(n-1)$ grahed equations of the form $y_2=
3910: \varphi_2 (x,y_1),\dots, y_n= \varphi_n (x,y_1)$, where
3911: the $\varphi_j$ are $\mathcal{ C}^{2, \alpha}$-smooth functions
3912: satisfying $\varphi_j(0)= \partial_{ x_k} \varphi_j(0)= \partial_{y_1}
3913: \varphi_j (0)= 0 $ for $j=2,\dots,n$, $k=1,\dots,n$ and
3914: we can represent $M^1$ by $n$ graphed equations $y_1= h_1(x), y_2=
3915: h_2(x), \dots,y_n= h_n(x)$, where the $h_j$ are $\mathcal{
3916: C}^{2,\alpha}$-smooth functions satisfying
3917: $h_j(0)=\partial_{x_k}h_j(0)=0$ for $j,k=1, \dots,n$.
3918: 
3919: \subsection*{5.22.~First order normalizations in Case 
3920: {\bf ($\mathbf{I_1}$)}} Thus, let us deal first with Case {\bf
3921: ($\mathbf{I_1}$)}. After a possible permutation of coordinates, we can
3922: assume that $T_{p_1}H^1$, which is a one-codimensional subspace of
3923: $T_{p_1}M^1$, is given by the equations
3924: \def\theequation{5.23}\begin{equation}
3925: x_1=b_2x_2+\cdots+b_nx_n, \ 
3926: y_1=0,y'=0,
3927: \end{equation}
3928: for some real numbers $b_2,\dots,b_n$. If we define the linear
3929: invertible transformation $\widehat{z}_1:=z_1-b_2z_2-\cdots-b_nz_n$,
3930: $\widehat{z}':=z'$, then the plane $T_{ p_1} H^1$ written
3931: in~\thetag{5.23} clearly transforms to the plane $\widehat{
3932: x}_1=\widehat{ y}_1= \widehat{ y}'=0$, and (fortunately) $T_{ p_1}M$
3933: and $T_{ p_1} M^1$ are left unchanged, namely $T_{p_1} \widehat{M}=\{
3934: \widehat{y}'=0\}$ and $T_{p_1}\widehat{ M^1}= \{\widehat{ y}_1=
3935: \widehat{ y}'=0\}$.
3936: 
3937: Dropping the hats on coordinates, we have $T_{p_1}M=\{y'=0\}$,
3938: $T_{p_1}M^1=\{y_1=y'=0\}$, $T_{p_1}H^1=\{x_1=y_1=y'=0\}$. Let ${\sf
3939: C}_1'\subset \subset {\sf C}_{p_1}'$ be the open infinite strictly
3940: convex linear cone introduced in \S5.14, which is contained in the
3941: real $(n-1)$-dimensional space $\{(0,x')\}$ and which is defined by
3942: $(n-1)$ strict inequalities $\ell_1'(x')>0,\dots,
3943: \ell_{n-1}'(x')>0$. By means of a real linear invertible
3944: transformation of the form $\widehat{ z}_1:= z_1$, $\widehat{ z}':=
3945: A'\cdot z'$, where $A'$ is an $(n-1)\times (n-1)$ real matrix, we can
3946: transform ${\sf C}_1'$ to a cone $\widehat{\sf C}_1'$ defined by the
3947: simpler inequalities $\widehat{ x}_2>0,\dots, \widehat{ x}_n>0$.
3948: Fortunately, this transformation stabilizes $T_{p_1}M$, $T_{p_1}M^1$
3949: and $T_{p_1}H^1$.
3950: 
3951: Dropping the hats on coordinates, we now have $T_{p_1}M=\{y'=0\}$,
3952: $T_{p_1}M^1=\{y_1=y'=0\}$, $T_{p_1}H^1=\{x_1=y_1=y'=0\}$ and ${\sf
3953: C}_1'=\{(0,x'): \ x_2>0,\dots,x_n>0\}$. Then the nonzero vector
3954: $v_1\in T_{p_1}H^1$ which belongs to ${\sf C}_1'$ has coordinates
3955: $v_1=(0,v_{2;1},\dots, v_{n;1})\in\R^n$, where
3956: $v_{2;1}>0,\dots,v_{n;1}>0$. By means of real dilatations or real
3957: contractions of the real axes $\R_{x_2},\dots, \R_{x_n}$ (a
3958: transformation which does not perturb the previously achieved
3959: normalizations), we can assume that $v_1=(0,1,\dots,1)$ and that
3960: $T_{p_1}(M^1)^+=\{y'=0, y_1>0\}$, $T_{p_1}(H^1)^+=\{y=0, \ x_1>0\}$.
3961: 
3962: Finally, the linear one-codimensional subspace $\sigma_1\subset
3963: T_{p_1}M^1$ introduced in \S5.14 which does not contain the
3964: characteristic direction $T_{p_1}M^1\cap T_{p_1}^cM \equiv \R_{x_1}$
3965: may be represented by an equation of the form
3966: $\sigma(x_1,x'):=x_1+a_2x_2+\cdots+a_nx_n=0$, for some real numbers
3967: $a_2,\dots,a_n$. By~\thetag{5.15}, the vector $v_1$ belongs to the
3968: cone ${\sf C}_1$, hence $a_2+\dots+a_n>0$. After a dilatation of the
3969: $x_1$-axis, we can even assume that $a_2+\dots+a_n=1$. We remind that
3970: by~\thetag{5.18}, the half-space $T_{p_1}(N^1)^+$ is given by
3971: $y_1+a_2y_2+\cdots+a_ny_n>0$, hence there exists a $\mathcal{
3972: C}^{2,\alpha}$-smooth function $\psi(x,y')$ with
3973: $\psi(0)=\partial_{x_k}\psi(0)= \partial_{y_j}\psi(0)=0$ for
3974: $k=1,\dots,n$ and $j=2,\dots,n$ such that $(N^1)^+$ is represented by
3975: the inequation $y_1+a_2y_2+\cdots+a_ny_n>
3976: \psi(x,y')$. Consequently, in this coordinate system, we may
3977: represent concretely the local half-wedge $\mathcal{ HW}_1^+\subset
3978: \mathcal{ HW}_{p_1}^+$ constructed in \S5.14 as
3979: \def\theequation{5.24}\begin{equation}
3980: \left\{
3981: \aligned
3982: {}
3983: &
3984: \mathcal{ HW}_1^+=
3985: \left\{
3986: (z_1,z')\in \C^n:
3987: \vert z_1 \vert < \delta_1, \
3988: \vert z' \vert < \delta_1, \right. \\
3989: & \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \
3990: y_1+a_2y_2+\dots+a_ny_n-\psi(x,y')>0,\\
3991: & \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \
3992: \left.
3993: y_2-\varphi_2(x,y_1) > 0, \dots, 
3994: y_n-\varphi_n(x,y_1) >0 \ 
3995: \right\}.
3996: \endaligned\right.
3997: \end{equation}
3998: For the continuation of the proof of Proposition~5.12, it will also be
3999: convenient to proceed to further second order normalizations of the
4000: totally real submanifolds $M^1$ and $H^1$. These normalizations will
4001: all be tangent to the identity tranformation, hence they will leave the
4002: previously achieved normalizations unchanged.
4003: 
4004: \subsection*{5.25.~Second order normalizations in Case 
4005: ($\mathbf{ I_1}$)}
4006: Let us then perform a second order Taylor development of the defining
4007: equations of $M^1$
4008: \def\theequation{5.26}\begin{equation}
4009: y=h(x)=\sum_{k_1,k_2=1}^n\, 
4010: a_{k_1,k_2}\, x_{k_1} x_{k_2}+o(\vert x \vert^2),
4011: \end{equation}
4012: where the $a_{k_1,k_2}=\frac{1}{2}
4013: \partial_{x_{k_1}}
4014: \partial_{x_{k_2}}h (0)$ are vectors of $\R^n$.
4015: If we define the quadratic invertible transformation
4016: \def\theequation{5.27}\begin{equation}
4017: \widehat{z}:= z-i\sum_{k_1,k_2=1}^n\, 
4018: a_{k_1,k_2}\, z_{k_1} z_{k_2}=\Phi(z), 
4019: \end{equation}
4020: which is tangent to the identity mapping at the origin, 
4021: then for $x+iy=x+ih(x)\in M^1$, 
4022: we have by replacing~\thetag{5.26} in the imaginary 
4023: part of $\widehat{z}$ given by~\thetag{5.27}
4024: \def\theequation{5.28}\begin{equation}
4025: \left\{
4026: \aligned
4027: \widehat{y}= 
4028: & \
4029: y - \sum_{k_1,k_2=1}^n\, 
4030: a_{k_1,k_2}\, x_{k_1}x_{k_2}+
4031: \sum_{k_1,k_2=1}^n\, a_{k_1,k_2}\, y_{k_1}y_{k_2} \\
4032: = 
4033: & \ 
4034: {\rm o}(\vert x \vert^2) \\
4035: = 
4036: & \
4037: {\rm o}\left(\left\vert
4038: {\rm Re}\, \Phi^{-1}(\widehat{z})
4039: \right\vert^2\right)= {\rm o}\left( \left\vert
4040: (\widehat{x}, \widehat{y})
4041: \right\vert^2 \right),
4042: \endaligned\right.
4043: \end{equation}
4044: whence by applying the $\mathcal{ C}^{2,\alpha}$ implicit function
4045: theorem to solve~\thetag{5.28} in terms of $\widehat{ y}$, we find
4046: that $\widehat{ M^1}:=\Phi( M^1)$ may be represented by an equation of
4047: the form $\widehat{ y}= \widehat{ h}(\widehat{ x})$, for some
4048: $\R^n$-valued local $\mathcal{ C}^{2,\alpha}$-smooth mapping
4049: $\widehat{ h}$ which satisfies $\widehat{ h}( \widehat{ x})= {\rm o}(
4050: \vert \widehat{ x} \vert^2)$.
4051: 
4052: Finally, dropping the hats on coordinates, we can assume that the
4053: functions $h_1, \dots, h_n$ vanish at the origin to second order.
4054: Since $T_{p_1}H^1=\{y=0, \ x_1=0\}$, there exists a $\mathcal{
4055: C}^{2,\alpha}$-smooth function $g(x')$ with $g(0)=\partial_{x_k}
4056: g(0)=0$ for $k=2,\dots,n$ such that $(H^1)^+$ is given by the equation
4057: $x_1 >g(x')$. We want to normalize also the defining equation
4058: $x_1=g(x')$ of $H^1$. Instead of requiring, similarly as for
4059: $h_1,\dots,h_n$, that $g$ vanishes to second order at the origin
4060: (which would be possible), we shall normalize $g$ in order that
4061: $g(x')=-x_1^2-\cdots - x_n^2+ {\rm o}\left(\vert x' \vert^{2+\alpha}
4062: \right)$ (which will also be possible, thanks to the total reality of
4063: $H^1$). The reason why we want $g$ to be strictly concave is a trick
4064: that will be useful in Section~8 below.
4065: 
4066: Thus, we now perform a second
4067: order Taylor development of the defining equations of $H^1$
4068: \def\theequation{5.29}\begin{equation}
4069: \left\{
4070: \aligned
4071: {}
4072: &
4073: x_1=g\left(x'\right)=\sum_{k_1,k_2=2}^n\, 
4074: b_{k_1,k_2} \, x_{k_1}x_{k_2}+ o (\left\vert
4075: x'
4076: \right\vert^2),\\
4077: &
4078: y=h\left(g(x'),x'\right)=:k\left(x'\right)=
4079: {\rm o}(\left\vert x' \right\vert^2),
4080: \endaligned\right.
4081: \end{equation}
4082: where the $b_{k_1,k_2}=\frac{1}{2}
4083: \partial_{x_{k_1}}
4084: \partial_{x_{k_2}} g(0)$
4085: are real numbers. If we define the quadratic invertible transformation
4086: \def\theequation{5.30}\begin{equation}
4087: \left\{
4088: \aligned
4089: {}
4090: &
4091: \widehat{z}_1:= z_1-\sum_{k_1,k_2=2}^n\, 
4092: b_{k_1,k_2}\, z_{k_1}z_{k_2} -z_2^2-\cdots-z_n^2, \\
4093: &
4094: \widehat{z}':= z', 
4095: \endaligned\right.
4096: \end{equation}
4097: which is tangent to the identity mapping, then 
4098: for $\left(g(x')+ik_1(x'),x'+ik'(x')\right)\in H^1$,
4099: we have by replacing~\thetag{5.29} in the 
4100: real part of $\widehat{z}_1$, given by~\thetag{5.30}:
4101: \def\theequation{5.31}\begin{equation}
4102: \left\{
4103: \aligned
4104: \widehat{x}_1=
4105: & \
4106: x_1-\sum_{k_1,k_2=2}^n\, b_{k_1,
4107: k_2}\, x_{k_1}x_{k_2}+\sum_{k_1,k_2=2}^n\, 
4108: b_{k_1,k_2}\, y_{k_1}y_{k_2}-
4109: \sum_{k=2}^n\, x_k^2+\sum_{k=2}^n\, 
4110: y_k^2, \\
4111: =
4112: & \
4113: -x_2^2-\cdots-x_n^2+{\rm o}\left(
4114: \left\vert x' \right\vert^2\right) \\
4115: =
4116: & \
4117: -\widehat{x}_2^2-\cdots-\widehat{x}_n^2-{\rm o}\left(\left\vert
4118: (\widehat{x},\widehat{y})\right\vert^2\right).
4119: \endaligned\right.
4120: \end{equation}
4121: Similarly (dropping the elementary computations), we may obtain for
4122: the imaginary part of $\widehat{z}_1$ and for the imaginary part of
4123: $\widehat{z}'$
4124: \def\theequation{5.32}\begin{equation}
4125: \widehat{y}_1={\rm o}\left(\left\vert
4126: (\widehat{x},\widehat{y})\right\vert^2\right) \ \ \ \ \ 
4127: {\rm and} \ \ \ \ \ 
4128: \widehat{y}'={\rm o}\left(\left\vert
4129: (\widehat{x},\widehat{y})\right\vert^2\right),
4130: \end{equation}
4131: whence by applying the $\mathcal{ C}^{2,\alpha}$ implicit function
4132: theorem to solve the system~\thetag{5.31}, \thetag{5.32} in terms
4133: of $\widehat{x}_1$, $\widehat{y}_1$ and $\widehat{y}'$, we find
4134: that $\widehat{ H}^1:=\Phi(H^1)$ may be represented 
4135: by equations
4136: of the form
4137: \def\theequation{5.33}\begin{equation}
4138: \left\{
4139: \aligned
4140: \widehat{x}_1=
4141: & \
4142: \widehat{g}\left(\widehat{x}'\right)=
4143: -\widehat{x}_2^2-\cdots-\widehat{x}_n^2+
4144: {\rm o}\left(
4145: \left\vert
4146: \widehat{x}'
4147: \right\vert^2
4148: \right), \\
4149: \widehat{y}=
4150: & \
4151: \widehat{k}\left(\widehat{x}'\right)=o\left(
4152: \left\vert
4153: \widehat{x}'
4154: \right\vert^2
4155: \right).
4156: \endaligned\right.
4157: \end{equation}
4158: It remains to check that the above transformation has not perturbed
4159: the previous second order normalizations of $h_1, \dots, h_n$ (this is
4160: important), which is easy: replacing $y$ by $h(x) = {\rm o} (\vert x
4161: \vert^2)$ in the imaginary parts of $\widehat{z}_1$ and of
4162: $\widehat{z}'$ defined by the transformation~\thetag{5.30}, we get
4163: firstly
4164: \def\theequation{5.34}\begin{equation}
4165: \left\{
4166: \aligned
4167: \widehat{y}_1 =
4168: & \
4169: y_1 -\sum_{k_1,k_2}^n\, 
4170: b_{k_1,k_2}\, \left(
4171: x_{k_1}y_{k_2}+y_{k_1}x_{k_2}
4172: \right)-2\sum_{k=2}^n\, 
4173: x_ky_k \\
4174: =
4175: & \
4176: {\rm o}(\vert x\vert^2) \\
4177: =
4178: & \
4179: {\rm o}\left(\left\vert
4180: {\rm Re}\, \Phi^{-1}(\widehat{z})
4181: \right\vert^2\right)= {\rm o}\left( \left\vert
4182: (\widehat{x}, \widehat{y})
4183: \right\vert^2 \right),
4184: \endaligned\right.
4185: \end{equation}
4186: and similarly 
4187: \def\theequation{5.35}\begin{equation}
4188: \widehat{y}'=o\left( \left\vert
4189: (\widehat{x}, \widehat{y})
4190: \right\vert^2 \right),
4191: \end{equation}
4192: whence by applying the $\mathcal{ C}^{2,\alpha}$ implicit function
4193: theorem to solve the system~\thetag{5.34}, \thetag{5.35} in terms of
4194: $\widehat{y}$, we find that $\widehat{ M}^1 :=\Phi(M^1)$ may be
4195: represented by equations of the form $\widehat{y}=\widehat{ h} \left(
4196: \widehat{x} \right)= {\rm o}\left(\left \vert \widehat{ x}\right \vert^2
4197: \right)$. Thus, after dropping the hats on coordinates, all the
4198: desired normalizations are satisfied. We shall now summarize these
4199: normalizations and we shall formulate just afterwards the analogous
4200: normalizations for Cases {\bf ($\mathbf{I_2}$)} and {\bf (II)}.
4201: 
4202: \subsection*{5.36.~Simultaneous normalization lemma}
4203: In the following lemma, the final choice of sufficiently small radii
4204: $\rho_1>0$ and $\delta_1>0$ is made after that all the biholomorphic
4205: changes of coordinates and all the applications of the implicit function
4206: theorem are achieved.
4207: 
4208: \def\thelemma{5.37}\begin{lemma}
4209: Let $M$, $M^1$, $p_1$, $H^1$, $(H^1)^+$, $\mathcal{ HW}_{p_1}^+$,
4210: ${\sf C}_{p_1}$ and ${\sf FC}_{p_1}$ be as in Proposition~5.12. Then
4211: there exists a sub-half-wedge $\mathcal{ HW}_1^+$ contained in
4212: $\mathcal{ HW}_{p_1}^+$ such that the following normalizations hold in
4213: each of the three cases {\bf ($\mathbf{I_1}$)}, {\bf
4214: ($\mathbf{I_2}$)} and
4215: {\bf (II)}{\rm :}
4216: 
4217: \smallskip
4218: \begin{itemize}
4219: \item[{\bf ($\mathbf{I_1}$)}]
4220: If $\dim_\R \left( T_{ p_1} H^1 \cap T_{p_1 }^cM \right) =0$, then
4221: there exists a system of holomorphic coordinates $z= (z_1, \dots,
4222: z_n)=(x_1+i y_1,\dots, x_n+iy_n)$ vanishing at $p_1$ with the vector
4223: $v_1$ equal to $(0,1,\dots,1)$, there exists positive numbers $\rho_1$
4224: and $\delta_1$ with $0< \delta_1 < \rho_1$, there exist $\mathcal{
4225: C}^{2,\alpha}$-smooth functions $\varphi_2,\dots, \varphi_n$, $h_1,
4226: \dots, h_n$, $g$, $k_1, \dots,k_n,\psi$, all defined in real cubes of
4227: edge $2 \rho_1$ and of the appropriate dimension, and there exist real
4228: numbers $a_1,\dots,a_n$ with $a_2+\cdots+a_n=1$, such that, if we
4229: denote $z':=(z_2, \dots, z_n)=x' +iy'$, then $M$, $M^1$, $(M^1)^+$,
4230: $H^1$, $(H^1)^+$ and $N^1$ are represented in the polydisc of radius
4231: $\rho_1$ centered at $p_1$ by the following graphed (in)equations and
4232: the sub-half-wedge $\mathcal{ HW}_1^+ \subset \mathcal{ HW}_{p_1}^+$ is
4233: represented in the polydisc of radius $\delta_1$ centered at $p_1$ by
4234: the following inequations
4235: \def\theequation{5.38}\begin{equation}
4236: \left\{
4237: \aligned
4238: M: \ \ \ \ \ 
4239: & \
4240: y_2=\varphi_2(x,y_1), \dots\dots, \
4241: y_n=\varphi_n(x,y_1), \\
4242: M^1: \ \ \ \ \ 
4243: & \ 
4244: y_1=h_1(x),y_2=h_2(x),\dots\dots, \ 
4245: y_n=h_n(x), \\
4246: (M^1)^+: \ \ \ \ \ 
4247: & \ 
4248: y_1>h_1(x),y_2=\varphi_2(x,y_1),\dots\dots, \ 
4249: y_n=\varphi_n(x,y_1), \\
4250: H^1: \ \ \ \ \ 
4251: & \
4252: x_1=g(x'), \
4253: y_1=k_1(x'), \dots\dots, \ 
4254: y_n=k_n(x'), \\
4255: (H^1)^+: \ \ \ \ \ 
4256: & \
4257: x_1>g(x'), \
4258: y_1=h_1(x),y_2=h_2(x), \dots\dots, \ 
4259: y_n=h_n(x), \\
4260: N^1 : \ \ \ \ \ 
4261: & \
4262: y_1+a_2y_2+\cdots+a_ny_n=\psi(x,y'), \\
4263: \mathcal{ HW}_1^+: \ \ \ \ \
4264: & \
4265: y_1+a_2y_2+\cdots+a_ny_n>\psi(x,y'), \\
4266: & \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ 
4267: y_2>\varphi_2(x,y_1),\dots, y_n>\varphi_n(x,y_1), 
4268: \endaligned\right.
4269: \end{equation}
4270: where we can assume that $M^1$ coincides with the intersection $M\cap
4271: \{y_1=h_1(x)\}$, that $H^1$ coincides with the intersection
4272: $M^1\cap \{x_1=g(x')\}$ and that $N^1$ contains $M^1$, which yields at
4273: the level of defining equations the following three collections of
4274: identities
4275: \def\theequation{5.39}\begin{equation}
4276: \left\{
4277: \aligned
4278: h_2(x)\equiv 
4279: & \
4280: \varphi_2(x,h_1(x)),\dots\dots, h_n(x)\equiv 
4281: \varphi_n(x,h_1(x)), \\
4282: k_1(x')\equiv
4283: & \
4284: h_1(g(x'),x'), \dots\dots,
4285: k_n(x')\equiv h_n(g(x'),x'), \\
4286: \psi(x,h'(x)) \equiv
4287: & \
4288: h_1(x)+a_2h_2(x)+\cdots+
4289: a_nh_n(x),
4290: \endaligned\right.
4291: \end{equation}
4292: and where the following normalizations hold (where $\delta_a^b$, equal
4293: to $0$ if $a\neq b$ and to $1$ if $a=b$, denotes Kronecker's
4294: symbol){\rm :}
4295: \def\theequation{5.40}\begin{equation}
4296: \left\{
4297: \aligned
4298: {}
4299: &
4300: \varphi_{j}(0)=\partial_{x_k}\varphi_j(0)
4301: =\partial_{y_1}\varphi_j(0)=0, \ \ \ \ \ j=2,\dots,n, \
4302: k=1,\dots,n,\\
4303: &
4304: h_j(0)=\partial_{x_k}
4305: h_j(0)=\partial_{x_{k_1}}
4306: \partial_{x_{k_2}}h_j(0)=0, \ \ \ \ \ \ \ \ \ \
4307: j,k,k_1,k_2=1,\dots,n, \\
4308: &
4309: g(0)=\partial_{x_k}
4310: g(0)=k_j(0)=\partial_{x_k}k_j(0)=0, \ \ \ \ \ 
4311: j=1,\dots,n, \ k=2,\dots,n, \\
4312: & \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \
4313: \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \
4314: \partial_{x_{k_1}}\partial_{x_{k_2}}
4315: g(0)=-\delta_{k_1}^{k_2}, \ \ \ \ \ \ \ 
4316: k_1,k_2=2,\dots,n, \\
4317: &
4318: \psi(0)=\partial_{x_k}\psi(0)=\partial_{y_j}\psi(0)=0, \ \ \ \ \
4319: k=1,\dots,n, \ j=2,\dots,n.
4320: \endaligned\right.
4321: \end{equation}
4322: In other words, $T_0M=\{y'=0\}$ (hence $T_0^cM$ coincides with the
4323: complex $z_1$-axis), $T_0N^1=\{ y_1+a_2y_2+\cdots+a_ny_n=0\}$ and the
4324: second order Taylor approximations of the defining equations of $M^1$,
4325: of $H^1$ and of $(H^1)^+$ are the quadrics
4326: \def\theequation{5.41}\begin{equation}
4327: \left\{
4328: \aligned
4329: T_{p_1}^{(2)}M^1: \ \ \ \ \ 
4330: & \
4331: y_1=0, \dots\dots, \ y_n=0,
4332: \\
4333: T_{p_1}^{(2)}H^1: \ \ \ \ \ 
4334: & \
4335: x_1=-x_2^2-\cdots-x_n^2, \ y_1=0,\dots\dots, \ y_n=0, \\
4336: T_{p_1}^{(2)}(H^1)^+: \ \ \ \ \ 
4337: & \
4338: x_1>-x_2^2-\cdots-x_n^2, \ y_1=0,\dots\dots, \ y_n=0.
4339: \endaligned\right.
4340: \end{equation}
4341: \item[{\bf ($\mathbf{I_2}$)}]
4342: Similarly, if $\dim_\R \left( T_{p_1}H^1\cap T_{p_1}^c M\right) = 1$
4343: and if $v_1$ is not complex tangent to $M$ {\rm (}this possibility can
4344: only occur in the case $n \geq 3${\rm )}, then there exists a system
4345: of holomorphic coordinates $z= (z_1,\dots,z_n)=
4346: (x_1+iy_1,\dots,x_n+iy_n)$ vanishing at $p_1$ with $v_1$ equal to
4347: $(1,\dots,1,0)$, there exists positive numbers $\rho_1$ and $\delta_1$
4348: with $0< \delta_1 < \rho_1$, there exist $\mathcal{ C}^{
4349: 2,\alpha}$-smoooth functions $\varphi_2,\dots,\varphi_n, h_1,\dots,
4350: h_n, g, k_1,\dots,k_n,\psi$ all defined in real cubes of edge
4351: $2\rho_1$ and of the appropriate dimension, such that if we denote
4352: $z'':=(z_1,\dots,z_{n-1})=x''+iy''$ and $z'=(z_2,\dots,z_n)=
4353: x'+iy'$, then $M$, $M^1$, $(M^1)^+$,
4354: $H^1$, $(H^1)^+$ and $N^1$ are represented in the polydisc of radius
4355: $\rho_1$ centered at $p_1$ by the following graphed (in)equations and
4356: the sub-half-wedge $\mathcal{ HW}_1^+\subset \mathcal{ HW}_{p_1}^+$ is
4357: represented in the polydisc of radius $\delta_1$ centered at $p_1$ by
4358: the following inequations
4359: \def\theequation{5.42}\begin{equation}
4360: \left\{
4361: \aligned
4362: M: \ \ \ \ \ 
4363: & \
4364: y_2=\varphi_2(x,y_1), 
4365: \dots\dots, \
4366: y_n=\varphi_n(x,y_1), \\
4367: M^1: \ \ \ \ \ 
4368: & \ 
4369: y_1=h_1(x),y_2=h_2(x),\dots\dots, \ 
4370: y_n=h_n(x), \\
4371: (M^1)^+: \ \ \ \ \ 
4372: & \ 
4373: y_1>h_1(x),y_2= \varphi_2
4374: (x,y_1), \dots\dots, \ 
4375: y_n=\varphi_n(x,y_1), \\
4376: H^1: \ \ \ \ \ 
4377: & \
4378: x_n=g(x''), \
4379: y_1=k_1(x''), \dots\dots, \ 
4380: y_n=k_n(x''), \\
4381: (H^1)^+: \ \ \ \ \ 
4382: & \
4383: x_n>g(x''), \
4384: y_1=h_1(x),y_2=h_2(x), \dots\dots, \ 
4385: y_n=h_n(x), \\
4386: N^1 : \ \ \ \ \ 
4387: & \
4388: y_2+\cdots+y_{n-1}-y_n=
4389: \psi(x,y'), \\
4390: \mathcal{ HW}_1^+: \ \ \ \ \
4391: & \
4392: y_2+\cdots+y_{n-1}-y_n>
4393: \psi(x,y'), \\
4394: & \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ 
4395: y_1>\varphi_1(x,y_1),\dots, 
4396: y_{n-1}>\varphi_{n-1}(x,y_1), 
4397: \endaligned\right.
4398: \end{equation}
4399: where we can assume that $M^1$ coincides with the intersection $M\cap
4400: \{y_1=h_1(x)\}$, that $H^1$ coincides with the intersection
4401: $M^1\cap \{x_1=g(x')\}$ and that $N^1$ contains $M^1$, which yields at
4402: the level of defining equations the following three collections of
4403: identities
4404: \def\theequation{5.43}\begin{equation}
4405: \left\{
4406: \aligned
4407: h_2(x)\equiv 
4408: & \
4409: \varphi_2(x,h_1(x)),\dots\dots, h_n(x)\equiv 
4410: \varphi_n(x,h_1(x)), \\
4411: k_1(x'')\equiv
4412: & \
4413: h_1(x'',g(x'')), \dots\dots,
4414: k_n(x'')\equiv h_n(x'',g(x'')), \\
4415: \psi(x,h'(x)) \equiv
4416: & \
4417: h_1(x)+h_2(x)+\cdots+h_{n-1}(x)-
4418: h_n(x),
4419: \endaligned\right.
4420: \end{equation}
4421: and where the following normalizations hold{\rm :}
4422: \def\theequation{5.44}\begin{equation}
4423: \left\{
4424: \aligned
4425: {}
4426: &
4427: \varphi_{j}(0)=\partial_{x_k}\varphi_j(0)
4428: =\partial_{y_1}\varphi_j(0)=0, \ \ \ \ \ j=2,\dots,n, \
4429: k=2,\dots,n, \\
4430: &
4431: h_j(0)=\partial_{x_k}
4432: h_j(0)=\partial_{x_{k_1}}
4433: \partial_{x_{k_2}}h_j(0)=0, \ \ \ \ \ \ \ \ \ \
4434: j,k,k_1,k_2=1,\dots,n, \\
4435: &
4436: g(0)=\partial_{x_k}
4437: g(0)=k_j(0)=\partial_{x_k}k_j(0)=0, \ \ \ \ \ 
4438: j=1,\dots,n, \ k=1,\dots,n-1, \\
4439: & \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \
4440: \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \
4441: \partial_{x_{k_1}}\partial_{x_{k_2}}
4442: g(0)=-\delta_{k_1}^{k_2}, \ \ \ \ \ \ \ 
4443: k_1,k_2=1,\dots,n-1, \\
4444: &
4445: \psi(0)=\partial_{x_k}\psi(0)=\partial_{y_j}\psi(0)=0, \ \ \ \ \
4446: k=1,\dots,n, \ j=2,\dots,n.
4447: \endaligned\right.
4448: \end{equation}
4449: In other words, $T_0M=\{y'=0\}$ {\rm (}hence $T_0^cM$ coincides with
4450: the complex $z_1$-axis{\rm )}, $T_0N^1=\{
4451: y_1+y_2+\cdots+y_{n-1}-y_n=0\}$ and the second order Taylor
4452: approximations of the defining equations of $M^1$, of $H^1$ and of
4453: $(H^1)^+$ are the quadrics
4454: \def\theequation{5.45}\begin{equation}
4455: \left\{
4456: \aligned
4457: T_{p_1}^{(2)}M^1: \ \ \ \ \ 
4458: & \
4459: y_1=0, \dots\dots, \ y_n=0,
4460: \\
4461: T_{p_1}^{(2)}H^1: \ \ \ \ \ 
4462: & \
4463: x_n=-x_1^2-\cdots-x_{n-1}^2, \ y_1=0,\dots\dots, \ y_n=0, \\
4464: T_{p_1}^{(2)}(H^1)^+: \ \ \ \ \ 
4465: & \
4466: x_n>-x_1^2-\cdots-x_{n-1}^2, \ y_1=0,\dots\dots, \ y_n=0.
4467: \endaligned\right.
4468: \end{equation}
4469: \item[{\bf (II)}]
4470: Finally, if $\dim_\R\left( T_{p_1}H^1\cap T_{p_1}^c M\right) = 1$ and
4471: if $v_1$ is complex tangent to $M$ {\rm (}this possibility can occur
4472: in all cases $n\geq 2${\rm )}, then there exists a system of holomorphic
4473: coordinates $z=(z_1,\dots,z_n)= (x_1+iy_1,\dots,x_n+iy_n)$ vanishing
4474: at $p_1$ with $v_1$ equal to $(1,0,\dots,0)$, there exist positive
4475: numbers $\rho_1$ and $\delta_1$ with $0< \delta_1 < \rho_1$, there
4476: exist $\mathcal{ C}^{ 2,\alpha}$-smoooth functions
4477: $\varphi_2,\dots,\varphi_n, h_1,\dots, h_n, g, k_1,\dots,k_n$ all
4478: defined in real cubes of edge $2\rho_1$ and of the appropriate
4479: dimension, such that if we denote $z'':=(z_1,\dots,z_{n-1})=x''+iy''$
4480: and $z'=(z_2,\dots,z_n)= x'+iy'$, then $M$, $M^1$, $(M^1)^+$, $H^1$
4481: and $(H^1)^+$ are represented in the polydisc of radius $\rho_1$
4482: centered at $p_1$ by the first five (in)equations of~\thetag{5.42}
4483: together with the normalizations~\thetag{5.45} and such that the local
4484: wedge $\mathcal{ W}_2 \subset \mathcal{ HW}_{p_1}^+$ of edge $M^1$
4485: at $p_1$ is represented in the polydisc of radius $\delta_1$ centered
4486: at $p_1$ by the following inequations
4487: \def\theequation{5.46}\begin{equation}
4488: \left\{
4489: \aligned
4490: \mathcal{ W}_2: \ \ \ \ \
4491: & \
4492: y_1-h_1(x)>-\left[y_2-h_2(x)\right], \dots\dots, \
4493: y_1-h_1(x)>-\left[y_n-h_n(x)\right], \\
4494: & \ \ \ 
4495: y_1-h_1(x)>y_2-h_2(x)+\cdots+
4496: y_n-h_n(x).
4497: \endaligned\right.
4498: \end{equation}
4499: \end{itemize}
4500: \end{lemma}
4501: 
4502: \subsection*{5.47.~Summarizing figure and proof of Lemma~5.37}
4503: As an illustration for this technical lemma, by specifying the value
4504: $n=3$, we have drawn in the following figure the cones ${\sf C}_1$ and
4505: ${\sf C}_2$ together with the vector $v_1$, the tangent plane
4506: $T_{p_1}H^1$ and the hyperplane $\Sigma^1$ in the three cases {\bf
4507: ($\mathbf{I_1}$)}, {\bf ($\mathbf{I_2}$)} and {\bf (II)}. In the left
4508: part of this figure, the cone ${\sf C}_1$ is given by $x_2>0$,
4509: $x_3>0$, $x_1>-\frac{1}{2}x_2-\frac{1}{2}x_3$, namely we have chosen
4510: the values $a_2=a_3=\frac{1}{2}$ for the drawing; in the central part,
4511: the cone ${\sf C}_1$ is given by $x_1>0$, $x_2>0$, $x_2>x_3$; in the
4512: right part, the cone ${\sf C}_2$ is given by $x_1>-x_2$, $x_1>-x_3$,
4513: $x_1>x_2+x_3$.
4514: 
4515: \bigskip
4516: \begin{center}
4517: \input three-cones.pstex_t
4518: \end{center}
4519: 
4520: \proof
4521: Case {\bf ($\mathbf{I_1}$)} has been completed before
4522: the statement of Lemma~5.37. 
4523: 
4524: For Case {\bf ($\mathbf{I_2}$)}, we reason similarly, as follows. We
4525: start with the normalizations $T_{p_1} M=\{y'=0\}$ and $T_{p_1} M^1=
4526: \{y=0\}$ as in the end of \S5.21. By assumption, $T_{p_1}H^1$ contains
4527: the characteristic direction, which coincides with the $x_1$-axis. By
4528: means of an elementary real linear transformation of the form
4529: $\widehat{ z}_1:= z_1$, $\widehat{ z}'=A'\cdot z'$, we may first
4530: normalize $T_{p_1} H^1$ to be the hyperplane (after dropping the hats
4531: on coordinates) $\{x_n=0, \ y=0\}$. 
4532: Similarly, we may normalize $v_1$ to be
4533: the vector $(1,1,\dots,1,0)$. Let again $\pi': (x_1,x') \mapsto x'$
4534: denote the canonical projection on the $x'$-space. Then
4535: $\pi'(v_1)=(1,\dots,1,0)$. Using again a real linear transformation of
4536: the form $\widehat{ z}_1:= z_1$, $\widehat{ z}'=A'\cdot z'$, we can
4537: assume that the proper subcone ${\sf C}_1'\subset \subset {\sf
4538: C}_{p_1}'\equiv \pi'({\sf C}_{p_1})$ which contains the vector $v_1$ is
4539: given (after dropping the hats on coordinates) by
4540: \def\theequation{5.48}\begin{equation}
4541: {\sf C}_1': \ \ \ \ \ 
4542: x_2>0,\dots,x_{n-1}>0, \ 
4543: x_2+\dots+x_{n-1}>x_n.
4544: \end{equation}
4545: Following \S5.14 ({\it cf.} {\sc Figure}~12), we choose a linear cone
4546: ${\sf C}_1\subset\subset {\sf FC}_{p_1}$ defined by the $(n-1)$
4547: inequations of ${\sf C}_1'$ plus one inequation of the form
4548: $x_1>a_2x_2+\cdots+a_nx_n$ with $1>a_2+\cdots+a_{n-1}$, since $v_1$
4549: belongs to ${\sf C}_1$. Then by means of a real linear transformation
4550: of the form $\widehat{ z}_1:=z_1 + a_2z_2+\cdots +a_nz_n$, $\widehat{
4551: z}':= z'$, which stabilizes $\pi'(v_1)$ and the
4552: inequations~\thetag{5.48} of ${\sf C}_1'$, we can assume that the
4553: supplementary inequation for ${\sf C}_1$, namely the inequation for
4554: $(\Sigma^1)^+$, is simply (after dropping the hats on coordinates)
4555: $x_1>0$. Then the vector $v_1$ is mapped to the vector of coordinates
4556: $(1-a_2-\cdots-a_n,1,\dots,1,0)$, which we map to the vector of
4557: coordinates $(1,1,\dots,1,0)$ by an obvious positive scaling of the
4558: $x_1$-axis. In conclusion, in the final system of coordinates,
4559: the cone ${\sf C}_1$ is given by
4560: \def\theequation{5.49}\begin{equation}
4561: {\sf C}_1: \ \ \ \ \ 
4562: x_1>0,x_2>0,\dots,x_{n-1}>0, \ 
4563: x_2+\cdots+x_{n-1}-x_n>0.
4564: \end{equation}
4565: This implies that the half-wedge $\mathcal{ HW}_1^+\subset \mathcal{
4566: HW}_{p_1}^+$ may be represented by the inequations of the last two line
4567: of~\thetag{5.42}. To conclude the proof of Case {\bf
4568: ($\mathbf{I_2}$)} of Lemma~5.37, it suffices to observe that, as in
4569: Case {\bf ($\mathbf{I_1}$)}, the further second order normalizations
4570: do not perturb the previously achieved first order normalizations,
4571: because the transformations are tangent to the identity mapping at the
4572: origin.
4573: 
4574: Finally, we treat Case {\bf (II)} of Lemma~5.37, starting with the
4575: system of coordinates $(z_1,\dots,z_n)$ of the end of \S5.21. After
4576: an elementary real linear transformation stabilizing the characteristic
4577: $x_1$-axis, we can assume that $v_1=(1,0,\dots,0)$ and that the
4578: convex infinite linear cone ${\sf C}_2$ introduced in \S5.19 which
4579: contains $v_1$ is given by the inequations
4580: \def\theequation{5.50}\begin{equation}
4581: x_1>-x_2,\dots\dots,\ x_1>-x_n,\ 
4582: x_1>x_2+\cdots+x_n.
4583: \end{equation}
4584: This implies that the local wedge $\mathcal{ W}_2\subset \mathcal{
4585: HW}_{p_1}^+$ of edge $M^1$ at $p_1$ introduced in \S5.19 may be
4586: represented by the inequations~\thetag{5.46}. Finally, the second
4587: order normalizations, which are tangent to the identity mapping, are
4588: achieved as in the two previous cases {\bf ($\mathbf{I_1}$)} and {\bf
4589: ($\mathbf{I_2}$)}.
4590: 
4591: The proof of Lemma~5.37 is complete.
4592: \endproof
4593: 
4594: \section*{\S6.~Three preparatory lemmas on H\"older spaces}
4595: 
4596: In this section, we first collect a few very elementary lemmas that
4597: will be useful in our geometric construction of half-attached analytic
4598: discs which will be achieved in Section~7 below. From now on, we
4599: shall admit the convenient index notation $g_{x_k}$ for the partial
4600: derivative which was denoted up to now by $\partial_{x_k}g$.
4601: 
4602: \subsection*{6.1.~Local growth of $\mathcal{ C}^{2,\alpha}$-smooth 
4603: mappings} Let $n\in \N$, $n \geq 1$ and let $x= (x_1, \dots, x_n) \in
4604: \R^n$. We shall use the norm $\vert x \vert := \max_{1 \leq k\leq n}\,
4605: \vert x_k \vert$. If $g= g(x)$ is an $\R^n$-valued $\mathcal{
4606: C}^1$-smooth mapping on the real cube $\{x\in \R^n: \ \vert x \vert <
4607: 2\rho_1\}$, for some $\rho_1>0$, and if $\vert x'\vert, \vert x''
4608: \vert \leq \rho$, for some $\rho<2\rho_1$, we have the trivial
4609: estimate
4610: \def\theequation{6.2}\begin{equation}
4611: \left\vert g(x')- g(x'') \right\vert \leq \left\vert 
4612: x' -x'' \right\vert \cdot \left(
4613: \sum_{k=1}^n \, \sup_{\vert x 
4614: \vert \leq \rho} \, \vert g_{x_k}(x)
4615: \vert \right),
4616: \end{equation} 
4617: where we denote by $g_{j,x_k}$ the partial derivative $\partial
4618: g_j/\partial x_k$. Notice that by the definition of the norm $\vert
4619: \cdot \vert$, we have in~\thetag{6.2} that $\vert g(x) \vert \equiv
4620: \max_{1\leq k\leq n} \, \vert g_j (x) \vert$ and that $\vert g_{x_k}
4621: (x)\vert \equiv \max_{1 \leq k \leq n}\, \vert g_{j,x_k}(x) \vert$.
4622: 
4623: Let $\alpha$ with $0 < \alpha < 1$ and let $h= h(x) = (h_1(x), \dots,
4624: h_n(x ))$ be an $\R^n$-valued mapping which is of class $\mathcal{
4625: C}^{2, \alpha}$ on the real cube $\{x\in \R^n: \ \vert x \vert <
4626: 2\rho_1\}$, for some $\rho_1 >0$. For every $\rho< 2 \rho_1$, we
4627: consider the $\mathcal{ C}^{2, \alpha}$ norm of $h$ over $\{ \vert x
4628: \vert \leq \rho\}$ which is defined precisely as:
4629: \def\theequation{6.3}\begin{equation}
4630: \left\{
4631: \aligned
4632: \vert \vert h \vert \vert_{\mathcal{ C}^{2,\alpha}(
4633: \{\vert x \vert \leq \rho\})}:= 
4634: & \
4635: \sup_{\vert x \vert \leq \rho}\, 
4636: \vert h(x)\vert+
4637: \sum_{k=1}^n\, 
4638: \sup_{\vert x \vert \leq \rho}\, 
4639: \left\vert
4640: h_{x_k}(x)\right\vert+
4641: \sum_{k_1,k_2=1}^n\, 
4642: \left\vert h_{x_{k_1}x_{k_2}}(x)\right\vert+\\
4643: & \
4644: +\sum_{k_1,k_2=1}^n\, 
4645: \sup_{\left\vert x'\right\vert, \,
4646: \left\vert x''\right\vert \leq \rho, \ 
4647: x'\neq x''}\ 
4648: \frac{
4649: \left\vert h_{x_{k_1}x_{k_2}}(x')-
4650: h_{x_{k_1}x_{k_2}}(x'')
4651: \right\vert
4652: }
4653: {\left\vert x' - x'' \right\vert^\alpha}<\infty,
4654: \endaligned\right.
4655: \end{equation}
4656: and which is finite. With these definitions at hand, the
4657: following lemma can easily be established by means of~\thetag{6.2}.
4658: 
4659: \def\thelemma{{\bf 6.4}}\begin{lemma}
4660: Under the above assumptions, let 
4661: \def\theequation{6.5}\begin{equation}
4662: K_1:= \vert \vert h \vert \vert_{
4663: \mathcal{ C}^{2, \alpha}(\{ \vert x \vert \leq \rho_1\})}<\infty
4664: \end{equation}
4665: be the $\mathcal{ C}^{2, \alpha}$ norm of $h$ over the cube $\{\vert x
4666: \vert \leq \rho_1\}$ and assume that $h_j (0)=0$, $h_{j, x_k} (0)=0$
4667: and $h_{j, x_{k_1}x_{k_2}}(0)=0$, for all $j,k, k_1,
4668: k_2=1,\dots,n$. Then the following three inequalities hold for $\vert
4669: x \vert \leq \rho_1${\rm :}
4670: \def\theequation{6.6}\begin{equation}
4671: \left\{
4672: \aligned
4673: {\bf [1]:}
4674: & \ \ \ \ \ 
4675: \vert h(x)\vert 
4676: \leq 
4677: \vert x\vert^{2+\alpha} \cdot K_1,\\
4678: {\bf [2]:} 
4679: & \ \ \ \ \ 
4680: \sum_{k=1}^n\, \left\vert
4681: h_{x_k}(x) \right\vert 
4682: \leq 
4683: \vert x \vert^{1+\alpha} \cdot K_1, \\
4684: {\bf [3]:} 
4685: & \ \ \ \ \ \
4686: \sum_{k_1,k_2=1}^n \, 
4687: \left\vert
4688: h_{x_{k_1}x_{k_2}}(x)
4689: \right\vert 
4690: \leq 
4691: \vert x \vert^\alpha \cdot K_1.
4692: \endaligned\right.
4693: \end{equation} 
4694: \end{lemma}
4695: 
4696: \subsection*{6.7.~A $\mathcal{ C}^{1,\alpha}$ estimate for
4697: composition of mappings} Recall that $\Delta$ is the open unit disc in
4698: $\C$ and that $\partial \Delta$ is its boundary, namely the unit
4699: circle. We shall constantly denote the complex variable in $\overline{
4700: \Delta}:= \Delta \cup \partial \Delta$ by $\zeta=\rho \, e^{i
4701: \theta}$, where $0 \leq \rho \leq 1$ and where $\vert \theta \vert
4702: \leq \pi$, except when we consider two points $\zeta'=e^{i\theta'}$,
4703: $\zeta''=e^{i\theta''}$, in which case we may obviously choose $\vert
4704: \theta ' \vert, \ \vert \theta'' \vert \leq 2\pi$ with $0\leq \vert
4705: \theta ' -\theta'' \vert \leq \pi$. Let now $X(\zeta)=\left(X_1
4706: (\zeta), \dots, X_n (\zeta)\right)$ be an $\R^n$-valued mapping which
4707: is of class $\mathcal{ C}^{1,\alpha}$ on the unit circle $\partial
4708: \Delta$. We define its $\mathcal{ C}^{1, \alpha}$-norm precisely by
4709: \def\theequation{6.8}\begin{equation}
4710: \left\{
4711: \aligned
4712: \vert \vert X \vert \vert_{\mathcal{ C}^{1,\alpha}(\partial 
4713: \Delta)}:= 
4714: & \
4715: \sup_{\vert \theta \vert \leq \pi}\, 
4716: \left\vert X\left(e^{i\theta} \right) \right\vert+ 
4717: \sup_{\vert \theta \vert \leq \pi}\, 
4718: \left\vert
4719: \frac{d X\left(e^{i\theta}\right)}{d \theta}
4720: \right\vert+ \\
4721: & \ \ \ \ \
4722: +\sup_{0 < \left\vert \theta' - \theta''\right\vert \leq \pi} \ 
4723: \frac{
4724: \left\vert\frac{dX(e^{i\theta'})}{d\theta}-
4725: \frac{dX(e^{i\theta''})}{d\theta}
4726: \right\vert
4727: }
4728: {\left\vert \theta' - \theta''\right\vert^\alpha},
4729: \endaligned\right.
4730: \end{equation}
4731: and we define its $\mathcal{ C}^1$-norm $\vert \vert X \vert
4732: \vert_{\mathcal{ C}^1(\partial \Delta)}$ by keeping only the first two
4733: terms. Let $h$ be as in Lemma~6.5.
4734: 
4735: \def\thelemma{{\bf 6.9}}\begin{lemma} Under the above assumptions, if
4736: moreover $\left\vert X \left(e^{i\theta}\right) \right\vert \leq\rho$
4737: for all $\theta$ with $\vert \theta \vert \leq \pi$, where $\rho \leq
4738: \rho_1$, then we have the following three estimates{\rm :}
4739: \def\theequation{6.10}\begin{equation}
4740: \small
4741: \left\{
4742: \aligned
4743: \vert \vert h(X) \vert \vert_{\mathcal{ 
4744: C}^{1,\alpha}(\partial \Delta)}
4745: \leq 
4746: & \
4747: \sup_{\vert x \vert \leq \rho} \, 
4748: \vert h(x) \vert +
4749: \left(
4750: \sum_{k=1}^n\, 
4751: \sup_{\vert x \vert \leq \rho}\, 
4752: \left\vert h_{x_k}(x) \right\vert
4753: \right)\cdot \vert \vert 
4754: X \vert \vert_{\mathcal{ C}^1(\partial \Delta)}+\\
4755: & \
4756: +\left(
4757: \sum_{k_1,k_2=1}^n\, \sup_{\vert x\vert\leq \rho}
4758: \left\vert h_{x_{k_1}x_{k_2}} (x) \right\vert
4759: \right)\cdot \pi^{1-\alpha} \cdot 
4760: \left[
4761: \vert \vert X \vert \vert_{\mathcal{ C}^1(\partial \Delta)}
4762: \right]^2+ \\
4763: & \
4764: +
4765: \left(
4766: \sum_{k=1}^n\, 
4767: \sup_{\vert x \vert\leq \rho} \, 
4768: \vert h_{x_k} (x) \vert
4769: \right)\cdot \vert \vert X 
4770: \vert \vert_{\mathcal{ C}^{1,\alpha}
4771: (\partial \Delta)}. \\
4772: \sum_{k=1}^n \, 
4773: \left\vert \left\vert h_{x_k}(X) \right\vert
4774: \right\vert_{\mathcal{ C}^\alpha
4775: (\partial \Delta)}\leq 
4776: & \
4777: \sum_{k=1}\, \sup_{\vert x \vert \leq \rho} \, 
4778: \left\vert h_{x_k} (x) \right\vert + \\
4779: & \ \
4780: +
4781: \left(
4782: \sum_{k_1,k_2=1}^n\, 
4783: \sup_{\vert x \vert \leq \rho} \, 
4784: \left\vert h_{x_{k_1}x_{k_2}}(x) \right\vert
4785: \right)\cdot \pi^{1-\alpha} \cdot
4786: \vert \vert X \vert \vert_{\mathcal{ C}^1(\partial 
4787: \Delta)},
4788: \\
4789: \sum_{k_1,k_2=1}^n \, 
4790: \left\vert 
4791: \left\vert
4792: h_{x_{k_1}x_{k_2}}(X)
4793: \right\vert
4794: \right\vert_{\mathcal{ C}^\alpha(\partial \Delta)} \leq
4795: & \
4796: \sum_{k_1,k_2=1}^n\, 
4797: \sup_{\vert x \vert \leq \rho} \,
4798: \left\vert
4799: h_{x_{k_1}x_{k_2}}(x)
4800: \right\vert + \\
4801: & \ \
4802: +\vert\vert
4803: h \vert \vert_{\mathcal{ C}^{2,\alpha}(
4804: \{\vert x \vert \leq \rho\})}\cdot
4805: \left(
4806: \vert \vert X \vert \vert_{\mathcal{C}^1(\partial \Delta)}
4807: \right)^\alpha.
4808: \endaligned\right.
4809: \end{equation}
4810: \end{lemma}
4811: 
4812: \proof
4813: We summarize the computations. Applying the definition~\thetag{6.8},
4814: using the chain rule for the calculation of
4815: $dh\left(X\left(e^{i\theta}\right)\right)/ d\theta$, and using the
4816: trivial inequality $\left\vert a'b'-a''b''\right\vert \leq \left\vert
4817: a' \right\vert \cdot \left\vert b'-b''\right\vert+ \left\vert b''
4818: \right\vert\cdot \left\vert a' -a'' \right\vert$, we may majorize
4819: \def\theequation{6.11}\begin{equation}
4820: \small
4821: \left\{
4822: \aligned
4823: \vert \vert h(X) \vert \vert_{\mathcal{ 
4824: C}^{1,\alpha}(\partial \Delta)}
4825: \leq 
4826: & \
4827: \sup_{\vert \theta \vert \leq \pi}\, 
4828: \left\vert h(X(e^{i\theta})) \right\vert + 
4829: \left(\sum_{k=1}^n\, 
4830: \sup_{\vert \theta \vert\leq \pi}\, 
4831: \left\vert h_{x_k}(X(e^{i\theta})) \right\vert
4832: \right)\cdot \max_{1\leq k\leq n}\, 
4833: \sup_{\vert \theta \vert \leq \pi}\, 
4834: \left\vert
4835: \frac{dX_k(e^{i\theta})}{d\theta}
4836: \right\vert+ \\
4837: & \
4838: \sup_{0 < \vert \theta' -\theta'' 
4839: \vert\leq \pi} \, \sum_{k=1}^n\, 
4840: \frac{
4841: \left\vert
4842: h_{x_k}(X(e^{i\theta'}))-
4843: h_{x_k}(X(e^{i\theta''}))
4844: \right\vert
4845: }{\vert \theta' -\theta ''\vert^\alpha} \cdot
4846: \max_{1\leq k\leq n}\, 
4847: \sup_{\vert \theta ' \vert \leq \pi}\, 
4848: \left\vert
4849: \frac{dX_k(e^{i\theta'})}{d\theta}
4850: \right\vert + \\
4851: & \
4852: \left(
4853: \sum_{k=1}^n\, \sup_{\vert \theta''\vert\leq \pi}\, 
4854: \left\vert h_{x_k}(e^{i\theta''}) \right\vert
4855: \right) \cdot \left(
4856: \max_{1\leq k\leq n}\, 
4857: \sup_{0< \vert \theta' -\theta''\vert\leq \pi}\,
4858: \frac{
4859: \left\vert
4860: \frac{dX_k(e^{i\theta'})}{d\theta}-
4861: \frac{dX_k(e^{i\theta''})}{d\theta}
4862: \right\vert
4863: }
4864: {\vert \theta' -\theta '' \vert^\alpha}\right),
4865: \endaligned\right.
4866: \end{equation}
4867: which yields the first inequality of~\thetag{ 6.10} after
4868: using~\thetag{ 6.2} for the second line of~\thetag{ 6.11} and the
4869: trivial majoration $\left\vert \theta' - \theta''
4870: \right\vert^{1-\alpha}\leq \pi^{1-\alpha}$. The second and the third
4871: inequalities of~\thetag{6.10} are established similarly, which
4872: completes the proof.
4873: \endproof
4874: 
4875: The following direct consequence will be strongly used in 
4876: Section~7 below.
4877: 
4878: \def\thelemma{{\bf 6.12}}\begin{lemma} Under the above assumptions,
4879: suppose that there exist constants $c_1>0$, $K_2>0$ with $c_1 K_2 \leq
4880: \rho_1$ such that for each $c\in\R$ with $0\leq c \leq c_1$, there
4881: exists $X_c\in\mathcal{ C}^{1,\alpha}(\partial \Delta,\R^n)$ with
4882: $\vert \vert X_c \vert \vert_{\mathcal{ C}^{1,\alpha}(\partial
4883: \Delta)}\leq c \cdot K_2$. Then there exists a constant $K_3>0$ such
4884: that the following three estimates hold{\rm :}
4885: \def\theequation{6.13}\begin{equation}
4886: \left\{
4887: \aligned
4888: \vert \vert h(X_c)\vert 
4889: \vert_{{\mathcal C}^{1,\alpha}(\partial 
4890: \Delta)} \leq 
4891: & \
4892: c^{2+\alpha} \cdot K_3, \\
4893: \sum_{k=1}^n\, 
4894: \vert \vert h_{x_k}(X_c) 
4895: \vert \vert_{\mathcal{ C}^\alpha(
4896: \partial \Delta)} \leq 
4897: & \
4898: c^{1+\alpha} \cdot K_3, \\
4899: \sum_{k_1,k_2=1}^n\, 
4900: \left\vert \left\vert
4901: h_{x_{k_1}x_{k_2}} (X_c)
4902: \right\vert \right\vert_{\mathcal{ 
4903: C}^\alpha(\partial \Delta)} \leq
4904: & \
4905: c^\alpha \cdot K_3.
4906: \endaligned\right.
4907: \end{equation}
4908: \end{lemma}
4909: 
4910: \proof
4911: Applying Lemmas~6.4 and~6.9, we see that it suffices to choose 
4912: \def\theequation{6.14}\begin{equation}
4913: K_3:=\max \left(
4914: K_1K_2^{2+\alpha}(3+\pi^{1-\alpha}), \ 
4915: K_1K_2^{1+\alpha}(1+\pi^{1-\alpha}), \ 
4916: 2K_1K_2^\alpha
4917: \right),
4918: \end{equation}
4919: which completes the proof.
4920: \endproof
4921: 
4922: Up to now, we have introduced three positive constants $K_1$, $K_2$,
4923: $K_3$. In Sections~7, 8 and~9 below, we shall introduce further
4924: positive constants $K_4$, $K_5$, $K_6$, $K_7$, $K_8$, $K_9$, $K_{10}$,
4925: $K_{11}$, $K_{12}$, $K_{13}$, $K_{14}$, $K_{15}$, $K_{16}$, $K_{17}$,
4926: $K_{18}$ and $K_{19}$, whose precise value will not be important.
4927: 
4928: \section*{\S7.~Families of analytic discs half-attached 
4929: to maximally real submanifolds}
4930: 
4931: \subsection*{7.1.~Preliminary}
4932: Let $E\subset \C^n$ be an arbitrary subset and let $A:
4933: \overline{\Delta}\to \C^n$ be a continuous mapping, holomorphic in
4934: $\Delta$. If $\partial^+\Delta:= \{\zeta\in\partial \Delta: {\rm Re}\,
4935: \zeta \geq 0\}$ denotes the {\sl positive half-boundary} of $\Delta$,
4936: we say that $A$ is {\sl half-attached} to $E$ if $A(\partial^+
4937: \Delta)\subset E$. Such analytic discs which are glued in part to a
4938: geometric object were studied by S.~Pinchuk in~\cite{p} to establish a
4939: boundary uniqueness principle about continuous functions on a
4940: maximally real submanifold of $\C^n$ which extend holomorphically to a
4941: wedge. Further works on the CR edge of the wedge theorem using discs
4942: partly attached to generic submanifolds were achieved by
4943: R.~Ayrapetian~\cite{a} and by A.~Tumanov~\cite{tu2}.
4944: 
4945: In this section, we shall construct local families of analytic discs
4946: $Z_{c,x,v}^1(\zeta): \overline{\Delta} \to \C^n$, where $c\in \R^+$ is
4947: small, where $x\in\R^n$ is small and where $v\in\R^n$ is small, which
4948: are half-attached to a $\mathcal{ C}^{2,\alpha}$-smooth maximally real
4949: submanifold $M^1$ of $\C^n$, which satisfy $Z_{c,0,v}^1(1)\equiv
4950: p_1\in M^1$, such that the boundary point $Z_{c,x,v}^1(1)$ covers a
4951: neighborhood of $p_1$ in $M^1$ as $x$ varies ($c$ and $v$ being fixed)
4952: and such that the tangent vector $\frac{\partial Z_{c,0,v}^1}{\partial
4953: \theta}(1)$ at the fixed point $p_1$ covers a cone in
4954: $T_{p_1}M^1$. These families will be used in Sections~8 and~9 below
4955: for the final steps in the proof of the main Proposition~5.12. With
4956: this choice, when $x$ varies, $v$ varies and $\zeta$ varies (but $c$
4957: is fixed), the set of points $Z_{c,x,v}^1(\zeta)$, covers a thin wedge
4958: of edge $M^1$ at $p_1$. Similar families of analytic discs were
4959: constructed in~\cite{ber} to reprove S.~Pinchuk's boundary uniqueness
4960: theorem, with $M^1$ of class $\mathcal{ C}^\infty$, using a method
4961: (implicit function theorem in Banach spaces) which in the case where
4962: $M$ is of class $\mathcal{ C}^{\kappa,\alpha}$ necessarily induces a
4963: loss of smoothness, yielding families of analytic discs which are only
4964: of class $\mathcal{ C}^{\kappa-1,\alpha}$. Since we want our families
4965: to be of class at least $\mathcal{ C}^2$ and since $M$ is only of
4966: class $\mathcal{ C}^{2,\alpha}$, we shall have to proceed differently.
4967: 
4968: To summarize symbolically the structure of the desired family:
4969: \def\theequation{7.2}\begin{equation}
4970: Z_{c,x,v}^1(\zeta): \left\{
4971: \aligned
4972: c= 
4973: & \
4974: \text{\rm small scaling factor}, \\
4975: x=
4976: & \
4977: \text{\rm translation parameter}, \\
4978: v=
4979: & \
4980: \text{\rm rotation parameter}, \\
4981: \zeta=
4982: & \
4983: \text{\rm unit disc variable}. \\
4984: \endaligned
4985: \right.
4986: \end{equation} 
4987: We shall begin our constructions in the ``flat'' case where the
4988: maximally real submanifold $M^1$ coincides with $\R^n$ and then
4989: perform a pertubation argument, using the scaling parameter $c$ in an
4990: essential way.
4991: 
4992: \subsection*{7.3.~A family of analytic
4993: discs sweeping $\R^n \subset \C^n$ with prescribed first order jets}
4994: We denote the coordinates over $\C^n$ by $z= x+iy= (x_1 + iy_1, \dots,
4995: x_n+ iy_n)$. Let $c\in \R$ with $c \geq 0$ be a ``scaling factor'',
4996: let $n \geq 2$, let $x= (x_1, \dots, x_n) \in \R^n$, let $v= (v_1,
4997: \dots, v_n) \in \R^n$ and consider the algebraically parametrized
4998: family of analytic discs defined by
4999: \def\theequation{7.4}\begin{equation} 
5000: B_{c,x,v}(s+it):=
5001: \left(x_1+cv_1(s+it),\dots, 
5002: x_n+cv_n(s+it)\right),
5003: \end{equation}
5004: where $s+it \in \C$ is the holomorphic variable. For $c \neq 0$, the
5005: map $B_{c, x, v}$ embeds the complex line $\C$ into $\C^n$ and sends
5006: $\R$ into $\R^n$ with {\it arbitrary first order jet at $0$}: center
5007: point $B_{c, x, v}(0) =x$ and tangent direction $\partial B_{c, x, v}
5008: (s) /\partial s \vert_{ s=0}= cv$.
5009: 
5010: To localize our family of analytic discs, we restrict the
5011: map~\thetag{7.4} to the following specific set of values: $0\leq c\leq
5012: c_0$ for some $c_0>0$; $\vert x \vert \leq c$~; $\vert v \vert \leq
5013: 2$~; and $\vert s+it \vert \leq 4$. To localize $\R^n$, we shall
5014: denote $M^0:= \{x\in\R^n: \,\vert x \vert \leq \rho_0\}$, where
5015: $\rho_0 >0$, and we notice that $B_{c,x,v}(\{\vert s+it \vert \leq
5016: 4\}) \subset M^0$ for all $c$, all $x$ and all $v$ provided that $c_0
5017: \leq \rho_0/9$.
5018: 
5019: We then consider the mapping $(s+it) \longmapsto B_{c,x, v} (s+it)$ as
5020: a local (nonsmooth) analytic disc defined on the rectangle
5021: $\{s+it\in\C: \, \vert s \vert \leq 4, 0\leq t\leq 4\}$ whose bottom
5022: boundary part $B_{c,x, v}([-4,4])$ is a small real segment contained
5023: in $\R^n$.
5024: 
5025: \subsection*{7.5.~A useful conformal equivalence}
5026: Next, we have to get rid of the corners of the rectangle $\{s+it\in
5027: \C: \, \vert s \vert \leq 4, 0\leq t \leq 4\}$. We proceed as
5028: follows. In the complex plane equipped with coordinates $s+it$, let
5029: $\mathcal{D}(i\sqrt{3},2)$ be the open disc of center $i\sqrt{3}$ and
5030: of radius $2$. Let $\mu: (-2,2)\to [0,1]$ be an {\it even}\,
5031: $\mathcal{ C}^\infty$-smooth function satisfying $\mu(s)=0$ for $0\leq
5032: s\leq 1$; $\mu(s)>0$ and $d\mu(s)/ds>0$ for $1< s < 2$; and $\mu(s)=
5033: \sqrt{3}-\sqrt{4-s^2}$ for $\sqrt{3} \leq s < 2$. The simply connected
5034: domain $C^+\subset \{t>0\}$ which is represented in {\sc Figure~1}
5035: just below may be formally defined as
5036: \def\theequation{7.6}\begin{equation}
5037: \left\{
5038: \aligned
5039: C^+\cap \{t\geq \sqrt{3}-1\}
5040: := \mathcal{D}(i\sqrt{3},2)\cap \{t\geq
5041: \sqrt{3}-1\}, \\
5042: C^+\cap \{0<t< \sqrt{3}-1\}
5043: := \{s+it\in\C: \, t>
5044: \mu(s)\}.
5045: \endaligned\right.
5046: \end{equation}
5047: 
5048: \bigskip
5049: \begin{center}
5050: \input conformal.pstex_t
5051: \end{center}
5052: 
5053: \noindent
5054: Let $\Psi: \Delta\to C^+$ be a conformal equivalence (Riemann's
5055: theorem). Since the boundary $\partial C^+$ is $\mathcal{
5056: C}^\infty$-smooth, the mapping $\Psi$ extends as a $\mathcal{
5057: C}^\infty$-smooth diffeomorphism $\partial \Delta \to \partial C^+$.
5058: Remind that $\partial^+ \Delta:=\{ \zeta \in \C: \vert \zeta \vert =1,
5059: \ {\rm Re}\, \zeta \geq 0\}$ is the positive half-boundary of
5060: $\Delta$. Then after a reparametrization of $\Delta$, we can (and we
5061: shall) assume that $\Psi(\partial^+\Delta)=[-1,1]$, $\Psi (1) =0$ and
5062: $\Psi(\pm i)=\pm 1$. It follows that $d\Psi \left(e^{i \theta}\right)
5063: /d \theta$ is a positive real number for all
5064: $e^{i\theta}\in\partial^+\Delta$. Although the precise shape of $C^+$
5065: and the specific expression of $\Psi$ will not be crucial in the
5066: sequel, it will be convenient to fix them once for all.
5067: 
5068: \subsection*{7.7.~Flat families of half-attached analytic discs}
5069: Thanks to $\Psi$, we can define a family of small analytic
5070: discs which are half-attached to the ``{\sl flat}'' maximally real
5071: manifold $M^0\equiv \{x\in \R^n: \ \vert x \vert \leq \rho_0
5072: \}$ as follows
5073: \def\theequation{7.8}\begin{equation}
5074: Z_{c,x,v}^0(\zeta):= 
5075: B_{c,x,v} \left (\Psi (\zeta) \right)=
5076: \left(
5077: x+cv\Psi(\zeta)
5078: \right).
5079: \end{equation}
5080: We then have $Z_{c, x, v}^0( \partial^+ \Delta) \subset M^0$ and $Z_{
5081: c, x, v }^0 (1) = x$. Notice that every disc $Z_{c,x,v}
5082: \left(\overline{\Delta}\right)$ 
5083: is contained in a single complex line. Starting
5084: with a maximally real submanifold of $\C^n$, as in Proposition~5.12, but
5085: dealing with the ``flat'' maximally real submanifold $M^0 \equiv
5086: \R^n$, we first construct a ``flat model'' of the desired family of
5087: analytic disc.
5088: 
5089: \def\thelemma{{\bf 7.9}}\begin{lemma} 
5090: Let $M^0=\{x\in \R^n: \ \vert x \vert \leq \rho_0\}$ be the ``flat''
5091: local maximally real submanifold defined above, let $p_0 \equiv 0 \in
5092: M^0$ denote the origin and let $v_0\in T_{p_0}M^0$ be a tangent vector
5093: with $\vert v^0\vert =1$. Then there exists a constant $\Lambda_0>0$
5094: and there exists a $\mathcal{ C }^\infty$-smooth family $A_{c,x,v}^0
5095: (\zeta)$ of analytic discs defined for $c\in \R$ with $0 \leq c \leq
5096: c_0$ for some $c_0>0$ satisfying $c_0\leq \rho_0/9$, for $x\in \R^n$
5097: with $\vert x \vert \leq c$ and for $v\in \R^n$ with $\vert v \vert
5098: \leq c$ which enjoy the following six properties{\rm :}
5099: 
5100: \smallskip
5101: \begin{itemize}
5102: \item[{\bf ($\mathbf{1_0}$)}] \
5103: $A_{c,0,v}^0(1)=p_0=0$ for all $c$ and all $v$.
5104: \item[{\bf ($\mathbf{2_0}$)}] \
5105: $A_{c,x,v}^0: \overline{\Delta}\to \C^n$ is
5106: an embedding and
5107: $\left\vert A_{c,x,v}^0(\zeta) \right \vert \leq c\cdot \Lambda_0$ 
5108: for all $c$, all $x$, all $v$ and all $\zeta$.
5109: \item[{\bf ($\mathbf{3_0}$)}] \
5110: $A_{c,x,v}^0(\partial^+\Delta)
5111: \subset M^0$ for all $c$, all $x$ and all $v$.
5112: \item[{\bf ($\mathbf{4_0}$)}] \
5113: $\frac{\partial A_{c,0,0}^0}{\partial 
5114: \theta}(1)$ is a positive multiple of 
5115: $v_0$ for all $c\neq 0$.
5116: \item[{\bf ($\mathbf{5_0}$)}] \
5117: For all $c$, all $v$ and all $e^{i\theta}\in \partial^+\Delta$, the
5118: mapping $x\longmapsto A_{c,x,v}^0\left(e^{i\theta}\right) \in M^0$ is
5119: of rank $n$.
5120: \item[{\bf ($\mathbf{6_0}$)}] \
5121: For all $e^{i \theta} \in \partial^+ \Delta$, all $c \neq 0$ and all
5122: $x$, the mapping $v \longmapsto \frac{ \partial A_{c, x, v}^0}{
5123: \partial \theta }\left(e^{ i\theta}\right)$ is of rank $n$ at $v=0$.
5124: Consequently, the positive half-lines $\R^+ \cdot \frac{ \partial
5125: A_{c,0,v}}{ \partial \theta} (1)$ describe an open infinite cone
5126: containing $v_0$ with
5127: vertex $p_0$ in $T_{p_0} M^0$ when $v$ varies.
5128: \end{itemize}
5129: \end{lemma}
5130: 
5131: \proof
5132: Proceeding as in the proof of Lemma~5.37, we can find a new affine
5133: coordinate system centered at $p_0$ and stabilizing $\R^n$, which we
5134: shall still denote by $(z_1,\dots,z_n)$, 
5135: in which the vector $v_0$
5136: has coordinates $(0,\dots,0,1)$. In this coordinate system, we then
5137: construct the family $Z_{c,x,v}^0(\zeta)$ as in~\thetag{7.8} above and
5138: we define the desired family simply as follows:
5139: \def\theequation{7.10}\begin{equation}
5140: A_{c,x,v}^0(\zeta):= Z_{c,x,v_0+v}^0(\zeta),
5141: \end{equation}
5142: where we restrict the variations of the parameter $v$ to $\vert v
5143: \vert \leq c$. Notice that every disc $A_{c,x,v}^0
5144: \left(\overline{\Delta}\right)$ is contained in a single complex line.
5145: All the properties are then elementary consequences of the explicit
5146: expression~\thetag{7.8} of $Z_{c,x,v}^0(\zeta)$.
5147: 
5148: Finally, we notice that it follows from properties {\bf
5149: ($\mathbf{5_0}$)} and {\bf ($\mathbf{6_0}$)} that the set of points
5150: $A_{c,x,v}^0(\zeta)$, where $c>0$ is fixed, where $x$ varies, where $v$
5151: varies and where $\zeta$ varies covers a local wedge of edge $M^0$ at
5152: $p_0$. The proof of Lemma~7.9 is complete.
5153: \endproof
5154: 
5155: \subsection*{7.11.~Curved families of half-attached analytic discs}
5156: Our main goal in this section is to obtain a statement similar to
5157: Lemma~7.9 after replacing the ``flat'' maximally real submanifold
5158: $M^0\cong \R^n$ by a ``curved'' $\mathcal{ C}^{2, \alpha}$-smooth
5159: maximally real submanifold $M^1$. We set up a formulation which will
5160: be appropriate for the achievement of the end of the proof of
5161: Proposition~5.12 in the next Sections~8 and~9. In particular, we shall
5162: have to shrink the family of half-disc $Z_{c, x,v}^1(\zeta)$ which we
5163: will construct as a perturbation of the family $Z_{c, x,v}^0(\zeta)$
5164: in \S7.50 below, and we shall construct discs of size $\leq c^2 \cdot
5165: \Lambda_1$ for some constant $\Lambda_1>0$, instead of requiring that
5166: their size is $\leq c \cdot \Lambda_1$, which would be the property
5167: analogous to {\bf ($\mathbf{2_0}$)}. Also, we shall loose the
5168: $\mathcal{ C}^{ 2, \alpha-0}$-smoothness with respect to the scaling
5169: parameter $c$.
5170: 
5171: \def\thelemma{7.12}\begin{lemma}
5172: Let $M^1$ be $\mathcal{ C}^{2,\alpha}$-smooth maximally real
5173: submanifold of $\C^n$, let $p_1\in M^1$ and let $v_1\in T_{p_1} M^1$
5174: be a tangent vector with $\vert v_1 \vert =1$. Then there exists
5175: a positive constant $\Lambda_1>0$ and there exists $c_1\in \R$ with
5176: $c_1>0$ such that for every $c\in \R$ with $0 < c \leq c_1$, there
5177: exists a family $A_{x,v:c}^1 (\zeta)$ of analytic discs defined for
5178: $x\in \R^n$ with $\vert x \vert \leq c^2$ and for $v\in \R^n$ with
5179: $\vert v \vert \leq c$ which is $\mathcal{ C }^{2,\alpha-0}$-smooth
5180: with respect to $(x,v,\zeta)$ and which enjoys the following six
5181: properties{\rm :}
5182: 
5183: \smallskip
5184: \begin{itemize}
5185: \item[{\bf ($\mathbf{1_1}$)}] \
5186: $A_{0,v:c}^1(1)=p_1$ for all $v$.
5187: \item[{\bf ($\mathbf{2_1}$)}] \
5188: $A_{x,v:c}^1: \overline{\Delta}\to \C^n$ is
5189: an embedding and
5190: $\left\vert A_{x,v:c}^1 (\zeta) \right \vert \leq c^2\cdot \Lambda_1$ 
5191: for all $x$, all $v$ and all $\zeta$.
5192: \item[{\bf ($\mathbf{3_1}$)}] \
5193: $A_{x,v:c}^1(\partial^+\Delta)
5194: \subset M^1$ for all $x$ and all $v$.
5195: \item[{\bf ($\mathbf{4_1}$)}] \
5196: $\frac{\partial A_{0,0:c}^1}{\partial 
5197: \theta}(1)$ is a positive multiple of 
5198: $v_1$ for all $c\neq 0$.
5199: \item[{\bf ($\mathbf{5_1}$)}] \ 
5200: The mapping $x\longmapsto A_{x,0:c}^1(1)\in M^1$ is of
5201: rank $n$.
5202: \item[{\bf ($\mathbf{6_1}$)}] \
5203: The mapping $v \longmapsto \frac{ \partial A_{0, v:c}^1}{ \partial
5204: \theta }\left(e^{ i\theta}\right)$ is of rank $n$ at $v=0$.
5205: Consequently, as $v$ varies, the positive half-lines $\R^+ \cdot
5206: \frac{ \partial A_{0,v:c}^1 }{ \partial \theta} (1)$ describe an open
5207: infinite cone containing $v_1$ with vertex $p_1$ in $T_{p_1} M^1$ and
5208: the set of points $A_{x,v:c}^1(\zeta)$, as $\vert x\vert \leq c$,
5209: $\vert v\vert \leq c$ and $\zeta\in \Delta$ vary, covers a wedge of
5210: edge $M^1$ at $(p_1,Jv_1)$.
5211: \end{itemize}
5212: \end{lemma}
5213: 
5214: In Figure~16 drawn in Section~8 after Lemma~8.3 below, we have drawn
5215: the property that the tangent direction $\frac{ \partial A_{0,v: c}^1}{
5216: \partial \theta} (1)$ describes an open cone in $T_{p_1}M^1$ with
5217: vertex $p_1$. The remainder of Section~3 is devoted to complete the
5218: proof of Proposition~7.12.
5219: 
5220: \subsection*{7.13.~Perturbed family of analytic discs half-attached to 
5221: a maximally real submanifold} Thus, let $M^1\subset \R^n$ be a locally
5222: defined maximally real $\mathcal{ C}^{2,\alpha}$-smooth submanifold
5223: passing through the origin. We can assume that $M^1$ is represented by
5224: $n$ Cartesian equations
5225: \def\theequation{7.14}\begin{equation}
5226: y_1=h_1(x_1,\dots,x_n),\cdots\cdots, y_n=h_n(x_1,\dots,x_n),
5227: \end{equation} 
5228: where $z_k= x_k+i y_k\in\C$, for $k= 1,\dots, n$, where $\vert x \vert
5229: \leq \rho_1$ for some $\rho_1 >0$, where $h= h(x)$ is of class
5230: $\mathcal{ C}^{2, \alpha}$ in $\{\vert x \vert < 2\rho_1\}$, and
5231: where, importantly, $h_j(0)=h_{j,x_k}(0)= h_{j, x_{k_1}
5232: x_{k_2}}(0)=0$, for all $j, k, k_1,k_2 =1, \dots,n$. As
5233: in~\thetag{6.5}, we set $K_1:=\vert\vert h \vert \vert_{ \mathcal{
5234: C}^{2,\alpha}(\{ \vert x \vert \leq \rho_1\})}$. Also, we can assume
5235: that $v_1=(0,\dots,0,1)$.
5236: 
5237: Our goal is to show that we can produce a $\mathcal{
5238: C}^{2,\alpha-0}$-smooth (remind $\mathcal{ C}^{2,\alpha-0} \equiv
5239: \bigcap_{ \beta < \alpha} \, \mathcal{ C}^{2, \beta}$) family of
5240: analytic discs $Z_{c,x,v}^1(\zeta)$ which is half-attached to $M^1$
5241: and which is sufficiently close, in $\mathcal{ C}^2$ norm, to the
5242: original family $Z_{c,x,v}^0(\zeta)$. After having constructed the
5243: family $Z_{c,x,v}^1(\zeta)$, we shall define the desired family
5244: $A_{x,v:c}^1(\zeta)$.
5245: 
5246: Let $d\in\R$ with $0\leq d \leq 1$ and let the maximally real
5247: submanifold $M^d$ (like ``$M$ {\it d}eformed'') be defined precisely
5248: as the set of $z= x+iy \in \C^n$ with $\vert x \vert \leq \rho_1$
5249: which satisfy the $n$ Cartesian equations
5250: \def\theequation{7.15}\begin{equation}
5251: y_1=d\cdot h_1(x_1,\dots,x_n),\cdots\cdots, \, 
5252: y_n=d\cdot h_n(x_1,\dots,x_n).
5253: \end{equation}
5254: Notice that $M^0 \equiv\{ x\in \R^n: \ \vert x \vert \leq \rho_1\}$ is
5255: essentially the same piece $M^0$ of $\R^n$ as in Lemma~7.9 (which
5256: contains the $M^0$ of Lemma~7.9 if we choose $\rho_0\leq \rho_1$) and
5257: notice that $\left. M^d\right \vert_{d=1} \equiv M^1$. Even better, we
5258: shall construct for each $d$ with $0 \leq d\leq 1$ a one-parameter
5259: family of analytic dics $Z_{c,x, v}^d (\zeta)$ which is of class at
5260: least $\mathcal{ C}^{2, \alpha-0}$ with respect to all variables and
5261: which is half-attached to $M^d$, by proceeding as follows.
5262: 
5263: First of all, the analytic disc $Z_{c,x, v}^d (\zeta)=: X_{c,x,
5264: v}^d (\zeta)+ iY_{c,x, v}^d( \zeta)$ is half-attached to $M^d$ if
5265: and only if
5266: \def\theequation{7.16}\begin{equation}
5267: Y_{c,x,v}^d(\zeta)=d\cdot h\left(
5268: X_{c,x,v}^d(\zeta)
5269: \right), \ \ \ \ \ 
5270: {\rm for} \ \zeta\in\partial^+\Delta.
5271: \end{equation}
5272: Furthermore, $Y_{c,x, v}^d$ should be a harmonic conjugate of
5273: $X_{c,x, v}^d$. However, the condition~\thetag{7.16} does not give any
5274: relation between $X_{ c,x, v}^d$ and $Y_{c,x, v}^d$ on the
5275: negative part $\partial^-\Delta$ of the unit circle. To fix this
5276: point, we shall assign the following more
5277: complete equation
5278: \def\theequation{7.17}\begin{equation}
5279: Y_{c,x,v}^d(\zeta)= d\cdot h\left(
5280: X_{c,x,v}^d(\zeta)\right)+Y_{c,x,v}^0(\zeta), 
5281: \ \ \ \ \
5282: \text{\rm for {\it all}} \ \zeta\in\partial\Delta,
5283: \end{equation}
5284: which coincides with~\thetag{7.16} for $\zeta\in
5285: \partial^+ \Delta$, since we have $Z_{c, x, v}^0( \partial^+
5286: \Delta) \subset \R^n$ by construction ({\it cf.}~\thetag{7.8}).
5287:  
5288: As in~\cite{tu2}, \cite{tu3}, \cite{mp1}, \cite{mp3}, we denote by
5289: $T_1$ the Hilbert transform (harmonic conjugate operator) on $\partial
5290: \Delta$ vanishing at $1$, namely $(T_1 X) (1)=0$, whence
5291: $T_1(T_1(X))=-X+X(1)$. By Privalov's theorem, for every integer
5292: $\kappa \geq 0$ and every $\alpha\in\R$ with $0< \alpha < 1$, its norm
5293: $\vert \vert \vert T_1 \vert \vert \vert_{ \kappa, \alpha}$ as an
5294: operator $\mathcal{ C}^{ \kappa, \alpha}( \partial \Delta, \R^n) \to
5295: \mathcal{ C}^{\kappa, \alpha} (\partial \Delta, \R^n)$ is finite and
5296: explodes as $\alpha$ tends either to $0$ or to $1$. Also, we shall
5297: require that $X_{c,x, v}^d (1) =x$, whence $Y_{ c,x, v}^d (1)= d
5298: \cdot h(x)$.
5299: 
5300: With this choice, the mapping $\zeta \mapsto Y_{c,x, v}^d (\zeta)$
5301: should necessarily coincide with the harmonic conjugate $\zeta \mapsto
5302: \left[T_1 X_{c,x, v}^d\right] 
5303: (\zeta)+ d\cdot h(x)$ (this property is already
5304: satisfied for $d=0$) and we deduce that $X_{c, x,v}^d( \zeta)$
5305: should satisfy the following Bishop type equation
5306: \def\theequation{7.18}\begin{equation}
5307: X_{c,x, v}^d(\zeta) =
5308: -T_1\left[
5309: d \cdot h\left(
5310: X_{c,x,v}^d
5311: \right)
5312: \right](\zeta)+X_{c,x,v}^0(\zeta), 
5313: \ \ \ \ \ 
5314: \text{\rm for all} \ \zeta\in\partial\Delta.
5315: \end{equation}
5316: Conversely, if $X_{c,x, v}^d$ is a solution of this functional
5317: equation, then setting $Y_{c,x, v}^d(\zeta):= T_1 X_{c,x, v}^d
5318: (\zeta)+ d\cdot h(x)$, it is easy to see that the analytic disc $Z_{c,x,
5319: v}^d (\zeta):= X_{c,x,v}^d (\zeta)+ iY_{c,x, v}^d(\zeta)$ is
5320: half-attached to $M^d$ and more precisely, satisfies the
5321: equation~\thetag{7.16}.
5322: 
5323: Applying now Theorem~1.2 of \cite{tu3}, we deduce that if the given
5324: positive number $c_1$ is sufficiently small, and if $c$ satisfies
5325: $0\leq c \leq c_1$, there exists a unique solution $X_{c,x,
5326: v}^d(\zeta)$ to~\thetag{7.18} which is of class $\mathcal{ C}^{2,
5327: \alpha}$ with respect to $\zeta$ and of class $\mathcal{ C}^{2,\alpha
5328: -0}$ with
5329: respect to all variables $(c,x, v, \zeta)$ with $0\leq c \leq c_1$,
5330: $\vert x\vert \leq c$, $\vert v \vert \leq 2$ and $\zeta\in
5331: \overline{\Delta}$. We shall now estimate the difference $\vert \vert
5332: Z_{c,x,v}^d-Z_{c,x,v}^0\vert \vert_{ \mathcal{ C}^{1,\alpha} (\partial
5333: \Delta)}$ and prove that it is bounded by a constant times
5334: $c^{2+\alpha}$. In particular, if $c_1$ is sufficiently small, this
5335: will imply that $Z_{c,x,v}^d$ is nonconstant.
5336: 
5337: \subsection*{7.19.~Size of the solution 
5338: $X_{c,x,v}^d(\zeta)$ in $\mathcal{ C}^{1,\alpha}$ norm} Following the
5339: beginning of the proof of Theorem~1.2 in~\cite{tu3}, we introduce the
5340: mapping
5341: \def\theequation{7.20}\begin{equation}
5342: F: \ X(\zeta)\longmapsto X_{c,x,v}^0(\zeta)
5343: -T_1\left[
5344: d\cdot h(X)
5345: \right](\zeta)
5346: \end{equation}
5347: from a neighborhood of $0$ in $\mathcal{ C}^{1,\alpha}( \partial
5348: \Delta, \R^n)$ to $\mathcal{ C}^{1,\alpha}(\partial \Delta, \R^n)$,
5349: and then, as in~\cite{b}, we introduce a Picard iteration processus by
5350: defining $X\{0\}_{c,x,v}^d(\zeta):= X_{c,x,v}^0(\zeta)$ and for every
5351: integer $\nu\geq 0$
5352: \def\theequation{7.21}\begin{equation}
5353: X\{\nu+1\}_{c,x,v}^d(\zeta):= 
5354: F\left(X\{\nu\}_{c,x,v}^d(\zeta)\right).
5355: \end{equation}
5356: In a first moment, A.~Tumanov proves in~\cite{tu3} that the sequence
5357: $\left( X\{\nu \}_{c,x, v}^d (\zeta) \right)_{ \nu \in \N}$ converges
5358: towards the unique solution $X_{c,x, v}^d(\zeta)$ of~\thetag{7.18} in
5359: $\mathcal{ C}^{1, \alpha}( \partial \Delta)$. Admitting this
5360: convergence result, we need to extract the supplementary information
5361: that $\left \vert \left \vert X_{c,x, v}^d \right \vert\right\vert_{
5362: \mathcal{ C}^{1, \alpha}( \partial \Delta)} \leq c\cdot K_2$ for some
5363: positive constant $K_2$, which will play the role of the constant
5364: $K_2$ of Lemma~6.12.
5365: 
5366: To get this information, we observe that by construction
5367: ({\it cf.}~\thetag{7.8}) there exists a constant $K_4>0$ such that
5368: \def\theequation{7.22}\begin{equation}
5369: \left\vert \left\vert X_{c,x,v}^0\right\vert 
5370: \right\vert_{\mathcal{
5371: C}^{2,\alpha}(\partial \Delta)}\leq c \cdot K_4.
5372: \end{equation}
5373: Also, we set $K_5:= K_1 (3+ \pi^{1-\alpha}) \vert \vert \vert T_1
5374: \vert \vert \vert_{ \mathcal{ C}^{1, \alpha}( \partial \Delta)}$.
5375: 
5376: \def\thelemma{{\bf 7.23}}\begin{lemma}
5377: With these notations, if 
5378: \def\theequation{7.24}\begin{equation}
5379: c_1\leq\min\left( 
5380: \frac{\rho_1}{2K_4}, \ 
5381: \left(
5382: \frac{1}
5383: {2^{2+\alpha}\, K_4^{1+\alpha} \, K_5}
5384: \right)^{\frac{1}{1+\alpha}}\right),
5385: \end{equation}
5386: then the solution of~\thetag{7.18} satisfies $\left\vert X_{c,x, v}^d
5387: \left( e^{i \theta }\right) \right\vert \leq \rho_1$ for all
5388: $e^{i\theta}\in\partial \Delta$ and there exists a constant $K_2>0$
5389: such that
5390: \def\theequation{7.25}\begin{equation}
5391: \left
5392: \vert \left \vert X_{c,x, v}^d 
5393: \right \vert \right \vert_{ \mathcal{
5394: C}^{1, \alpha} ( \partial \Delta)} \leq c \cdot K_2. 
5395: \end{equation}
5396: In fact, it
5397: suffices to choose $K_2:= 2 K_4$.
5398: \end{lemma}
5399: 
5400: \proof
5401: Indeed, using Lemmas~6.4 and~6.9, if $X \in \mathcal{ C}^{1, \alpha}(
5402: \partial \Delta, \R^n)$ satisfies $\left\vert 
5403: X\left( e^{i \theta }\right) \right\vert \leq
5404: \rho_1$ for all $e^{i\theta}\in\partial\Delta$ and $\vert \vert X \vert
5405: \vert_{ \mathcal{ C}^{1, \alpha} (\partial \Delta )} \leq c \cdot
5406: 2K_4$ for all $c \leq c_1$, where $c_1$ is as in~\thetag{7.24}, we
5407: may estimate (remind $0 \leq d \leq 1$)
5408: \def\theequation{7.26}\begin{equation}
5409: \left\{
5410: \aligned
5411: \vert \vert F(X) \vert \vert_{\mathcal{ C}^{1, 
5412: \alpha}(\partial \Delta)} 
5413: & \
5414: \leq
5415: \left\vert \left\vert X_{c,x,v}^0 \right\vert 
5416: \right\vert_{\mathcal{ C}^{1,\alpha}
5417: (\partial \Delta)}+
5418: \vert \vert \vert T_1 \vert \vert \vert_{\mathcal{ C}^{1, \alpha}
5419: (\partial \Delta)} \cdot
5420: \vert \vert h(X) \vert \vert_{\mathcal{ C}^{1, \alpha}
5421: (\partial \Delta)} \\
5422: & \
5423: \leq 
5424: c\cdot K_4+
5425: \vert \vert \vert T_1 \vert \vert \vert_{\mathcal{ C}^{1, \alpha}
5426: (\partial \Delta)} \cdot K_1
5427: ( c \cdot 2K_4)^{2+\alpha}(
5428: 3+\pi^{1-\alpha}) \\
5429: & \
5430: = 
5431: c\cdot \left(
5432: K_4+c^{1+\alpha} 2^{2+\alpha} K_4^{2+\alpha} K_5
5433: \right) \\
5434: & \
5435: \leq 
5436: c\cdot (K_4+c_1^{1+\alpha} 
5437: 2^{2+\alpha}K_4^{2+\alpha} K_5) \\
5438: & \
5439: \leq c\cdot 2K_4.
5440: \endaligned\right.
5441: \end{equation}
5442: Notice that from the last inequality, it also follows that $\left\vert
5443: F\left( X\left( e^{i\theta} \right) \right)\right\vert \leq \rho_1$
5444: for all $e^{i\theta}\in\partial \Delta$. Consequently, the processus
5445: of successive approximations~\thetag{7.21} is well defined for each
5446: $\nu\in\N$ and from the inequality~\thetag{7.26}, we deduce that the
5447: limit $X_{c,x,v}^d$ satisfies the desired estimate $\left\vert
5448: \left\vert X_{c,x, v}^d \right\vert \right\vert_{ \mathcal{ C}^{1,
5449: \alpha}( \partial \Delta )}\leq c \cdot 2K_4$, which completes the
5450: proof.
5451: \endproof
5452: 
5453: \def\thecorollary{{\bf 7.27}}\begin{corollary}
5454: Under the above assumptions, there exists a constant
5455: $K_6>0$ such that
5456: \def\theequation{7.28}\begin{equation}
5457: \left\vert \left\vert X_{c,x,v}^d - X_{c,x,v}^0 
5458: \right\vert \right\vert_{
5459: \mathcal{ C}^{1,\alpha}(\partial \Delta)} \leq
5460: c^{2+\alpha} \cdot K_6.
5461: \end{equation}
5462: \end{corollary}
5463: 
5464: \proof
5465: We estimate
5466: \def\theequation{7.29}\begin{equation}
5467: \left\{
5468: \aligned
5469: \left\vert \left\vert X_{c,x,v}^d - X_{c,x,v}^0 
5470: \right\vert \right\vert_{
5471: \mathcal{ C}^{1,\alpha}(\partial \Delta)} 
5472: & \
5473: \leq
5474: \vert \vert \vert T_1 \vert \vert \vert_{\mathcal{ C}^{1,\alpha}
5475: (\partial \Delta)} \cdot 
5476: \left\vert \left\vert h
5477: \left(X_{c,x,v}^d\right) \right\vert 
5478: \right\vert_{\mathcal{ C}^{1,\alpha}
5479: (\partial \Delta)} \\
5480: & \ 
5481: \leq 
5482: \vert \vert \vert T_1 \vert \vert \vert_{\mathcal{ C}^{1,\alpha}
5483: (\partial \Delta)} \cdot 
5484: K_1 (c\cdot 2K_4)^{2+\alpha}(3+\pi^{1-\alpha}) \\
5485: & \
5486: \leq
5487: c^{2+\alpha} \cdot K_5 (2K_4)^{2+\alpha}.
5488: \endaligned\right.
5489: \end{equation}
5490: so that it suffices to set 
5491: $K_6:= K_5(2K_4)^{2+\alpha}$.
5492: \endproof
5493: 
5494: \subsection*{7.30.~Smallness of the deformation in $\mathcal{ C}^2$
5495: norm} As was already noticed (and admitted), the solution
5496: $X_{c,x,v}^d(\zeta)$ is in fact $\mathcal{ C}^{2,\alpha}$-smooth with
5497: respect to $\zeta$ and $\mathcal{ C}^{2,\alpha-0}$-smooth with respect
5498: to all variables $(d,c,x,v,\zeta)$. We can therefore differentiate
5499: twice Bishop's equation~\thetag{7.18}. First of all, if $X\in\mathcal{
5500: C}^{2,\alpha-0}( \partial \Delta, \R^n)$, we remind the commutation
5501: relation $\frac{\partial }{\partial \theta} (TX)= T\left(
5502: \frac{\partial X}{\partial \theta} \right)$, whence
5503: \def\theequation{7.31}\begin{equation}
5504: \frac{\partial}{\partial \theta} \left(
5505: T_1X
5506: \right)= 
5507: T\left(
5508: \frac{\partial X}{\partial \theta}
5509: \right),
5510: \end{equation}
5511: since $T_1X= TX-TX(1)$. We may then compute the first order 
5512: derivative of~\thetag{7.18} with respect 
5513: to $\theta$:
5514: \def\theequation{7.32}\begin{equation}
5515: \left\{
5516: \aligned
5517: \frac{\partial }{\partial \theta} 
5518: X_{c,x,v}^d\left(e^{i\theta}\right) -
5519: & \
5520: \frac{\partial }{\partial \theta}
5521: X_{c,x,v}^0\left(e^{i\theta}\right)=
5522: -T\left[
5523: d\cdot \sum_{l=1}^n\, 
5524: \frac{\partial h}{\partial x_l} \left(X_{c,x,v}^d\right) \
5525: \frac{\partial X_{l;c,x,v}^d}{\partial \theta}
5526: \right]\left(e^{i\theta}\right).
5527: \endaligned\right.
5528: \end{equation}
5529: and then its second order partial derivatives $\partial^2 / \partial
5530: v_k \partial \theta$, for $k=1,\dots,n$, without writing the argument
5531: $e^{i\theta}$:
5532: \def\theequation{7.33}\begin{equation}
5533: \left\{
5534: \aligned
5535: \frac{\partial^2 X_{c,x,v}^d}{\partial v_k\partial \theta}
5536: -
5537: \frac{\partial^2X_{c,x,v}^0}{\partial v_k
5538: \partial \theta}
5539: = 
5540: & \
5541: -T\left[
5542: d\cdot \sum_{l_1,l_2=1}^n\,
5543: \frac{\partial^2 h}{\partial x_{l_1} \partial x_{l_2}}\left(
5544: X_{c,x,v}^d\right) \
5545: \frac{\partial X_{l_1; c,x,v}^d}{\partial v_k} \,
5546: \frac{\partial X_{l_2; c,x,v}^d}{\partial \theta} + \right. \\
5547: & \
5548: \left. \ \ \ \ \ \ \ \ \ \
5549: +d\cdot \sum_{l=1}^n\, 
5550: \frac{\partial h_j}{\partial x_l} \left(
5551: X_{c,x,v}^d
5552: \right) \
5553: \frac{\partial^2 X_{l;c,x,v}^d}{\partial v_k 
5554: \partial \theta}
5555: \right].
5556: \endaligned\right.
5557: \end{equation}
5558: Let now $K_2$ be as in~\thetag{7.25} and let 
5559: $K_3$ be as in Lemma~6.12, applied to 
5560: $X_{c,x,v}^d(\zeta)$.
5561: 
5562: \def\thelemma{7.34}\begin{lemma}
5563: If in addition to the inequality~\thetag{7.24}, the constant
5564: $c_1$ satisfies the inequality
5565: \def\theequation{7.35}\begin{equation}
5566: c_1\leq \left(
5567: \frac{1}{2K_3 \vert \vert \vert 
5568: T \vert \vert \vert_{
5569: \mathcal{ C}^\alpha(\partial \Delta)}}
5570: \right)^{\frac{1}{1+\alpha}},
5571: \end{equation}
5572: then there exists a positive constant $K_7>0$
5573: such that for all $d$, all $c$, all $x$, all $v$, and for
5574: $k=1,\dots,n$, the following two estimates hold{\rm :}
5575: \def\theequation{7.36}\begin{equation}
5576: \left\{
5577: \aligned
5578: \left\vert\left\vert
5579: \frac{\partial^2 X_{c,x,v}^d}{
5580: \partial v_k \partial \theta}-
5581: \frac{\partial^2 X_{c,x,v}^0}{
5582: \partial v_k \partial \theta}
5583: \right\vert\right\vert_{\mathcal{ 
5584: C}^\alpha(\partial \Delta)}\leq 
5585: & \
5586: c^{2+\alpha} \cdot K_7, \\
5587: \left\vert\left\vert
5588: \frac{\partial^2 X_{c,x,v}^d
5589: }{\partial \theta^2}-
5590: \frac{\partial^2 X_{c,x,v}^0
5591: }{\partial \theta^2}
5592: \right\vert\right\vert_{\mathcal{ 
5593: C}^\alpha(\partial \Delta)}\leq 
5594: & \
5595: c^{2+\alpha} \cdot K_7.
5596: \endaligned\right.
5597: \end{equation}
5598: \end{lemma}
5599: 
5600: \proof
5601: We check only the first inequality, the proof of the second being
5602: totally similar. According to Lemma~1.6 in~\cite{ tu3}, 
5603: there exists a solution 
5604: $\frac{\partial^2 X_{c,x,v}^d}{
5605: \partial v_k \partial \theta}$
5606: to the linearized Bishop equation~\thetag{ 7.33},
5607: hence it suffices to make an estimate.
5608: 
5609: Introducing for the second line of~\thetag{7.33} a
5610: new simplified notation $\mathcal{ R}:= -T\left[ d\cdot
5611: \sum_{l_1,l_2=1}^n\, \frac{\partial^2 h}{ \partial x_{l_1} \partial
5612: x_{l_2} }\left( X_{c,x,v}^d \right) \ \frac{\partial X_{l_1;
5613: c,x,v}^d}{\partial v_k} \, \frac{\partial X_{l_2; c,x,v}^d}{ \partial
5614: \theta}\right]$ and setting further obvious simplifying changes of
5615: notation, we can rewrite~\thetag{7.33} more concisely as
5616: \def\theequation{7.37}\begin{equation}
5617: \mathcal{ X}^d-\mathcal{ X}^0= 
5618: \mathcal{ R}- T\left[
5619: d\cdot \mathcal{ H} \mathcal{ X}^d
5620: \right].
5621: \end{equation}
5622: Here, thanks to the inequality $\left\vert \left\vert X_{c,x, v}^d
5623: \right\vert \right\vert_{\mathcal{ C}^{1, \alpha} (\partial
5624: \Delta)}\leq c\cdot K_2$ already established in Lemma~7.23 and thanks
5625: to Lemma~6.12, we know that the vector $\mathcal{ R}\in \mathcal{
5626: C}^{\alpha }( \partial \Delta, \R^n)$ and the matrix $\mathcal{ H}\in
5627: \mathcal{ C}^{1, \alpha}( \partial \Delta, \mathcal{ M}_{n\times
5628: n}(\R))$ are small and more precisely, they satisfy the following two
5629: estimates
5630: \def\theequation{7.38}\begin{equation}
5631: \left\{
5632: \aligned
5633: \vert \vert \mathcal{ R} \vert \vert_{
5634: \mathcal{ C}^\alpha(\partial \Delta)}\leq
5635: & \
5636: c^{2+\alpha} \cdot 
5637: \vert \vert \vert T \vert \vert \vert_{\mathcal{ C}^\alpha
5638: (\partial \Delta)} K_3 (K_2)^2
5639: \\
5640: \vert \vert
5641: \mathcal{ H} \vert \vert_{\mathcal{ C}^\alpha(
5642: \partial \Delta)}\leq 
5643: & \
5644: c^{1+\alpha} \cdot K_3.
5645: \endaligned\right.
5646: \end{equation}
5647: We can rewrite~\thetag{7.37} under the form
5648: \def\theequation{7.39}\begin{equation}
5649: \mathcal{ X}^d -\mathcal{ X}^0 =
5650: \mathcal{ S} -T\left[
5651: d\cdot \mathcal{ H}(\mathcal{ X}^d-
5652: \mathcal{ X}^0)
5653: \right],
5654: \end{equation}
5655: with $\mathcal{ S}:= \mathcal{ R}-T \left[ d\cdot \mathcal{ H}
5656: \mathcal{ X}^0 \right]$. Using the inequality $\left\vert \left\vert
5657: \mathcal{ X}^0 \right\vert \right\vert_{ \mathcal{ C}^\alpha( \partial
5658: \Delta)} \leq c\cdot K_4$ which is a direct consequence
5659: of~\thetag{7.22} and taking the previous estimates~\thetag{7.38}
5660: into account, we deduce the inequality
5661: \def\theequation{7.40}\begin{equation}
5662: \vert \vert \mathcal{ S}
5663: \vert \vert_{\mathcal{ C}^\alpha(\partial \Delta)}\leq
5664: c^{2+\alpha}\cdot
5665: \vert \vert \vert T \vert \vert \vert_{\mathcal{ C}^\alpha
5666: (\partial \Delta)}\left[
5667: K_3(K_2)^2+
5668: K_3K_4
5669: \right].
5670: \end{equation} 
5671: Taking the $\mathcal{ C}^\alpha(\partial \Delta)$ norm of both sides
5672: of~\thetag{7.40}, we deduce the estimate
5673: \def\theequation{7.41}\begin{equation}
5674: \left\{
5675: \aligned
5676: \left\vert \left\vert 
5677: \mathcal{ X}^d - \mathcal{ X}^0 \right\vert 
5678: \right\vert_{
5679: \mathcal{ C}^\alpha(\partial \Delta)} \leq
5680: & \ 
5681: c^{2+\alpha} \cdot \frac{
5682: \vert \vert \vert T \vert \vert \vert_{\mathcal{ C}^\alpha
5683: (\partial \Delta)}\left[
5684: K_3(K_2)^2+
5685: K_3K_4
5686: \right]
5687: }{
5688: 1- c^{1+\alpha}
5689: \cdot \vert \vert \vert T \vert \vert \vert_{\mathcal{ C}^\alpha
5690: (\partial \Delta)} K_3} \\
5691: \leq
5692: & \
5693: c^{2+\alpha} \cdot 2\vert 
5694: \vert \vert T \vert \vert \vert_{\mathcal{ C}^\alpha
5695: (\partial \Delta)}\left[
5696: K_3(K_2)^2+
5697: K_3K_4
5698: \right]
5699: \endaligned\right.
5700: \end{equation}
5701: where we use the assumption~\thetag{7.35} on $c_1$
5702: to obtain the second second inequality. It suffices
5703: to set $K_7:= 2\vert 
5704: \vert \vert T \vert \vert \vert_{\mathcal{ C}^\alpha
5705: (\partial \Delta)}\left[
5706: K_3(K_2)^2+
5707: K_3K_4
5708: \right]$, which completes the proof.
5709: \endproof
5710: 
5711: \subsection*{7.42.~Adjustment of the tangent vector}
5712: Let $v_1\in T_{p_1}M^1$ with $\vert v_1 \vert =1$, as in Lemma~7.12.
5713: Coming back to the first family $Z_{c,x, v}^0 (\zeta)$ defined
5714: by~\thetag{7.8}, we observe that
5715: \def\theequation{7.43}\begin{equation}
5716: \left\{
5717: \aligned
5718: \frac{\partial Z_{j;c,0,v_1}^0}{\partial x_k}
5719: (1)=
5720: & \
5721: \delta_k^j, 
5722: \ \ \ \ \ 
5723: j,k=1,\dots,n, \\
5724: \frac{\partial^2 Z_{j;
5725: c,0,v_1}^0}{\partial v_k 
5726: \partial \theta}(1)=
5727: & \
5728: c\, \frac{\partial \Psi}{\partial \theta}
5729: \left(e^{i\theta}\right) \, 
5730: \delta_k^j, 
5731: \ \ \ \ \ 
5732: j,k=1,\dots,n.
5733: \endaligned\right.
5734: \end{equation}
5735: From now on, we shall set $d=1$ and we shall only consider the family
5736: $Z_{c,x,v}^1(\zeta)$. Thanks to the estimates~\thetag{7.28}
5737: and~\thetag{7.36}, we deduce that if $c_1$ is sufficiently small, then
5738: for all $c$ with $0< c \leq c_1$, the two Jacobian matrices
5739: \def\theequation{7.44}\begin{equation}
5740: \left(
5741: \frac{\partial Z_{j;c,0,v_1}^1}{\partial x_k}(1)
5742: \right)_{1\leq j,k\leq n} \ \ \ \ \
5743: {\rm and} \ \ \ \ \ \
5744: \left(
5745: \frac{\partial^2 Z_{j;c,0,v_1}^1}{\partial v_k\partial \theta}(1)
5746: \right)_{1\leq j,k\leq n}
5747: \end{equation}
5748: are invertible. It would follow that if we would set $A_{x, v:c}^1
5749: (\zeta):= Z_{c, x, v_1 +v}^1(\zeta)$, similarly as in~\thetag{7.10},
5750: then the disc $A_{x, v:c}^1 (\zeta)$ would satisfy the two rank
5751: properties {\bf ($\mathbf{ 5_1}$)} and {\bf ($\mathbf{ 6_1}$)} of
5752: Lemma~7.12. However, the tangency condition {\bf ($\mathbf{ 4_1}$)}
5753: would certainly not be satisfied, because as $d$ varies from $0$ to
5754: $1$, the disc $Z_{c,x,v}^d(\zeta)$ undergo a nontrivial deformation.
5755: 
5756: Consequently, for every $c$ with $0< c \leq c_1$, 
5757: we have to adjust the ``cone
5758: parameter'' $v$ in order to maintain the tangency 
5759: condition.
5760: 
5761: \def\thelemma{7.45}\begin{lemma}
5762: For every $c$ with $0< c \leq c_1$, there exists a
5763: vector $v(c)\in \R^n$ such that
5764: \def\theequation{7.46}\begin{equation}
5765: \frac{\partial Z_{c,0,v_1
5766: +v(c)}^1}{\partial \theta} 
5767: (1)=
5768: \frac{\partial Z_{c,0,
5769: v_1}^0}{\partial \theta}(1)=c \cdot
5770: \frac{\partial \Psi}{\partial \theta}(1) \cdot v_1.
5771: \end{equation}
5772: Furthermore, there exists a constant $K_8>0$ such that
5773: $\vert v(c) \vert \leq c^{1+\alpha} \cdot K_8$.
5774: \end{lemma}
5775: 
5776: \proof
5777: Unfortunately, we cannot apply the implicit function theorem, because
5778: the mapping $Z_{c,x,v}^1$ is identically zero when $c=0$, so we have
5779: to proceed differently. First, we set
5780: \def\theequation{7.47}\begin{equation}
5781: C_1:= \frac{\partial \Psi}{\partial \theta}(1), \ \ \ \ \ 
5782: {\rm and}
5783: \ \ \ \ \
5784: C_2:= \vert\vert
5785: \Psi \vert\vert_{\mathcal{ C}^2\left(\overline{\Delta}\right)}.
5786: \end{equation}
5787: The constant $C_2$ will be used only in Section~8 below.
5788: Choose $K_8\geq \frac{2K_6}{C_1}$. According to the explicit
5789: expression~\thetag{7.8}, the set of points
5790: \def\theequation{7.48}\begin{equation}
5791: \left\{
5792: \frac{\partial X_{c,0,v_1+v}^0}{\partial \theta}(1)\in \R^n: \ 
5793: \vert v \vert \leq c^{1+\alpha} \cdot K_8
5794: \right\}
5795: \end{equation}
5796: covers a cube in $\R^n$ centered at the point $\frac{ \partial X_{
5797: c,0, v_1 }^0}{\partial \theta}(1)$ of radius $c^{2+ \alpha} \cdot C_1
5798: K_8$.
5799: Thanks to the estimate~\thetag{7.28}, we deduce
5800: that the (deformed) set of points 
5801: \def\theequation{7.49}\begin{equation}
5802: \left\{
5803: \frac{\partial X_{c,0,v_1+v}^1}{\partial \theta}(1)\in \R^n: \ 
5804: \vert v \vert \leq c^{1+\alpha} \cdot K_8
5805: \right\}
5806: \end{equation}
5807: covers a cube in $\R^n$ centered at the same
5808: point $\frac{ \partial X_{
5809: c,0, v_1 }^0}{ \partial \theta}(1)$, but of radius 
5810: \def\theequation{7.50}\begin{equation}
5811: c^{2+ \alpha} \cdot C_1
5812: K_8-c^{2+\alpha} \cdot K_6\geq c^{2+\alpha} \cdot K_6.
5813: \end{equation}
5814: Consequently, there exists at least one $v(c) \in \R^n$ with $\vert
5815: v(c) \vert \leq c^{ 1+\alpha} \cdot K_8$ such that~\thetag{7.46}
5816: holds, which completes the proof. 
5817: \endproof
5818: 
5819: \subsection*{7.51.~Construction of the family $A_{x, v:c}^1 (\zeta)$}
5820: We can now complete the proof of the main Lemma~7.12 of the present
5821: section. First of all, with $\Psi (\zeta)$ as in \S7.5 ans in {\sc
5822: Figure~14}, we consider the composed conformal mapping
5823: \def\theequation{7.52}\begin{equation}
5824: \zeta \longmapsto c\Psi(\zeta) \longmapsto 
5825: \frac{i-c\Psi(\zeta)}{i+c\Psi(\zeta)}=:
5826: \Phi_c(\zeta).
5827: \end{equation}
5828: The image $\Phi_c ( \zeta)$ of the unit disc is a small domain
5829: contained in $\Delta$ and concentrated near $1$. More precisely,
5830: assuming that $c$ satifies $0< c \leq c_1$ with $c_1<<1$ as in the
5831: previous paragraphs, and taking account of the definition of $\Psi
5832: (\zeta)$, it can be checked easily that $\Phi_c(1) =1$, that $\Phi_c
5833: (\partial^+ \Delta)$ is contained in $\{e^{ i\theta} \in
5834: \partial^+\Delta: \ \vert \theta \vert < 10c\}$, and that
5835: \def\theequation{7.53}\begin{equation}
5836: \Phi_c\left(\overline{\Delta}\backslash 
5837: \partial^+\Delta\right)\subset 
5838: \{ \zeta \in \Delta : \ 
5839: \vert \zeta -1 \vert < 8c\} \subset 
5840: \{\rho e^{i\theta} \in \Delta : \
5841: \vert \theta \vert < 10c, \
5842: 1-10c < \rho < 1\}.
5843: \end{equation}
5844: the second inclusion being trivial. Here is an illustration:
5845: 
5846: \bigskip
5847: \begin{center}
5848: \input sub-half-disc.pstex_t
5849: \end{center}
5850: 
5851: We can now define the final desired family of analytic discs, 
5852: writing the parameter $c$ after a semi-colon, since we lose the
5853: $\mathcal{ C}^{2,\alpha-0}$-smoothness with respect to 
5854: $c$ after the application of Lemma~7.45:
5855: \def\theequation{7.54}\begin{equation}
5856: A_{x,v:c}^1(\zeta):= Z_{c,x,v_1+v(c)+v}^1\left(
5857: \Phi_c(\zeta)\right).
5858: \end{equation}
5859: We restrict the variation of the parameters $x$ to $\vert x \vert \leq
5860: c^2$ and $v$ to $\vert v \vert\leq c$. Property {\bf ($\mathbf{
5861: 4_1}$)} holds immediately, thanks to the choice of $v(c)$. Properties
5862: {\bf ($\mathbf{1_1}$)}, {\bf ($\mathbf{ 3_1}$)}, {\bf ($\mathbf{
5863: 5_1}$)} and {\bf ($\mathbf{ 6_1}$)} as well as the embedding property
5864: in {\bf ($\mathbf{2_1}$)} are direct consequences of the similar
5865: properties~\thetag{7.44} satisfied by $Z_{ c,x, v_1+ v(c)+v}^1 (\zeta)$,
5866: using the chain rule and the nonvanishing of the partial derivative
5867: $\frac{ \partial \Phi_c }{ \partial \theta}(1)$. The size estimate in
5868: {\bf ($\mathbf{2_1}$)} follows from~\thetag{7.25}, from~\thetag{7.28},
5869: from the restriction of the domains of variation of $x$ and of $v$ and
5870: from~\thetag{7.53}. This completes the proof of Lemma~7.12.
5871: \hfill 
5872: \qed
5873: 
5874: \section*{\S8.~Geometric properties of families of half-attached 
5875: analytic discs}
5876: 
5877: \subsection*{8.1.~Preliminary}
5878: By Lemma~7.12, for every $c$ with $0 < c \leq c_1$, the family of
5879: half-attached analytic discs $A_{x, v:c}^1 (\zeta)$ covers a local
5880: wedge of edge $M^1$ at $p_1$. However, not only we want the family
5881: $A_{x, v:c}^1$ to cover a local wedge of edge $M^1$ at $p_1$, but we
5882: certainly want to remove the point $p_1$ by means of the continuity
5883: principle, under the assumptions of the main Proposition~5.12, a final
5884: task which will be achieved in Section~9 below. Consequently, in each
5885: one of the three geometric situations {\bf ($\mathbf{I_1}$)}, {\bf
5886: ($\mathbf{I_2}$)} and {\bf (II)} which we have normalized in
5887: Lemma~5.37 above, we shall firstly deduce from the tangency condition
5888: {\bf ($\mathbf{4_1}$)} of Lemma~7.12 that the blunt half-boundary
5889: $A_{0,0:c}^1( \partial^+ \Delta \backslash \{1\})$ is contained in the
5890: open side $(H^1)^+$ (this is why we have normalized in Lemma~5.37 the
5891: second order terms of the supporting hypersurface $H^1$ in order that
5892: $(H^1)^+$ is strictly concave; the reason why we require that
5893: $A_{0,0:c}^1(\partial^+\Delta\backslash \{1\})$ is contained in
5894: $(H^1)^+$ will be clear in Section~9 below). Secondly, we shall show
5895: that for all $x$ with $\vert x \vert \leq c^2$, the disc interior
5896: $A_{x,0:c}(\Delta)$ is contained in the local half-wedge $\mathcal{
5897: HW}_1^+$ in the cases {\bf ($\mathbf{I_1}$)}, {\bf ($\mathbf{I_2}$)}
5898: and is contained in the wedge $\mathcal{ W}_2$ in case {\bf (II)}.
5899: 
5900: \subsection*{8.2.~Geometric disposition of the
5901: discs with respect to $H^1$ and to $\mathcal{ HW}_1^+$ or to
5902: $\mathcal{ W}_2$} We remember that the positive $c_1$ of Lemmas~7.12,
5903: 7.23 and 7.34 was shrunk explicitely, in terms of the constants
5904: $K_1,K_2,K_3,\dots$. In this section, we shall again shrink $c_1$
5905: several times, but without mentioning all the similar explicit
5906: inequalities which will appear. The precise statement of the main
5907: lemma of this section, which is a continuation of Lemma~7.12, is as
5908: follows; whereas we can essentially gather the three cases in the
5909: formal statement of the lemma, it is necessary to treat them
5910: separately in the proof, because the normalizations of
5911: Lemma~5.37 differ.
5912: 
5913: \def\thelemma{8.3}\begin{lemma}
5914: Let $M$, let $M^1$, let $p_1$, let $H^1$, let $v_1$, let $(H^1)^+$,
5915: let $\mathcal{ HW }_1^+$ or let $\mathcal{ HW }_2$ and let a
5916: coordinate system $z= (z_1, \dots, z_n)$ vanishing at $p_1$ be as in
5917: Case {\bf ($\mathbf{I_1}$)}, as in Case {\bf ($\mathbf{ I_2 }$)} or as
5918: in Case {\bf (II)} of Lemma~5.37. Choose as a local one-dimensional
5919: submanifold $T^1 \subset M^1$ transversal to $H^1$ in $M^1$ and
5920: passing through $p_1$ the submanifold $T_1:=\{\left( x_1,0, \dots,0) +
5921: ih(x_1, 0, \dots, 0) \right)\}$ in Case {\bf
5922: ($\mathbf{I_1}$)} and the submanifold $T_1:=\{\left( 0, \dots,0,x_n) +
5923: ih(0, \dots, 0,x_n) \right)\}$ in Cases {\bf
5924: ($\mathbf{I_2}$)} and {\bf (II)}. For every $c$ with $0 < c \leq
5925: c_1$, let $A_{x,v:c}^1(\zeta)$ be the family of analytic discs
5926: satisfying properties {\bf ($\mathbf{1_1}$)}, {\bf ($\mathbf{2_1}$)},
5927: {\bf ($\mathbf{3_1}$)}, {\bf ($\mathbf{4_1}$)}, {\bf ($\mathbf{5_1}$)}
5928: and {\bf ($\mathbf{6_1}$)} of Lemma~7.12. Shrinking $c_1$ if
5929: necessary, then for every $c$ with $0< c\leq c_1$, the following three
5930: further properties hold
5931: \begin{itemize}
5932: \item[{\bf ($\mathbf{7_1}$)}] $A_{0, 0:c}^1( \partial^+ \Delta
5933: \backslash\{ 1\}) \subset (H^1)^+$.
5934: \item[{\bf ($\mathbf{8_1}$)}]
5935: $A_{x,0:c}^1(\partial^+\Delta)$ is contained
5936: in $(H^1)^+$ for all $x$ such that the point 
5937: $A_{x,0:c}^1(1)$ belongs to 
5938: $T^1\cap (H^1)^+$.
5939: \item[{\bf ($\mathbf{9_1}$)}]
5940: $A_{x,v:c}^1(\overline{\Delta}\backslash\partial^+\Delta)$ is
5941: contained in the half-wedge
5942: $\mathcal{ HW}_1^+$ or in the wedge $\mathcal{ W}_2$ for all
5943: $x$ and all $v$.
5944: \end{itemize}
5945: \end{lemma}
5946: 
5947: \proof
5948: For the three new properties {\bf ($\mathbf{ 7_1}$)}, {\bf ($\mathbf{
5949: 8_1}$)}, and {\bf ($\mathbf{ 9_1}$)}, we study thoroughly only Case
5950: {\bf ($\mathbf{I_1}$)}, because the other two cases can be treated in
5951: a totally similar way. {\sc Figure~16} just below illustrates
5952: properties {\bf ($\mathbf{ 7_1}$)} and {\bf ($\mathbf{ 8_1}$)} and
5953: also properties {\bf ($\mathbf{ 1_1}$)}, {\bf ($\mathbf{ 5_1}$)} and
5954: {\bf ($\mathbf{ 6_1}$)} of Lemma~7.12.
5955: 
5956: \bigskip
5957: \begin{center}
5958: \input half-boundaries.pstex_t
5959: \end{center}
5960: 
5961: Intuitively, the reason why property {\bf ($\mathbf{ 7_1}$)} holds
5962: true is clear: the open set $(H^1)^+$ is strictly concave and the
5963: small segment $A_{0,0:c}^1(\partial^+\Delta)$ is tangent to $H^1$ at
5964: $p_1$; also, the reason why property {\bf ($\mathbf{ 8_1}$)} holds
5965: true is equally clear: when $x$ varies, the small segments
5966: $A_{x,0:c}^1(\partial^+\Delta)$ are essentially translated (inside
5967: $M^1$) from $p_1$ by the vector $x\in\R^n$; and finally, the reason
5968: why property {\bf ($\mathbf{ 8_1}$)} holds true has a simple geometric
5969: interpretation: if the scaling parameter $c_1$ is small enough, the
5970: small analytic disc $A_{x,v:c}^1(\overline{\Delta})$ is essentially a
5971: slightly deformed small part of the straight complex line
5972: $\C\cdot(v_1+Jv_1)$, where the half-wedge $\mathcal{ HW}_1^+$ or the
5973: wedge $\mathcal{ W}_2$ is directed by the vector $Jv_1$ according to
5974: Lemma~5.37. The next paragraphs are devoted to some elementary
5975: estimates which will establish these properties rigorously.
5976: 
5977: Firstly, let us prove property {\bf ($\mathbf{ 7_1}$)} in Case {\bf
5978: ($\mathbf{I_1}$)}. According to Lemma~5.37, the vector $v_1$ is given
5979: by $(0,1,\dots,1)$ and the side $(H^1)^+\subset M^1$ is defined by
5980: $x_1>g(x')=-x_2^2-\cdots-x_n^2+ \widehat{ g}(x')$, where the
5981: $\mathcal{ C}^{2,\alpha}$-smooth function $\widehat{ g}(x')$ vanishes
5982: to second order at the origin, thanks to the normalization
5983: conditions~\thetag{5.40}. By Lemma~6.4, the remainder $\widehat{
5984: g}(x')$ then satisfies an inequality of the form $\vert \widehat{
5985: g}(x') \vert \leq K_9 \cdot \vert x' \vert^{2+\alpha}\leq K_9 \cdot
5986: \left( \sum_{j=2}^n\, x_j^2 \right)^{\frac{ \alpha+2}{2}}$, for some
5987: constant $K_9>0$. Since the strictly concave open subset
5988: $(\widetilde{ H}^1)^+$ of $M^1$ with $\mathcal{ C}^{2,\alpha}$-smooth
5989: boundary defined by the inequality $x_1>-x_1^2-\cdots-x_n^2 + K_9
5990: \cdot \left( \sum_{j=2}^n \, x_j^2 \right)^{\frac{2+\alpha}{2}}$ is
5991: contained in $(H^1)^+$, it suffices to prove property {\bf ($\mathbf{
5992: 7_1}$)} with $(H^1)^+$ replaced by $(\widetilde{ H}^1)^+$.
5993: 
5994: By construction, the disc boundary $A_{0,0:c}(\partial\Delta)$ is
5995: tangent at $p_1$ to $H^1$, hence also to $\widetilde{
5996: H}^1$. Intuitively, it is clear that the blunt disc half-boundary
5997: $A_{0,0:c}(\partial^+\Delta\backslash\{1\})$ should then be contained
5998: in the strictly concave open subset $(\widetilde{ H}^1)^+$, 
5999: {\it see}\, {\sc Figure~16} above. 
6000: 
6001: To proceed rigorously, we shall come back to the definition~\thetag{
6002: 7.53} which yields $A_{0, 0:c}^1( \zeta) \equiv Z_{c,0, v_1+ v(c)}^1
6003: \left( \Phi_c(\zeta)\right)$, with the tangency condition~\thetag{
6004: 7.46} satisfied. First of all, denoting the $n$ components of $v(c)$
6005: by $(v_1(c), \dots, v_n(c))$, we may compute the second order
6006: derivatives of the similar discs attached to $M^0$:
6007: \def\theequation{8.4}\begin{equation}
6008: \left\{
6009: \aligned
6010: \frac{\partial^2 Z_{1;c,0,
6011: v_1+v(c)}^0}{\partial \theta^2}(1)= 
6012: & \
6013: c\cdot \frac{\partial^2\Psi}{
6014: \partial \theta^2}\left(e^{i\theta}\right)\cdot
6015: v_1(c), \\
6016: \frac{\partial^2 Z_{j;c,0,
6017: v_1+v(c)}^0}{\partial \theta^2}(1)= 
6018: & \
6019: c\cdot \frac{\partial^2\Psi}{\partial 
6020: \theta^2}\left(e^{i\theta}\right)\cdot
6021: (1+v_j(c)), \ \ \ \ \ \ \ 
6022: j=2,\dots,n.
6023: \endaligned\right.
6024: \end{equation}
6025: Using the definition~\thetag{ 7.47}, the inequality
6026: $\vert v(c) \vert \leq c^{1+\alpha} \cdot K_8$ and the 
6027: second estimate~\thetag{7.36}, we deduce that
6028: \def\theequation{8.5}\begin{equation}
6029: \left\{
6030: \aligned
6031: \left\vert
6032: \frac{\partial^2 Z_{1;c,0,
6033: v_1+v(c)}^1}{\partial \theta^2}(1)\right\vert \leq
6034: & \
6035: c^{2+\alpha}\cdot K_7+c^{2+\alpha}\cdot C_2 K_8
6036: =: c^{2+\alpha} \cdot 2K_{10} 
6037: \\
6038: \left\vert
6039: \frac{\partial^2 Z_{j;c,0,
6040: v_1+v(c)}^1}{\partial \theta^2}(1)\right\vert \leq
6041: & \
6042: c\cdot 2C_2, 
6043: \ \ \ \ \ \ \
6044: j=2,\dots,n.
6045: \endaligned\right.
6046: \end{equation}
6047: Applying then Taylor's integral formula $F (\theta)=F(0) +\theta \cdot
6048: F'(0)+ \int_0^\theta \, (\theta- \theta') \cdot \partial_\theta
6049: \partial_\theta F(\theta') \cdot d \theta'$ to $F(\theta):= X_{1;c,
6050: 0,v_1 +v(c)}^1\left(e^{i\theta}\right)$ and afterwards to $F(\theta):=
6051: X_{j;c, 0,v_1 +v(c)}^1\left(e^{i\theta}\right)$ for $j=2,\dots,n$,
6052: taking account of the tangency conditions
6053: \def\theequation{8.6}\begin{equation}
6054: \frac{\partial X_{1;c,0,v_1+v(c)}^1}{\partial \theta}(1)=0, 
6055: \ \ \ \ \ \ \ \ \ \ \ \ 
6056: \frac{\partial X_{j;c,0,v_1+v(c)}^1}{\partial \theta}(1)=
6057: c\cdot C_1, \ \ \ \ \ \ \ 
6058: j=2,\dots,n,
6059: \end{equation}
6060: (a simple rephrasing of~\thetag{7.46}) 
6061: and using the inequalities~\thetag{8.5}, 
6062: we deduce that
6063: \def\theequation{8.7}\begin{equation}
6064: \left\{
6065: \aligned
6066: \left\vert
6067: X_{1;c,0,v_1+v(c)}^1\left(e^{i\theta}\right)
6068: \right\vert \leq
6069: & \
6070: \theta^2\cdot c^{2+\alpha} \cdot K_{10}, 
6071: \\
6072: \left\vert
6073: X_{j;c,0,v_1+v(c)}^1\left(e^{i\theta}\right) -\theta \cdot c \cdot C_1
6074: \right\vert \leq
6075: & \
6076: \theta^2 \cdot
6077: c\cdot C_2, 
6078: \ \ \ \ \ 
6079: j=2,\dots,n.
6080: \endaligned\right.
6081: \end{equation}
6082: Recall that
6083: \def\theequation{8.8}\begin{equation}
6084: x_1>\widetilde{ g}(x'):= -x_2^2-\cdots-x_n^2+K_9\left(
6085: \sum_{j=2}^n\, x_j^2
6086: \right)^{\frac{2+\alpha}{2}}
6087: \end{equation}
6088: denotes the equation of $(\widetilde{ H}^1)^+$. We now claim that if
6089: $c_1$ is sufficiently small, then for every $\theta$ with $0<\vert
6090: \theta \vert < 10c$, we have
6091: \def\theequation{8.9}\begin{equation}
6092: X_{1;c,0,v_1+v(c)}^1\left(e^{i\theta}\right) >
6093: \widetilde{ g}\left(
6094: X_{2;c,0,v_1+v(c)}^1\left(e^{i\theta}\right),\dots\dots,
6095: X_{n; c,0,v_1+v(c)}^1\left(e^{i\theta}\right)
6096: \right).
6097: \end{equation}
6098: Since $\Phi_c( \partial^+\Delta)$ is 
6099: contained in $\{e^{ i\theta}\in \partial^+
6100: \Delta: \ \vert \theta \vert <
6101: 10c\}$, this will imply the desired inclusion
6102: for proving {\bf ($\mathbf{ 7_1}$)}:
6103: \def\theequation{8.10}\begin{equation}
6104: \left\{
6105: \aligned
6106: A_{x,v:c}^1(\partial^+\Delta\backslash \{1\})=
6107: & \
6108: Z_{c,0,v_1+v(c)}^1\left(
6109: \Phi_c(\partial^+\Delta\backslash\{1\})
6110: \right)\subset \\
6111: \subset 
6112: & \
6113: Z_{c,0,v_1+v(c)}^1\left(
6114: \{e^{i\theta}\in \partial^+\Delta: \ 0 <\vert \theta \vert \leq 10c\}
6115: \right)
6116: \subset (\widetilde{ H}^1)^+.
6117: \endaligned\right.
6118: \end{equation}
6119: To prove the claim, we notice a minoration 
6120: of the left hand side of~\thetag{8.9}, using~\thetag{8.7}
6121: \def\theequation{8.11}\begin{equation}
6122: X_{1;c,0,v_1+v(c)}^1\left(e^{i\theta}\right) 
6123: \geq -\theta^2\cdot c^{2+\alpha}
6124: \cdot K_{10}.
6125: \end{equation}
6126: On the other hand, using two
6127: inequalities which are direct consequences of
6128: the second line of~\thetag{8.7}, provided that 
6129: $10c_1\cdot C_2\leq \frac{C_1}{2}$:
6130: \def\theequation{8.12}\begin{equation}
6131: \left\{
6132: \aligned
6133: \left\vert
6134: X_{j;c,0,v_1+v(c)}^1\left(e^{i\theta}\right)
6135: \right \vert \leq 
6136: & \
6137: \vert \theta\vert \cdot c \cdot \left(
6138: C_1+ \vert \theta \vert \cdot C_2\right)
6139: \leq \vert \theta \vert 
6140: \cdot c\cdot \frac{ 3C_1}{2}, \\
6141: \left[
6142: X_{j;c,0,v_1+v(c)}^1
6143: \right]^2\geq
6144: & \
6145: \theta^2\cdot c^2\cdot 
6146: (C_1-\vert \theta\vert \cdot C_2)^2 \geq
6147: \theta^2\cdot c^2 \cdot
6148: \frac{ C_1^2}{4}, 
6149: \endaligned\right.
6150: \end{equation}
6151: for $j=2,\dots,n$, we deduce the following
6152: majoration of the 
6153: right hand side of~\thetag{8.9}
6154: \def\theequation{8.13}\begin{equation}
6155: \left\{
6156: \aligned
6157: \widetilde{ g}
6158: & \
6159: \left(
6160: X_{2;c,0,v_1+v(c)}^1\left(e^{i\theta}\right),\dots\dots,
6161: X_{n;c,0,v_1+v(c)}^1\left(e^{i\theta}\right)
6162: \right)= \\
6163: & \
6164: =-\sum_{j=2}^n\, \left[
6165: X_{j;c,0,v_1+v(c)}^1
6166: \right]^2+K_9\left(
6167: \sum_{j=2}^n\, 
6168: \left[
6169: X_{j;c,0,v_1+v(c)}^1\left(e^{i\theta}\right)\right]^2
6170: \right)^{\frac{2+\alpha}{2}} \\
6171: & \
6172: \leq
6173: -\theta^2\cdot c^2 \cdot \frac{C_1^2}{4}
6174: (n-1)+
6175: \vert \theta \vert^{2+\alpha} \cdot c^{2+\alpha} \cdot
6176: \left(\frac{(n-1)9C_1^2}{4}\right)^{\frac{2+\alpha}{2}} K_9 \\
6177: & \
6178: \leq 
6179: -\theta^2\cdot c^2\left(
6180: \frac{ C_1^2}{4}(n-1)-
6181: c^\alpha\cdot \left(\frac{(n-1)9C_1^2}{4}\right)^{
6182: \frac{2+\alpha}{2}} K_9
6183: \right).
6184: \endaligned\right.
6185: \end{equation}
6186: Thanks to the minoration~\thetag{8.11} and to the
6187: majoration~\thetag{8.13}, in order that the inequality~\thetag{8.9}
6188: holds for all $\theta$ with $0< \vert \theta \vert \leq 10c$, it
6189: suffices that the right hand side of~\thetag{8.11} be greater than the
6190: last line of~\thetag{8.13}. By writing this strict inequality
6191: and clearing the factor $\theta^2\cdot c^2$, we see
6192: that it suffices that
6193: \def\theequation{8.14}\begin{equation}
6194: -K_{10}\cdot c^\alpha > -\left( \frac{ C_1^2}{4}
6195: (n-1)-c^\alpha\cdot
6196: \left(
6197: \frac{(n-1)9C_1^2}{4}
6198: \right)^{\frac{2+\alpha}{2}}K_9
6199: \right),
6200: \end{equation}
6201: or equivalently
6202: \def\theequation{8.15}\begin{equation}
6203: c_1 <
6204: \left(
6205: \frac{
6206: \frac{ C_1^2}{4}(n-1)
6207: }{
6208: K_{10}+\left(\frac{ (n-1)9C_1^2}{4}\right)^{
6209: \frac{2+\alpha}{2}} K_9
6210: }
6211: \right)^{\frac{1}{\alpha}}.
6212: \end{equation}
6213: This completes the proof of property 
6214: {\bf ($\mathbf{7_1}$)}.
6215: 
6216: Secondly, let us prove property {\bf ($\mathbf{ 8_1}$)} in Case {\bf
6217: ($\mathbf{I_1}$)}, proceeding similarly. As above, we come back to
6218: the definition $A_{x, 0 :c}^1 (\zeta):= Z_{c, x,v_1 +v(c)}^1 \left(
6219: \Phi_c (\zeta) \right)$ and we remind that $A_{x, 0: c}^1 (1)= Z_{c,
6220: x,v_1+ v(c)}^1(1)= x+ih(x)$, which follows by putting $d=1$ and
6221: $\zeta=1$ in~\thetag{ 7.18}. Thanks to the inclusion $\Phi_c(
6222: \partial^+ \Delta)\subset \{e^{i\theta}\in \partial^+\Delta : \ \vert
6223: \theta \vert < 10c\}$, it suffices to prove that the segment $Z_{c, x,
6224: v_1+v(c)}\left( \left\{ e^{i\theta}: \ \vert \theta \vert < 10c
6225: \right\} \right)$ is contained in the open side $( \widetilde{ H}^1)^+
6226: \subset (H^1 )^+$ defined by the inequation~\thetag{8.8}, if the point
6227: $x+ i h(x)$ belongs to the transversal half-submanifold $T^1 \cap
6228: (H^1)^+$, namely if $x=(x_1, 0, \dots,0)$ with $x_1>0$. In the sequel,
6229: we shall denote the disc $Z_{c,x, v_1+ v(c)}^1(\zeta)$ by $Z_{c, x_1,
6230: x',v_1+ v(c) }^1( \zeta)$, emphasizing the decomposition $x= (x_1,
6231: x')\in \R \times \R^{n-1}$, and we shall also use the convenient
6232: notation
6233: \def\theequation{8.16}\begin{equation}
6234: Z_{c,x_1,x',v_1+v(c)}^{'1}\left(\rho e^{i\theta}\right):=
6235: \left(
6236: Z_{2;c,x_1,x',v_1+v(c)}^1\left(\rho e^{i\theta}\right),\dots\dots, 
6237: Z_{n;c,x_1,x',v_1+v(c)}^1\left(\rho e^{i\theta}\right)
6238: \right).
6239: \end{equation}
6240: So, we have to show that for all $c$ with $0 < c \leq c_1$, for all
6241: $x_1$ with $0 < x_1 \leq c^2$ and for all $\theta$ with $\vert \theta
6242: \vert < 10c$, then the following strict inequality holds true
6243: \def\theequation{8.17}\begin{equation}
6244: X_{1;c,x_1,0,v_1+v(c)}^1\left(e^{i\theta}\right) > 
6245: \widetilde{ g}\left(
6246: X_{c;x_1,0,v_1+v(c)}^{1'}\left(e^{i\theta}\right)
6247: \right),
6248: \end{equation}
6249: First of all, coming back to the family of discs attached to $M^0$, we
6250: see by differentiating~\thetag{ 7.8} twice with respect to $x_1$ that
6251: $\frac{\partial^2 Z_{ c,x_1,0,v_1+v(c)}^0}{\partial x_1^2}(\zeta)
6252: \equiv 0$. Next, by differentiating twice Bishop's equation~\thetag{
6253: 7.18} with respect to $x_1$ and by reasoning as in Lemma~7.34, we
6254: deduce the estimate
6255: \def\theequation{8.18}\begin{equation}
6256: \left\vert
6257: \left\vert
6258: \frac{\partial^2 Z_{c,x_1,0,v_1+v(c)}^1}{
6259: \partial x_1^2}
6260: \right\vert
6261: \right\vert_{\mathcal{ C}^\alpha(\partial \Delta)}
6262: \leq c^{2+\alpha} \cdot K_7,
6263: \end{equation}
6264: say, with the same constant $K_7>0$ as in Lemma~7.34, after enlarging
6265: it if necessary. Applying then Taylor's integral formula $F(x_1)=
6266: F(0) +x_1 \cdot \partial_{x_1}F(0)+ \int_0^{ x_1}(x_1- \widetilde{
6267: x}_1) \cdot \partial_{x_1} \partial_{ x_1} F( \widetilde{ x}_1) \cdot
6268: d\widetilde{ x}_1$ to the function $F(x_1 ):= X_{1;c, x_1,0, v_1+
6269: v(c)}^1 \left(e^{ i\theta}\right)$, we deduce the minoration
6270: \def\theequation{8.19}\begin{equation}
6271: X_{1;c,x_1,0,v_1+v(c)}^1\left(e^{i\theta}\right)\geq 
6272: X_{1;c,0,0,v_1+v(c)}^1\left(e^{i\theta}\right)+
6273: x_1 \cdot
6274: \frac{\partial X_{1;c,0,0,v_1+v(c)}^1}{
6275: \partial x_1}\left(e^{i\theta}\right)
6276: - x_1^2 \cdot c^{2+\alpha} \cdot \frac{K_7}{2}.
6277: \end{equation}
6278: On the other hand, by differentiating Bishop's equation~\thetag{7.18}
6279: with respect to $x_1$ at $x=0$, the derivative $\partial_{x_1} x$
6280: yields the vector $(1,0,\dots,0)$ and we obtain
6281: \def\theequation{8.20}\begin{equation}
6282: \small
6283: \left\{
6284: \aligned
6285: \frac{\partial X_{c,0,0,v_1+v(c)}}{\partial x_1}
6286: \left(e^{i\theta} \right)=
6287: & \
6288: -T_1\left[
6289: \sum_{l=1}^n\, 
6290: \frac{\partial h}{\partial x_l}
6291: \left(
6292: X_{c,0,0,v_1+v(c)}^1(\cdot)
6293: \right) \, 
6294: \frac{\partial X_{l;c,0,0,v_1+v(c)}^1}{\partial x_1}(\cdot)
6295: \right](e^{i\theta})+ \\
6296: & \ \
6297: +
6298: (1,0,\dots,0).
6299: \endaligned\right.
6300: \end{equation}
6301: Using then the second inequality~\thetag{ 6.13} and
6302: the estimate~\thetag{ 7.25}, we deduce from~\thetag{ 8.20}
6303: \def\theequation{8.21}\begin{equation}
6304: \left\{
6305: \aligned
6306: \left\vert\left\vert
6307: \frac{\partial X_{1;c,0,0,v_1+v(c)}^1}{\partial x_1}(\cdot)-1
6308: \right\vert\right\vert_{
6309: \mathcal{ C}^\alpha(\partial \Delta)}\leq
6310: & \
6311: c^{2+\alpha} \cdot \vert\vert\vert
6312: T_1 \vert \vert \vert_{\mathcal{ C}^\alpha(\partial \Delta)} 
6313: K_2 K_3, \\
6314: \left\vert\left\vert
6315: \frac{\partial X_{j;c,0,0,v_1+v(c)}^1}{\partial x_1}(\cdot)
6316: \right\vert\right\vert_{
6317: \mathcal{ C}^\alpha(\partial \Delta)}\leq
6318: & \
6319: c^{2+\alpha} \cdot \vert\vert\vert
6320: T_1 \vert \vert \vert_{\mathcal{ C}^\alpha(\partial \Delta)} 
6321: K_2 K_3, \ \ \ \ \ \ \ 
6322: j=2,\dots,n.
6323: \endaligned\right.
6324: \end{equation}
6325: Thanks to the first line of~\thetag{8.21}, we can refine the
6326: minoration~\thetag{8.19} by replacing the first order partial
6327: derivative $\frac{ \partial X_{1;c, 0,0, v_1+v(c)}^1 }{ \partial x_1
6328: }\left (e^{i \theta} \right)$ in the right hand side of~\thetag{8.19}
6329: by the constant $1$, modulo an error term and also, we can use the
6330: trivial minoration $-x_1^2\geq -x_1$, which yields a new, more
6331: interesting minoration of the form
6332: \def\theequation{8.22}\begin{equation}
6333: X_{1;c,x_1,0,v_1+v(c)}^1\left(e^{i\theta}\right)\geq 
6334: X_{1;c,0,0,v_1+v(c)}^1\left(e^{i\theta}\right)+
6335: x_1 - x_1 \cdot c^{2+\alpha} \cdot K_{11}, 
6336: \end{equation}
6337: for some constant $K_{ 11}>0$. On the other hand, using the
6338: inequalities $\vert \partial_{ x_j} \widetilde{ g} (x') \vert \leq
6339: \vert x' \vert+ K_9 \cdot \vert x' \vert^{1+\alpha} \cdot \left(1+
6340: \frac{ \alpha }{2}\right) (n-1)^{ \frac{ \alpha }{2}}$ for $j=
6341: 2,\dots,n$, using the estimate~\thetag{ 7.25} and using~\thetag{ 6.2},
6342: we deduce an inequality of the form
6343: \def\theequation{8.23}\begin{equation}
6344: \widetilde{ g}\left(
6345: X_{c,x_1,0,v_1+v(c)}^{'1}\left(e^{i\theta}\right)
6346: \right)\leq 
6347: \widetilde{ g}\left(
6348: X_{c,0,0,v_1+v(c)}^{'1}\left(e^{i\theta}\right)
6349: \right)+ x_1 \cdot c \cdot K_{12}, 
6350: \end{equation}
6351: for some constant $K_{12}>0$. Finally, putting together the two
6352: inequalities~\thetag{ 8.22} and~\thetag{ 8.23}, and using the
6353: following inequality, which is an immediate consequence
6354: of the strict inequality~\thetag{ 8.9}:
6355: \def\theequation{8.24}\begin{equation}
6356: X_{1;c,0,0,v_1+v(c)}\left(e^{i\theta}\right)
6357: \geq \widetilde{ g}\left(
6358: X_{c,0,0,v_1+v(c)}^{'1}\left(
6359: e^{i\theta}
6360: \right)
6361: \right),
6362: \end{equation}
6363: valuable for all $\theta$ with $\vert \theta \vert < 10c$, we deduce
6364: the desired inequality~\thetag{ 8.17} as follows:
6365: \def\theequation{8.25}\begin{equation}
6366: \left\{
6367: \aligned
6368: X_{1;c,x_1,0,v_1+v(c)}^1\left(e^{i\theta} \right) 
6369: \geq 
6370: & \
6371: X_{1;c,0,0,v_1+v(c)}^1\left(e^{i\theta} \right)+
6372: x_1-x_1\cdot c^{2+\alpha} \cdot K_{11} \\
6373: \geq 
6374: & \ \widetilde{ g}\left(
6375: X_{c,0,0,v_1+v(c)}^{'1}\left(
6376: e^{i\theta}
6377: \right)
6378: \right)+ x_1-x_1\cdot c^{2+\alpha} \cdot K_{11} \\
6379: \geq 
6380: & \
6381: \widetilde{ g}\left(
6382: X_{c,x_1,0,v_1+v(c)}^{'1}\left(
6383: e^{i\theta}
6384: \right)
6385: \right)+ x_1-x_1\cdot c \cdot K_{11}-x_1\cdot c \cdot K_{12} \\
6386: > 
6387: & \
6388: \widetilde{ g}\left(
6389: X_{c,x_1,0,v_1+v(c)}^{'1}\left(
6390: e^{i\theta}
6391: \right)
6392: \right),
6393: \endaligned\right.
6394: \end{equation}
6395: for all $x_1$ with $0< x_1 \leq c^2$, all $\theta$ 
6396: with $\vert \theta \vert < 10c$ and all $c$ with 
6397: $0 < c \leq c_1$, 
6398: provided
6399: \def\theequation{8.26}\begin{equation}
6400: c_1 \leq \frac{ 1/2}{K_{11}+K_{12}}.
6401: \end{equation}
6402: This completes the proof of property {\bf ($\mathbf{ 8_1}$)}.
6403: 
6404: Thirdly, let us prove property {\bf ($\mathbf{ 9_1}$)} in Case {\bf
6405: ($\mathbf{I_1}$)}. The half-wedge $\mathcal{ HW}_1^+$ is defined by
6406: the $n$ inequalities of the last two lines of~\thetag{5.38}, where
6407: $a_2+\cdots+a_n=1$. For notational reasons, it will be convenient to
6408: set $a_1:=1$ and to write the first inequality defining $\mathcal{
6409: HW}_1^+$ simply as $\sum_{j=1}^n\, a_j y_j> \psi(x,y')$.
6410: 
6411: Because $\Phi_c \left( \overline{ \Delta} \backslash \partial^+ \Delta
6412: \right)$ is contained in the open sector $\{\rho e^{i \theta} \in
6413: \overline{ \Delta}: \ \vert \theta \vert < 10c, \ 1-10c < \rho < 1$,
6414: taking account of the definition~\thetag{7.53} of $A_{x,v
6415: :c}^1(\zeta)$, in order to check property {\bf ($\mathbf{ 9_1}$)}, it
6416: clearly suffices to show that $Z_{c,x, v_1+ v(c)+ v}^1 \left( \left\{
6417: \rho e^{i\theta}\in \Delta : \ 1-10c < \rho < 1, \ \vert \theta \vert
6418: < 10c \right\} \right)$ is contained in $\mathcal{ HW}_1^+$, which
6419: amounts to establish that for all $x$ with $\vert x \vert \leq c^2$,
6420: all $v$ with $\vert v \vert \leq c$, all $\rho e^{i \theta}$ with
6421: $1-10c < \rho < 1$ and with $\vert \theta \vert < 10c$, the following
6422: two collections of strict inequalities hold true
6423: \def\theequation{8.27}\begin{equation}
6424: \left\{
6425: \aligned
6426: \sum_{k=1}^n\, a_k 
6427: Y_{j;c,x,v_1+v(c)+v}^1\left(\rho e^{i\theta}\right) 
6428: > 
6429: & \
6430: \psi\left(
6431: X_{c,x,v_1+v(c)+v}^1\left(\rho e^{i\theta}\right), \
6432: Y_{c,x,v_1+v(c)+v}^{'1}\left(\rho e^{i\theta}\right)
6433: \right), \\
6434: Y_{j; c,x,v_1+v(c)+v}^1\left(
6435: \rho e^{i\theta}
6436: \right) > 
6437: & \
6438: \varphi_j\left(
6439: X_{c,x,v_1+v(c)+v}^1\left(
6440: \rho e^{i\theta}
6441: \right), \ 
6442: Y_{1;c,x,v_1+v(c)+v}^1\left(
6443: \rho e^{i\theta}
6444: \right)
6445: \right),
6446: \endaligned\right.
6447: \end{equation}
6448: for $j=2,\dots,n$, 
6449: provided $c_1$ is sufficiently small, where we use the
6450: notation~\thetag{ 8.16}.
6451: 
6452: We first treat the collection of $(n-1)$ strict inequalities
6453: in the second line of~\thetag{ 8.27}.
6454: First of all, by differentiating~\thetag{ 7.8}
6455: twice with respect to $\theta$, we obtain
6456: \def\theequation{8.28}\begin{equation}
6457: \frac{\partial^2 Z_{c,x,v_1+v(c)+v}^0}{\partial \theta^2}\left(
6458: e^{i\theta}
6459: \right)= c\cdot \frac{
6460: \partial^2\Psi}{\partial \theta^2}
6461: \left(
6462: e^{i\theta}
6463: \right) \cdot
6464: \left[
6465: v_1+v(c)+v
6466: \right].
6467: \end{equation}
6468: Using the second estimate~\thetag{7.36}, we deduce that
6469: there exists a constant $K_{13}>0$ such that
6470: \def\theequation{8.29}\begin{equation}
6471: \left\vert
6472: \frac{\partial^2 Z_{c,x,v_1+v(c)+v}^1}{
6473: \partial \theta^2}\left(e^{i\theta}\right)
6474: \right\vert\leq c\cdot K_{13}.
6475: \end{equation}
6476: Using the inequality~\thetag{ 6.2}, using~\thetag{ 8.29}, and then
6477: taking account of the inequalities $\vert \theta \vert < 10c$, $\vert
6478: x \vert \leq c^2$ and $\vert v \vert < c$, we deduce the following
6479: inequality
6480: \def\theequation{8.30}\begin{equation}
6481: \left\{
6482: \aligned
6483: \left\vert
6484: \frac{\partial Z_{c,x,v_1+v(c)+v}^1}{\partial \theta}
6485: \left( e^{i\theta}
6486: \right)-
6487: \frac{\partial Z_{c,0,v_1+v(c)}^1}{\partial \theta} (1)
6488: \right\vert
6489: \leq 
6490: & \
6491: c \cdot \left(
6492: \vert \theta \vert + \vert x \vert + \vert v \vert
6493: \right) \\
6494: \leq 
6495: & \
6496: c^2\cdot K_{14},
6497: \endaligned\right.
6498: \end{equation}
6499: for some constant $K_{14} >0$. On the other hand, by
6500: differentiating~\thetag{ 7.8} with respect to $\theta$ at $\theta=0$
6501: and applying the inequality~\thetag{ 7.28}, we obtain
6502: \def\theequation{8.31}\begin{equation}
6503: \left\vert
6504: \frac{\partial Z_{c,0,v_1+v(c)}^1}{\partial \theta} (1)
6505: -c\cdot C_1\cdot (0,1,\dots,1)
6506: \right\vert\leq c^{2+\alpha}\cdot K_6, 
6507: \end{equation}
6508: where $C_1=\frac{ \partial \Psi }{ \partial \theta}(1)$, as defined
6509: in~\thetag{7.47}. We remind that for every $\mathcal{ C}^1$-smooth
6510: function $Z$ on $\overline{ \Delta}$ which is holomorphic in $\Delta$,
6511: we have $i \frac{ \partial }{ \partial \theta} Z ( e^{i\theta})=
6512: -\frac{ \partial }{\partial \rho} Z( e^{i\theta})$. Consequently, we
6513: deduce from~\thetag{ 8.30} the following first (among three)
6514: interesting inequality
6515: \def\theequation{8.32}\begin{equation}
6516: \left\vert
6517: -\frac{\partial Z_{c,x,v_1+v(c)+v}^1}{\partial \rho}
6518: \left(
6519: e^{i\theta}
6520: \right)- c\cdot C_1 \cdot (0,i,\dots,i)
6521: \right\vert\leq 
6522: c^2 \cdot K_{15}, 
6523: \end{equation}
6524: for some constant $K_{15}>0$.
6525: 
6526: According to the definition~\thetag{ 7.8}, 
6527: we may compute 
6528: \def\theequation{8.33}\begin{equation}
6529: \frac{ \partial^2 Z_{c,x,v_1+v(c)+v}^0}{\partial \rho^2}
6530: \left(
6531: \rho e^{i\theta}\right)= c\cdot 
6532: \frac{\partial^2 \Psi}{\partial \rho^2}
6533: \left(
6534: \rho e^{i\theta} 
6535: \right)\cdot
6536: (v_1+v(c)+v)
6537: \end{equation}
6538: By reasoning as in the proof of Lemma~7.34, we may obtain an
6539: inequality similar to~\thetag{ 7.36}, with the second order partial
6540: derivative $\partial^2/ \partial \theta^2$ replaced by the second
6541: order partial derivative $\partial^2 / \partial \rho^2$. Putting this
6542: together with~\thetag{ 8.33}, we deduce that there exists a constant
6543: $K_{16} >0$ such that
6544: \def\theequation{8.34}\begin{equation}
6545: \left\vert
6546: \frac{\partial^2 Z_{c,x,v_1+
6547: v(c)+v}^1}{\partial \rho^2}
6548: \left(
6549: \rho e^{i\theta}
6550: \right)
6551: \right\vert\leq 
6552: c \cdot 2K_{16},
6553: \end{equation}
6554: for some constant $K_{16}>0$. Applying then Taylor's integral formula
6555: $F (\rho) = F(1) + (\rho -1) \cdot \partial_\rho F(1)+ \int_1^\rho (
6556: \rho- \widetilde{ \rho})\cdot \partial_\rho \partial_\rho F(
6557: \widetilde{ \rho}) \cdot d\widetilde{ \rho}$ to the functions $F(
6558: \rho):= Y_{k;c, x,v_1 +v(c) +v}^1\left( \rho e^{ i\theta} \right)$ for
6559: $k=1,\dots,n$, we deduce the second interesting collection of
6560: inequalities
6561: \def\theequation{8.35}\begin{equation}
6562: \left\{
6563: \aligned
6564: {}
6565: &
6566: \left\vert
6567: Y_{k;c,x,v_1+v(c)+v}^1
6568: \left(
6569: \rho e^{i\theta}
6570: \right)
6571: -Y_{k;c,x,v_1+v(c)+v}^1\left(
6572: e^{i\theta} 
6573: \right) - \right. \\
6574: & \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ 
6575: \left.
6576: -(\rho-1)\cdot
6577: \frac{\partial Y_{k;c,x,v_1+v(c)+v}^1}{\partial \rho }
6578: \left(
6579: e^{i\theta}
6580: \right)
6581: \right\vert 
6582: \leq (1-\rho)^2 \cdot c \cdot K_{16},
6583: \endaligned\right.
6584: \end{equation}
6585: for $k=1,\dots,n$.
6586: 
6587: On the other hand, thanks to the normalizations of the functions
6588: $\varphi_j (x, y_1)$ given in~\thetag{ 5.40}, namely $\varphi_j (0)=
6589: \partial_{ x_k} \varphi_j (0)= \partial_{ y_1} \varphi_j (0)=0$, $j=2,
6590: \dots, n$, $k=1, \dots, n$, we see that, possibly after increasing the
6591: constant $K_1 >0$ of Lemma~6.4, we have inequalities of the form
6592: \def\theequation{8.36}\begin{equation}
6593: \left\{
6594: \aligned
6595: {}
6596: &
6597: \sum_{k=1}^n\, 
6598: \left\vert
6599: \varphi_{j,x_k}(x,y_1)
6600: \right\vert+
6601: \left\vert
6602: \varphi_{j,y_1}(x,y_1)
6603: \right\vert\leq (\vert x \vert 
6604: + \vert y_1 \vert) \cdot
6605: K_1, \\
6606: &
6607: \left\vert \varphi_j(x,y_1) - 
6608: \varphi_j\left(\widetilde{ x}, 
6609: \widetilde{ y}_1\right)
6610: \right\vert \leq 
6611: \left(
6612: \left\vert 
6613: x-\widetilde{ x}
6614: \right\vert+
6615: \left\vert
6616: y_1- \widetilde{ y}_1 
6617: \right\vert
6618: \right)\cdot
6619: \left(
6620: \sum_{k=1}^n\, 
6621: \sup_{\vert x \vert, \, 
6622: \vert y_1 \vert \leq c\cdot K_2}
6623: \left\vert
6624: \varphi_{j,x_k}(x,y_1)
6625: \right\vert+ \right. \\
6626: & \ \ \ \ \ \ \ \ \ \ \ \ \ \ \
6627: \left. 
6628: +\sup_{\vert x \vert, \, 
6629: \vert y_1 \vert \leq c\cdot K_2}
6630: \left\vert
6631: \varphi_{j,y_1}(x,y_1)
6632: \right\vert
6633: \right),
6634: \endaligned\right.
6635: \end{equation}
6636: for $j=2, \dots, n$, provided $\vert x\vert, \, \vert \widetilde{ x}
6637: \vert, \, \vert y_1 \vert, \, \vert \widetilde{ y_1} \vert \leq c\cdot
6638: K_2$. On the other hand, computing $\frac{ \partial Z_{c,x, v_1+
6639: v(c)+v}^0}{\partial \rho} \left( \rho e^{ i\theta} \right)$
6640: in~\thetag{ 7.8}, using~\thetag{ 7.25}, \thetag{ 7.28} and an
6641: inequality of the form~\thetag{ 6.2}, we deduce that there exists a
6642: constant $K_{17}>0$ such that
6643: \def\theequation{8.37}\begin{equation}
6644: \left\vert
6645: Z_{c,x,v_1+v(c)+v}^1\left(
6646: \rho e^{i\theta} 
6647: \right)-
6648: Z_{c,x,v_1+v(c)+v}^1\left(
6649: e^{i\theta} 
6650: \right) 
6651: \right\vert\leq
6652: (1-\rho) \cdot c \cdot K_{17}.
6653: \end{equation}
6654: Finally, using the inequality $\vert Z_{ c,x, v_1+ v(c)+v}^1\left(
6655: \rho e^{i \theta} \right) \vert \leq c \cdot K_2$ obtained in~\thetag{
6656: 7.25}, using the collection of inequalities~\thetag{ 8.36} and using
6657: the inequality~\thetag{ 8.37}, we may deduce the third (and last)
6658: interesting inequality for $j=2, \dots, n$:
6659: \def\theequation{8.38}\begin{equation}
6660: \left\{
6661: \aligned
6662: {}
6663: &
6664: \left\vert
6665: \varphi_j\left(
6666: X_{c,x, v_1+v(c)+v}^1\left(
6667: \rho e^{i\theta}
6668: \right), \
6669: Y_{1;c,x, v_1+v(c)+v}^1
6670: \left(\rho e^{i\theta} \right)
6671: \right)- \right. \\
6672: &
6673: \ \ \ \ \ \ \ \ \ \ \ \ \ 
6674: \left.
6675: -\varphi_j\left(
6676: X_{c,x, v_1+v(c)+v}^1\left(
6677: e^{i\theta}
6678: \right), \
6679: Y_{1;c,x, v_1+v(c)+v}^1
6680: \left(e^{i\theta} \right)
6681: \right)
6682: \right\vert\leq \\
6683: &
6684: \left(
6685: \left\vert
6686: X_{c,x,v_1+v(c)+v}^1\left(
6687: \rho e^{i\theta}
6688: \right)- X_{c,x,v_1+v(c)+v}^1\left(
6689: e^{i\theta}\right)
6690: \right\vert+ \right. \\
6691: & \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \
6692: \left.
6693: +\left\vert
6694: Y_{1;c,x, v_1+v(c)+v}^1\left(
6695: \rho e^{i\theta}
6696: \right)- Y_{1;c,x, v_1+v(c)+v}^1\left(
6697: e^{i\theta}\right)
6698: \right\vert
6699: \right)\cdot \\
6700: & \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \
6701: \cdot 
6702: \left(
6703: \sum_{k=1}^n\, 
6704: \sup_{\vert x \vert, \, 
6705: \vert y_1 \vert \leq c\cdot K_2}
6706: \left\vert
6707: \varphi_{j,x_k}(x,y_1)
6708: \right\vert
6709: +\sup_{\vert x \vert, \, 
6710: \vert y_1 \vert \leq c\cdot K_2}
6711: \left\vert
6712: \varphi_{j,y_1}(x,y_1)
6713: \right\vert
6714: \right) \leq \\
6715: & 
6716: \leq (1-\rho) \cdot c^2 \cdot K_{18}, 
6717: \endaligned\right.
6718: \end{equation}
6719: for some constant $K_{18}>0$.
6720: 
6721: We can now complete the proof of the collection of inequalities in the
6722: second line of~\thetag{ 8.27}. As before, 
6723: let $c$ with $0 < c \leq c_1$, let $\rho$ with $10c < \rho
6724: < 1$, let $\theta$ with $\vert \theta \vert < 10 c$, 
6725: let $x$ with $\vert x \vert \leq c^2$, let $v$ with
6726: $\vert v \vert \leq c$ and let $j=2, \dots, n$. Starting
6727: with~\thetag{ 8.35}, using~\thetag{ 8.32}, using the fact that
6728: $Z_{c,x,v_1+ v(c)+v}^1 \left( \partial^+\Delta \right) \subset M^1
6729: \subset M$ and using~\thetag{ 8.38}, we have
6730: \def\theequation{8.39}\begin{equation}
6731: \left\{
6732: \aligned
6733: {}
6734: &
6735: Y_{j;c,x,v_1+v(c)+v}^1\left(
6736: \rho e^{i\theta} 
6737: \right) \geq \\
6738: & \ \ \ \
6739: \geq Y_{j;c,x,v_1+v(c)+v}^1
6740: \left(
6741: e^{i\theta} 
6742: \right)+ \\
6743: & \ \ \ \ \ \ \ 
6744: +(\rho-1) \cdot 
6745: \frac{\partial Y_{j;c,x,
6746: v_1+v(c)+v}^1}{\partial
6747: \rho}\left(
6748: e^{i\theta}
6749: \right)- (1-\rho)^2\cdot c \cdot K_{16} \geq \\
6750: & \ \ \ \ \
6751: \geq 
6752: Y_{j;c,x,v_1+v(c)+v}^1
6753: \left(
6754: e^{i\theta} 
6755: \right)+
6756: (1-\rho)\cdot
6757: c\cdot C_1
6758: -(1-\rho) \cdot c^2 \cdot K_{15} - 
6759: (1-\rho)^2\cdot c \cdot K_{16} \\
6760: & \ \ \ \ \
6761: = 
6762: \varphi_j\left(
6763: X_{1;c,x,v_1+v(c)+v}^1\left(
6764: e^{i\theta} 
6765: \right), \
6766: Y_{1;c,x,v_1+v(c)+v}^1\left(
6767: e^{i\theta}
6768: \right)
6769: \right)+
6770: (1-\rho)\cdot
6771: c\cdot C_1 - \\
6772: & \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ 
6773: \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \
6774: \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \
6775: \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \
6776: -(1-\rho)\cdot c^2 
6777: \cdot K_{15} - (1-\rho)^2\cdot c \cdot K_{16} \\
6778: & \ \ \ \ \ 
6779: \geq 
6780: \varphi_j\left(
6781: X_{1;c,x,v_1+v(c)+v}^1\left(
6782: \rho e^{i\theta} 
6783: \right), \
6784: Y_{1;c,x,v_1+v(c)+v}^1\left(
6785: \rho e^{i\theta}
6786: \right)
6787: \right)+
6788: (1-\rho)\cdot
6789: c\cdot C_1 - \\
6790: & \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ 
6791: \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \
6792: \ \ \ \ \
6793: -(1-\rho) \cdot c^2 \cdot K_{15} - (1-\rho)^2\cdot c \cdot K_{16}-
6794: (1-\rho)\cdot c^2\cdot K_{18} \\
6795: & \ \ \ \ \ 
6796: \geq 
6797: \varphi_j\left(
6798: X_{1;c,x,v_1+v(c)+v}^1\left(
6799: \rho e^{i\theta} 
6800: \right), \
6801: Y_{1;c,x,v_1+v(c)+v}^1\left(
6802: \rho e^{i\theta}
6803: \right)
6804: \right)+
6805: (1-\rho)\cdot
6806: c\cdot \left[C_1 - \right.\\
6807: & \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ 
6808: \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \
6809: \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \
6810: \ \ \ \ \ \ \ \ \ \
6811: \left.
6812: -c \cdot K_{15} - 10c \cdot K_{16}-
6813: c\cdot K_{18}\right]\\
6814: & \ \ \ \ \
6815: \geq 
6816: \varphi_j\left(
6817: X_{1;c,x,v_1+v(c)+v}^1\left(
6818: \rho e^{i\theta} 
6819: \right), \
6820: Y_{1;c,x,v_1+v(c)+v}^1\left(
6821: \rho e^{i\theta}
6822: \right)
6823: \right)+
6824: (1-\rho)\cdot
6825: c\cdot \frac{ C_1}{2} \\
6826: & \ \ \ \ \ 
6827: > 
6828: \varphi_j\left(
6829: X_{1;c,x,v_1+v(c)+v}^1\left(
6830: \rho e^{i\theta} 
6831: \right), \
6832: Y_{1;c,x,v_1+v(c)+v}^1\left(
6833: \rho e^{i\theta}
6834: \right)
6835: \right),
6836: \endaligned\right.
6837: \end{equation}
6838: provided that
6839: \def\theequation{8.40}\begin{equation}
6840: c_1 \leq 
6841: \frac{ C_1/2}{K_{15}+10K_{16}+K_{18}}.
6842: \end{equation}
6843: This yields the collection of
6844: inequalities in the second line of~\thetag{ 8.27}.
6845: 
6846: For the first inequality~\thetag{ 8.27}, we proceed similarly. Recall
6847: that $v_1= (0,1, \dots,1)$, that $a_1=1$ and that $a_2+ \cdots+a_n=1$.
6848: Since $Z_{c, x, v_1+ v(c)+v }^1( \partial^+\Delta)\subset M^1 \subset
6849: N^1$, we have for all $\theta$ with $\vert \theta \vert \leq
6850: \frac{\pi}{2}$ the following relation
6851: \def\theequation{8.41}\begin{equation}
6852: \sum_{k=1}^n\, 
6853: a_k\, 
6854: Y_{k;c,x,v_1+v(c)+v}^1\left(
6855: e^{i\theta}
6856: \right)= \psi\left(
6857: X_{c,x,v_1+v(c)+v}^1\left(
6858: e^{i\theta}
6859: \right), \ 
6860: Y_{c,x,v_1+v(c)+v}^{'1} \left(
6861: e^{i\theta}\right)
6862: \right).
6863: \end{equation}
6864: Using that $\psi$ vanishes to order one at the origin by the
6865: normalization conditions~\thetag{ 5.40} and proceeding as in the
6866: previous paragraph concerning the functions $\varphi_j$, we obtain an
6867: inequality similar to~\thetag{ 8.38}:
6868: \def\theequation{8.42}\begin{equation}
6869: \left\{
6870: \aligned
6871: {}
6872: &
6873: \left\vert
6874: \psi\left(X_{c,x,v_1+v(c)+v}^1\left(
6875: \rho e^{i\theta} 
6876: \right), \
6877: Y_{c,x,v_1+v(c)+v}^{'1}\left(
6878: \rho e^{i\theta}
6879: \right)\right)- \right . \\
6880: & \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \
6881: \left.
6882: -\psi\left(X_{c,x,v_1+v(c)+v}^1\left(
6883: e^{i\theta} 
6884: \right), \
6885: Y_{c,x,v_1+v(c)+v}^{'1}\left(
6886: e^{i\theta}
6887: \right)\right)
6888: \right\vert\leq 
6889: (1-\rho)\cdot c^2 \cdot K_{19}, 
6890: \endaligned\right.
6891: \end{equation}
6892: for some constant $K_{19}>0$.
6893: 
6894: As before, let $c$ with $0 < c \leq c_1$, let $\rho$ with $10c < \rho
6895: < 1$, let $\theta$ with $\vert \theta \vert < 10 c$, let $x$ with
6896: $\vert x \vert \leq c^2$ and let $v$ with $\vert v \vert \leq c$. Using
6897: then~\thetag{ 8.35}, \thetag{ 8.32}, \thetag{ 8.41} and~\thetag{
6898: 8.42}, we deduce the desired strict inequality
6899: \def\theequation{8.43}\begin{equation}
6900: \small
6901: \left\{
6902: \aligned
6903: {}
6904: &
6905: \sum_{k=1}^n\, 
6906: a_k \, 
6907: Y_{k; c,x,v_1+v(c)+v}^1\left(
6908: \rho e^{i\theta}
6909: \right)\geq 
6910: \sum_{k=1}^n\, 
6911: a_k \, 
6912: Y_{k; c,x,v_1+v(c)+v}^1\left(
6913: e^{i\theta}
6914: \right) +\\
6915: & \ \ \ \ \ \ \ \ 
6916: +(1-\rho) \left[
6917: \sum_{k=1}^n\, 
6918: a_k\left(
6919: -\frac{\partial Y_{k;c,x,v_1+v(c)+v}^1}{\partial
6920: \rho}\left(
6921: e^{i\theta}
6922: \right)
6923: \right)
6924: \right]-
6925: (1-\rho)^2 \cdot c \cdot \left( \sum_{k=1}^n\, a_k\right)K_{16} \\
6926: & \ \ \ \ \ \ \ \ \ 
6927: \geq 
6928: \sum_{k=1}^n\, 
6929: a_k \, 
6930: Y_{k; c,x,v_1+v(c)+v}^1\left(
6931: e^{i\theta}
6932: \right) +
6933: (1-\rho) \left[
6934: \sum_{j=2}^n\,
6935: a_j\cdot 
6936: c\cdot C_1 - \sum_{k=1}^n\, a_k \cdot c^2 \cdot
6937: K_{15} 
6938: \right] - \\
6939: & \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ 
6940: \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \
6941: -(1-\rho)^2\cdot c \cdot 2K_{16} \\
6942: & \ \ \ \ \ \ \ \ \ 
6943: \geq 
6944: \sum_{k=1}^n\, 
6945: a_k \, 
6946: Y_{k; c,x,v_1+v(c)+v}^1\left(
6947: e^{i\theta}
6948: \right) +
6949: (1-\rho)\cdot c \cdot C_1- \\
6950: & \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \
6951: \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ 
6952: \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \
6953: -(1-\rho)\cdot c^2
6954: \cdot 2K_{15}- (1-\rho)^2 \cdot c \cdot 2K_{16} \\
6955: & \ \ \ \ \ \ \ \ \ 
6956: =
6957: \psi\left(
6958: X_{c,x,v_1+v(c)+v}^1\left(
6959: e^{i\theta}
6960: \right), \ 
6961: Y_{c,x,v_1+v(c)+v}^{'1} \left(
6962: e^{i\theta}\right)
6963: \right) + \\
6964: & \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \
6965: \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ 
6966: \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \
6967: + (1-\rho) \cdot c \cdot \left[
6968: C_1-c \cdot 2K_{15}- 10c \cdot 2K_{16}
6969: \right] \\
6970: & \ \ \ \ \ \ \ \ \ 
6971: \geq 
6972: \psi\left(
6973: X_{c,x,v_1+v(c)+v}^1\left(
6974: \rho e^{i\theta}
6975: \right), \ 
6976: Y_{c,x,v_1+v(c)+v}^{'1} \left(
6977: \rho e^{i\theta}\right)
6978: \right) + \\
6979: & \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \
6980: \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ 
6981: \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \
6982: + (1-\rho) \cdot c \cdot \left[
6983: C_1-c \cdot 2K_{15}- 10c \cdot 2K_{16}-c\cdot K_{19}
6984: \right] \\
6985: & \ \ \ \ \ \ \ \ \ 
6986: \geq 
6987: \psi\left(
6988: X_{c,x,v_1+v(c)+v}^1\left(
6989: \rho e^{i\theta}
6990: \right), \ 
6991: Y_{c,x,v_1+v(c)+v}^{'1} \left(
6992: \rho e^{i\theta}\right)
6993: \right) +
6994: (1-\rho) \cdot c \cdot \frac{ C_1}{2} \\
6995: & \ \ \ \ \ \ \ \ \ 
6996: >
6997: \psi\left(
6998: X_{c,x,v_1+v(c)+v}^1\left(
6999: \rho e^{i\theta}
7000: \right), \ 
7001: Y_{c,x,v_1+v(c)+v}^{'1} \left(
7002: \rho e^{i\theta}\right)
7003: \right), 
7004: \endaligned\right.
7005: \end{equation}
7006: provided
7007: \def\theequation{8.44}\begin{equation}
7008: c_1\leq \frac{ C_1/2}{2K_{15}+20K_{16}+K_{19}}
7009: \end{equation}
7010: This yields the first inequality of~\thetag{ 8.27} and
7011: completes the proof of {\bf ($\mathbf{ 9_1}$)} in 
7012: Case {\bf ($\mathbf{ I_1}$)}.
7013: 
7014: The proof of Lemma~8.3 is complete, because by following the
7015: normalizations of Lemma~5.37 and by formulating analogous inequalities,
7016: cases {\bf ($\mathbf{ I_2}$)} and {\bf (II)} are achieved in a totally
7017: similar way.
7018: \endproof
7019: 
7020: \section*{\S9.~End of proof of Theorem~1.2': application
7021: of the continuity principle}
7022: 
7023: \subsection*{9.1.~Preliminary}
7024: In this section, we shall now complete the proof of Proposition~5.12
7025: (at last!), hence the proof of Theorem~3.19 {\bf (i)} (which is a
7026: direct consequence of Proposition~5.12, as was explained in Section~5)
7027: and hence also the proof of Theorem~1.2', modulo supplementary
7028: arguments postponed to \S9.27 below. By means of a deformation
7029: $A_{x,v,u:c}^1(\zeta)$ (we add a real parameter
7030: $u$) of the family of analytic discs
7031: $A_{x,v:c}^1(\zeta)$ satisfying properties {\bf ($\mathbf{ 1_1}$)} to
7032: {\bf ($\mathbf{ 9_1}$)} of Lemmas~7.12 and~8.3, and by means of the
7033: continuity principe, we shall show that, in Cases {\bf ($\mathbf{
7034: I_1}$)} and {\bf ($\mathbf{ I_2}$)}, there exists a local wedge
7035: $\mathcal{ W}_{p_1}$ of edge $M$ at $p_1$ to which all holomorphic
7036: functions in $\mathcal{ O}\left( \Omega \cup \mathcal{ HW}_1^+\right)$
7037: extend holomorphically and we shall show that in Case {\bf (II)},
7038: there exists a neighborhood $\omega_{p_1}$ of $p_1$ in $\C^n$ to which
7039: all holomorphic functions in $\mathcal{ O}\left( \Omega \cup \mathcal{
7040: W}_2\right)$ extend holomorphically. To organize this last main
7041: step of the proof of Proposition~5.12, we shall consider jointly
7042: Cases~{\bf ($\mathbf{ I_1}$)}, {\bf ($\mathbf{ I_2}$)} and then
7043: afterwards Case~{\bf (II)} separately in \S9.22 below.
7044: 
7045: \subsection*{9.2.~Isotopies of analytic
7046: discs and continuity principle} To begin with, we shall formulate a
7047: convenient version of the continuity principle. If $E\subset \C^n$ is
7048: an arbitrary subset, we denote by
7049: \def\theequation{9.3}\begin{equation}
7050: \mathcal{V}_{\C^n}(E,\rho)=\bigcup_{p\in E}\, 
7051: \{z\in \C^n: \vert z - p \vert < \rho\}
7052: \end{equation}
7053: the union of polydiscs of radius $\rho>0$ centered at points of $E$.
7054: We then have the following lemma, extracted from~\cite{m2}, which
7055: applies to families of analytic discs $A_\tau(\zeta)$ which are
7056: embeddings of $\overline{\Delta}$ into $\C^n$:
7057: 
7058: \def\thelemma{9.4}\begin{lemma}
7059: \text{\rm (\cite{m2}, Proposition~3.3)}
7060: Let $\mathcal{ D}$ be a nonempty domain in $\C^n$ and let $A_\tau:
7061: \overline{\Delta} \to \C^n$ be a one-parameter family of analytic
7062: discs, where $\tau\in\R$ satisfies $0\leq \tau \leq 1$. Assume that
7063: there exist constants $c_\tau$ and $C_\tau$ with $0 < c_\tau < C_\tau$
7064: such that 
7065: \def\theequation{9.5}\begin{equation} c_\tau \vert
7066: \zeta_1 - \zeta_2 \vert < \vert A_\tau(\zeta_1) 
7067: - A_\tau (\zeta_2) \vert <
7068: C_\tau \vert \zeta_1 - \zeta_2 \vert,
7069: \end{equation}
7070: for all distinct points $\zeta_1,\zeta_2 \in \overline{ \Delta}$ and
7071: all $0\leq \tau \leq 1$. Assume that $A_1(\overline{ \Delta} )
7072: \subset \mathcal{ D}$, set $\rho_\tau := \inf\{ \vert t - A_\tau
7073: (\zeta)\vert: \, t\in \partial \mathcal{ D},\, \zeta \in \partial
7074: \Delta\}$, namely $\rho_\tau$ is the polydisc distance between $A_\tau
7075: (\partial \Delta)$ and $\partial \mathcal{ D}$, assume
7076: $\rho_\tau>0$ for all $\tau$ with $0\leq \tau \leq 1$,
7077: and set $\sigma_\tau
7078: := \rho_\tau c_\tau / 2C_\tau$. Then for every holomorphic function
7079: $f \in \mathcal{O} (\mathcal{ D})$, and for every $\tau\in [0, 1]$,
7080: there exists a holomorphic function $F_\tau \in \mathcal{ O}\left(
7081: \mathcal{ V}_{\C^n} (A_\tau\left( \overline{ \Delta}\right),
7082: \sigma_\tau)\right)$ such that $F_\tau =f$ in $\mathcal{ V}_{\C^n}
7083: (A_\tau ( \partial \Delta), \sigma_\tau)\subset \mathcal{ D}$.
7084: \end{lemma}
7085: 
7086: Two analytic discs $A', A'': \overline{\Delta} \to \C^n$ which are of
7087: class $\mathcal{ C}^1$ over $\overline{\Delta}$ and holomorphic in
7088: $\Delta$ and which are both {\it embeddings}\, of $\overline{\Delta}$
7089: into $\C^n$ are said to be {\sl analytically isotopic} if there exists
7090: a $\mathcal{ C}^1$-smooth family of analytic discs $A_\tau: \overline{
7091: \Delta}\to \C^n$ which are of class $\mathcal{ C}^1$ over
7092: $\overline{\Delta}$ and holomorphic in $\Delta$ such that $A_0=A'$,
7093: such that $A_1=A''$ and such that $A_\tau$ is an {\it embedding}\, of
7094: $\overline{ \Delta}$ into $\C^n$ for all $\tau$ with $0\leq
7095: \tau \leq 1$.
7096: 
7097: \subsection*{9.6.~Translations of $M^1$ in $M$}
7098: According to Lemma~5.37, in Case~{\bf ($\mathbf{ I_1}$)}, the
7099: one-codimensional submanifold $M^1\subset M$ is given by the equations
7100: $y'=\varphi'(x,y_1)$ and $x_1=g(x')$. If $u\in \R$ is a small real
7101: parameter, we may define a ``translation'' $M_u^1$ of $M^1$ in $M$ by
7102: the $n$ equations
7103: \def\theequation{9.7}\begin{equation}
7104: M_u^1: \ \ \ \ \ 
7105: y'=\varphi'(x,y_1), \ \ \ \ \ 
7106: x_1=g(x')+u.
7107: \end{equation}
7108: Clearly, we have $M_0^1 \equiv M^1$, we have $M_u^1 \subset (M^1 )^+$
7109: if $u>0$ and we have $M_u^1 \subset (M^1)^-$ if $u <0$. We may perturb
7110: the family of analytic discs $Z_{c, x,v}^d ( \zeta)$ attached to $M^1$
7111: and satisfying Bishop's equation~\thetag{ 7.18} by requiring that it
7112: is attached to $M_u^1$. We then obtain a new family of analytic discs
7113: $Z_{c,x, v,u}^d (\zeta)$ which is half-attached to $M_u^1$ and which
7114: is of class $\mathcal{ C}^{2,\alpha-0}$ with respect to all variables
7115: $(c,x, v, u, \zeta)$, thanks to the stability under perturbation of
7116: the solutions to Bishop's. For $u=0$, this solution coincides with the
7117: family $Z_{c, x, v}^d ( \zeta)$ constructed in \S7.13. Using a similar
7118: definition as in~\thetag{7.53}, namely setting $A_{x,v, u:c}^1
7119: (\zeta):= Z_{c,x, v_1+v(c) +v,u}^1 \left( \Phi_c (\zeta) \right)$, we
7120: obtain a new family of analytic discs which coincides, for $u=0$, with
7121: the family of analytic discs $A_{x, v: c}^1 (\zeta)$ of Lemmas~7.12
7122: and~8.3. Similarly, in Case~{\bf ($\mathbf{ I_2}$)}, taking account
7123: of the normalizations stated in Lemma~5.37, we can also construct an
7124: analogous family of analytic discs $A_{x,v,u:c}^1(\zeta)$. From now on,
7125: we shall fix the scaling parameter $c$ with $0 < c \leq c_1$, so that
7126: the nine properties {\bf ($\mathbf{ 1_1}$)} to {\bf ($\mathbf{ 9_1}$)}
7127: of Lemmas~7.12 and~8.3 are satisfied.
7128: 
7129: \subsection*{9.8.~Definition of a local wedge of edge $M$ at $p_1$
7130: in Cases~($\mathbf{ I_1}$) and~($\mathbf{I_2}$)} First of all, in
7131: Cases~{\bf ($\mathbf{I_1}$)} and~{\bf ($\mathbf{ I_2}$)}, we shall
7132: restrict the variation of the parameter $v$ to a certain
7133: $(n-2)$-dimensional linear subspace $V_2$ of $T_{p_1} \R^n \cong \R^n$
7134: as follows. By hypothesis, the vector $v_1$ does not belong to the
7135: characteristic direction $T_{p_1}M^1 \cap T_{p_1}^cM$, so the real
7136: vector space $\left( \R \cdot v_1 \right) \oplus \left( T_{p_1}M^1
7137: \cap T_{p_1 }^cM \right) \subset T_{ p_1} M^1$ is $2$-dimensional. We
7138: choose an arbitrary $(n-2)$-dimensional real vector subspace
7139: $V_2\subset T_{p_1}M^1$ which is a supplementary in $T_{p_1}M^1$ to
7140: $\left( \R\cdot v_1\right) \oplus \left( T_{p_1} M^1 \cap T_{ p_1}^cM
7141: \right)$ and we shall let the parameter $v$ vary
7142: only in $V_2$.
7143: 
7144: From the rank properties {\bf
7145: ($\mathbf{ 5_1}$)} and {\bf ($\mathbf{ 6_1}$)} of Lemma~7.12 and from
7146: the definition of $V_2$, we deduce that there exists $\varepsilon>0$
7147: small enough with $\varepsilon < < c^2$ such that the mapping
7148: \def\theequation{9.9}\begin{equation}
7149: (x,v,u,\rho)\longmapsto 
7150: A_{x,v,u:c}^1(\rho)
7151: \end{equation}
7152: is a {\it one-to-one immersion}\, from the open set $\{(x,v,u,\rho)\in
7153: \R^n\times V_2 \times \R\times \R: \ \vert x \vert < \varepsilon,
7154: \ \vert v \vert < \varepsilon, \ \vert u \vert < \varepsilon, \
7155: 1-\varepsilon < \rho < 1\}$ into $\C^n$. This property will be
7156: important for uniqueness of the holomorphic extension in our
7157: application of the continuity principle to be conducted in Lemma~9.20
7158: below. Furthermore, shrinking $\varepsilon >0$ if necessary, we can
7159: insure that the open subset
7160: \def\theequation{9.10}\begin{equation}
7161: \left\{
7162: \aligned
7163: {}
7164: &
7165: \mathcal{ W}_{p_1} :=\left\{
7166: A_{x,v,u:c}^1(\rho)\in \C^n: 
7167: (x,v,u,\rho)\in
7168: \R^n\times V_2 \times \R\times \R, \right. \\
7169: & \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \
7170: \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \
7171: \left.
7172: \vert x \vert < \varepsilon,
7173: \ \vert v \vert < \varepsilon, \ 
7174: \vert u \vert < \varepsilon, \
7175: 1-\varepsilon < \rho < 1
7176: \right\}
7177: \endaligned\right.
7178: \end{equation}
7179: is a local wedge of edge $M$ at $(p_1, Jv_1)$, with $\mathcal{ W}_{
7180: p_1} \cap M=\emptyset$.
7181: 
7182: Let the closed subset $C$ with $p_1 \in C$ and $C \backslash \{ p_1\}
7183: \subset (H^1)^-$, let the neighborhood $\Omega$ of $M \backslash C$ in
7184: $\C^n$, let the half-wedge $\mathcal{ HW }_{p_1 }^+$ be as in
7185: Proposition~5.12, and let the sub-half-wedge $\mathcal{ HW}_1^+\subset
7186: \mathcal{ HW}_{ p_1 }^+$ be as in \S5.14 and Lemma~5.37. 
7187: In Cases {\bf ($\mathbf{ I_1
7188: }$)} and {\bf ($\mathbf{ I_2}$)}, we shall consider the envelope of
7189: holomorphy of the open subset $\Omega \cup \mathcal{ HW }_1^+ $. We
7190: shall prove in the next paragraphs that, after possibly shrinking it a
7191: little bit, its envelope of holomorphy contains the wedge $\mathcal{ W}_{ p_1 }$.
7192: 
7193: \subsection*{9.11.~Boundaries of analytic discs}
7194: Since we want to apply the continuity principle Lemma~9.4, we have to
7195: show that most discs $A_{x, v, u:c}^1 ( \zeta)$ have their boundaries
7196: in $\Omega \cup \mathcal{ HW}_1^1+$. 
7197: To this aim, it will be useful to decompose the boundary
7198: $\partial \Delta$ in three closed parts $\partial \Delta = \partial^1
7199: \Delta \cup \partial^2 \Delta \cup \partial^3 \Delta$, where
7200: \def\theequation{9.12}\begin{equation}
7201: \left\{
7202: \aligned
7203: \partial^1\Delta := 
7204: & \
7205: \left\{
7206: e^{i\theta} \in \partial\Delta : \ 
7207: \vert \theta \vert \leq \pi/2 -\varepsilon
7208: \right\} \subset \partial^+\Delta, \\
7209: \partial^2\Delta := 
7210: & \
7211: \left\{
7212: e^{i\theta} \in \partial\Delta : \ 
7213: \pi/2 +\varepsilon \leq
7214: \vert \theta \vert \leq \pi
7215: \right\} \subset \partial^-\Delta, \\
7216: \partial^3\Delta := 
7217: & \
7218: \left\{
7219: e^{i\theta} \in \partial\Delta : \ 
7220: \pi/2-\varepsilon \leq 
7221: \vert \theta \vert \leq \pi/2 +\varepsilon
7222: \right\} \subset \partial\Delta, \\
7223: \endaligned\right.
7224: \end{equation}
7225: where $\varepsilon$ with $0< \varepsilon < < c^2$ is as in \S9.8 just
7226: above. This decomposition is illustrated in the left hand side of {\sc
7227: Figure~17} below. Next, we observe that the two points $A_{0,
7228: 0,0:c}^1(i)$ and $A_{ 0,0, 0:c}^1(-i)$ belong to $(H^1 )^-\subset M
7229: \backslash C\subset \Omega$, hence there exists a fixed open
7230: neighborhood of these two points which is contained in $\Omega$. We
7231: shall denote by $\omega^3$ such a (disconnected) neighborhood, for 
7232: instance the union of two small open polydiscs centered at 
7233: these two points. To
7234: proceed further, we need a crucial geometric information about the
7235: boundaries of the analytic discs $A_{x,v,u:c}^1 (\zeta)$ with $u\neq
7236: 0$.
7237: 
7238: \def\thelemma{9.13}\begin{lemma}
7239: Under the assumptions of Cases~{\bf ($\mathbf{ I_1}$)} and~{\bf
7240: ($\mathbf{ I_2}$)} of Proposition~5.12, after
7241: shrinking $\varepsilon >0$ if necessary, then 
7242: \def\theequation{9.14}\begin{equation}
7243: A_{x,v,u:c}^1\left(\partial \Delta
7244: \right)\subset \Omega \cup 
7245: \mathcal{ HW}_1^+,
7246: \end{equation}
7247: for all $x$ with $\vert x \vert < \varepsilon$, for all $v$ with
7248: $\vert v \vert < \varepsilon$ and for all {\rm nonzero} $u\neq 0$ with
7249: $\vert u \vert < \varepsilon$. 
7250: \end{lemma}
7251: 
7252: \proof 
7253: Firstly, since $A_{0, 0, 0:c}^1 (\pm i) \in \omega^3$, it
7254: follows just by continuity of the family $A_{x, v, u: c}^1(\zeta)$
7255: that, after possibly shrinking $\varepsilon >0$, the closed arc $A_{x,
7256: v,u: c}^1 \left( \partial^3 \Delta \right)$ is contained in
7257: $\omega^3$, for all $x$ with $\vert x \vert < \varepsilon$, for all
7258: $v$ with $\vert v \vert < \varepsilon$ and for all $u$ with $\vert u
7259: \vert < \varepsilon$. Secondly, since $A_{0, 0, 0:c}^1( \partial^2
7260: \Delta) \subset A_{0, 0, 0: c}^1 \left( \partial^-\Delta \backslash
7261: \{i,-i\}\right) \subset \mathcal{ HW }_1^+$, then by property {\bf
7262: ($\mathbf{ 9_1 }$)} of Lemma~8.3, it follows just by continuity of the
7263: family $A_{x, v, u:c}^1(\zeta)$ that, after possibly shrinking
7264: $\varepsilon >0$, the closed arc $A_{x, v, u:c}^1 \left( \partial^2
7265: \Delta \right)$ is contained in $\mathcal{ HW}_1^+$, for all $x$ with
7266: $\vert x \vert < \varepsilon$, for all $v$ with $\vert v \vert <
7267: \varepsilon$ and for all $u$ with $\vert u \vert < \varepsilon$.
7268: Thirdly, it follows from the inclusion $A_{x, v, u: c}^1 \left(
7269: \partial^1 \Delta \right) \subset A_{x, v,u:c}^1 \left( \partial^+
7270: \Delta \right) \subset M_u^1$ and from the inclusion $M_u^1 \subset
7271: \Omega$ for all $u\neq 0$ that, after possibly shrinking
7272: $\varepsilon>0$, the closed arc $A_{x, v, u:c}^1 \left( \partial^1
7273: \Delta \right)$ is contained in $\Omega$, for all $x$ with $\vert x
7274: \vert < \varepsilon$, for all $v$ with $\vert v \vert < \varepsilon$
7275: and for all $u$ with $\vert u \vert < \varepsilon$ and $u \neq 0$. 
7276: This completes the proof of Lemma~9.13.
7277: \endproof
7278: 
7279: \bigskip
7280: \begin{center}
7281: \input three-boundary.pstex_t
7282: \end{center}
7283: 
7284: In addition to this lemma, we notice that it follows immediately from
7285: properties {\bf ($\mathbf{8_1}$)} and {\bf ($\mathbf{9_1}$)} of
7286: Lemma~8.3 that $A_{x,0,0:c}^1(\overline{\Delta})$ is contained in
7287: $\Omega \cup \mathcal{ HW}_1^+$ for all $x$ with $\vert x \vert <
7288: \varepsilon$ such that $A_{x,0,0:c}^1(1)\in T^1\cap (H^1)^+$. This
7289: property and Lemma~9.13 are illustrated in the right hand side of {\sc
7290: Figure~17} just above.
7291: 
7292: \subsection*{9.15.~Analytic isotopies}
7293: Next, in Case {\bf ($\mathbf{ I_1}$)}, we fix some $x_0= (x_{1; 0},0,
7294: \dots, 0) \in \R^n$ with $0< x_{1;0} < \varepsilon$. Then $A_{x_0, 0,
7295: 0:c }^1 (1)= x_0+ i h(x_0)$ belongs to $T^1 \cap (
7296: H^1)^+$. Analogously, in Cases {\bf ($\mathbf{ I_2}$)}, we fix some
7297: $x_0=(0,\dots,0, x_{ n;0})\in \R^n$ with $0< x_{n;}0 <
7298: \varepsilon$. Then in this second case, the point $A_{x_0, 0,
7299: 0:c}^1(1)=x_0+ ih(x_0)$ also belongs to $T^1 \cap (H^1 )^+$. We fix
7300: the disc $A_{x_0, 0,0:c}^1( \zeta )$, which satisfies $A_{x_0,
7301: 0,0:c}^1( \overline{ \Delta })\subset \Omega \cup \mathcal{ HW 
7302: }_1^+$.
7303: 
7304: \def\thelemma{9.16}\begin{lemma}
7305: In Cases~{\bf ($\mathbf{ I_1})$} and~{\bf ($\mathbf{ I_2}$)}, every
7306: disc $A_{x, v, u:c}^1 (\zeta)$ with $\vert x \vert < \varepsilon$,
7307: $\vert v \vert < \varepsilon$, $\vert u \vert < \varepsilon$ and
7308: $u\neq 0$ is analytically isotopic to the disc $A_{x_0, 0,
7309: 0:c}^1(\zeta)$, with the boundaries of the analytic discs of the
7310: isotopy being all contained in $\Omega \cup \mathcal{ HW}_1^+$.
7311: \end{lemma}
7312: 
7313: \proof
7314: Indeed, since the set $\{u=0\}$ is a hyperplane, there clearly exists
7315: a $\mathcal{ C}^{2, \alpha- 0}$-smooth curve $\tau \mapsto (x (\tau),v
7316: (\tau),u (\tau))$ in the parameter space which joins a given arbitrary
7317: point $(x^*, v^*,u^*)$ with $u^*\neq 0$ to the point $(x_0, 0,0)$
7318: without meeting the hyperplane $\{u= 0\}$, except at its endpoint
7319: $(x_0, 0,0)$. According to the previous Lemma~9.13, each boundary
7320: $A_{x (\tau), v (\tau), u (\tau) :c}^1( \partial \Delta)$ is then
7321: automatically contained in $\Omega \cup \mathcal{ HW }_1^+$, which
7322: completes the proof.
7323: \endproof
7324: 
7325: \subsection*{9.17.~holomorphic extension 
7326: to a local wedge of edge $M$ at $p_1$} In Cases {\bf ($\mathbf{
7327: I_1}$)} and~{\bf ($\mathbf{ I_2}$)}, we define the
7328: following $\mathcal{ C}^{2,\alpha-0}$-smooth connected
7329: hypersurface of $\mathcal{ W}_{p_1}$:
7330: \def\theequation{9.18}\begin{equation}
7331: \left\{
7332: \aligned
7333: \mathcal{ M}_{p_1}:= 
7334: \left\{
7335: A_{x,v,0:c}^1(\rho): \ 
7336: (x,v,\rho)\in \R^n\times V_2 \times \R, \right. \\
7337: \left.
7338: \vert x \vert < \varepsilon, \ 
7339: \vert v \vert < \varepsilon, \
7340: 1-\varepsilon < \rho < \varepsilon
7341: \right\}, 
7342: \endaligned\right.
7343: \end{equation}
7344: together with the following closed subset of 
7345: $\mathcal{ M}_{p_1}$:
7346: \def\theequation{9.19}\begin{equation}
7347: \left\{
7348: \aligned
7349: \mathcal{ C}_{p_1} := 
7350: & \
7351: \left\{
7352: A_{x,v,0:c}^1(\rho): \ 
7353: (x,v,\rho)\in \R^n\times V_2 \times \R, \right. \\
7354: & \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \
7355: \left.
7356: A_{x,v,0:c}^1(\partial^+ \Delta) \cap 
7357: C= \emptyset, \ 
7358: \vert x \vert < \varepsilon, \ 
7359: \vert v \vert < \varepsilon, \
7360: 1-\varepsilon < \rho < \varepsilon
7361: \right\}.
7362: \endaligned\right.
7363: \end{equation}
7364: Since $A_{x, 0, 0: c}^1( \partial^+ \Delta)$ is contained in $(
7365: H^1)^+$ for all $x$ such that $A_{ x, 0, 0}^1(1) \in T^1 \cap (
7366: H^1)^+$, the closed subset $\mathcal{ C}_{ p_1}$ of $\mathcal{ M}_{
7367: p_1}$ is a {\it proper}\, closed subset of $\mathcal{ M}_{ p_1}$.
7368: The following figure provides a geometric illustration.
7369: 
7370: \bigskip
7371: \begin{center}
7372: \input wedge-half-wedge.pstex_t
7373: \end{center}
7374: 
7375: We can now state the main lemma of this section, which will complete
7376: the proof of Proposition~5.12 in Cases~{\bf ($\mathbf{ I_1}$)}
7377: and~{\bf ($\mathbf{ I_2}$)}.
7378: 
7379: \def\thelemma{9.20}\begin{lemma}
7380: After possibly shrinking $\Omega$ in a small neighborhood of $p_1$ and
7381: after possibly shrinking $\varepsilon >0$, the set $\mathcal{ W}_{
7382: p_1} \cap \left[ \Omega \cup \mathcal{ HW}_1^+ \right]$ is connected
7383: and for every holomorphic function $f \in \mathcal{ O}\left( \Omega
7384: \cup \mathcal{ HW }_1^+ \right)$, there exists a holomorphic function
7385: $F \in \mathcal{ O} \left( \Omega \cup \mathcal{ HW }_1^+ \cup
7386: \mathcal{ W}_{p_1} \right)$ such that $F \vert_{ \Omega \cup \mathcal{
7387: HW}_1^+}=f$.
7388: \end{lemma}
7389: 
7390: \proof
7391: Remind that $\varepsilon < < c^2$ and remind that the wedge $\mathcal{
7392: W}_{p_1}$ with $\mathcal{ W}_{p_1}\cap M= \emptyset$ in the two cases
7393: is of size ${\rm O}( \varepsilon)$. Since $C$ is contained in
7394: $(H^1)^-\cup \{ p_1\}\subset M^1$, we observe that $M \backslash C$ is
7395: locally connected at $p_1$. Since the half-wedge $\mathcal{ HW }_1^+$
7396: defined in Lemma~5.37 by simple inequalities is of size ${\rm O} (
7397: \delta_1)$, if moreover $\varepsilon < < \delta_1$, after shrinking
7398: $\Omega$ if necessary in a smal neighborhood of $p_1$ whose size
7399: is ${\rm
7400: O}(\varepsilon)$, it follows that we can assume that $\mathcal{
7401: W}_{p_1} \cap \left[ \Omega \cup \mathcal{ HW}_1^+ \right]$ is
7402: connected. However, in {\sc Figure~18} above, because we draw $M$ as
7403: if it were one-dimensional, the intersection $\mathcal{ W}_{p_1} \cap
7404: \left[ \Omega \cup \mathcal{ HW }_1^+ \right]$ appears to be
7405: disconnected, which is a slight incorrection.
7406: 
7407: Let $f$ be an arbitrary holomorphic function in $\mathcal{ O} \left(
7408: \Omega \cup \mathcal{ HW}_1^+ \right)$. Thanks to the isotopy
7409: Lemma~9.16 and thanks to the continuity principle Lemma~9.4, we deduce
7410: that $f$ extends holomorphically to a neighborhood in $\C^n$ of every
7411: disc $A_{x, v, u: c}( \overline{ \Delta})$ whose boundary $A_{ x, v,
7412: u:c} (\partial \Delta)$ is contained in $\Omega \cup \mathcal{
7413: HW}_1^+$. Using the fact that the mapping~\thetag{ 9.9} is one-to
7414: one, we deduce that we can extend $f$ uniquely by means of Cauchy's
7415: formula at points of the form $A_{x,v,u:c}^1(\rho)$ with 
7416: such values $\vert x
7417: \vert < \varepsilon$, $\vert v \vert < \varepsilon$, $\vert u \vert <
7418: \varepsilon$ and $1-\varepsilon < \rho < \varepsilon$ for 
7419: which $A_{x,v,u:c}^1(\partial \Delta) \subset
7420: \Omega \cup \mathcal{ HW}_1^+$, simply as follows
7421: \def\theequation{9.21}\begin{equation}
7422: f\left(
7423: A_{x,v,u:c}^1(\rho)
7424: \right):=\int_{\partial \Delta} \, 
7425: \frac{ f\left(
7426: A_{x,v,u:c}^1(\widetilde{ \zeta})
7427: \right)}{\widetilde{ \zeta} -\rho} \, 
7428: d \widetilde{ \zeta}.
7429: \end{equation}
7430: With this definition, we extend $f$ holomorphically and uniquely to
7431: the domain $\mathcal{ W}_{ p_1} \backslash \mathcal{ C}_{p_1}$, where
7432: $\mathcal{ C}_{p_1}$ is the proper closed subset, defined by~\thetag{
7433: 9.21}, of the $\mathcal{ C}^{2,\alpha-0}$-smooth hypersurface
7434: $\mathcal{ M}_{ p_1 }$ defined by~\thetag{ 9.20}. Let $F \in \mathcal{
7435: O} \left(\mathcal{ W}_{p_1} \backslash \mathcal{ C}_{p_1} \right)$
7436: denote this holomorphic extension. Since $\mathcal{ W}_{p_1} \cap
7437: \left[ \Omega \cup \mathcal{ HW}_1^+ \right]$ is connected, it follows
7438: that $\left[ \mathcal{ W}_{p_1} \backslash \mathcal{ C}_{p_1}\right]
7439: \cap \left[ \Omega \cup \mathcal{ HW }_1^+ \right]$ is also
7440: connected. By Lemma~9.4, the function $f$ and its holomorphic
7441: extension $F$ coincide in a neighborhood of every boundary
7442: $A_{x,v,u:c}^1 (\partial \Delta)$ which is contained in the domain
7443: $\Omega \cup \mathcal{ HW}_1^+$. From the analytic continuation
7444: principle, we deduce that there exists a well-defined function, still
7445: denoted by $F$, which is holomorphic in $\left[ \mathcal{ W}_{p_1}
7446: \backslash \mathcal{ C}_{p_1}\right] \cap \left[ \Omega \cup \mathcal{
7447: HW}_1^+ \right]$ and which extends $f$, namely $F\vert_{\Omega \cup
7448: \mathcal{ HW}_1^+}=f$.
7449: 
7450: To conclude the proof of Proposition~5.12 in Cases {\bf ($\mathbf{
7451: I_1}$)} and~{\bf ($\mathbf{ I_2}$)}, it suffices to extend $F$
7452: holomorphically through the closed subset $\mathcal{ C}_{p_1}$ of the
7453: connected hypersurface $\mathcal{ M}_{p_1}$ in the domain $\mathcal{
7454: W}_{p_1} \subset \C^n$. Since $n\geq 2$, we notice that we are
7455: exactly in the situation of Theorem~1.4 in the special, much simpler
7456: case where the generic submanifold $M$ is replaced by a domain of
7457: $\C^n$. It may therefore appear to be quite satisfactory to have
7458: reduced Theorem~1.2' to the CR dimension $\geq 2$ version Theorem~1.4,
7459: in an open subset of $\C^n$ (notice however that Theorem~1.4 as well
7460: as Theorems~1.2 and~1.2' are stated in positive codimension, since the
7461: case where $M$ is replaced by a domain of $\C^n$ is relatively trivial
7462: in comparison). The removability of such a proper closed subset
7463: contained in a connected hypersurface of a domain in $\C^n$ is known,
7464: follows from~\cite{j4} and is explicitely stated and proved as
7465: Lemma~2.10, p.~842 in~\cite{ mp1}. However, we shall still provide
7466: another different geometric proof of this simple removability result
7467: in Lemma~10.10 below, using fully the techniques developed in the
7468: previous sections.
7469: 
7470: The proofs of Lemma~9.20 together with Cases~{\bf ($\mathbf{I_1}$)}
7471: and~{\bf ($\mathbf{ I_2}$)} of Proposition~5.12 are complete now.
7472: \endproof
7473: 
7474: \subsection*{9.22.~End of proof of Proposition~5.12 in 
7475: Case {\bf (II)}} According to Lemma~5.37, in Case {\bf (II)}, the
7476: one-codimensional totally real submanifold $M^1 \subset M$ is given by
7477: the equations $y'=\varphi'(x,y_1)$ and $x_n=g(x'')$. If $u\in \R$ is
7478: a small real parameter, we may define a ``translation'' $M_u^1$ of
7479: $M^1$ in $M$ by the equations
7480: \def\theequation{9.23}\begin{equation}
7481: y'=\varphi'(x,y_1), \ \ \ \ \ \ \ \ 
7482: x_n=g\left(x''\right)+u.
7483: \end{equation}
7484: Similarly as in \S9.6, we may construct a family of analytic discs
7485: $A_{x,v,u:c}^1(\zeta)$ half-attached to $M_u^1$. We then we fix a
7486: small scaling parameter $c$ with $0 < c \leq c_1$ so that properties
7487: {\bf ($\mathbf{ 1_1}$)} to {\bf ($\mathbf{ 9_1}$)} of Lemmas~7.12
7488: and~8.3 hold true. Similarly as in \S9.8, we shall restrict 
7489: the variation
7490: of the parameter $v$ to an arbitrary $(n-1)$-dimensional subspace
7491: $V_1$ of $T_{p_1}M^1\cong \R^n$ which is supplementary to the real
7492: line $\R\cdot v_1$ in $T_{p_1}M^1$. If $\varepsilon>0$ is small
7493: enough with $\varepsilon < < c^2$, it follows that the mapping
7494: \def\theequation{9.24}\begin{equation}
7495: (x,v,u,\rho) \longmapsto 
7496: A_{x,v,u:c}^1(\rho) 
7497: \end{equation}
7498: is a {\it one-to-one immersion}\, from the open set $\{(x,v,\rho)\in
7499: \R^n\times V_1 \times \R\times \R: \ \vert x \vert < \varepsilon, \
7500: \vert v \vert < \varepsilon, \ 1-\varepsilon < \rho < 1\}$ into
7501: $\C^n$. Thanks to the choice of the linear subspace $V_1$, shrinking
7502: $\varepsilon >0$ if necessary, it follows that for every $u$ with
7503: $\vert u \vert < \varepsilon$, the open subset
7504: \def\theequation{9.25}\begin{equation}
7505: \left\{
7506: \aligned
7507: {}
7508: &
7509: \mathcal{ W}_u^1 :=\left\{
7510: A_{x,v,u:c}^1(\rho)\in \C^n: 
7511: (x,v,\rho)\in
7512: \R^n \times \R\times \R, \right. \\
7513: & \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \
7514: \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \
7515: \left.
7516: \vert x \vert < \varepsilon,
7517: \ \vert v \vert < \varepsilon, \ 
7518: 1-\varepsilon < \rho < 1
7519: \right\}
7520: \endaligned\right.
7521: \end{equation}
7522: is a local wedge of edge $M_u^1$. Clearly, this wedge 
7523: $\mathcal{ W}_u^1$ depends $\mathcal{ C}^{2,\alpha-0}$-smoothly 
7524: with respect to $u$. 
7525: 
7526: Using the fact that in Case~{\bf (II)} we have
7527: \def\theequation{9.26}\begin{equation}
7528: \frac{\partial A_{0,0,0:c}^1}{\partial \theta}(1)=
7529: v_1=(1,0,\dots,0)\in T_{p_1}M^1\cap T_{p_1}^cM,
7530: \end{equation}
7531: one can prove that Lemma~9.13 holds true with $\mathcal{ HW}_1^+$
7532: replaced by $\mathcal{ W}_2$ in~\thetag{ 9.14} and also that
7533: Lemma~9.16 holds true, again with $\mathcal{ HW}_1^+$ replaced by
7534: $\mathcal{ W}_2$. Similarly as in the proof of Lemma~9.20, applying
7535: then the continuity principle and using the fact that the
7536: mapping~\thetag{ 9.24} is one-to-one, after possibly shrinking
7537: $\Omega$ in a neighborhood of $p_1$, and shrinking $\varepsilon>0$, we
7538: deduce that for each $u\neq 0$, there exists a holomorphic function
7539: $F\in \mathcal{ O}\left(\Omega \cup \mathcal{ W}_2 \cup
7540: \mathcal{ W}_u^1 \right)$ with 
7541: $F_{\vert \Omega \cup \mathcal{ W}_2}=f$.
7542: 
7543: To conclude the proof of Proposition~5.12 in Case {\bf (II)}, it
7544: suffices to observe that for every fixed small $u$ with $
7545: -\varepsilon < < u < 0$, the wedge $\mathcal{ W}_u^1$ contains in fact
7546: a neighborhood $\omega_{p_1}$ of $p_1$ in $\C^n$.
7547: 
7548: The proofs of Proposition~5.12 and of Theorem~3.19 {\bf (i)} 
7549: are complete now. \qed
7550: 
7551: \subsection*{9.27.~End of proof of Theorem~1.2'}
7552: In order to derive Theorem~1.2' from Theorem~3.19 {\bf (i)}, we now
7553: remind the necessity of supplementary arguments about the stability of
7554: our constructions under deformation. Coming back to the strategy
7555: developped in \S3.16, we had a wedge $\mathcal{ W}_1$ attached to $M
7556: \backslash C_{\rm nr}$. Using a partition of unity, we may introduce a
7557: one-parameter $\mathcal{ C}^{ 2, \alpha}$-smooth family of generic
7558: submanifolds $M^d$, $d\in \R$, $d\geq 0$, with $M^0 \equiv M$, with
7559: $M^d$ containing $C_{\rm nr}$ and with $M^d \backslash C_{ \rm nr}$
7560: contained in $\mathcal{ W}_1$. In the proof of Theorem~3.19 {\bf (i)},
7561: thanks to this deformation, the wedge $\mathcal{ W}_1$ was replaced by
7562: a neighborhood $\Omega$ of $M\backslash C_{\rm nr}$ in $\C^n$.
7563: 
7564: In Sections~4 and~5, we constructed an important semi-local half-wedge
7565: $(\mathcal{ HW}_\gamma^+)^d$ attached to a one-sided neighborhood of
7566: $(M^1)^d$ in $M^d$ along a characteristic segment $\gamma^d$ of
7567: $M^d$. Now, we make the crucial claim that, after possibly adapting
7568: the deformation $M^d$, we may achieve that the geometric extent of
7569: this semi-local half-wedge be uniform as $d>0$ tends to zero, namely
7570: $(\mathcal{ HW}_\gamma^+)^d$ tends to a semi-local half-wedge
7571: $(\mathcal{ HW}_\gamma^+)^0$ attached to a one-sided neighborhood of
7572: $M^1$ in $M$ along $\gamma$, as $d$ tends to zero. Indeed, in
7573: Section~4 we have constructed a family of analytic discs $(\mathcal{
7574: Z}_{t,\chi,\nu:s}(\zeta))^d$ ({\it cf.}~\thetag{ 4.61}) which covers
7575: the half-wedge $(\mathcal{ HW}_\gamma^+)^d$. Thanks to the stability
7576: of Bishop's equation under $\mathcal{ C}^{2,\alpha}$-smooth
7577: perturbations, the deformed family $(\mathcal{
7578: Z}_{t,\chi,\nu:s}(\zeta))^d=: \mathcal{ Z}_{t,\chi,\nu:s}^d(\zeta)$ is
7579: also of class $\mathcal{ C}^{2,\alpha-0}$ with respect to the
7580: parameter $d$. We remind that for every $d>0$, the family $\mathcal{
7581: Z}_{t,\chi,\nu:s}^d(\zeta)$ was in fact constructed by means of a
7582: family $\widehat{ Z}_{ r_0,t,\tau,\chi,\nu:s}^d(\zeta)$ obtained by
7583: solving Bishop's equation~\thetag{ 4.40}, where we now add the
7584: parameter $d$ in the function $\Phi'$. In order to construct the
7585: semi-local attached half-wedge, we have used the rank property stated
7586: in Lemma~4.34. This rank property relied on the possibility of
7587: deforming the disc $\widehat{ Z}_{r_0,t:s}(\zeta)$ near the point
7588: $\widehat{ Z}_{r_0,t:s}^d(-1)$ in the open neighborhood
7589: $\Phi_s(\Omega) \equiv \Phi_s\left(\mathcal{ W}_1 \right)$ of $
7590: \Phi_s\left(M^d \right)$. As $d>0$ tends to zero, if $M^d$ tends to
7591: $M$, the size of the neighborhood $\Phi_s\left(\mathcal{ W}_1 \right)$
7592: shrinks to zero, hence it could seem that the we have no control on
7593: the semi-local attached half-wege $(\mathcal{ HW}_\gamma^+)^d$ as
7594: $d>0$ tends to zero. Fortunately, since the points $\widehat{
7595: Z}_{r_0,0:s}^d(-1)$ in a neighborhood of which we introduce the
7596: deformations~\thetag{ 4.30} are at a uniformly positive distance
7597: $\delta >0$ from $\gamma$, we may choose the deformation $M^d$ of $M$
7598: to tend to $M$ as $d$ tends to zero only in a small neighborhood of
7599: $\gamma$, whose size is small in comparison to this distance
7600: $\delta$. By smoothness with respect to $d$ of the family $\mathcal{
7601: Z}_{t,\chi,\nu:s}^d (\zeta)$, we then deduce that the semi-local
7602: half-wedge $(\mathcal{ HW}_\gamma^+)^d$ tends to a nontrivial
7603: semi-local half-wedge $(\mathcal{ HW}_\gamma^+)^0$ as $d$ tends to
7604: zero, which proves the claim.
7605: 
7606: Next, again thanks to the stability of Bishop's equation under
7607: perturbation, all the constructions of Sections~5, 6, 7, 8 and~9 above
7608: may be achieved to depend $\mathcal{ C}^{2, \alpha-0}$-smoothly, hence
7609: uniformly, with respect to $d$. Importantly, we observe that if the
7610: deformation $M^d$ is chosen so that $M^d$ tends to $M$ only in a small
7611: neighborhood of $p_1$ of size $ < < \varepsilon$, then the shrinking
7612: of $\varepsilon$ which occurs in Lemma~9.13 may be achieved to be
7613: uniform as $d$ tends to zero, because the part
7614: $A_{x,v,u:c}(\partial^3\Delta)$ stays in a uniform compact subset of
7615: $\Omega$, as $d$ tends to zero. At the end of the proof of
7616: Proposition~5.12, we then obtain univalent holomorphic extension to a
7617: local wedge $\mathcal{ W}_{p_1}^d$ of edge $M^d$ or to a neighborhood
7618: $\omega_{ p_1 }^d$ of $M^d$ in $\C^n$, and they tend smoothly to a
7619: wedge $\mathcal{ W }_{ p_1}^0$ of edge $M$ at $p_1$ or to a
7620: neighborhood $\omega_{ p_1 }$ of $p_1$ in $\C^n$.
7621: 
7622: The proof of Theorem~1.2' is complete.
7623: \qed
7624: 
7625: \section*{\S10.~Three proofs of Theorem~1.4}
7626: 
7627: \subsection*{10.1.~Preliminary}
7628: Theorem~1.4 may be established by means of the processus of
7629: minimalization of generic submanifods developed by B.~J\"oricke
7630: in~\cite{j2}, as is done effectively in~\cite{j4} in the hypersurface
7631: case and then in~\cite{p2} in arbitrary codimension. In this section,
7632: we shall suggest three more different proofs of Theorem~1.4. As
7633: explained in Section~3, it suffices to treat $\mathcal{
7634: W}$-removability, and essentially to prove Proposition~3.22. The first
7635: proof appears already in~\cite{ m2} and also in~\cite{p1}. The second
7636: proof consists in repeating some of the constructions of the previous
7637: sections, using the fact that $M^1$ is of positive CR dimension in
7638: order to simplify substantially the reasonings. The third proof
7639: consists of a slicing argument showing that {\it Theorem~1.4 is in
7640: fact a logical consequence of Theorem~1.2'}. In fact, because these
7641: three proofs are already written elsewhere or very close to the
7642: constructions developed in the previous sections, we shall only
7643: provide summaries here.
7644: 
7645: \subsection*{10.2.~Normal deformation of analytic discs attached 
7646: to $M^1$} Firstly, in the situation of Proposition~3.22, because $M^1$
7647: is of positive CR dimension, we can construct a small analytic disc $A
7648: (\zeta)$ attached to $M^1$ which satisfies $A(1) = p_1 \in M^1$ and $A
7649: (\partial \Delta \backslash \{1\}) \subset (H^1)^+$. As in~\cite{m2},
7650: \cite{ mp1}, using normal deformations of $A$ near $A(-1)$, we may
7651: include $A$ in a $\mathcal{ C}^{ 2, \alpha-0}$-smooth parametrized
7652: family $A_v (\zeta)$ of analytic discs attached to $M^1$, where $v \in
7653: \R^{d+1}$ is small, so that the rank at $v=0$ of the mapping $v
7654: \longmapsto -\frac{ \partial A_v}{ \partial \rho}(1) \in T_{ p_1} \C^n
7655: \ {\rm mod} \ T_{p_1} M^1 \cong \R^{ d+1}$ is maximal equal to $(d
7656: +1)$, the codimension of $M^1$ in $\C^n$. For this, we use a
7657: deformation lemma which is essentially due to A.~Tumanov~\cite{tu3},
7658: which appears as Lemma~2.7 in~\cite{ mp1} and which was already used
7659: above (with a supplementary parameter $s$) in Lemma~4.34. Then we add
7660: a ``translation'' parameter $x\in \R^{ 2m+d-1}$, getting a family of
7661: analytic discs $A_{x,v} (\zeta)$ with $A_{x,v} (\partial
7662: \Delta)\subset M^1$ so that the rank at $x=0$ of the mapping $x
7663: \longmapsto A_{x,0} (1) \in M^1$ is maximal equal to $(2m+d-1)$, the
7664: dimension of $M^1$. Finally, we introduce some ``translations''
7665: $M_u^1$ of $M^1$ in $M$, where $u \in \R$, and we obtain a family $A_{
7666: x, v,u} (\zeta)$ of analytic discs attached to $M_u^1$. For $u\neq 0$,
7667: since the discs $A_{x,v,u}(\zeta)$ are attached to $M_u^1$, their
7668: boundaries are contained in $M\backslash C$. Applying the
7669: approximation Lemma~4.8, we deduce that for every $u\neq 0$, all
7670: holomorphic functions in the open set $\Omega$ of Proposition~3.22
7671: (which contains $M\backslash C$) extend holomorphically to the local
7672: wedge $\mathcal{ W }_u := \{A_{x, v,u} (\rho): \ \vert x \vert <
7673: \varepsilon, \ \vert v \vert < \varepsilon, \ 1-\varepsilon < \rho <
7674: 1\}$ of edge $M_u^1$. To control the univalence of the holomorphic
7675: extension, it suffices to shrink $\Omega$ a little bit in a
7676: neighborhood of $p_1$. To conclude the proof, one observes as
7677: in~\cite{ j4}, \cite{ cs} that the union $\bigcup_{ u \neq 0} \,
7678: \mathcal{ W}_u^1$ contains a local wedge of edge $M$ at $p_1$. This
7679: processus is sometimes called ``sweeping out by wedges''. We notice
7680: that this proof is geometrically much simpler than the proof of
7681: Theorem~1.2' achieved in the previous sections.
7682: 
7683: \subsection*{10.3.~Half-attached analytic discs}
7684: Secondly, we may generalize our constructions achieved in the previous
7685: sections from the CR dimension $m=1$ case to the CR dimension $m\geq
7686: 2$ case, as follows. Let $M$, $M^1$, $p_1$, $H^1$ and $C$ be as in
7687: Proposition~3.22. The fact that $M^1$ is of positive CR dimension
7688: will provide substantial geometric simplifications for essentially two
7689: reasons. Indeed, as the hypersruface $H^1\subset M^1$ is generic and
7690: at least of real dimension $n$, there exists a (in fact infinitely many)
7691: maximally real submanifold $K^1$ passing through $p_1$ which is
7692: contained in $H^1$. Then
7693: \begin{itemize}
7694: \item[{\bf (i)}]
7695: Applying the considerations of Section~7, we may construct families of
7696: small analytic discs $A_{x,v} (\zeta)$ which are half-attached to
7697: $K^1$ and which cover a local wedge of edge $K^1$ at $p_1$, when the
7698: translation parameter $x$ and the rotation parameter $v$ vary. We
7699: notice that this would be impossible in the case $m=1$, because in
7700: this case $H^1$ is of real dimension $(n-1)$, hence does {\it not}\,
7701: contain any maximally real submanifold.
7702: \item[{\bf (ii)}]
7703: We can even prescribe the direction $\frac{ \partial A_{ 0,0} }{
7704: \partial \theta} (1)$ as an arbitrary given nonzero vector $v_1\in T_{
7705: p_1}K^1$. Again, this would be impossible in the CR dimension $m=1$
7706: case.
7707: \end{itemize}
7708: Let $\mathcal{ HW }_1^+$ be a local half-wedge of edge $(M^1)^+$ at
7709: $p_1$, whose construction is suggested in \S4.64. Since $K^1$ is
7710: generic, we may choose a nonzero vector $v_1\in T_{ p_1}K^1$ with the
7711: property that $\mathcal{ HW }_1^+$ is directed by $J
7712: v_1$. Generalizing Lemma~8.3, we see that $A_{x,v} \left( \overline{
7713: \Delta } \backslash \partial^+ \Delta \right)$ is contained in
7714: $\mathcal{ HW}_1^+$. We notice that in the CR dimension $m=1$ case,
7715: $H^1$ is {\it not}\, generic, and we remember that the choice of a
7716: special point $p_1$ to be removed locally and the choice of a
7717: supporting hypersurface $H^1 \subset M^1$ was much more subtle,
7718: because we had to insure that there exists a vector $v_1 \in T_{
7719: p_1}H^1$ such that $\mathcal{ HW }_1^+$ is directed by $J v_1$.
7720: 
7721: Next, we can translate $K^1$ in $M^1$ by means of a small parameter
7722: $t\in \R^{d-1}$ and then $M^1$ in $M$ by means of a small parameter
7723: $u\in \R$. By stability of Bishop's equation, we get a family of
7724: analytic discs $A_{x,v,t,u}(\zeta)$ half-attached to the translation
7725: $K_{t,u}^1$ of $K^1$. Nine properties analogous to properties {\bf
7726: ($\mathbf{ 1_1}$)} to {\bf ($\mathbf{ 9_1}$)} of Lemmas~7.12 and~8.3
7727: are then satisfied and we conclude the proofs of Proposition~3.22 and
7728: of Theorem~1.4 in essentially the same way as in Section~9
7729: above. We shall not write down all the details.
7730: 
7731: We notice that this second strategy of proof is much more complicated
7732: than the first (known) proof summarized in \S10.2 just above.
7733: 
7734: \subsection*{10.4.~Slicing argument: reduction of Theorem~1.4 to 
7735: Theorem~1.2'} Thirdly, we may provide a new proof of the central
7736: Theorem 1.4 valid in CR dimension $m\geq 2$, in order to illustrate
7737: how our results in CR dimension 1 can be applied to the general case
7738: via slicing techniques. We shall see that Theorem~1.4 is a logical 
7739: consequence of Theorem~1.2'.
7740: 
7741: Let $M$, $M^1$ and $C$ be as in Theorem~1.4. To begin with, we shall
7742: treat the three notions of removability (CR-, $L^{\sf p}$- and
7743: $\mathcal{ W}$-) commonly. However, we remind that CR-removability is
7744: immediately reduced to $\mathcal{ W}$-removability thanks to
7745: Lemma~3.5, hence it suffices to consider only $L^{\sf p}$- and
7746: $\mathcal{ W}$-removability.
7747: 
7748: Arguing by contradiction, we see as in the
7749: previous parts of this paper that we lose essentially nothing if we
7750: consider $C$ to be the minimal nonremovable subset. Also, we may
7751: assume holomorphic extension to a wedge attached to $M\backslash C$. 
7752: As explained in \S3.16 it is enough to remove one single point of $C$.
7753: 
7754: As in \S3.16 ({\it see} especially the statement of Proposition~3.22),
7755: we can show that there is a $\mathcal{ C}^{2,\alpha}$-smooth
7756: hypersurface $H$ of $M$ which is generic in $\C^n$, transversal to
7757: $M^1$, and which has the following property: $H$ contains some point
7758: $p_1\in C$, and we can choose a small neighborhood of $H$ in $M$ in
7759: which $H$ has two connected open sides $H^+$ and $H^-$ such that $C$
7760: is contained in $H^-\cup\{p_1\}$ locally in a neighborhood of
7761: $p_1$. Notice that the hypersurface $H^1$ constructed in Lemma~3.21 is
7762: simply the intersection of $M^1$ with such a hypersurface $H$.
7763: 
7764: Next, in a neighborhood of $p_1$, we construct a local foliation of
7765: $M$ by generic submanifolds of CR dimension 1 (hence of dimension
7766: $n+1$) as follows. We notice that $\dim_\R H=2m+d-1$. We first choose
7767: a local $\mathcal{ C}^{2,\alpha}$-smooth foliation of $H$ by generic
7768: submanifolds of CR dimension 1 (hence of real dimension $n+1$) which
7769: we denote by $\widetilde{ M}_s$, where the transversal parameters $s=
7770: (s_1,\ldots,s_{m-2})\in \I_{m-2}(\rho_1)$ belong to a cube in
7771: $\R^{m-2}$ of radius some $\rho_1>0$. Afterwards, we extend this
7772: foliation to a local $\mathcal{ C}^{2,\alpha}$-smooth foliation
7773: $\widetilde{ M}_{s,t}$ of a neighborhood of $p_1$ in $M$, where $s\in
7774: \I_{m-2} (\rho_1)$, where $t\in \I(\rho_1)$ and where $\widetilde{
7775: M}_{ s,0}\equiv \widetilde{ M}_s$; if $m=2$, we notice that the
7776: parameter $s$ disappears and that we have only one real parameter $t$.
7777: Also, we can assume that $\widetilde{ M}_{s,t}$ is contained in 
7778: $H^+$ if and only if $t>0$.
7779: 
7780: By genericity, the submanifolds $\widetilde{ M}_{s,t}$ may be chosen
7781: in addition to be transversal to the one-codimensional submanifold
7782: $M^1\subset M$ containing the singularity $C$ with of course $p_1\in
7783: \widetilde{ M}_{0,0}$. Then for all parameters $s$ and $t$ the
7784: intersections $\widetilde{ M}_{s,t}^1:=M^1\cap \widetilde{ M}_{s,t}$
7785: are maximally real submanifolds of $\C^n$. Considering $\widetilde{
7786: M}_{s,t}^1$ as a maximally real one-codimensional submanifold of
7787: $\widetilde{ M}_{ s,t}$, a characteristic foliation is induced on each
7788: $\widetilde{ M}_{s,t}^1$. After contraction around $p_1$ we can
7789: assume that these characteristic foliations all have trivial topology:
7790: their leaves are the level sets of an $\R^{m-1}$-valued submersion. As
7791: the intersection $\widetilde{ M}_{0,0}^1 \cap C$ is the singleton
7792: $\{p_1\}$ (by construction of $H$), there exists $\varepsilon>0$ such
7793: that for all $s$ with $\vert s \vert < \varepsilon$ and all $t$ with
7794: $\vert t \vert < \varepsilon$, the closed subsets
7795: $C_{s,t}:=\widetilde{ M}_{s, t}^1\cap C$ are compact in $\widetilde{
7796: M}_{s, t}^1$. Notice that $C_{s,t}$ is even empty if 
7797: $t>0$, because $C$ is contained in $H^-\cup \{p_1\}$.
7798: The following simple fact shows that $C_{s,t}$ satisfies
7799: the nontransversality condition $\mathcal{ F}^c_{
7800: \widetilde{ M}_{s,t}^1}\{C_{s,t}\}$ of
7801: Theorem~1.2', for all $s$ with $\vert s \vert < \varepsilon$ and all
7802: $t$ with $-\varepsilon < t \leq 0$.
7803: 
7804: \def\thelemma{10.5}\begin{lemma}
7805: Let $\mathcal{ F }_\mathcal{ M}$ be a $\mathcal{ C}^{ 1,
7806: \alpha}$-smooth foliation by curves on some $m$-dimensional $\mathcal{
7807: C}^{ 2, \alpha}$-smooth real manifold $\mathcal{ M}$ defined by a
7808: surjective $\mathcal{ C}^{ 1, \alpha}$-smooth submersion $F: \mathcal{
7809: M} \rightarrow \I_{ m-1}( \rho_1)$. Then every compact set $\mathcal{
7810: C} \subset \mathcal{ M}$ satisfies the nontransversality condition
7811: $\mathcal{ F}_\mathcal{ M}\{ \mathcal{ C} \}$.
7812: \end{lemma}
7813: 
7814: \proof
7815: Let $\mathcal{ C}'$ be an arbitrary compact subset of $\mathcal{ C}$.
7816: As $\mathcal{ C}'$ is compact, there exists the smallest $\rho_2<
7817: \rho_1$ with $\mathcal{ C}' \subset F^{-1}( \overline{ \I_{ m-1}(
7818: \rho_2)})$. The semi-local projection $\pi_{ \mathcal{ F}_{\mathcal{
7819: M}}}$ along the leaves of $\mathcal{ F }$ may of course be identified
7820: with $F$. Thus, $\pi_{ \mathcal{ F}_{ \mathcal{ M } }} ( \mathcal{
7821: C}')$ is contained in $\overline{ \I_{m-1 }( \rho_2)}$ and meets the
7822: boundary $\partial \I_{ m-1} (\rho_2 )$ of the cube $\I_{m-1}
7823: (\rho_2)$. Also, by compactness, the set
7824: $\mathcal{ C}'$ cannot contain a fiber of $F$ in the whole. 
7825: This completes the proof.
7826: \endproof
7827: 
7828: We can now show that Theorem~1.4 is a logical consequence of
7829: Theorem~1.2'. In fact, we cannot insure that the generic submanifolds
7830: $\widetilde{ M}_{s,t}$ of CR dimension $1$ defined above are all
7831: globally minimal, hence it seems that Theorem~1.2' itself does not
7832: apply. However, we notice that the wedge attached to $M\backslash C$
7833: restricts to a wedge attached to $\widetilde{ M}_{s,t}\backslash
7834: C_{s,t}$, for all $s$ and $t$. Hence, we can observe that everything
7835: that was needed in the proof of Theorem~1.2' was the existence of a
7836: wedge attached to $M\backslash C$ to which holomorphic extension is
7837: already assumed. One can even formulate a slightly more general
7838: version of Theorem~1.2', where the global minimality assumption is
7839: replaced by the assumption of holomorphic extension to a wedge
7840: attached to $M\backslash C$. Of course, with this more general
7841: assumption, $M$ may consists of several CR orbits, but thanks to
7842: Lemma~3.5 about stability of CR orbits, one may check that the proof
7843: of the main Theorem~3.19 {\bf (i)} and of the main Proposition~5.12
7844: remain unchanged, in the case where holomorphic extension is assumed
7845: in a wedgelike domain over $M\backslash C$ and not only in wedgelike
7846: domains attached to the CR orbits of $M\backslash C$. 
7847: 
7848: Thus this slight generalization of Theorem~1.2' together with the
7849: observation made in Lemma~10.5 just above yield that for all $s$ with
7850: $\vert s \vert < \varepsilon$ and for all $t$ with $-\varepsilon < t
7851: \leq 0$, the closed subset $C_{ s,t}$ is $\mathcal{ W}$-removable in
7852: the generic submanifold $\widetilde{ M}_{s, t}$. We deduce that for
7853: every $(s, t)$ with $\vert s \vert < \varepsilon$ and $-\varepsilon <
7854: t \leq 0$, we get holomorphic extension from the given restricted
7855: wedge attached to $\widetilde{ M}_{ s, t} \backslash C_{ s, t}$ into
7856: an open wedge $\widetilde{ \mathcal{ W}}_{ s,t}$ attached $\widetilde{
7857: M}_{ s, t}$. Notice that this does not immediately achieves the proof,
7858: since the direction of the $\widetilde{ \mathcal{ W }}_{s, t}$ need
7859: not depend continuously on $(s, t)$. In fact, the proof of the slight
7860: generalization of Theorem~1.2' contains arguments (for example the
7861: localization near a very special point) which do not depend nicely on
7862: external parameters. Hence the attached wedges $\widetilde{ \mathcal{
7863: W}}_{s,t}$ may well be completely unrelated.
7864: 
7865: To overcome this difficulty we proceed in the following way, already
7866: argued in~\cite{ m2}, \cite{ mp1} (Lemma~2.7) in slightly different
7867: contexts. We first construct a regular family $A_{ x, v}( \zeta)$ of
7868: analytic discs attached to $\widetilde{ M}_{ 0, 0 }\cup \Omega$ whose
7869: size is small in comparison to the basis of the wedge $\widetilde{
7870: \mathcal{ W} }_{0, 0}$ and which sweep out a local wedge $\mathcal{ W}
7871: (A_{ x, v})$ of edge $\widetilde{ M}_{ 0, 0}$ at $p_1$. Here, the
7872: parameter $x\in \R^{ n+ 1}$ corresponds to translations in
7873: $\widetilde{ M}_{ 0, 0}$ and $v \in \R^{ n-2}$ to normal deformations
7874: in a neighborhood of the point $A_{ 0, 0} (-1)\in \Omega$. Deforming
7875: this family thanks to the flexibility of Bishop's equation, we
7876: construct a family $A_{x, v,s,t} (\zeta)$ attached to $\widetilde{
7877: M}_{ s,t} \cup \Omega$, still sweeping out a local wedge $\mathcal{ W
7878: }(A_{ x, v, s, t})$ of edge $\widetilde{ M}_{ s, t}$. This family is
7879: of class $\mathcal{ C}^{2,\alpha-0}$ with respect to all
7880: parameters. Using the $\mathcal{ W}$-removability of $C_{s,t}$, we can
7881: introduce for every $(s, t)$ a one-parameter deformation $\widetilde{ M
7882: }_{ s,t}^d$ of $\widetilde{ M}_{s,t}$ which is contained in the
7883: attached wedge $\widetilde{ W}_{ s,t}$ whenever $d>0$ and which
7884: coincides with $\widetilde{ M}_{s,t}$ when $d=0$. Thanks to the
7885: flexibility of Bishop's equation with parameters, we get a deformed
7886: family $A_{x,v, s, t: d}(\zeta)$ of analytic discs. Since the wedges
7887: $\widetilde{ W}_{ s,t}$ are {\it a priori}\, unrelated, we loose the
7888: smoothness with respect to all variables, including $d$. Fortunately,
7889: by an application of the continuity principle, for every $d>0$, we
7890: deduce holomorphic extension to the wedge generated by the family
7891: $A_{x, v,s, t:d} (\zeta)$. If we let $d$ tend to zero, fixing $(s,t)$,
7892: we obtain univalent holomorphic extension to the wedge generated by
7893: the family $A_{x,v,s,t}(\zeta)$. Finally, as $(s,t)$ varies, the
7894: wedges $\mathcal{ W}( A_{x,v,s,t})$ varies smoothly and covers a local
7895: wedge $\mathcal{ W}$ of edge $M$ at $p_1$. By the continuity
7896: principle, we may verify that we obtain univalent holomorphic
7897: extension to $\mathcal{ W}$.
7898: 
7899: Secondly, we explain how $L^{\sf p}$-removability of the
7900: point $p_1\in C$ in $M$ follows logically from the $L^{\sf
7901: p}$-removability of every $C_{s,t}$ in $\widetilde{ M}_{s,t}$. The
7902: main argument relies on the following simple but useful fact: if $M$
7903: is a generic CR submanifold of $\C^n$ and if $N\subset M$ is a lower
7904: dimensional submanifold which is itself a generic CR submanifold of
7905: $\C^n$ of positive CR dimension, then differentiable CR functions on
7906: $M$ obviously restrict to CR functions on $N$. More generally, a
7907: foliated version of this observation with lower regularity assumptions
7908: is as follows.
7909: 
7910: \def\thelemma{10.6}\begin{lemma}
7911: Let $M\subset\C^n$ be a generic submanifold of class $\mathcal{
7912: C}^{2,\alpha}$ and of CR dimension $m\geq 2$.
7913: 
7914: \begin{itemize}
7915: \item[{\bf (a)}]
7916: If $p_1\in M$ and
7917: if $M$ carries a local $\mathcal{ C }^{ 2, \alpha}$-smooth foliation
7918: by a family $N_u$ of generic submanifolds of CR dimension $1$ where
7919: $u \in \R^{m-1}$ is a small parameter and where
7920: $p_1 \in N_0$, then for every CR function $f
7921: \in L_{ loc}^{\sf p} (M)$, ${\sf p}\geq 1$, and for almost every $u\in
7922: \R^{ m-1}$, its restriction $f|_{ N_u}$ is an $L_{ loc}^{\sf p} ( N_u)$
7923: function which is CR on $N_u$.
7924: \item[{\bf (b)}]
7925: Conversely, if $p_1 \in M$ and if $M$ carries $m$ local $\mathcal{
7926: C}^{2, \alpha}$-smooth foliations by families $N_{u_j}^j$,
7927: $j=1, \dots,m$, $u_j\in \R^{m-1}$, of generic submanifolds of CR
7928: dimension $1$ satisfying $p_1\in N_0^j$ for $j=1,\dots,m$ and
7929: \def\theequation{10.7}\begin{equation}
7930: T_{p_1}N_0^1+\cdots + T_{p_1}N_0^m =T_{p_1}M, 
7931: \end{equation}
7932: then a function $f\in L_{ loc}^{ \sf p }(M)$ is CR in a neighborhood of
7933: $p_1$ if and only if for all $j =1, \dots,m$ and for almost every
7934: $u_j \in \R^{ m-1}$, its restriction $f\vert_{ N_{ u_j}^j}$ is CR on
7935: $N_{ u_j }^j$.
7936: \end{itemize}
7937: \end{lemma}
7938: 
7939: \proof 
7940: Of course, property {\bf (a)} only makes sense for an everywhere
7941: defined representative of $f$ and the nullset of excluded parameters
7942: $t$ depends on the choice of the representative of $f$.
7943: 
7944: To establish {\bf (a)}, we choose a small box-neighborhood $U\cong
7945: \I_{ m-1}( \rho_1)\times \widetilde{ N}$, where $\I_{ m-1}(\rho_1)$ is
7946: a cube of some positive radius $\rho_1 >0$ in $\R^{ m-1}$, such that
7947: every plaque $\{ \upsilon \}\times \widetilde{N}$ is an open subset of
7948: some leaf $N_u$. By the $L^{\sf p}$ version of the
7949: approximation theorem ({\it cf.}~\cite{j5}, \cite{p1}, \cite{mp1}),
7950: the restriction $f|_U$ is the limit in the $L^{ \sf p}$ norm of the
7951: restrictions of holomorphic polynomials $(P_\nu )_{ \nu \in
7952: \N}$. Thanks to Fubini's theorem, we deduce that for almost every
7953: $\upsilon \in \I_k ( \rho_1)$, the restriction $P_\nu |_{ \{\upsilon
7954: \} \times \widetilde{ N }}$ converges in $L^{\sf p}$ norm to $f|_{ \{
7955: \upsilon \} \times \widetilde{ N }}$. Hence for such parameters
7956: $\upsilon$, the restriction $f \vert_{\{\upsilon \} \times \widetilde{
7957: N}}$ is CR, which completes the proof of {\bf (a)}.
7958: 
7959: To establish {\bf (b)}, we observe first that the ``only if'' part is
7960: a direct consequence of {\bf (a)}. To prove the ``if'' part, we may
7961: introduce for every $j=1,\dots,m$ and for every $u_j\in \R^{m-1}$ a
7962: $(0,1)$ vector field $\overline{L}_{u_j}^j$ tangent to $N_{u_j}^j$ and
7963: $\mathcal{ C}^{1,\alpha}$-smoothly parameterized by $u_j$. The
7964: geometric assumption~\thetag{ 10.7} entails that the $m$ vector fields
7965: $\overline{L}_{u_1}^1,\dots,\overline{L}_{u_m}^m$ generate the CR
7966: bundle $T^{0,1}M$ in a neighborhood of $p_1$. By assumption, the
7967: $L_{loc}^{\sf p}$ function $f$ is annihilated in the distributional
7968: sense by these $m$ vector field, hence it is CR. This completes the
7969: proof of {\bf (b)}.
7970: \endproof
7971: 
7972: We can now prove that the $L^{ \sf p}$-removability of $C$ in $M$
7973: follows from an application of Theorem~1.2'. Let a function $f\in
7974: L^{\sf p}_{ loc}(M)$ which is CR on $M \backslash C$. Coming back to
7975: the construction of the submanifolds $\widetilde{ M}_{s,t}$ achieved
7976: in the paragraphs before Lemma~10.5, it is clear that for almost every
7977: $(s,t) \in \R^{ m-1}$, the restriction $f\vert_{ \widetilde{
7978: M}_{s,t}}$ is $L_{ loc}^{ \sf p}$-integrable. More generally,
7979: proceeding as in the paragraph before Lemma~10.5, we may construct $m$
7980: such families $\widetilde{ M}_{j; s_j,t_j}$ for $j=1, \dots, m$ with
7981: $p_1\in \widetilde{ M}_{ j; 0,0}$ and
7982: \def\theequation{10.8}\begin{equation}
7983: T_{p_1}\widetilde{ M}_{1;s_1,t_1}+\cdots+
7984: T_{p_1}\widetilde{ M}_{m;s_m,t_m}=
7985: T_{p_1}M,
7986: \end{equation}
7987: without changing the conclusion that the corresponding closed subsets
7988: $C_{s_j, t_j}^j$ are $L^{\sf p}$-removable in $\widetilde{ M }_{j;
7989: s_j,t_j}$ for $j =1, \dots, m$. Applying Lemma~10.7 just above, we
7990: finally deduce that $f$ is CR on $M$, as desired.
7991: 
7992: This completes the description of the reduction of
7993: Theorem~1.4 to Theorem~1.2' via a slicing argument.
7994: 
7995: \subsection*{10.9.~Version of Theorem~1.4 in an open subset 
7996: of $\C^n$} To conclude this section, we remind that in the end of the
7997: proof of Proposition~5.12 in Cases {\bf ($\mathbf{ I_1}$)} and {\bf
7998: ($\mathbf{ I_2}$)}, we came down to the removability of a proper
7999: closed subset $\mathcal{ C}_{p_1}$ of a one-codimensional submanifold
8000: $\mathcal{ M}_{p_1}$ of an open subset of $\C^n$ ($n\geq 2$), namely
8001: the wedge $\mathcal{ W}_{p_1}$ (remind {\sc Figure~18} above), which
8002: amounts exactly to prove Theorem~1.4 in the case where the generic
8003: submanifold $M$ is replaced by an open subset of $\C^n$. We may
8004: formulate this result as the following lemma. To our knowledge, its
8005: first known proof is given in~\cite{ j4}. Here, we provide a slightly
8006: different proof, using half-attached analytic discs.
8007: 
8008: \def\thelemma{10.10}\begin{lemma}
8009: Let $\mathcal{ D}\subset \C^n$ be a domain, let $\mathcal{ M}^1
8010: \subset \mathcal{ D}$ be a connected $\mathcal{ C}^{2,\beta}$-smooth
8011: hypersurface with $0 < \beta < 1$ and let $\mathcal{ C}$ be a proper
8012: closed subset of $\mathcal{ M}^1$ which does not contain any CR orbit
8013: of $\mathcal{ M}^1$. Then for every holomorphic function $f \in
8014: \mathcal{ O} \left( \mathcal{ D} \backslash \mathcal{ C} \right)$,
8015: there exists a holomorphic function $F \in \mathcal{ O}(\mathcal{ D})$
8016: such that $F \vert_{\mathcal{ D} \backslash \mathcal{ C }} = f$.
8017: \end{lemma}
8018: 
8019: \proof
8020: We summarize the proof, which anyway is very similar to the proof of
8021: Proposition~3.22 delineated in \S10.3 above. Reasoning by
8022: contradiction and constructing a supporting hypersurface, we come down
8023: to the local removability of a single point $p_1$ in a geometric
8024: situation analogous to the one described in Proposition~3.22, with $M$
8025: replaced by the domain $\mathcal{ D}$, with $M^1$ replaced by
8026: $\mathcal{ M}^1$ and with a generic $\mathcal{ C}^{2,\beta}$-smooth
8027: submanifold $\mathcal{ H}^1\subset \mathcal{ M}^1$ such that, locally
8028: in a neighborhood of $p_1$, we have $\mathcal{ C}\subset (\mathcal{
8029: H}^1)^-\cup \{p_1\}$, where we use the same notation $\mathcal{ C}$
8030: for the smallest non-removable subset of the original $\mathcal{ C}$.
8031: 
8032: Notice that $\mathcal{ H}^1$ is of codimension $2$. Let $\mathcal{
8033: K}^1\subset \mathcal{ H}^1$ be a maximally real submanifold passing
8034: through $p_1$. We may translate $\mathcal{ K}^1$ in $\mathcal{ M}^1$
8035: by means of a small parameter $t\in \R^{n-1}$ and then $\mathcal{
8036: M}^1$ in $\mathcal{ D}$ by means of a small parameter $u\in \R$.
8037: Following Section~7, we then construct a small family of analytic
8038: discs $\mathcal{ A}_{x,v,t,u}^1(\zeta)$ half-attached to the
8039: ``translations'' $\mathcal{ K}_{t,u}^1$. Nine properties analogous to
8040: properties {\bf ($\mathbf{ 1_1}$)} to {\bf ($\mathbf{ 9_1}$)} of
8041: Lemmas~7.12 and~8.3 are then satisfied and we conclude the proof 
8042: in essentially the same way as in Section~9 above.
8043: \endproof
8044: 
8045: \section*{\S11~$\mathcal{ W}$-removability implies
8046: $L^{\sf p}$-removability}
8047: 
8048: \subsection*{11.1.~Preliminary}
8049: This section is devoted to prove Lemma~3.15 about $L^{\sf
8050: p}$-removability of the proper closed subset $C\subset M^1$, granted
8051: it is $\mathcal{ W}$-removable. More generally, we shall establish the
8052: $L^{\sf p}$-removability of certain proper closed subsets $\Phi$ of
8053: $M$ that are nullsets with respect to the Lebesgue measure of $M$.
8054: 
8055: As a preliminary, we remind that if $M'$ is a globally minimal
8056: $\mathcal{ C}^{2,\alpha}$-smooth generic submanifold of $\C^n$ of CR
8057: dimension $m \geq 1$ and of codimension $d= n-m\geq 1$, there exists a
8058: wedge $\mathcal{ W}'$ attached to $M'$ constructed by means of
8059: analytic discs successively glued to $M'$ and to conelike submanifolds
8060: attached to $M'$ consisting of parametrized families of pieces of
8061: analytic discs. By means of the approximation theorem of~\cite{ bt},
8062: one deduces classically that continuous CR functions on $M'$ extend
8063: holomorphically to $\mathcal{ W}'$, and continuously to $M'\cup
8064: \mathcal{ W}'$.
8065: 
8066: For the holomorphic extension of the $L_{loc}^{\sf p}$ CR functions to
8067: a wedge attached to $M'$, some supplementary routine, though not
8068: obvious, work has to be achieved. Firstly, using a convolution with
8069: Gauss' kernel as in~\cite{bt}, one shows that on a $\mathcal{
8070: C}^2$-smooth generic submanifold $M'$ of $\C^n$, every $L_{loc}^{\sf
8071: p}$ CR function on $M'$ is locally the limit, in the $L^{\sf p}$ norm,
8072: of a sequence of polynomials ({\it see} Lemma~3.3 in~\cite{j5}). In
8073: the case where $M'$ is a hypersurface, studied in~\cite{j5}, the wedge
8074: $\mathcal{ W}'$ is in fact a one-sided neighborhood attached to $M'$,
8075: which we will denote by $\mathcal{ S}'$. The theory of Hardy spaces on
8076: the unit disc transfers to parameterized families of small analytic
8077: discs glued to $M'$ which cover local one-sided neighborhoods of a
8078: hypersurface, provided the boundaries of these discs foliate an open
8079: subset of $M'$. Using in an essential way L.~Carleson's imbedding
8080: theorem, B.~J\"oricke established in~\cite{j5} that every $L^{\sf
8081: p}_{loc}$ CR function on a globally minimal $\mathcal{ C}^2$-smooth
8082: hypersurface $M'$ extends holomorphically in the Hardy space $H^{\sf
8083: p}(\mathcal{ S}')$ of holomorphic functions defined in the one-sided
8084: neighborhood $\mathcal{ S}'$, with
8085: $L^{\sf p}$ boundary values on the hypersurface $M'$. 
8086: In his thesis~\cite{p1}, the second author
8087: of the present paper has built the theory in higher codimension,
8088: introducing the Hardy space $H^{\sf p}(\mathcal{ W}')$ of functions
8089: holomorphic in the wedge $\mathcal{ W}'$ attached to $M'$, with
8090: $L^{\sf p}$ boundary values on the edge $M'$.
8091: 
8092: At present, these background statements about holomorphic
8093: extendability of $L_{loc}^{\sf p}$ CR functions on globally minimal
8094: generic submanifolds may be reproved in a more elegant way than by
8095: going through the rather complicated technology dispersed in the
8096: articles~\cite{tu1}, \cite{tu2}, \cite{m1}, \cite{j2}, thanks to a
8097: simplification of the wedge extendability theorem obtained recently by
8098: the second author of this paper, which treats in an unified way local
8099: and global minimality. We refer the reader to the work in
8100: preparation~\cite{p3} for a substantial cleaning of the theory.
8101: 
8102: \subsection*{11.2.~$L^{\sf p}$-removability of nullsets}
8103: Let us say that a subset $\Phi$ of a $\mathcal{ C}^{2,\alpha}$-smooth
8104: generic submanifold is {\sl stably $\mathcal{ W}$-removable} if it is
8105: $\mathcal{ W}$-removable on every compactly supported sufficiently
8106: small $\mathcal{ C}^{2,\alpha}$-smooth deformation $M^d$ of $M$
8107: leaving $\Phi$ fixed. In the situations of Theorems~1.2' and~1.4, the
8108: assumptions of Lemma~11.3 just below are satisfied with $\Phi=C$,
8109: taking account of the fact that we have already established the
8110: $\mathcal{ W}$-removability of $C$ and that for logical reasons only,
8111: the closed set $C$ in the statements of Theorems~1.2' and~1.4 is
8112: obviously stably removable.
8113: 
8114: \def\thelemma{11.3}\begin{lemma}
8115: Let $M$ be a $\mathcal{ C}^{2, \alpha}$-smooth generic submanifold of
8116: $\C^n$ of CR dimension $m \geq 1$ and of codimension $d= n-m\geq 1$,
8117: hence of dimension $(2m+ d)$, let $\Phi \subset M$ be a nonempty
8118: proper closed subset whose $(2m +d)$-dimensional Hausdorff measure is
8119: equal to zero. Assume that $M \backslash \Phi$ is globally minimal and
8120: let $\mathcal{ W}$ be a wedge attached to $M\backslash \Phi$ such that
8121: every function in $L_{loc}^{\sf p}(M) \cap CR(M\backslash \Phi)$
8122: extends holomorphically as a function in the Hardy space $H^{\sf
8123: p}(\mathcal{ W})$. If $\Phi$ is stably 
8124: $\mathcal{ W}$-removable, then $\Phi$
8125: is $L^{\sf p}$-removable.
8126: \end{lemma}
8127: 
8128: Before giving the proof,
8129: let us summarize intuitively the reason why this strong $L^{\sf
8130: p}$-removability result Lemma~11.3 holds. Indeed, let $f\in L_{loc
8131: }^{\sf p}(M) \cap CR (M \backslash \Phi)$. As soon as wedge extension
8132: over points of $\Phi$ is known, thanks to the fact that we can deform
8133: $M$ over $\Phi$ in the wedgelike domain, thus erasing the singularity
8134: $\Phi$, we get a $L_{loc}^{ \sf p}$ CR function $f^d$ on the deformed
8135: manifold $M^d$, without singularities anymore, and in additition, we
8136: can let the deformation $M^d$ tend to $M$ with a uniform with $L^{\sf
8137: p}$ control of the extension $f^d$, which therefore tends to a CR
8138: extension of $f$ through $\Phi$.
8139: 
8140: \proof
8141: First of all, we remind that for every ${\sf p}$ with $1\leq {\sf p}
8142: \leq \infty$, the space $L_{loc}^{\sf p}(M)$ is contained in
8143: $L_{loc}^1(M)$. We claim that it follows that $\Phi$ is $L^{\sf
8144: p}$-removable for every ${\sf p}$ with $1\leq {\sf p} \leq \infty$ if
8145: and only if $\Phi$ is $L^1$-removable. Indeed, suppose that $\Phi$ is
8146: $L^1$-removable, namely for every function $f\in L_{loc}^1(M)\cap {\rm
8147: CR}(M\backslash \Phi)$, and every $\mathcal{ C}^1$-smooth
8148: $(n,m-1)$-form with compact support, we have $\int_M \, f \cdot
8149: \overline{\partial} \psi =0$. In particular, since $L_{loc}^{\sf p}$
8150: is contained in $L_{loc}^1$ by H\"older's inequality, this property
8151: holds for every function $g\in L_{loc}^{\sf p}(M) \cap CR(M\backslash
8152: \Phi)$, hence $\Phi$ is $L^{\sf p}$-removable, as
8153: claimed. Consequently, it suffices to show that $\mathcal{
8154: W}$-removability implies $L^1$-removability.
8155: 
8156: Let $f\in L_{loc}^1(M\backslash \Phi)\cap L^1(M)$ be an arbitrary
8157: function. The goal is to show that $f$ is in fact CR on $\Phi$. Of
8158: course, it suffices to show that $f$ is CR locally at every point of
8159: $\Phi$. So, we fix an arbitrary point $q\in \Phi$. If $\psi$ is an
8160: arbitrary $(n,m-1)$-form of class $\mathcal{ C}^1$ supported in a
8161: sufficiently small neighborhood of $q$, we have to prove that $\int_M
8162: \, f \cdot \overline{\partial} \psi =0$.
8163: 
8164: We may also fix a small open polydisc $\mathcal{ V}_q$ centered at
8165: $q$. We shall first argue that we can assume that the $L_{loc}^1$
8166: function $f$ is holomorphic in a neighborhood of $\left(M\backslash
8167: \Phi\right)\cap \mathcal{ V}_q$ in $\C^n$. Indeed, since $M \backslash
8168: \Phi$ is globally minimal, there exists a wedge $\mathcal{ W}$
8169: attached to $M\backslash \Phi$ such that every $L_{loc}^1$ CR function
8170: on $M \backslash \Phi$, and in particular $f$, extends holomorphically
8171: as a function which belongs to the Hardy space $H^1(\mathcal{ W})$. By
8172: slightly deforming $\left(M\backslash \Phi\right) \cap \mathcal{ V}_q$
8173: into $\mathcal{ W}$ along Bishop discs glued to $M\backslash \Phi$,
8174: keeping $\Phi$ fixed, using the theory of Hardy spaces in wedges
8175: developed in~\cite{p1}, we may obtain the following deformation result
8176: with $L^1$ control, a statement which is a particular case of
8177: Proposition~1.16 in~\cite{mp1}.
8178: 
8179: \def\theproposition{11.4}\begin{proposition}
8180: For every $\varepsilon > 0$, every $\beta <\alpha$, there exists a
8181: small $\mathcal{ C}^{ 2, \beta}$-smooth deformation $M^d$ of $M$ with
8182: support contained in $\overline{ \mathcal{ V }_q}$ and there exists a
8183: function $f^d \in L_{ loc}^1 \left(M^d \right)\cap CR \left( M^d
8184: \backslash \Phi \right)$, such that
8185: \begin{itemize}
8186: \item[{\bf (1)}]
8187: $M^d \cap \mathcal{V}_q \supset \Phi \cap \mathcal{ V}_q\ni q$.
8188: \item[{\bf (2)}]
8189: $\left(M^d\backslash \Phi \right)\cap \mathcal{ V}_q 
8190: \subset \mathcal{ W}\cap \mathcal{ V}_q$.
8191: \item[{\bf (3)}]
8192: $f^d$ is holomorphic in the neighborhood
8193: $\mathcal{ W}\cap \mathcal{ V}_q$ of $(M^d\backslash \Phi)\cap 
8194: \mathcal{ V}_q$ in $\C^n$.
8195: \item[{\bf (4)}]
8196: $M\cap \mathcal{ V}_q$ and $M^d\cap \mathcal{ V}_q$ are graphed over
8197: the same $(2m+d)$ linear real subspace and $\left\vert\left\vert M^d\cap
8198: \mathcal{ V}_q - M \cap \mathcal{ V}_q
8199: \right\vert\right\vert_{\mathcal{ C}^{2,\beta}} \leq \varepsilon$.
8200: \item[{\bf (5)}]
8201: The volume forms of $M\cap \mathcal{ V}_q$ and of $M^d \cap \mathcal{
8202: V}_q$ may be identified and $\left\vert f-f^d
8203: \right\vert_{L^1(M\cap \mathcal{ V}_q)} \leq \varepsilon$. 
8204: \end{itemize}
8205: \end{proposition}
8206: 
8207: Since it will suffice to have a control of the
8208: deformation $M^d$ only in $\mathcal{ C}^2$ norm, we shall
8209: replace $\mathcal{ C}^{2,\beta}$ and $\mathcal{ C}^{1,\beta}$
8210: by $\mathcal{ C}^2$ and $\mathcal{ C}^1$ in the sequel.
8211: 
8212: Let us be more explicit about conditions {\bf (4)} and {\bf
8213: (5)}. Without loss of generality, we can assume that in coordinates
8214: $(z,w)= (x+iy, u+iv)\in \C^m\times \C^d$ centered at $q$, we have
8215: $T_qM= \{v=0\}$, hence the generic submanifolds $M$ and $M^d$ are
8216: represented locally by vectorial equations $v=\varphi(x,y,u)$ and $v=
8217: \varphi^d (x,y,u)$, where $\varphi$ and $\varphi^d$ are defined in the
8218: real cube $\I_{2m+d}(2\rho_1)$, for some small $\rho_1>0$ and that
8219: $\mathcal{ V}_q$ is the polydisc $\Delta_n(\rho_1)$ of radius
8220: $\rho_1$. Then condition {\bf (4)} simply means that $\vert \vert
8221: \varphi^d - \varphi \vert \vert_{\mathcal{ C}^2 (\I_{2m+d}
8222: (\rho_1) )}\leq \varepsilon$ and condition {\bf (5)} is clear if we
8223: choose $dxdydu$ as the volume form on $M$ and on $M^d$.
8224: 
8225: Suppose that for every $\varepsilon>0$ and for every 
8226: deformation $M^d$, we can show that
8227: the function $L_{loc}^1$ function $f^d$ on $M^d$ is in fact CR over
8228: $M^d\cap \Delta_n(\rho_1)$. Then we claim that $f$ is CR in a
8229: neighborhood of $q$.
8230: 
8231: Indeed, to begin with, let us denote by
8232: $\overline{L}_1,\dots,\overline{L}_m$ a basis of $(0,1)$ vector fields
8233: tangent to $M$, having coefficients depending on the first order
8234: derivatives of $\varphi$. More precisely, in slightly abusive matrix
8235: notation, we can choose the basis $\overline{L}:=
8236: \frac{\partial}{\partial \bar z}+2(i-\varphi_u)^{-1} \, \varphi_{\bar
8237: z} \frac{\partial}{\partial \bar w}$. Let us denote this basis
8238: vectorially by $\overline{L}=\frac{\partial}{\partial \bar z}+ A \,
8239: \frac{\partial}{\partial \bar w}$. To compute the formal adjoint of
8240: $\overline{L}$ with respect to the local Lebesgue measure $dx dy du$
8241: on $M$, we choose two $\mathcal{ C}^1$-smooth functions $\psi$, $\chi$
8242: of $(x,y,u)$ with compact support in $\I_{2m+d}(\rho_1)$. Then the
8243: integration by part $\int \overline{L} (\psi) \cdot \chi \cdot dxdydu=
8244: \int\, \psi\cdot {}^T \overline{L}
8245: (\chi)\cdot dx dy du$ yields the explicit
8246: expression ${}^T\overline{L} (\chi):= -\overline{L}(\chi)-A_{\bar w}
8247: \cdot \chi$ of the formal adjoint of $\overline{L}$.
8248: 
8249: It follows immediately that if we denote by ${}^T(\overline{L}^d)$ the
8250: formal adjoint of the basis of CR vector fields tangent to $M^d$, then
8251: we have an estimate of the form $\vert \vert {}^T(\overline{L}^d) -
8252: {}^T(\overline{L})\vert \vert_{ \mathcal{ C}^1} \leq {\rm C} \cdot
8253: \varepsilon$, for some constant $C>0$. Recall that $f^d$ is assumed to
8254: be CR in $M^d\cap \Delta_n(\rho_1)$. Equivalently, we have $\int \,
8255: f^d \cdot \, {}^T(\overline{L}^d)(\psi)\cdot dxdydu=0$ for every
8256: $\mathcal{ C}^1$-smooth function $\psi$ with compact support in the
8257: cube $\I_{2m+d}(\rho_1)$. Then we deduce that (some explanation
8258: follows)
8259: \def\theequation{11.5}\begin{equation}
8260: \left\{
8261: \aligned
8262: {}
8263: &
8264: \left\vert
8265: \int\, f\cdot {}^T\overline{L}(\psi) \cdot dxdydu \right\vert=
8266: \left\vert
8267: \int\, 
8268: \left[
8269: f\cdot {}^T\overline{L}(\psi)-
8270: f^d\cdot {}^T(\overline{L}^d)(\psi)
8271: \right] \cdot dxdydu \right\vert \\
8272: & \
8273: \leq 
8274: \left\vert
8275: \int\, 
8276: \left[
8277: f\cdot {}^T\overline{L}(\psi)-f\cdot {}^T(\overline{L}^d)(\psi)+
8278: f\cdot {}^T(\overline{L}^d)(\psi)-
8279: f^d\cdot {}^T(\overline{L}^d)(\psi)
8280: \right] \cdot dxdydu \right\vert
8281: \\
8282: & \
8283: \leq 
8284: C_1(\psi) \cdot \varepsilon \cdot 
8285: \int_{\I_{2m+d}(\rho_1)} \, 
8286: \vert f \vert \cdot dxdydu + C_2(\psi) \cdot
8287: \int_{\I_{2m+d}(\rho_1)}\, 
8288: \vert f - f^d \vert \cdot dxdydu
8289: \\
8290: & \
8291: \leq C(\psi,f,\rho_1)\cdot \varepsilon,
8292: \endaligned\right. 
8293: \end{equation}
8294: taking account of property {\bf (5)} of Proposition~11.4 for the
8295: passage from the third to the fourth line, where $C(\psi,f,\rho_1)$ is
8296: a positive constant depending only on $\psi$, $f$ and $\rho_1$. As
8297: $\varepsilon$ was arbitrarily small, it follows that $\int\, f\cdot
8298: {}^T\overline{L}(\psi) \cdot dxdydu =0 $ for every $\psi$, namely $f$
8299: is CR on $M\cap \Delta_n(\rho_1)$, as was claimed.
8300: 
8301: It remains to show that $f^d$ is CR on $M^d\cap \Delta_n(\rho_1)$.
8302: First of all, we need some observations. For every compactly
8303: supported small deformation $M^d$ stabilizing $\Phi$, the wedge
8304: $\mathcal{ W}$ attached to $M\backslash \Phi$ is still a wedge
8305: attached to $M^d\backslash \Phi$. In addition, this wedge contains a
8306: neighborhood of $\left(M^d \backslash \Phi\right)\cap
8307: \Delta_n(\rho_1)$ in $\C^n$ by property {\bf (3)} of Proposition~11.4.
8308: As $\Phi$ was supposed to be stably removable, it follows that there
8309: exists a wedge $\mathcal{ W}_1$ attached to $M^d$ (including points of
8310: $\Phi$) to which holomorphic functions in $\mathcal{ W}$ extend
8311: holomorphically.
8312: 
8313: Consequently, replacing $M^d\cap \Delta_n(\rho_1)$ by $M$, we are led
8314: to prove the following lemma, which, on the geometric side, is totally
8315: similar to Lemma~11.3, except that the wedge $\mathcal{ W}$ attached
8316: to $M\backslash \Phi$ appearing in the formulation of Lemma~11.3 is
8317: now replaced by a neighborhood $\Omega$ of $M\backslash \Phi$ in
8318: $\C^n$.
8319: 
8320: \def\thelemma{11.6}\begin{lemma} 
8321: Let $M$ be a $\mathcal{ C}^{2, \alpha}$-smooth generic submanifold of
8322: $\C^n$ of CR dimension $m \geq 1$ and of codimension $d= n-m\geq 1$, 
8323: let $\Phi \subset M$ be a nonempty
8324: proper closed subset whose $(2m +d)$-dimensional Hausdorff measure is
8325: equal to zero. Let $\Omega$ be a neighborhood of $M\backslash \Phi$ in
8326: $\C^n$ and let $\mathcal{ W}_1$ be a wedge attached to $M$, including
8327: points of $\Phi$.
8328: Let $f\in L_{loc}^1(M)$ and assume that its restriction to
8329: $M\backslash \Phi$ extends as a holomorphic function $f'\in \mathcal{
8330: O}(\Omega \cup \mathcal{ W}_1)$. Then $f$ is CR all over $M$.
8331: \end{lemma}
8332: 
8333: \proof
8334: It suffices to prove that $f$ is CR at every point of $\Phi$. Let
8335: $q\in \Phi$ be arbitrary and let $\mathcal{ W}_q$ be a local wedge of
8336: edge $M$ at $q$ which is contained in $\mathcal{ W}_1$. Without loss
8337: of generality, we can assume that in coordinates
8338: $(z,w)=(x+iy,u+iv)\in\C^m\times \C^d$ vanishing at $q$ with
8339: $T_qM=\{v=0\}$, the generic submanifold $M$ is represented locally in
8340: the polydisc $\Delta_n(\rho_1)$ by $v=\varphi(x,y,u)$ for some
8341: $\mathcal{ C}^{2,\alpha}$-smooth $\R^d$-valued mapping $\varphi$
8342: defined on the real cube on $\I_{2m+d}(\rho_1)$. First of all, we
8343: construct a family of analytic discs half attached to $M$ whose
8344: interior is contained in the local wedge $\mathcal{ W}_q\subset
8345: \mathcal{ W}_1$.
8346: 
8347: \def\thelemma{11.7}\begin{lemma}
8348: There exists a family of analytic
8349: discs $A_s(\zeta)$, with $s\in \R^{2m+d-1}$, $\vert s \vert \leq
8350: 2\delta$ for some $\delta>0$, and $\zeta \in\overline{\Delta}$, which
8351: is of class $\mathcal{ C}^{2,\alpha-0}$ with respect to all variables,
8352: such that
8353: \begin{itemize}
8354: \item[{\bf (1)}]
8355: $A_0(1)=q$.
8356: \item[{\bf (2)}]
8357: $A_s(\overline{\Delta})\subset \Delta_n(\rho_1)$. 
8358: \item[{\bf (3)}]
8359: $A_s(\Delta) \subset \mathcal{ W}_q\cap \Delta_n(\rho_1)$.
8360: \item[{\bf (4)}]
8361: $A_s(\partial^+\Delta)\subset M$.
8362: \item[{\bf (5)}]
8363: $A_s(i)\in M\backslash \Phi$ and $A_s(-i)\in M\backslash\Phi$
8364: for all $s$. 
8365: \item[{\bf (6)}]
8366: The mapping
8367: $[-2\delta,2\delta]^{2m+d-1}\times [-\pi/2,\pi/2]
8368: \ni (s,\theta)\longmapsto 
8369: A_s(e^{i\theta})\in M$ is an embedding onto 
8370: a neighborhood of $q$ in $M$.
8371: \item[{\bf (7)}]
8372: There exists $\rho_2>0$ such that the image of
8373: $[- \delta, \delta]^{ 2m+d-1} \times [-\pi /4,\pi /4]$ through 
8374: this mapping contains $M \cap \Delta_n (\rho_2)$.
8375: \end{itemize}
8376: \end{lemma} 
8377: 
8378: \proof
8379: Let $M^1$ be a $\mathcal{ C}^{2,\alpha}$-smooth maximally real
8380: submanifold of $M$ passing through $q$ such that $M^1\cap \Phi$ is of
8381: zero measure with respect to the Lebesgue measure of $M^1$. Let $t\in
8382: \R^d$ and include $M^1$ in a parametrized family of maximally real
8383: submanifolds $M_t^1$ which foliates a neighborhood of $q$ in $M$.
8384: Starting with a family of analytic discs $A_{c,x,v}^1(\zeta)$ which
8385: are half-attached to $M^1$ as constructed in Lemma~7.12 above, 
8386: we first choose the rotation
8387: parameter $v_0$ and a sufficiently small scaling factor $c_0$ in order
8388: that $A_{c_0,0,v_0}^1(\pm i)$ does not belong to $\Phi$. In fact, this
8389: can be done for almost every $(c_0,v_0)$, because the
8390: mapping $(c,v)\mapsto A_{c,0,v}^1(\pm i)$ is of rank $n$ at
8391: every point $(c,v)$ with $c\neq 0$ and $v\neq 0$. In addition, we
8392: adjust the rotation parameter $v_0$ in order that the vector $Jv_0$
8393: points inside a proper subcone of the cone which defines the wedge
8394: $\mathcal{W}_q$. If the scaling parameter $c$ is sufficiently small,
8395: this implies that $A_{c_0,0,v_0}^1(\Delta)$ is contained in $\mathcal{
8396: W}_q\cap \Delta_n(\rho_1)$, as in Lemma~8.3 above. The translation
8397: parameter $x$ runs in $\R^n$ and we may select a $(n-1)$-dimensional
8398: parameter subspace $x'$ which is transversal in $M^1$ to the half
8399: boundary $A_{c_0, 0,v_0}^1(\partial^+ \Delta)$. With such a choice,
8400: there exists $\delta>0$ such that the mapping $[-2 \delta,
8401: 2\delta]^{n-1}\times [-\pi/2, \pi/2 ]\ni (x', \theta) \longmapsto
8402: A_{c_0, x', v_0}^1(e^{i\theta})$ is a diffeomorphism onto a
8403: neighborhood of $q$ in $M^1$. Finally, using the stability of Bishop's
8404: equation under perturbations, we can deform this family of discs by
8405: requiring that it is half attached to $M_t^1$, thus obtaining a family
8406: $A_s(\zeta) := A_{c_0, x',v_0, t}^1( \zeta)$ with $s:= (x',t) \in\R^{
8407: 2m+d-1}$. Shrinking $\delta$ if necessary, we can check as
8408: in the proof of Lemma~8.3 {\bf ($\mathbf{9_1}$)} that
8409: condition {\bf (5)} holds. This completes the proof.
8410: \endproof
8411: 
8412: Let now $f \in L_{ loc}^1 (M)$ and let $f' \in \mathcal{ O} ( \Omega
8413: \cup \mathcal{ W }_1)$. Thanks to the foliation propery {\bf (6)} of
8414: Lemma~11.7, it follows from Fubini's theorem that for almost every
8415: translation parameter $s$, the mapping $e^{ i\theta} \mapsto f \left(
8416: A_s (e^{ i\theta })\right)$ defines a $L^1$ function on $\partial^+
8417: \Delta$. In addition, the restriction of the function $f' \in
8418: \mathcal{ O}( \Omega\cup \mathcal{ W}_1)$ to the disc $A_s( \Delta)
8419: \subset \mathcal{ W}_q \subset \mathcal{ W}_1$ yields a holomorphic
8420: function $f' \left( A_s( \zeta) \right)$ in $\Delta$.
8421: 
8422: \def\thelemma{11.8}\begin{lemma}
8423: For almost every $s$ with $\vert s \vert \leq 2 \delta$, the function
8424: $f' \left( A_s (\zeta) \right)$ belongs to the Hardy space $H^1 (
8425: \Delta )$.
8426: \end{lemma}
8427: 
8428: \proof
8429: Indeed, for almost every $s$, the restriction $f\left (A_s(e^{
8430: i\theta})\right)$ belongs to $L^1(\partial^+ \Delta)$. We can also
8431: assume that for almost every $s$, the intersection $\Phi \cap
8432: A_s(\partial^+ \Delta)$ is of zero one-dimensional measure. By the
8433: assumption of Lemma~11.6, the restriction of $f\circ A_s$ and of
8434: $f'\circ A_s$ to $\partial^+\Delta \backslash \Phi$ coincide. Recall
8435: that $\partial^-\Delta=\{\zeta\in \partial \Delta: \, {\rm Re}\, \zeta
8436: \leq 0\}$. Since $A_s( \pm i)$ does not belong to $\Phi$ and since
8437: $A_s \left(e^{i\theta}\right)$ belongs to $\mathcal{ W}_q$ for all
8438: $\theta$ with $\pi/2 < \vert \theta \vert \leq \pi$, it follows that
8439: $f \circ A_s \vert_{\partial^+ \Delta}$ and $f' \circ A_s \vert_{
8440: \partial^-\Delta}$ (which is holomorphic in a neighborhood of
8441: $\partial^-\Delta$ in $\C$) match together in a function which is
8442: $L^1$ on $\partial \Delta$. Let us denote this function by $f_s$.
8443: Furthermore, $f_s$ extends holomorphically to $\Delta$ as $f'\circ A_s
8444: \vert_\Delta$. Consequently, $f'\circ A_s\vert_\Delta$ belongs to the
8445: Hardy space $H^1( \Delta)$, which proves the lemma.
8446: \endproof
8447: 
8448: Since we have now established that the boundary value of
8449: $f'$ on $M\backslash \Phi$ along the family
8450: of discs $A_s(\zeta)$ coincides with $f$, we can
8451: now denote both functions by the same letter $f$. 
8452: 
8453: For $\varepsilon\geq 0$ small, let now $\chi_\varepsilon \left(s,e^{i
8454: \theta} \right)$ be a $\mathcal{ C}^2$-smooth function on $[ -2\delta,
8455: 2\delta]\times \partial \Delta$ which equals $\varepsilon$ for $\vert
8456: s \vert \leq \delta$ and for $\theta \in [-\pi/4,\pi/4]$ and which
8457: equals $0$ if either $\pi/2 \leq \vert \theta \vert \leq \pi$ or
8458: $\vert s \vert \geq 2\delta/3$. We may require in addition that $\vert
8459: \vert \chi_\varepsilon \vert \vert_{\mathcal{ C}^2}\leq
8460: \varepsilon$. We define a deformation $M^\varepsilon$ of $M$ compactly
8461: supported in a neighborhood of $q$ by pushing $M$ inside $\mathcal{
8462: W}_q$ along the family of discs $A_s(\zeta)$ as follows:
8463: \def\theequation{11.9}\begin{equation}
8464: M^\varepsilon:= \{A_s\left([1-\chi_\varepsilon (s,e^{i\theta})]\, 
8465: e^{i\theta}\right): \, 
8466: \vert \theta \vert \leq \pi/2, \ 
8467: \vert s \vert \leq 2\delta\}.
8468: \end{equation}
8469: Notice that $M^\varepsilon$ coincides with $M$ outside a small
8470: neighborhood of $q$. Then we have $\vert \vert M^\varepsilon - M \vert
8471: \vert_{\mathcal{ C}^2}\leq C\cdot \varepsilon$, for some constant
8472: $C>0$ which depends only on the $\mathcal{ C}^2$ norms of $A_s(\zeta)$
8473: and of $\chi_\varepsilon(s,e^{i\theta})$. If the radius $\rho_2$ is as
8474: in Property {\bf (7)} of Lemma~11.7 above, the deformation
8475: $M^\varepsilon \cap \Delta_n(\rho_2)$ is entirely contained in
8476: $\mathcal{ W}_q$ and since $f$ is holomorphic in $\mathcal{ W}_q$,
8477: its restriction to $M^\varepsilon \cap \Delta_n (\rho_2)$ is obviously
8478: CR. An illustration is provided in the right hand side of the
8479: following figure.
8480: 
8481: \bigskip
8482: \begin{center}
8483: \input M-deformation-discs.pstex_t
8484: \end{center}
8485: 
8486: As in~\cite{j5}, \cite{p1}, \cite{mp1}, we notice that for every $s$
8487: and every $\varepsilon$, the one-dimensional Lebesgue measure on the
8488: arc 
8489: \def\theequation{11.10}\begin{equation}
8490: \Gamma_{\varepsilon,s}:=\{[1-\chi_\varepsilon
8491: (s,e^{i\theta})]e^{i\theta}\in \Delta: \vert
8492: \theta \vert \leq \pi\}
8493: \end{equation}
8494: is a Carleson measure. Thanks to the geometric uniformity of these
8495: arcs $\Gamma_{\varepsilon,s}$, it follows from an inspection of the
8496: proof of L.~Carleson's imbedding theorem that there exists a (uniform)
8497: constant $C$ such that for all $s$ with $\vert s \vert \leq 2\delta$
8498: and all $\varepsilon$, one has the estimate
8499: \def\theequation{11.11}\begin{equation}
8500: \int_{\Gamma_{\varepsilon,s}}\, 
8501: \left\vert f\left(
8502: A_s\left(
8503: \left[
8504: 1-\chi_\varepsilon(s,e^{i\theta})
8505: \right]\, e^{i\theta}
8506: \right)
8507: \right)\right\vert \cdot d\theta
8508: \,
8509: \leq \, C 
8510: \int_{\partial\Delta}\, 
8511: \vert f \vert \cdot d\theta.
8512: \end{equation} 
8513: 
8514: We are now ready to complete the proof of Lemma~11.6. Let
8515: $\pi_{x,y,u}$ denote the projection parallel to the $v$-space onto the
8516: $(x, y,u)$-space. The mapping $(s, \theta) \mapsto \pi_{x, y,u }\left(
8517: A_s ( \theta) \right)$ may be used to define new coordinates in a
8518: neighborhood of the origin in $\C^m\times \R^d$, an open subset above
8519: which $M$ and $M^\varepsilon$ are graphed. We shall now work with
8520: these coordinates. With respect to the coordinates $(s,\theta)$, on
8521: $M$ and on $M^\varepsilon$, we have formal adjoints ${}^T \overline{
8522: L}$ and ${}^T( \overline{L }^\varepsilon)$ of the basis of CR vector
8523: fields with an estimation of the form $\left\vert \left\vert {}^T(
8524: \overline{ L }^\varepsilon)- {}^T \overline{ L} \right \vert \right
8525: \vert_{ \mathcal{ C}^1}\leq C \cdot \varepsilon$, for some constant
8526: $C>0$. Let now $\psi= \psi (s,\theta)$ be $\mathcal{ C}^1$-smooth
8527: function with compact support in the set $\{\vert s \vert < \delta, \,
8528: \vert \theta \vert \leq \pi/4\}$. By construction, the subpart of
8529: $M^\varepsilon$ defined by $\widetilde{ M }^\varepsilon:= \{A_s
8530: \left([1-\chi_\varepsilon (s,e^{ i \theta })]\, e^{ i\theta} \right):
8531: \, \vert \theta \vert \leq \pi/4, \ \vert s \vert \leq \delta\}$ is
8532: contained in the wedge $\mathcal{ W}_q$, hence the restriction of the
8533: holomorphic function $f\in \mathcal{ W}_q$ to $\widetilde{
8534: M}^\varepsilon$ is obviously CR on $\widetilde{ M}^\varepsilon$.
8535: 
8536: For simplicity of notation, we shall denote $f \left(A_s(e^{ i\theta})
8537: \right)$ by $f_s(\theta)$ and $f\left( A_s \left( \left[ 1-
8538: \chi_\varepsilon(s, e^{i\theta }) \right]\, e^{ i\theta} \right)
8539: \right)$ by $f_s^\varepsilon ( \theta)$. Since by construction
8540: for every $\varepsilon >0$, the $L^1$ function $(s,\theta)\mapsto
8541: f_s^\varepsilon (\theta)$ is annihilated in the distributional sense
8542: by the CR vector fields $\overline{ L}^\varepsilon$ on $\widetilde{
8543: M}^\varepsilon$, we may compute (not writing the arguments $(s,
8544: \theta)$ of $\psi$)
8545: \def\theequation{11.12}\begin{equation}
8546: {\small
8547: \left\{
8548: \aligned
8549: {}
8550: &
8551: \left\vert
8552: \int_{\vert s \vert \leq \delta}\,\int_{\vert
8553: \theta \vert \leq \pi/4}\, 
8554: f_s(\theta)\cdot 
8555: {}^T\overline{L}(\psi) \cdot dsd\theta \right\vert=\\
8556: & \
8557: =\left\vert
8558: \int_{\vert s \vert \leq \delta}\,\int_{\vert
8559: \theta \vert \leq \pi/4}\, 
8560: \left[
8561: f_s(\theta)
8562: \cdot {}^T\overline{L}(\psi)-
8563: f_s^\varepsilon(\theta)
8564: \cdot {}^T(\overline{L}^\varepsilon)(\psi)
8565: \right] \cdot dsd\theta \right\vert \\
8566: & \
8567: \leq 
8568: \left\vert
8569: \int_{\vert s \vert \leq \delta} \left(
8570: \int_{\vert \theta \vert \leq \pi/4}\,
8571: \left[
8572: f_s(\theta)\cdot {}^T\overline{L}(\psi)-f_s(\theta)\cdot 
8573: {}^T(\overline{L}^\varepsilon)(\psi)+ \right.\right.\right.\\
8574: & \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \
8575: \ \ \ \ \ \ \
8576: \left.\left.\left.
8577: +f_s(\theta)\cdot {}^T(
8578: \overline{L}^\varepsilon)(\psi)-
8579: f_s^\varepsilon(\theta)
8580: \cdot {}^T(\overline{L}^\varepsilon)(\psi)
8581: \right] \cdot d\theta \right)\cdot ds \right\vert
8582: \\
8583: & \
8584: \leq 
8585: C_1(\psi) \cdot \varepsilon \cdot 
8586: \int_{\vert s \vert \leq \delta}\,\int_{\vert
8587: \theta \vert \leq \pi/4}\, 
8588: \vert f_s(\theta) \vert \cdot dsd\theta + \\
8589: & \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ 
8590: +C_2(\psi) \cdot
8591: \int_{\vert s \vert \leq \delta}\,\int_{\vert
8592: \theta \vert \leq \pi/4}\,
8593: \vert f_s(\theta) - f_s^\varepsilon(\theta) \vert \cdot dsd\theta
8594: \\
8595: & \
8596: \leq C_1(\psi,f,\delta)\cdot \varepsilon+
8597: C_2(\psi,\delta)\cdot \max_{\vert s \vert \leq \delta}\, \int_{\vert
8598: \theta \vert \leq \pi/4} \, 
8599: \vert f_s(\theta) - f_s^\varepsilon(\theta) \vert \cdot dsd\theta.
8600: \endaligned\right. 
8601: }
8602: \end{equation}
8603: However, thanks to the estimate~\thetag{11.11} and thanks to
8604: Lebesgue's dominated convergence theorem, the last integral tends to
8605: zero as $\varepsilon$ tends to zero. It follows that the integral in
8606: the first line of~\thetag{11.12} can be made arbitrarily small, hence
8607: it vanishes. This proves that $f$ is CR in a neighborhood
8608: of $q$ and completes the proof of Lemma~11.6.
8609: \endproof
8610: 
8611: The proof of Lemma~11.3 is complete.
8612: \endproof
8613: 
8614: \section*{\S12.~Proofs of Theorem~1.1 and of Theorem~1.3}
8615: 
8616: \subsection*{12.1.~Tree of separatrices linking hyperbolic points}
8617: Let $M\subset \C^2$ be a globally minimal $\mathcal{
8618: C}^{2,\alpha}$-smooth hypersurface, let $S\subset M$ be a $\mathcal{
8619: C}^{2,\alpha}$-smooth open surface (without boundary) and let
8620: $K\subset S$ be a proper compact subset of $S$. Assume that $S$ is
8621: totally real outside a discrete subset of complex tangencies which are
8622: hyperbolic in the sense of E.~Bishop. Since we aim to remove the
8623: compact subset $K$ of $S$, we can shrink the open surface
8624: $S$ around $K$ in order that $S$ contains {\it 
8625: only finitely many}\, such hyperbolic
8626: complex tangencies, which we shall
8627: denote by $\{h_1,\dots,h_\lambda\}$, where
8628: $\lambda$ is some integer, possibly zero. Furthermore, we can
8629: assume that $\partial S$ is of class $\mathcal{ C}^{2,\alpha}$. As a
8630: corollary of the qualitative theory of planar vector fields, due to
8631: H.~Poincar\'e and I.~Bendixson, we know that
8632: \begin{itemize}
8633: \item[{\bf (i)}]
8634: The hyperbolic points $h_1, \dots, h_\lambda$ are singularities of the
8635: characteristic foliation $\mathcal{ F}_S^c$.
8636: \item[{\bf (ii)}]
8637: Incoming to every hyperbolic point $h_1,\dots,h_\lambda$, there
8638: are exactly four $\mathcal{ C}^{2,\alpha}$-smooth open {\sl
8639: separatrices} (to be defined precisely below).
8640: \item[{\bf (iii)}]
8641: After perturbing slightly the boundary $\partial S$ if necessary,
8642: these separatrices are all transversal to $\partial S$ and the union of
8643: all separatrices together with all hyperbolic points makes a {\sl
8644: finite tree without cycles in $S$} (to be defined below).
8645: \end{itemize} 
8646: 
8647: Precisely, by an (open) {\sl separatrix}, we mean a $\mathcal{
8648: C}^{2,\alpha}$-smooth curve $\tau: (0,1) \to S$ with $\frac{ d
8649: \tau}{ds}(s) \in T_{\tau(s)}S \cap T_{\tau(s)}^c M\backslash \{0\}$
8650: for every $s\in (0,1)$, namely its tangent vectors are all nonzero and
8651: characteristic, such that one limit point, say $\lim_{s\to 0} \,
8652: \tau(s)$ is a hyperbolic point, and the other $\lim_{s\to 1} \,
8653: \tau(s)$ either belong to the boundary $\partial S$ or is a second
8654: hyperbolic point.
8655: 
8656: From the local study of saddle phase diagrams ({\it cf.}~\cite{
8657: ha}), we get in addition:
8658: 
8659: \begin{itemize}
8660: \item[{\bf (iv)}] 
8661: There exists $\varepsilon >0$ and for every $l = 1, \dots, \lambda$,
8662: there exist two curves $\gamma_l^1, \gamma_l^2 : (- \varepsilon,
8663: \varepsilon ) \to S$ which are of class $\mathcal{ C }^{ 1, \alpha}$,
8664: {\it not more}, with $\gamma_l^i (0) = h_l$ and $\frac{ d \gamma_l^1
8665: }{ dt} (s) \in T_{ \gamma_l (s)} S \cap T_{ \gamma_l(s) }^c M
8666: \backslash \{0\}$ for every $s \in (- \varepsilon, \varepsilon)$ and
8667: for $i= 1,2$, such that the four open segments $\gamma_l^1 (
8668: -\varepsilon,0)$, $\gamma_l^1 ( 0, \varepsilon )$, $\gamma_l^2 (
8669: -\varepsilon, 0)$ and $\gamma_l^2 ( 0, \varepsilon )$ cover the four
8670: pieces of open separatrices incoming at $h_l$.
8671: \end{itemize}
8672: 
8673: Let $\tau_1,\dots,\tau_\mu: (0, 1) \to S$ denote all the separatrices of
8674: $S$, where $\mu$ is some integer, possibly equal to zero. By the
8675: {\sl finite hyperbolic tree $T_S$ of $S$}, we mean:
8676: \def\theequation{12.2}\begin{equation}
8677: T_S := \{h_1,\dots,h_\lambda\} 
8678: \bigcup_{1\leq k\leq \mu} \ \tau_k(0,1).
8679: \end{equation} 
8680: We say that $T_S$ has {\sl no cycle} if it does not contain any subset
8681: homeomorphic to the unit circle. For instance, in the case where
8682: $S\equiv D$ is diffeomorphic to a real disc (as in the assumptions of
8683: Theorem~1.1), its hyperbolic tree $T_D$ necessarily has no
8684: cycle. However, in the case where $S$ is an annulus (for instance),
8685: there is a trivial example of a characteristic foliation with two
8686: hyperbolic points and a circle in the hyperbolic tree.
8687: 
8688: \subsection*{12.3.~Hyperbolic decomposition in the disc case}
8689: Let the real disc $D$ and the compact subset $K \subset D$ be as in
8690: Theorem~1.1. As in \S12.1 just above, we shrink $D$ slightly and
8691: smooth out its boundary, so that its hyperbolic tree
8692: $T_D$ is finite and has
8693: no cycle. We may decompose $D$ as the disjoint union
8694: \def\theequation{12.4}\begin{equation}
8695: D= T_D \cup D_o,
8696: \end{equation}
8697: where the complement of the hyperbolic tree $D_o: = D \backslash T_D$
8698: is an open subset of $D$ entirely contained in the totally real part
8699: of $D$. Then $D_o$ has finitely many connected
8700: components $D_1,\dots,D_\nu$, the {\sl hyperbolic sectors of $D$}.
8701: Then, for $j=1,\dots,\nu$, we define the proper closed subsets $C_j:=
8702: D_j \cap K$ of $D_j$ as illustrated in the left hand side of the
8703: following figure.
8704: 
8705: \bigskip
8706: \begin{center}
8707: \input hyperbolic.pstex_t
8708: \end{center}
8709: 
8710: Again from H.~Poincar\'e and I.~Bendixson's 
8711: theory, we know that for every
8712: component $D_j$ (in which the characteristic foliation is
8713: nonsingular), the proper closed subset $C_j$ satisfies the nontransversality
8714: condition $\mathcal{ F}_{ D_j}^c\{ C_j\}$ formulated in Theorem~1.2.
8715: In {\sc Figure~20} just above, we have drawn the characteristic curves
8716: only for the two sectors $D_4$ and $D_6$. One may observe that
8717: $\mathcal{ F}_{D_4}\{ C_4\}$ and $\mathcal{ F}_{D_6}\{ C_6\}$ hold
8718: true. Also, $K\cap T_D$ is a proper closed subset of the hyperbolic
8719: tree of $D$.
8720: 
8721: \subsection*{12.5.~Global minimality of some complements}
8722: Before proceeding to the deduction of Theorems~1.1 and~1.3 from
8723: Theorem~1.2, we must verify that the complement $M\backslash K$ is
8724: also globally minimal. Here, we state a generalization of Lemma~3.5 to
8725: the case where some hyperbolic complex tangencies are allowed. Its
8726: proof is not immediate.
8727: 
8728: \def\thelemma{12.6}\begin{lemma}
8729: Let $M$ be a $\mathcal{ C}^{2,\alpha}$-smooth hypersurface in $\C^2$
8730: and let $S\subset M$ be $\mathcal{ C}^{2,\alpha}$-smooth surface which
8731: is totally real outside a discrete subset of hyperbolic complex
8732: tangencies. Assume that the hyperbolic tree $T_S$ of $S$ has no cycle.
8733: Then for every compact subset $K\subset S$ and for
8734: an arbitrary point $p\in
8735: M\backslash K$, its CR orbit in $M\backslash K$ coincides with its CR
8736: orbit in $M$, minus $K$, namely
8737: \def\theequation{12.7}\begin{equation}
8738: \mathcal{ O}_{CR}(M\backslash K, p)= 
8739: \mathcal{ O}_{CR}(M, p) \backslash K.
8740: \end{equation}
8741: \end{lemma}
8742: 
8743: \proof
8744: Of course, we may assume that $S$ coincides with the shrinking of a
8745: slightly larger surface and has finitely many hyperbolic points
8746: $\{h_1,\dots,h_\lambda\}$, as described in \S12.1, with the same
8747: notation. Let $K_{T_S}:= K \cap T_S$ be the track of $K$ on the
8748: hyperbolic tree $T_S$. Since the intersection of $K_{T_S}$ with any
8749: open separatrix may in general coincide
8750: with any arbitrary closed subset of an
8751: interval, in order to fix ideas, it will be convenient to deal with an
8752: enlargement $\overline{K}$ of $K_{T_S}$, simply defined by
8753: filling the possible holes of $K_{T_S}$ in $T_S$: more
8754: precisely, $\overline{K}$
8755: should contain all hyperbolic points together with all separatrices
8756: joining them and for every separatrix $\tau_k(0,1)$ with right limit
8757: point $\lim_{s\to 1} \, \tau_k(s)$ belonging to the boundary of $S$,
8758: we require that $\overline{K}$ contains the segment $\tau_k[0, r_1]$,
8759: where $r_1<1$ is close enough to $1$ in order that $\overline{ K}$
8760: effectively contains $K_{T_S}$. 
8761: 
8762: Obviously, from the inclusions
8763: \def\theequation{12.8}\begin{equation}
8764: K_{T_S} \subset \overline{ K} \subset K,
8765: \end{equation}
8766: we deduce that for every point $p\in M\backslash K$, we have
8767: the reverse inclusions
8768: \def\theequation{12.9}\begin{equation}
8769: \mathcal{ O}_{CR}(M\backslash K,p) \subset
8770: \mathcal{ O}_{CR}(M\backslash \overline{ K},p) \subset
8771: \mathcal{ O}_{CR}(M\backslash K_{T_S},p).
8772: \end{equation}
8773: The main step in the proof of Lemma~12.6 will be to establish the
8774: following two assertions, implying the third, desired
8775: assertion, already stated as~\thetag{12.7}.
8776: 
8777: \begin{itemize}
8778: \item[{\bf (A1)}]
8779: For every point $q\in M\backslash \overline{ K}$, we have
8780: $\mathcal{ O}_{CR}(M\backslash \overline{ K}, q)= 
8781: \mathcal{ O}_{CR}(M, q) \backslash \overline{K}$.
8782: \item[{\bf (A2)}]
8783: For every point $r\in M\backslash K_{T_S}$, we have $\mathcal{
8784: O}_{CR}(M\backslash K_{T_S}, r)= \mathcal{ O}_{CR}(M, r) \backslash
8785: K_{T_S}$.
8786: \end{itemize}
8787: 
8788: Indeed, taking these two assertions for granted, let us conclude the
8789: proof of Lemma~12.6. Let $p\in M\backslash K$ and decompose $K$ as a
8790: disjoint union $K= K_{T_S}\cup C'$, where $C':= K\backslash K_{T_S}$
8791: is a relatively closed subset of the hypersurface $M':= M\backslash
8792: K_{T_S}$. Notice that $C'$ is contained in the totally real 
8793: part of $S$.
8794: Again thanks to foliation theory, we see that
8795: the assumption that $K_{T_S}$ does not contain any cycle
8796: entails that $C'$ does not contain maximal characteristic lines of the
8797: totally real surface $S\backslash K_{T_S}$. Consequently, all the
8798: assumptions of Lemma~3.5 are satisfied, hence by applying it to $p$, we
8799: deduce that $\mathcal{ O}_{CR}(M' \backslash C', p) = \mathcal{
8800: O}_{CR}(M', p) \backslash C'$. By developing in length this identity
8801: between sets, we get
8802: \def\theequation{12.10}\begin{equation}
8803: \aligned
8804: \mathcal{ O}_{CR}(M\backslash K, p)
8805: & \ 
8806: = 
8807: \mathcal{ O}_{CR}\left((M\backslash K_{T_S})
8808: \backslash C', p\right) \\
8809: & \
8810: =
8811: \left[\mathcal{ O}_{CR}(M\backslash K_{T_S},p)
8812: \right]\backslash C' \\
8813: & \ 
8814: = 
8815: \left[\mathcal{ O}_{CR}(M, p)\backslash K_{T_S}
8816: \right]\backslash C' \\
8817: & \ 
8818: =
8819: \mathcal{ O}_{CR}(M,p) \backslash K,
8820: \endaligned
8821: \end{equation}
8822: where, for the passage from the second to the third line, we
8823: use {\bf (A2)}. This is~\thetag{ 12.7}, as desired. 
8824: 
8825: Thus, the (main) remaining task is to establish the 
8826: assertion {\bf (A2)}, with {\bf (A1)} being a preliminary step.
8827: 
8828: First of all, we show how to deduce {\bf (A2)} from {\bf (A1)}. Pick
8829: an arbitrary point $r \in M\backslash K_{ T_S}$. Of course, we have
8830: the trivial inclusion $\mathcal{ O}_{ CR}(M \backslash K_{ T_S},r)
8831: \subset \mathcal{ O}_{CR} (M, r) \backslash K_{ T_S}$ and we want an
8832: equality. As $\overline{ K}$ contains $K_{ T_S}$, we have either $r
8833: \in M \backslash \overline{ K}$ (first case) or $r \in \overline{ K}
8834: \backslash K_{T_S}$ (second case), {\it see}\, {\sc Figure~21} just
8835: below, where $r$ is located in $\overline{ K}\backslash K_{T_S}$.
8836: 
8837: \bigskip
8838: \begin{center}
8839: \input kts.pstex_t
8840: \end{center}
8841: 
8842: Since the second case makes it impossible to apply {\bf (A1)}, we need
8843: to find another point $r'$ in the CR orbit of $r$ in $M\backslash
8844: K_{T_S}$ such that $r'$ belongs to $M\backslash \overline{ K}$. This
8845: is elementary. We make a dichotomy: either $r=h_l$ is a hyperbolic
8846: point or it belongs to an open separatrix $\tau_k(0,1)$. If $r=h_l\in
8847: M\backslash K_{T_S}$ is a hyperbolic point, we may use one of the two
8848: $\mathcal{ C}^{2,\alpha}$-smooth complex tangent curves $\gamma_l^1$
8849: or $\gamma_l^2$ passing through $h_l$ and running in $M\backslash
8850: K_{T_S}$ to join $r$ with another point which belongs to an open
8851: separatrix and which obviously lies in the same CR orbit $\mathcal{
8852: O}_{CR }( M \backslash K_{T_S}, r)$. Hence, we may assume that $r\in
8853: \overline{ K} \backslash K_{T_S}$ belongs to an open separatrix. Since
8854: $S$ is now maximally real near $r$, we may choose a $T^cM$-tangent
8855: vector field $Y$ defined in a neighborhood of $r$ which is transversal
8856: to $S$ at $r$. Then for all $\delta>0$ small enough, the point $r':
8857: =\exp (\delta Y)(r)$ is outside $S$, hence does not belong to
8858: $\overline{ K}$ and clearly lies in the same CR orbit $\mathcal{ O
8859: }_{CR }( M \backslash K_{ T_S}, r)$.
8860: 
8861: In summary, when $r \in \overline{ K} \backslash K_{ T_S}$, we have
8862: exhibited a point $r'\in \mathcal{ O }_{CR}(M \backslash K_{T_S}, r)$
8863: with $r '\in M \backslash \overline{ K}$ so that it suffices now to
8864: show that for every point $r\in M \backslash \overline{ K}$, we have
8865: $\mathcal{ O}_{ CR}(M \backslash K_{T_S}, r)= \mathcal{ O}_{ CR} (M,r)
8866: \backslash K_{ T_S}$.
8867: 
8868: Using a trivial inclusion and
8869: applying {\bf (A1)}, we deduce that
8870: \def\theequation{12.11}\begin{equation}
8871: \mathcal{ O}_{CR}(M\backslash K_{T_S}, r) \supset
8872: \mathcal{ O}_{CR}(M\backslash \overline{ K}, r) = 
8873: \mathcal{ O}_{CR}(M, r) \backslash \overline{ K}.
8874: \end{equation}
8875: Unfortunately, there may well exist points of $\mathcal{ O}_{CR}(M,
8876: r)$ belonging to $\overline{ K} \backslash K_{T_S}$, so that it
8877: remains to show that {\it every}\, point $r'\in \mathcal{ O}_{CR}(M,
8878: r)\cap [ \overline{ K} \backslash K_{T_S}]$ also belong to $\mathcal{
8879: O}_{ CR}(M \backslash K_{T_S}, r)$. Again, this last step is
8880: elementary and totally analogous to the above argument: we first claim
8881: that we can join such a point $r'$ to a point $r'''\in M\backslash
8882: \overline{ K}$ by means of a piecewise smooth CR curve running in
8883: $M\backslash K_{T_S}$. Indeed, if $r'=h_l$ is a hyperbolic point, we
8884: may first use one of the two $\mathcal{ C}^{2, \alpha}$-smooth complex
8885: tangent curves $\gamma_l^1$ or $\gamma_l^2$ passing through $h_l$ and
8886: running in $M \backslash K_{ T_S}$ to join $r'$ with another nearby
8887: point $r''$ which belongs to an open separatrix. If $r'$ already
8888: belongs to an open separatrix, we simply set $r'':= r'$. Since $S$ is
8889: now maximally real near $r''$, we may choose a $T^cM$-tangent vector
8890: field $Y$ defined in a neighborhood of $r''$ which is transversal to
8891: $S$ at $r''$. Then for all $\delta>0$ small enough, the point
8892: $r''':=\exp( \delta Y )(r'')$ satisfies $r''' \not \in S$, whence
8893: $r''' \not \in \overline{ K}$. Of course, by choosing $r''$ and $r'''$
8894: sufficiently close to $r'$, it follows that the piecewise smooth CR
8895: curve joining them does not meet $K_{T_S}$. We deduce that $r'\in
8896: \mathcal{O}_{CR} (M\backslash K_{ T_S}, r''')$.
8897: 
8898: Since by assumption $r' \in \mathcal{ O}_{CR}(M, r)$, it follows that
8899: $r'''\in \mathcal{ O}_{CR}(M,r)$ and then $r'''\in \mathcal{ O}_{
8900: CR}(M,r)\backslash \overline{ K}$. By means of the
8901: supclusion~\thetag{ 12.11}, we deduce that $r'''\in \mathcal{
8902: O}_{CR}(M\backslash K_{T_S},r)$.
8903: 
8904: Finally, from the two relations $r'''\in \mathcal{ O}_{CR}(M\backslash
8905: K_{T_S},r)$ and $r'\in \mathcal{O}_{CR} (M\backslash K_{ T_S}, r''')$,
8906: we conclude immediately that $r'\in \mathcal{ O}_{CR}(M\backslash
8907: K_{T_S}, r)$. We have thus shown that the supclusion in~\thetag{
8908: 12.11} is an equality.
8909: 
8910: This completes the deduction of {\bf (A2)}
8911: from {\bf (A1)}.
8912: 
8913: It remains now to establish {\bf (A1)}. We remind that in 
8914: Section~3, we derived Lemma~3.5 from Lemma~3.7.
8915: By means of a totally similar argument, which we shall
8916: not repeat, one deduces {\bf (A1)} from the following
8917: assertion. Remind that as $M$ is a hypersurface in 
8918: $\C^2$, its CR orbits are of dimension either $2$ or
8919: $3$.
8920: 
8921: \def\thelemma{12.12}\begin{lemma}
8922: Let $M$, $S$, $T_S$ and $\overline{ K} \subset T_S$ be as above.
8923: There exists a connected submanifold $\Omega$ embedded in $M$
8924: containing the hyperbolic tree $T_S$ such that
8925: \begin{itemize}
8926: \item[{\bf (1)}]
8927: $\Omega$ is a $T^c M$-integral manifold, namely $T_p^cM \subset
8928: T_p \Omega$ for all $p \in \Omega$.
8929: \item[{\bf (2)}]
8930: $\Omega$ is contained in a single CR orbit of $M$.
8931: \item[{\bf (3)}]
8932: $\Omega\backslash \overline{ K}$ is also 
8933: contained in a single CR orbit of $M \backslash
8934: \overline{ K}$.
8935: \end{itemize}
8936: More precisely, $\Omega$ is an open neighborhood of $T_S$ if it is of
8937: real dimension $3$ and a complex curve surrounding $T_S$ if it is of
8938: dimension $2$.
8939: \end{lemma}
8940: 
8941: \proof
8942: We shall construct $\Omega$ by means of a flowing procedure, 
8943: starting from a local piece of it. We start
8944: locally in a neighborhood of a fixed point
8945: $p_0 \in \overline{ K} \backslash \{ h_1, \dots, h_\lambda \}$,
8946: whose precise
8947: choice does not matter. Since $S$ is totally real in a neighborhood of
8948: $p_0$, there exists a locally defined $T^cM$-tangent vector field $Y$
8949: which is transversal to $S$ at $p_0$. Consequently, for $\delta >0$
8950: small enough, the small segment $I_0:= 
8951: \{ \exp(sY) (p_0): \, - \delta < s < \delta\}$ is transversal to 
8952: $S$ at $p_0$ and moreover,
8953: the
8954: two half-segments
8955: \def\theequation{12.13}\begin{equation}
8956: I_0^\pm:= \left\{
8957: \exp(s Y)(p_0): \, 
8958: 0 < \pm s < \delta 
8959: \right\}
8960: \end{equation}
8961: lie in $M\backslash S$. Since $p_0$ belongs to some $T^cM$-tangent
8962: open separatrix $\tau_k(0,1)$, there exists a $\mathcal{ C}^{1,
8963: \alpha}$-smooth vector field $X$ defined in a neighborhood of $p_0$ in
8964: $M$ which is tangent to $S$ and whose integral curve passing through
8965: $p_0$ is a piece of $\tau_k(0,1)$. Since $Y$ is transversal to $S$ at
8966: $p_0$, it follows that the set $\omega_0 := \{\exp(s_2 X)(\exp ( s_1
8967: Y) (p_0)): \, -\delta < s_1, \, s_2 < \delta\}$ is a well-defined
8968: $\mathcal{ C}^{1, \alpha}$-smooth codimension one small submanifold
8969: passing through $p_0$ which is transversal to $S$ at $p_0$. Clearly,
8970: we even have $T_{p_0}\omega_0 = T_{ p_0}^cM$. Thanks to the fact that
8971: the flow of $X$ stabilizes $S$, we see that the integral curves $s_2
8972: \mapsto \exp(s_2 X)(\exp (s_1 Y) (p_0))$ are contained in $M
8973: \backslash S$ for every starting point $\exp (s_1 Y) (p_0)$ in the
8974: segment $I_0$ which does not lie in $S$, namely for all $s_1 \neq
8975: 0$. We deduce that the two open halves of $\omega_0$ defined by
8976: \def\theequation{12.14}\begin{equation}
8977: \omega_0^\pm:=\{
8978: \exp(s_2
8979: X)(\exp (s_1 Y)(p_0)): \, 
8980: 0 < \pm s_1 < \delta, \ 
8981: -\delta < s_2 < \delta
8982: \}
8983: \end{equation}
8984: are contained in a single CR orbit of $M \backslash \overline{ K}$.
8985: 
8986: To begin with, assume that the CR orbit in $M \backslash \overline{
8987: K}$ the point $q_0^+:= \exp\left(\frac{ \delta}{2} Y \right) (p_0)$
8988: which belong to $\omega_0^+$, as drawn in {\sc Figure~22} below,
8989: is of real dimension $2$. Afterwards, we shall treat the case where
8990: its CR orbit in $M\backslash \overline{ K}$ is of dimension $3$.
8991: 
8992: Since $M$ is a hypersurface in $\C^2$ and since we have just proved
8993: that the CR orbit $\mathcal{ O}_{ CR}(M \backslash \overline{ K},
8994: q_0^+)$ already contains the $2$-dimensional half piece $\omega_0^+$,
8995: we deduce that $\omega_0^+$ is a piece of complex curve whose boundary
8996: $\partial \omega_0^+$ is (by construction) contained in the separatrix
8997: $\tau_k (0,1)$. Since $\tau_k (0,1)$ is an embedded segment, we may
8998: suppose from the beginning that the vector field $X$ is defined in a
8999: neighborhood of $\tau_k (0,1)$ in $M$. Using then the flow of $X$, we
9000: may easily prolong the small piece $\omega_0^+$ to get a semi-local
9001: $\mathcal{ C }^{1, \alpha}$-smooth submanifold $\omega_k^+$ stretched
9002: along $\tau_k(0,1)$, which constitutes its boundary. Again, this
9003: piece $\omega_k^+$ is (by construction) contained in the CR orbit of
9004: $q_0^+$ in $M\backslash \overline{K}$. By the fundamental stability
9005: property of CR orbits under flows, we deduce that $\omega_k^+$ is in
9006: fact a piece of complex curve with boundary $\tau_k(0,1)$.
9007: 
9008: Remind that by definition of separatrices, the point $\tau_k(0)$ is
9009: always a hyperbolic point. There is a dichotomy: either $\tau_k(1)$ is
9010: also a hyperbolic point or it lies in $\partial S$. If $\tau_k(1)$ is
9011: a hyperbolic point, then by the definition of $\overline{ K}$, the
9012: complete boundary $\partial \omega_k^+ =\tau_k(0,1)$ is contained in
9013: $\overline{ K}$, hence it may {\it not}\, be crossed by means of a CR
9014: curve running in $M \backslash \overline{ K}$. Since the piece
9015: $\omega_k^+$ will be flowed all around $T_S$, our filling $\overline{
9016: K}$ of $K_{T_S}$ was motivated by the desire of simplifying the
9017: geometric situation without having to discuss whether $K_{T_S}$
9018: contains or does not contain the whole segment $\tau_k (0,1)$, for
9019: each $k=1, \dots, \mu$.
9020: 
9021: Before studying the case where $\tau_k (1) \in \partial S$, 
9022: let us analyze the local situation in a neighborhood of the
9023: hyperbolic point $\tau_k(0)=:h_l$, for some $l$ with $1\leq l\leq
9024: \lambda$. 
9025: 
9026: As a preliminary, in order to understand clearly the situation, 
9027: let us assume that the two characteristic curves $\gamma_l^1$ and
9028: $\gamma_l^2$ passing through $h_l$ are of class $\mathcal{
9029: C}^{ 2,\alpha}$, an assumption which would be satisfied if
9030: we had assumed that $M$ and $S$ are of class $\mathcal{ C}^{3,
9031: \alpha}$. By a straightening of $\gamma_l^1$ and of $\gamma_l^2$, 
9032: which induces a loss of one derivative,
9033: we easily show that there exist two linearly independent $\mathcal{
9034: C}^{ 1,\alpha}$-smooth $T^cM$-tangent vector fields $X_1$ and $X_2$
9035: whose integral curves issued from $h_l$ coincide with $\gamma_l^1$ and
9036: $\gamma_l^2$. After possibly renumbering and reversing $X_1$ and $X_2$
9037: and also reparametrizing $\gamma_l^1$ and
9038: $\gamma_l^2$, we may assume that
9039: $\gamma_l^1(s) =\exp(sX_1) (h_l)$ and that 
9040: $\gamma_l^2(s) =\exp(sX_2) (h_l)$, for all small
9041: $s>0$. Furthermore, we may assume that the direction of $X_2$ in a
9042: neighborhood of $h_l$ is the same as the direction from 
9043: $\partial \omega_k^+$ to $\omega_k^+$.
9044: 
9045: \bigskip
9046: \begin{center}
9047: \input tree-orbit.pstex_t
9048: \end{center}
9049: 
9050: As in {\sc Figure~22} just above, let $\tau_j(0,1)$ be the separatrix
9051: issued from $h_l$ in the positive direction of $X_2$. 
9052: We may assume that $\tau_j (0) = h_l$.
9053: Thanks to the
9054: flow of the vector field $X_2$, we may now propagate the piece of
9055: complex curve $\omega_k^+$ by stretching it along $\tau_j(0,1)$ in a
9056: neighborhood of $h_l$. Using then the flow of a semi-locally defined
9057: complex tangent vector field defined in a neighborhood of
9058: $\tau_j(0,1)$, we may extend this local piece as a complex curve
9059: $\omega_j^+$ with boundary $\tau_j(0,1)$. Finally, $\omega_k^+$ and
9060: $\omega_j^+$ glue together as a complex curve with boundary
9061: $\tau_k(0,1) \cup \{h_l\} \cup \tau_j(0,1)$ and corner $h_l$.
9062: 
9063: However, by {\bf (iv)} above, $\gamma_l^1$ and $\gamma_l^2$ are only
9064: of class $\mathcal{ C}^{1, \alpha}$. Examples for which this
9065: regularity is optimal are easily found. Straightening them is again
9066: possible, but the vector fields $X_1$ and $X_2$ would be of class
9067: $\mathcal{ C}^\alpha$, and we would lose the uniqueness of their
9068: integral curves as well as the regularity of their flow. Consequently,
9069: to prove that $\omega_k^+$ propagates along the second separatrix, with
9070: its boundary contained in it, we must proceed differently~:
9071: the proof is longer and we need one more diagram.
9072: 
9073: In {\sc Figure~23} just below, we draw the sadlle-looking surface $S$
9074: in the $3$-dimensional space $M$~; the horizontal plane passing
9075: through $h_l$ is thought to be the complex tangent plane $T_{
9076: h_l}^cM$.
9077: 
9078: \bigskip
9079: \begin{center}
9080: \input saddle-orbit.pstex_t
9081: \end{center}
9082: 
9083: Let us introduce two $T^cM$-tangent vector fields $X_1$ and $X_2$
9084: defined in a neighborhood of $h_l$ with $X_1 (h_l)$ directed along
9085: $\tau_k$ in the sense of increasing $s$ and $X_2 (h_l)$ directed along
9086: $\tau_j$ in the sense of increasing $s$. Let $Z$ denote the vector
9087: field $X_1 +X_2$, as shown in the top of the left hand side of {\sc
9088: Figure~23} above. Using the flow of $Z$ we can begin by extending the
9089: banana-looking piece $\omega_k^+$ of complex curve by introducing the
9090: submanifold $\omega$ consisting of points
9091: \def\theequation{12.15}\begin{equation}
9092: \exp(s_2 Z ) (\tau_k(s_1)),
9093: \end{equation}
9094: where $0 < s_1 < \delta$ and $0 < s_2 < \delta$, for some small
9095: $\delta > 0$. One checks that all these points stay in $M \backslash
9096: S$, hence are contained in the same CR orbit as
9097: $\omega_k^+$ in $M \backslash \overline{ K}$. By the stability
9098: property of CR orbits, it follows of course that $\omega$ is a piece
9099: of complex curve contained in $M$.
9100: 
9101: For $0 < s < \delta$, let $\mu(s) := \exp (sZ) (h_l)$ denote the CR
9102: curve lying ``between'' $\tau_k$ and $\tau_j$ and which
9103: constitutes a part of the boundary of $\omega$. Let $p$ be an
9104: arbitrary point of this curve, close
9105: to $h_l$. 
9106: 
9107: \def\thelemma{12.16}\begin{lemma}
9108: The integral curve $s\mapsto
9109: \exp (-s X_1)(p)$ of $-X_1$ issued from $p$ necessarily
9110: intersects $S$ at a point $q$ close to $h_l$ and close to $\tau_j$
9111: {\rm (}{\it cf.} {\sc Figure~23}{\rm )}.
9112: \end{lemma}
9113: 
9114: \proof
9115: First of all, we need some preliminary.
9116: 
9117: Thanks to the existence of a ``$1/8$ piece'' of complex curve $\omega$
9118: with $h_l\in \overline{ \omega}$ which is contained in the
9119: hypersurface $M$, we see that $M$ is necessarily Levi-degenerate at
9120: $h_l$.
9121: 
9122: Next, we introduce local holomorphic coordinates $(z, w) = (x+ iy,
9123: u+iv)\in \C^2$ vanishing at $h_l$ in which the hypersurface $M$ is
9124: given as the graph $v= \varphi (x, y, u)$, where $\varphi$ is a
9125: $\mathcal{ C}^{ 2, \alpha}$-smooth function. Since $M$ is Levi
9126: degenerate at $h_l$, we may assume that $\varphi (x, y, u) \vert \leq
9127: C \cdot \left(\vert x \vert + \vert y \vert + \vert u \vert \right)^{
9128: 2+ \alpha}$. We may also assume that the surface $S$, as a subset of
9129: $M$, is represented by one supplementary equation of the form $u = h
9130: (x,y)$, where the $\mathcal{ C}^{ 2,\alpha}$-smooth
9131: function $h$ satisfies
9132: \def\theequation{12.17}\begin{equation}
9133: \left\{
9134: \aligned
9135: h(x,y)
9136: & \
9137: = z \bar z+ \gamma \, (z^2+ \bar z^2)+ 
9138: {\rm O}\left( \vert z \vert^{ 2+ \alpha} \right) \\
9139: & \
9140: =
9141: (2\gamma + 1) \, x^2 - 
9142: (2\gamma -1) \, y^2+ {\rm O} \left( \vert z \vert^{ 2+ \alpha}
9143: \right),
9144: \endaligned\right.
9145: \end{equation}
9146: and where $\gamma >\frac{ 1}{2}$ is E.~Bishop's invariant. Then the
9147: tangents at $h_l$ to the two half-separatrices $\tau_k$ and $\tau_j$
9148: are given respectively by the linear (in)equations $x>0$, $y = -
9149: \frac{ 2\gamma + 1}{ 2\gamma -1} \, x$, $u=0$ and $x< 0$, $y = -
9150: \frac{ 2\gamma + 1}{ 2\gamma -1} \, x$, $u=0$. In {\sc Figure~23},
9151: where we do not draw the axes, the $u$-axis is vertical, the $y$ axis
9152: points behind $h_l$ and the $x$-axis is horizontal, from left to
9153: right.
9154: 
9155: Expressing the two $T^cM$-tangent vector fields $X_1$ and $X_2$ in the
9156: (natural) real coordinates $(x,y,u)$ over $M$, we may write them as
9157: \def\theequation{12.18}\begin{equation}
9158: \left\{
9159: \aligned
9160: X_1
9161: & \
9162:  = 
9163: \frac{ \partial}{ \partial x} -
9164: \left( 
9165: \frac{ 2\gamma + 1}{ 2\gamma -1}\right) \, 
9166: \frac{ \partial }{ \partial y} + 
9167: A_1 (x,y,u)\, 
9168: \frac{ \partial }{ \partial u}, \\ 
9169: X_2 
9170: & \
9171: = 
9172: -
9173: \frac{ \partial}{ \partial x} -
9174: \left( 
9175: \frac{ 2\gamma + 1}{ 2\gamma -1}\right) \, 
9176: \frac{ \partial }{ \partial y} + 
9177: A_2 (x,y,u)\, 
9178: \frac{ \partial }{ \partial u}. 
9179: \endaligned\right.
9180: \end{equation}
9181: Since $\varphi$ vanishes to second order at $h_l$, the two
9182: $\mathcal{ C}^{ 1,\alpha}$-smooth
9183: coefficients $A_1$ and $A_2$ satisfy an estimate
9184: of the form
9185: \def\theequation{12.19}\begin{equation}
9186: \left\vert 
9187: A_1, A_2 (x,y,u) 
9188: \right\vert < C \cdot \left(
9189: \vert x \vert + \vert y \vert + 
9190: \vert u \vert
9191: \right)^{ 1+ \alpha}.
9192: \end{equation}
9193: 
9194: Now, we come back to the integral curve of Lemma~12.16. It is
9195: contained in the real $2$-surface passing through $h_l$ defined by
9196: \def\theequation{12.20}\begin{equation}
9197: \Sigma:= \{\exp(-s_2 X_1)
9198: (\exp(s_1 Z)(h_l)): \, -\delta < s_1, s_2 <
9199: \delta\},
9200: \end{equation}
9201: for some $\delta >0$. Because the vector fields $X_1$, $X_2$ and $Z=
9202: X_1 + X_2$ have $\mathcal{ C}^{ 1,\alpha}$-smooth coefficients, the
9203: surface $\Sigma$ is only $\mathcal{ C}^{ 1,\alpha}$-smooth in
9204: general. In $M$ equipped with the three real coordinates $(x,y,u)$, we
9205: may parametrize $\Sigma$ by a mapping of the form
9206: \def\theequation{12.21}\begin{equation}
9207: (s_1, s_2) \longmapsto 
9208: \left(
9209: s_2 - 
9210: 2s_1 \left(
9211: \frac{ 2\gamma +1}{ 2\gamma -1}
9212: \right), \ s_2
9213: \left(
9214: \frac{ 2\gamma+ 1}{ 2\gamma -1}
9215: \right), \ 
9216: u( s_1,s_2)
9217: \right), 
9218: \end{equation}
9219: where $u$ is of class $\mathcal{ C}^{ 1,\alpha}$. It is clear that
9220: $u(0)= \partial_{ s_1} u(0) = \partial_{ s_2} u( 0) =0$, so that
9221: there is a constant $C$ such that
9222: \def\theequation{12.22}\begin{equation}
9223: \left \vert u (s_1,\, s_2) 
9224: \right \vert < C \cdot
9225: \left(
9226: \vert s_1 \vert + \vert s_2 \vert
9227: \right)^{ 1+ \alpha},
9228: \end{equation}
9229: since $u$ is of class $\mathcal{ C}^{ 1,\alpha}$. Furthermore, by
9230: inspecting the flows appearing in~\thetag{ 12.20}, taking account of
9231: the estimates~\thetag{ 12.19}, we claim that $u$ satisfies the
9232: better estimate
9233: \def\theequation{12.23}\begin{equation}
9234: \vert u (s_1, s_2) \vert < C \cdot
9235: \left(
9236: \vert s_1 \vert + \vert s_2 \vert 
9237: \right)^{ 2+ \alpha},
9238: \end{equation}
9239: for some constant $C>0$.
9240: In other words, $\Sigma$ osculates the complex tangent plane $T_{
9241: h_l}^cM$ to second order at $h_l$~: $\Sigma$ is more flat than $S$ at
9242: $h_l$. One may check the estimate~\thetag{ 12.23} is sufficient to
9243: establish Lemma~12.16, because the second jet of the saddle function
9244: $h(x,y)$ does not vanish at the origin.
9245: 
9246: To prove the claim, we formulate the main argument as an independent
9247: assertion. Mild modifications of this argument apply to our case, but
9248: we shall not provide all the details.
9249: 
9250: Let $L_1 := \frac{ \partial }{ \partial x} + A_1 (x,y, u) \, \frac{
9251: \partial }{ \partial u}$ and $L_2 := \frac{ \partial }{ \partial y} +
9252: A_2 (x,y, u) \, \frac{ \partial}{ \partial u}$ be two vector fields
9253: having $\mathcal{ C}^{ 1,\alpha }$-smooth coefficients
9254: satisfying~\thetag{ 12.19}. Denote 
9255: by $s_1 \longmapsto (s_1, \lambda (s_1), \mu(s_1))$
9256: the integral curve of $L_1$ passing through the origin. It is
9257: $\mathcal{ C}^{ 2,\alpha}$-smooth and
9258: we have 
9259: \def\theequation{12.24}\begin{equation}
9260: \left \vert
9261: \lambda (s_1)
9262: \right \vert < C \cdot \vert s_1 \vert^{ 2+\alpha}
9263: \ \ \ \ \ \ \
9264: {\rm and} 
9265: \ \ \ \ \ \ \
9266: \left \vert
9267: \mu (s_1)
9268: \right \vert < C \cdot \vert s_1 \vert^{ 2+\alpha},
9269: \end{equation}
9270: for some constant $C>0$. Consider the composition of flows $\exp (s_2
9271: L_2) (\exp (s_1 L_1)(0))$. We have to solve the system of ordinary
9272: differential equations
9273: \def\theequation{12.25}\begin{equation}
9274: \frac{ dx}{ ds_2}= 0, 
9275: \ \ \ \ \ \ \
9276: \frac{ dy}{ ds_2} = 1, 
9277: \ \ \ \ \ \ \
9278: \frac{ du}{ ds_2}= A_2 (x,y,u)
9279: \end{equation}
9280: with initial conditions
9281: \def\theequation{12.26}\begin{equation}
9282: x(0)= s_1, 
9283: \ \ \ \ \ \ \
9284: y(0) = \lambda( s_1), 
9285: \ \ \ \ \ \ \ 
9286: u(0) = \mu(s_1).
9287: \end{equation}
9288: This yields $x(s_1,s_2) =s_1$, $y(s_1,s_2)= \lambda (s_1)+ s_2$ and
9289: the integral equation
9290: \def\theequation{12.27}\begin{equation}
9291: u(s_1,s_2) = 
9292: \mu(s_1)+ 
9293: \int_0^{ s_2} \, 
9294: A_2 \left(
9295: s_1, s_2'+ \lambda (s_1), u(s_1, s_2')
9296: \right) \, ds_2'.
9297: \end{equation}
9298: Since $u$ is at least $\mathcal{ C}^{ 1,\alpha}$-smooth and vanishes
9299: to order $1$ at $(s_1,s_2)= (0,0)$, we know already that it
9300: satisfies~\thetag{ 12.22}. Using~\thetag{ 12.19}, it is now
9301: elementary to provide an upper estimate of the right hand side
9302: of~\thetag{ 12.27} which yields the desired estimate~\thetag{ 12.23}.
9303: 
9304: The proof of Lemma~12.16 is complete.
9305: \endproof
9306: 
9307: So, for various points $p= \mu(s)$ close to $h_l$ the intersection
9308: points $q\in S$ exist. If all points $q$ belong to $\tau_j$, we are
9309: done: the piece $\omega$ extends a $1/4$ piece of complex curve with
9310: boundary $\tau_k \cup \tau_j$ near $h_l$ and corner $h_l$.
9311: 
9312: Assume therefore that one such point $q$ does not belong to $\tau_j$,
9313: as drawn in the left hand side of {\sc Figure~23} above. Suppose that
9314: $q$ lies above $\tau_j$, the case where $q$ lies under $\tau_j$ being
9315: similar and in fact simpler. The characteristic curve $\gamma'\subset
9316: S$ passing through $q$ stays above $\tau_j$ and is nonsingular.
9317: Propagating the complex curve $\omega$ in $M\backslash \overline{ K}$
9318: by means of the flow of $-X_1$, we deduce that there exists at $q$ a
9319: local piece $\omega_q^+$ of complex curve with boundary contained in
9320: $\gamma'$ which is contained in the same CR orbit as $\omega$. Using
9321: then the flow of a CR vector having $\gamma'$ as an integral curve, we
9322: can propagate $\omega_q^+$ along $\gamma'$, which yields a long thin
9323: banana-looking complex curve with boundary in $\gamma'$. However, this
9324: piece may remain too thin. Fortunately, thanks to the flow of $X_1 -
9325: X_2$, we can extend it as a piece $\omega'$ of complex curve with
9326: boundary $\gamma'$ which goes over $h_l$, with respect to a
9327: complex projection onto $T_{h_l}^cM$, as illustrated in {\sc
9328: Figure~23} above. We claim that this yields a contradiction.
9329: 
9330: Indeed, as $\omega$ and $\omega'$ are complex curves, they are locally
9331: defined as graphs of holomorphic functions $g$ and $g'$ defined in
9332: domains $D$ and $D'$ in the complex line $T_{h_l}^cM$. By
9333: construction, there exists a point in $r \in D\cap D'$ at which the
9334: values of $g$ and $g'$ are distinct. However, since by construction
9335: $g$ and $g'$ coincide in a neighborhood of the CR curve joining $p$ to
9336: $q$, they must coincide at $r$ because of the principle of analytic
9337: continuation: this is a contradiction. In conclusion, the CR orbit
9338: passes through the hyperbolic point $h_l$, in a neighborhood of which
9339: it consists of a cornered complex curve with boundary $\tau_k\cup
9340: \tau_j$.
9341: 
9342: We can now continue the proof. Since the hyperbolic tree $T_S$ does
9343: not contain any cycle, by proceeding this way we claim that the small
9344: piece of complex curve $\omega_0^+$ propagates all around $T_S$ and
9345: matches up as a smooth complex curve $\Omega$ containing the
9346: hyperbolic tree. Indeed, in the case where $\tau_k( 1)$ is not a
9347: hyperbolic point, recall that we arranged at the beginning that
9348: $\overline{ K}\cap \tau_k(0,1)= \tau_k(0,r_1]$, where $r_1<1$. It is
9349: then crucial that when a limit point $\tau_k(1)$ belongs to $\partial
9350: S$, we escape from $\overline{ K}$ and using a local CR vector field
9351: $Y$ transversal to $S$, we may cross the separatrix $\tau_k(0,1)$ at a
9352: point $\tau_k(r_2)$ where $r_2$ satisfies $r_1 < r_2 < 1$. Hence, we
9353: pass to the other side of $S$ in $M$ and then, by means of a futher
9354: flowing, we turn around to the other side of $\tau_k(0,1)$. Also, the
9355: two pieces in either side of $\tau_k(0,1)$ match up at least
9356: $\mathcal{ C}^{1,\alpha}$-smoothly. Then thanks to the stability
9357: property of orbits under flows, we deduce that these two pieces match
9358: up as a piece of complex curve containing $\tau_k(0,1)$ in its
9359: interior.
9360: 
9361: We thus construct the complex curve $\Omega$ surrounding $T_S$, which
9362: is obviously contained in a single CR orbit of $M$. Also, by
9363: construction, $\Omega\backslash \overline{ K}$ is contained in a
9364: single CR orbit of $M\backslash \overline{ K}$. Thus, we have
9365: established Lemma~3.12 under the assumption that the CR orbit of
9366: $q_0^+$ is two-dimensional.
9367: 
9368: Assume finally that the CR orbit of $q_0^+$ is $3$-dimensional. By a
9369: similar propagation procedure, we easily construct a neighborhood
9370: $\Omega$ in $M$ of the hyperbolic tree satisfying conditions {\bf
9371: (1)}, {\bf (2)} and {\bf (3)} of Lemma~12.12. This complete its proof.
9372: \endproof
9373: 
9374: The proof of Lemma~12.6 is complete.
9375: \endproof
9376: 
9377: \subsection*{12.28.~Proofs of Theorems~1.1 and 1.3}
9378: We can now prove Theorem~1.1. In fact, we shall directly prove the
9379: more general version stated as Theorem~1.3 which implies Theorem~1.1
9380: as a corollary, thanks to the geometric observations of \S12.3.
9381: 
9382: First of all, we notice that as $M\backslash K$ is globally minimal,
9383: there exists a wedge attached to $M\backslash K$ to which continuous
9384: CR functions on $M\backslash K$ extend holomorphically. Hence, the
9385: CR-removability of $K$ is a consequence of its $\mathcal{
9386: W}$-removability. Also, Lemma~11.3 shows that the 
9387: $L^{\sf p}$-removability of $K$ is a consequence of
9388: its $\mathcal{ W}$-removability. Consequently, it suffices
9389: to establish that $K$ is $\mathcal{ W}$-removable in 
9390: Theorem~1.3.
9391: 
9392: Let $T_S$ be the hyperbolic tree of (a suitable shrinking of) $S$,
9393: which contains no cycle by assumption. Let $\omega_1$ be a one-sided
9394: neighborhood of $M \backslash K$ in $\C^2$. Because the
9395: nontransversality condition $\mathcal{ F }_{ S \backslash T_S}^c \{ K
9396: \cap (S\backslash T_S) \}$ holds true by assumption, we may apply
9397: Theorem~1.2 to the totally real surface $S \backslash T_S$ in the
9398: globally minimal (thanks to Lemma~12.6) hypersurface $M \backslash K_{
9399: T_S}$ to remove the proper closed subset $K\cap (S \backslash
9400: T_S)$. We deduce that there exists a one-sided neighborhood $\omega_2$
9401: of $M\backslash K_{ T_S}$ in $\C^2$ such that (after shrinking
9402: $\omega_1$ if necessary), holomorphic functions in $\omega_1$ extend
9403: holomorphically to $\omega_2$. Then we slightly deform $M$ inside
9404: $\omega_2$ over points of $K\cap (S \backslash T_S)$. We obtain a
9405: $\mathcal{ C}^{2, \alpha}$-smooth hypersurface $M^d$ with
9406: $M^d\backslash K_{ T_S} \subset \omega_2$. Also, by stability of
9407: global minimality under small perturbations, we can assume that $M^d$
9408: is also globally minimal. By construction, we obtain holomorphic
9409: functions in the neighborhood $\omega_2$ of $M^d \backslash K_{T_S}$
9410: in $\C^2$.
9411: 
9412: Since $M$ and $M^d$ are of codimension $1$, the union of a one-sided
9413: neighborhood $\omega^d$ of $M^d$ in $\C^2$ together with $\omega_2$
9414: constitutes a one-sided neighborhood of $M$ in $\C^2$. To conclude the
9415: proof of Theorem~1.1, it suffices therefore to show that the closed
9416: set $K_{T_S}$ is $\mathcal{ W}$-removable.
9417: The reader is referred to {\sc Figure~20} above for
9418: an illustration.
9419:  
9420: Reasoning by contradiction (as for the proof of Theorem~1.2'), let
9421: $K_{\rm nr}\subset K_{T_S}$ denote the smallest nonremovable subset of
9422: $K_{T_S}$. If $K_{\rm nr}$ is empty, we are done, gratuitously.
9423: Assume therefore that $K_{\rm nr}$ is nonempty. Let $T'$ be a
9424: connected component of the minimal subtree of $T$ containing $K_{\rm
9425: nr}$. By a {\sl subtree} of a tree $T$ defined as in~\thetag{12.2}
9426: above, we mean of course a finite union of some of the separatrices
9427: $\tau_k(0,1)$ together with all hyperbolic points which are endpoints
9428: of separatrices. Since $T'$ does not contain any subset homeomorphic
9429: to the unit circle, there exists at least one extremal branch of $T'$,
9430: say $\tau_1(0,1)$ after renumbering, with $\tau_1(1)\in \partial
9431: S$. To reach a contradiction, we shall show that at least one point of
9432: the nonempty set $K_{\rm nr} \cap T'$ is in fact $\mathcal{
9433: W}$-removable.
9434: 
9435: If the subtree $T'$ consists of the single branch $\tau_1 (0,1)$
9436: together with the single elliptic point $\tau_1 (0)$, thanks to
9437: properties {\bf (iii)} and {\bf (iv)} of \S12.1, we can enlarge a
9438: little bit this branch by prolongating the curve $\tau_1 (0,1)$ to an
9439: open $\mathcal{ C}^{ 2, \alpha}$-smooth Jordan arc $\tau_1
9440: (-\varepsilon, 1+ \varepsilon)$, for some $\varepsilon>0$, with the
9441: appendix $\tau_1(1,1+\varepsilon)$ outside $S$ (in the slightly larger
9442: surface containing $S$). Then we remind that by a special case of
9443: Theorem~4 (ii) of~\cite{mp1}, every proper closed subset of
9444: $\tau_1(-\varepsilon, 1+\varepsilon)$ is $\mathcal{ W}$-removable. It
9445: follows that the proper closed subset $K_{\rm nr}$ of the Jordan arc
9446: $\tau_1(-\varepsilon, 1+\varepsilon)$ is removable, which yields the
9447: desired contradiction in the case where $T'$ consists of a single
9448: branch together with a single elliptic point.
9449: 
9450: If $T'$ consists of at least two branches, again with $\tau_1(1)\in
9451: \partial S$, then applying Theorem~4 (ii) of~\cite{mp1}, we may at
9452: least deduce that $K_{\rm nr} \cap \tau_1(0,1)$ is $\mathcal{
9453: W}$-removable, since $K_{\rm nr}\cap \tau_1(0,1)$ is contained in
9454: $\tau_1(0, r_1]$, for some $r_1<1$. But possibly, this set $K_{\rm nr}
9455: \cap \tau_1(0,1)$ could be empty.
9456: 
9457: However, we claim that it is
9458: nonempty. Indeed, otherwise, if $K_{ \rm nr} \cap \tau_1[0,1)$
9459: consists of the single point $\tau_1(0)$, which is a hyperbolic point,
9460: then $K_{ \rm nr}$ is in fact contained in the smaller subtree $T''$
9461: defined by $T'' := T' \backslash \tau_1 (0,1)$ (here, we use that
9462: $\tau_1 (0)$ is a hyperbolic point, hence there exists another branch
9463: $\tau_k (0,1)$ with $\tau_k (1)= \tau_1 (0)$ or $\tau_k (0)=
9464: \tau_1(0)$). This contradicts the assumption that $T'$ is the minimal
9465: subtree containing $K_{\rm nr}$. Then $K_{\rm nr} \cap \tau_1 (0,1)$
9466: is nonempty and removable, which contradict the assumption that
9467: $K_{\rm nr}$ is the smallest nonremovable subset of $K$.
9468: 
9469: The proofs of Theorems~1.1 and~1.3 are complete.
9470: \endproof
9471: 
9472: \section*{\S13.~Applications to the edge of the wedge theorem}
9473: 
9474: In this section we formulate three versions of the edge of the wedge
9475: theorem for holomorphic and meromorphic functions, two of which are
9476: based upon an application of our removable singularities theorems.
9477: Let us begin with some definition.
9478: 
9479: \subsection*{13.1.~Preliminary}
9480: Let $E$ be a generic submanifold of $\C^n$, which may be maximally
9481: real. By a {\sl double wedge attached to $E$}, we mean a pair
9482: $(\mathcal{ W}_1, \mathcal{ W }_2)$ of disjoint wedges attached to $E$
9483: which admit a nowhere vanishing continuous vector field $v: E
9484: \rightarrow T \C^n\vert_E / T E$ such that $ Jv ( p)$ points into
9485: $\mathcal{ W }_1$ and $-Jv(p)$ into $\mathcal{ W }_2$, for every $p \in
9486: E$.
9487: 
9488: In the case where $E=\R^n$, the classical edge of the wedge theorem
9489: states that there exists a neighborhood $\mathcal{ D}$ of $E$ in
9490: $\C^n$ such that every function which is continuous on 
9491: $\mathcal{ W}_1 \cup E \cup 
9492: \mathcal{ W }_2$ and holomorphic in $\mathcal{
9493: W}_1 \cup \mathcal{ W }_2$ extends holomorphically to $\mathcal{ D}$.
9494: Also, generalizations are known in the case where $f|_{ \mathcal{
9495: W}_1}$ and $f|_{\mathcal{ W }_2}$ have coinciding distributional
9496: boundary values on $E$.
9497: 
9498: The assumption about the matching up of boundary values along $E$ from
9499: $\mathcal{ W}_1$ and from $\mathcal{ W}_1$ is really needed, even if
9500: the two boundary values coincide on a thick subset of $E$. To support
9501: this observation, consider the following elementary example: the
9502: complex hyperplane $H := \{z_n=0\} \subset \C^n$ and the maximally
9503: real plane $E:= \R^n \subset \C^n$ intersect transversally in the
9504: $(n-1)$-dimensional totally real plane $C:= \{y=0, \ x_n =0\}$; the
9505: pair of wedges $\mathcal{ W}_1 := \{y_1 >0, \dots, y_n >0\}$ and
9506: $\mathcal{ W }_2:= \{y_1<0, \dots, y_n <0\}$ clearly form a double wedge
9507: attached to $E$; the function $\exp(-1 /z_n)$ restricted to the two
9508: wedges is holomorphic there, has coinciding boundary values on the
9509: thick set $E \backslash C$, but does not extend holomorphically to a
9510: neighborhood of $E$ in $\C^n$. Evidently, the envelope of holomorphy
9511: of the union of $\mathcal{ W}_1 \cup \mathcal{ W}_2$ together with a
9512: thin neighborhood of $E \backslash C$ in $\C^n$ does
9513: {\it not}\, contain any neighborhood of $E$ in $\C^n$.
9514: 
9515: Thus, in order to apply our removability theorems (which are
9516: essentially statements about envelopes of holomorphy), the first
9517: question is how to impose coincidence of boundary values on the
9518: edge. We shall first see that the fact that $C\subset E$ is exactly of
9519: codimension one in the above example is the limiting case for the
9520: obstruction to holomorphic extension. Since we want to treat
9521: also meromorphic extension, let us remind some definitions.
9522: 
9523: \subsection*{13.2.~Meromorphic functions and envelopes}
9524: Let $U$ be a domain in $\C^n$. A meromorphic function $f \in \mathcal{
9525: M} ( U)$ is a collection of equivalence classes of quotients of
9526: locally defined holomorphic functions. It defines a $P_1 (\C)$-valued
9527: function, which is single-valued only on some Zariski dense open
9528: subset $D_f \subset U$. More geometrically, we may represent $f$ by
9529: the closure $\Gamma_f$ of its graph $f|_{ D_f}$ over $D_f$, which
9530: always constitutes an irreducible $n$-dimensional complex analytic
9531: subset of $U \times P_1 (\C)$ with surjective, almost everywhere
9532: biholomorphic projection onto $U$ (equivalent definition). It is well
9533: known that the {\sl indeterminacy set} of $f$, namely the set of $z
9534: \in U$ over which the whole fiber $\{ z\} \times P_1 (\C)$ is
9535: contained in $\Gamma_f$, is an analytic subset of $U$ of codimension
9536: at least 2. It is the only set where $f$ is multivalued.
9537: 
9538: We shall constantly apply a theorem due to P.~Thullen
9539: (generalized by S.~Ivashkovitch in~\cite{ i} in the
9540: context of K\"ahler manifolds) according to which 
9541: the envelope of holomorphy of a domain in $\C^n$
9542: coincides with its envelope of meromorphy. As holomorphic
9543: functions are meromorphic, we shall state
9544: Lemma~13.4, Corollary~13.8 and Corollary~13.11 below
9545: directly for meromorphic functions.
9546: 
9547: \subsection*{13.3.~Edge of the wedge theorem over a maximally 
9548: real edge} Let $E \subset \C^n$ $(n \geq 2)$ be a real analytic
9549: maximally real submanifold, let $( \mathcal{ W }_1, \mathcal{
9550: W}_2)$ be a double wedge attached to $E$ and let $f_1$, $f_2$ be two
9551: meromorphic functions in $\mathcal{ W}_1$, $\mathcal{ W}_2$.
9552: 
9553: We need an assumption which tames the behaviour of their indeterminacy
9554: sets, as one tends towards the edge $E$ from either $\mathcal{ W}_1$
9555: or $\mathcal{ W}_2$. It will be sufficient to impose a matching up of
9556: their boundary values on the complement of a closed subset $C$ whose
9557: $(n-1)$-dimensional Hausdorff measure vanishes. Let ${\sf H}_d$ denote
9558: the $d$-dimensional Hausdorff measure.
9559: 
9560: \def\thelemma{13.4}\begin{lemma} If there is a closed subset $C\subset
9561: E$ with ${\sf H }_{ n-1} (C) =0$ such that both $\overline{
9562: \Gamma_{f_1 }}\cap [(E \backslash C) \times P_1(\C)]$ and $\overline{
9563: \Gamma_{ f_2 }} \cap [( E \backslash C)\times P_1(\C)]$ coincide with
9564: the graph of a {\rm (}single{\rm )} continuous mapping from $E
9565: \backslash C$ to $ P_1 (\C)$, then there exists a neighborhood
9566: $\mathcal{ D}$ of $E$ in $\C^n$ which depends only on $(\mathcal{ W}_1, 
9567: \mathcal{ W}_2)$ and a meromorphic function 
9568: \def\theequation{13.5}\begin{equation}
9569: f\in \mathcal{ M}\left(\mathcal{ D} \cup
9570: \mathcal{ W}_1 \cup \mathcal{ W}_2\right), 
9571: \end{equation}
9572: extending the $f_j$, namely such that $f|_{ \mathcal{ W }_j}
9573: =f_j$, for $j= 1,2$.
9574: \end{lemma}
9575: 
9576: \proof
9577: First of all, the assumption of continuous coincidence of boundary
9578: values enables us to apply the classical edge of the wedge theorem at
9579: each point of $E \backslash C$. This yields a neighborhood $\mathcal{
9580: D }_0$ of $E \backslash C$ in $\C^n$ and a meromorphic extension $f_0
9581: \in \mathcal{ M}( \mathcal{ D}_0 \cup \mathcal{ W}_1 \cup \mathcal{
9582: W}_2)$. We claim that the envelope of meromorphy of
9583: $\mathcal{ D}_0 \cup \mathcal{ W}_1 \cup \mathcal{ W}_2$ contains a
9584: neighborhood $\mathcal{ D}_1$ of $E$ in $\C^n$.
9585: 
9586: Indeed, this follows from a very elementary application of the
9587: continuity principle. Let $p\in C$ be arbitrary. After a local
9588: straightening, we may insure that $p$ is the origin, that $E=\R^n$, that
9589: $\mathcal{ W}_1$ contains $\{y_1 >0, \dots, y_n >0 \}$ and that
9590: $\mathcal{ W}_2$ contains $\{y_1 < 0, \dots, y_n < 0\}$.
9591: 
9592: Let us introduce the trivial family of analytic discs
9593: \def\theequation{13.6}\begin{equation}
9594: A_{c,x,v}( \zeta) :=
9595: \left(x_1+c(1+v_1)\zeta,\dots,
9596: x_n+(1+v_n)\zeta\right), 
9597: \end{equation}
9598: where $c >0$ is a sufficiently small fixed scaling factor, where $x
9599: \in \R^n$ is a small translation parameter and where $v\in \R^n$ is a
9600: small pivoting parameter. Clearly, $A_{c, x,v}( \partial^+ \Delta)$ is
9601: contained in $\mathcal{ W }_1$ and $A_{c, x,v} (\partial^- \Delta)$ is
9602: contained in $\mathcal{ W }_2$. However $A_{c, x,v}( \pm 1)$ may
9603: encounter $C$.
9604: 
9605: First of all, using the submersiveness of the two mappings $v \mapsto
9606: A_{c,0,v}(\pm 1)\in E$, we may find $v_0$ arbitrarily 
9607: close to the origin in $\R^n$ such that $A_{c,0,v_0}(\pm
9608: 1)$ does not belong to $C$. It follows that for all small
9609: translation vectors $q\in \C^n$, the disc boundary
9610: $A_{c,0,v_0}(\partial \Delta)+q$ is contained in the domain $\mathcal{
9611: D}_0\cup \mathcal{ W}_1 \cup \mathcal{ W}_2$.
9612: 
9613: Furthermore, because $C$ is of Hausdorff $(n-1)$-dimensional measure
9614: zero, for almost all $x\in \R^n$, the segment $A_{c,x,v_0}([-1,1])$
9615: does not meet $C$. It follows that for such $x$, the disc
9616: $A_{c,x,v_0}(\overline{ \Delta})$ is contained in the domain
9617: $\mathcal{ D}_0\cup \mathcal{ W}_1 \cup \mathcal{ W}_2$. We deduce
9618: that every disc $A_{c,0,v_0}(\partial \Delta)+q$ is analytically
9619: isotopic to a point in $\mathcal{ D}_0\cup \mathcal{ W}_1 \cup
9620: \mathcal{ W}_2$. An application of the continuity principle yields
9621: meromorphic extension to a neighborhood of 
9622: $p=A_{c,0,0}(0)\in C$.
9623: 
9624: In sum, we have constructed a neighborhood $\mathcal{ D}_1$ of $E$ in
9625: $\C^n$ and a meromorphic extension $f\in \mathcal{ M}( \mathcal{
9626: D}_1)$. But $\mathcal{ D}_1$ is not independent of $(f_1,f_2)$, since
9627: it depends on $C$. Fortunately, once we know meromorphic extension to
9628: a neighborhood $\mathcal{ D}_1$ of $E$ in $\C^n$, we may reemploy the
9629: analytic disc technique of the classical edge of the wedge theorem to
9630: describe a neighborhood $\mathcal{ D}$ of $E$ in $\C^n$ which depends
9631: only on $(\mathcal{ W}_1, \mathcal{ W}_2)$ ({\it see}\, the end of the
9632: proof of Corollary~13.8 below for more arguments). This completes the
9633: proof of Lemma~13.4.
9634: \endproof
9635: 
9636: \subsection*{13.7.~Edge of the wedge theorem over an edge of positive
9637: CR dimension} Let $M$ be a $\mathcal{ C}^{2,\alpha}$-smooth generic
9638: submanifold of $\C^n$ of positive CR dimension and let $C$ be a proper
9639: closed subset of $M$ such that $M$ and $M\backslash C$ are globally
9640: minimal. In~\cite{ mp3}, Theorem~1.1, it was shown as a main theorem
9641: that every such closed subset $C$ of $M$ is $CR$-, $\mathcal{ W}$- and
9642: $L^{\sf p}$-removable. We may formulate the following application,
9643: where, for simplicity, we assume local minimality at every point.
9644: 
9645: \def\thecorollary{13.8}\begin{corollary}
9646: Let $E \subset \C^n$ $( n \geq 2)$ be a generic manifold of class
9647: $\mathcal{ C }^{ 2, \alpha}$ of positive CR dimension which is locally
9648: minimal at every point, let $(\mathcal{ W}_1, \mathcal{ W}_2)$ a
9649: double wedge attached to $E$ and let two meromorphic functions $f_j\in
9650: \mathcal{ M} ( \mathcal{ W}_j)$ for $j=1,2$. If there is a closed
9651: subset $C\subset E$ with ${\sf H }_{ n-1} (C) =0$ such that both
9652: $\overline{ \Gamma_{f_1 }} \cap [(E \backslash C) \times P_1( \C)]$
9653: and $\overline{ \Gamma_{ f_2 }} \cap [( E \backslash C)\times P_1(
9654: \C)]$ coincide with the graph of a {\rm (}single{\rm )} continuous
9655: mapping from $E \backslash C$ to $ P_1 (\C)$, then there exists a
9656: neighborhood $\mathcal{ D }$ of $E$ in $\C^n$ which depends only on
9657: $(\mathcal{ W}_1, \mathcal{ W}_2)$ and a meromorphic function
9658: \def\theequation{13.9}\begin{equation}
9659: f\in \mathcal{ M} \left(\mathcal{ W}_1 \cup
9660: \mathcal{ D} \cup \mathcal{ W}_2\right), 
9661: \end{equation}
9662: extending the $f_j$, namely such that $f|_{ \mathcal{ W }_j}
9663: =f_j$, for $j = 1,2$.
9664: \end{corollary}
9665: 
9666: \proof
9667: Applying the classical edge of the wedge theorem, we get a meromorphic
9668: extension $f_0 \in \mathcal{ M}( \mathcal{ W}_1 \cup \mathcal{ D}_0
9669: \cup \mathcal{ W}_2)$, where $\mathcal{ D}_0$ is some open
9670: neighborhood of $E \backslash C$ in $\C^n$. Next, we include $E$ in a
9671: CR manifold $M$ with $M \subset \mathcal{ W}_1 \cup E \cup \mathcal{
9672: W}_2$ and $\dim_\R M = 1 + \dim_\R E$, as shown in the following
9673: figure.
9674: 
9675: \bigskip
9676: \begin{center}
9677: \input double-wedge.pstex_t
9678: \end{center}
9679: 
9680: Of course, the domain $\mathcal{ W}_1 \cup \mathcal{ D}_0 \cup
9681: \mathcal{ W}_2$ constitues a (rather thick) wedge attached to
9682: $M\backslash C$. Since $E$ is locally minimal at every point and
9683: since one CR tangential direction of $M$ is transversal to $E$, both
9684: $M$ and $M \backslash \Sigma$ are both globally minimal. Applying
9685: Theorem~1.1 in~\cite{ mp2}, we deduce that there exists a wedge
9686: $\mathcal{ W}$ attached to $M$ which is contained in the envelope of
9687: meromorphy of $\mathcal{ W}_1 \cup \mathcal{ D}_0 \cup \mathcal{
9688: W}_2$. We then claim that there exists a neighborhood $\mathcal{ D}$
9689: of $p$ in $\C^n$, which depends only on $(\mathcal{ W}_1, \mathcal{
9690: W}_2)$ such that $\mathcal{ D}$ is contained in the envelope of
9691: meromorphy of $\mathcal{ W}_1 \cup \mathcal{ W} \cup \mathcal{ W}_2$.
9692: 
9693: Indeed, by deforming slightly $M$ inside $\mathcal{ W}$ near $E$, we
9694: get a $\mathcal{ C}^{2, \alpha}$-smooth generic submanifold $M^d
9695: \subset \mathcal{ W}_1 \cup \mathcal{ W} \cup \mathcal{ W}_2$. Instead
9696: of functions meromorphic in the disconnected open set $\mathcal{ W}_1
9697: \cup \mathcal{ W}_2$, we now consider meromorphic functions in the
9698: {\it connected}\, open set $\mathcal{ W}_1 \cup \mathcal{ W} \cup
9699: \mathcal{ W}_2$, which is a neighborhood of $M^d$ in $\C^n$. Then by
9700: following the proof of the edge of the wedge theorem given in~\cite{a}
9701: and applying the continuity principle, one deduces meromorphic
9702: extension to a neighborhood $\mathcal{ D }^d$ in $\C^n$ of the
9703: deformed submanifold $M^d$. Since the size of $\mathcal{ W }_1$ and
9704: the size of $\mathcal{ W }_2$ are uniform with respect to $d$, the
9705: size of the domain $\mathcal{ D }^d$ is also uniform with respect to
9706: $d$, as follows from the stability of the edge of the wedge theorem
9707: established in~\cite{ a}, since it relies on E.~Bishop's equation.
9708: Hence for $M^d$ sufficiently close to $M$, the domain $\mathcal{ D
9709: }^d$ contains a neighborhood of $p$ in $\C^n$. This completes the
9710: proof of Corollary~13.8.
9711: \endproof
9712: 
9713: \subsection*{13.10.~Hartogs-Bochner phenomenon and edge of the wedge 
9714: theorem} Next we turn to the question whether a Hartogs-Bochner
9715: phenomenon holds in presence of a double wedge. More
9716: precisely we ask when it is sufficient to require coincidence of
9717: boundary values only outside some compact $K\subset E$. Let us first
9718: look at a prototypical case where the answer is particularly neat,
9719: thanks to Theorem~1.1. Obviously, the proof is totally similar to the
9720: proof of Corollary~13.8 and will not be repeated.
9721: 
9722: \def\thecorollary{13.11}\begin{corollary}
9723: Let $E \subset \C^2$ be an embedded real analytic totally real disc,
9724: let $(\mathcal{ W }_1, \mathcal{ W }_2)$ a double wedge attached to
9725: $E$ and let two meromorphic functions $f_j\in \mathcal{ M} ( \mathcal{
9726: W}_j)$ for $j=1,2$. If there is a compact subset $K\subset E$ such
9727: that both $\overline{ \Gamma_{f_1 }} \cap [(E \backslash K) \times
9728: P_1( \C)]$ and $\overline{ \Gamma_{ f_2 }} \cap [( E \backslash
9729: K)\times P_1( \C)]$ coincide with the graph of a {\rm (}single{\rm )}
9730: continuous mapping from $E \backslash K$ to $ P_1 (\C)$, then there
9731: exists a neighborhood $\mathcal{ D}$ of $E$ in $\C^2$ which depends
9732: only on $(\mathcal{ W}_1,\mathcal{ W}_2)$ and a meromorphic function
9733: \def\theequation{13.12}\begin{equation}
9734: f\in \mathcal{ M}\left(\mathcal{ W}_1 \cup
9735: \mathcal{ D} \cup \mathcal{ W}_2\right), 
9736: \end{equation}
9737: extending the $f_j$, namely such that $f|_{ \mathcal{ W }_j}
9738: =f_j$, for $j = 1,2$.
9739: \end{corollary}
9740: 
9741: In order to find the most general application of Theorems~1.2 and
9742: 1.2', we first remark that in {\sc Figure~24}, we have a considerable
9743: freedom in the choice of the generic submanifolds $M$ of CR dimension
9744: $1$ with $E\subset M \subset \mathcal{ W}_1 \cup E \cup \mathcal{ W}_2$,
9745: depending on the aperture of $\mathcal{ W }_1$ and $\mathcal{ W
9746: }_2$. Since $\dim_\R M = 1+ \dim_\R E$, the tangent space
9747: to such an $M$ at
9748: a point $p\in E$ is uniquely determined by some nonzero vector $v_p\in
9749: T_p \C^n / T_p E$. In order that $M$ is locally contained in
9750: $\mathcal{ W}_1 \cup E \cup \mathcal{ W}_2$, it is necessary and
9751: sufficient that either $Jv$ points in $\mathcal{ W}_1$ and $-Jv$
9752: points in $\mathcal{ W}_2$, or vice versa, depending on the
9753: orientations of $\mathcal{ W}_1$ and $\mathcal{ W}_2$ with respect to
9754: $E$. Without loss of generality, after a possible shrinking, we can
9755: therefore assume that the cones of $\mathcal{ W}_1$ and of $\mathcal{
9756: W}_2$ are exactly opposite to each other at every point of $E$;
9757: indeed, it would be impossible to construct an $M$ locally contained
9758: in $\mathcal{ W}_1 \cup E \cup \mathcal{ W}_2$ which satisfies $T_p M
9759: = T_p E \oplus \R v_p$ at a point $p\in E$, in the case where the
9760: vector $Jv_p$ points in the cone at $p$ of $\mathcal{ W}_1$ but $-J
9761: v_p$ lies outside the cone at $p$ of $\mathcal{ W}_2$, or
9762: vice versa.
9763: 
9764: Assuming $\mathcal{ W}_2$ to be opposite to $\mathcal{ W}_1$, let us
9765: define an induced field of open cones $p\mapsto {\sf C }_p^{ \mathcal{
9766: W }_1, \mathcal{ W}_2 }$ as follows: a nonzero vector $v_p \in T_p E
9767: \backslash \{0\}$ belongs to ${\sf C }_p^{ \mathcal{ W }_1, \mathcal{
9768: W}_2 }$ if either $Jv$ or $-Jv$ points into $\mathcal{ W }_1$. A
9769: nowhere vanishing vector field $p \mapsto v (p)$ is said to be {\sl
9770: directed by ${\sf C }_p^{ \mathcal{ W }_1, \mathcal{ W}_2 }$} if $v(p)
9771: \in {\sf C}_p^{ \mathcal{ W }_1, \mathcal{ W}_2 }$ for every $p \in
9772: E$. Clearly, for every vector field $p\mapsto v(p)$ directed by $
9773: p\mapsto {\sf C }_p^{ \mathcal{ W }_1, \mathcal{ W}_2 }$, we may
9774: construct a $\mathcal{ C}^{2,\alpha}$-smooth semi-local generic
9775: submanifold $M$ containing $E$, contained in $\mathcal{ W}_1 \cup E
9776: \cup \mathcal{ W}_2$ which satisfies $T_p M = T_p E \oplus J v(p)$ at
9777: every point $p\in E$.
9778: 
9779: In the statement of Theorem~1.2', we defined the condition $\mathcal{
9780: F}_{M^1}^c\{C\}$ with respect to some CR manifold $M$ containing the
9781: totally real manifold $M^1$. But $M$ entered in the definition only
9782: via the characteristic foliation induced on $M^1$. Hence its
9783: reasonable to define a more general nontransversality property, by
9784: replacing the characteristic foliation by the foliation induced by any
9785: vector field directed by the field of cones $p\mapsto {\sf
9786: C}_p^{\mathcal{ W}_1, \mathcal{ W}_2}$, as follows:
9787: 
9788: \smallskip
9789: \begin{itemize}
9790: \item[
9791: $\mathcal{ F}_{ \mathcal{ W}_1, \mathcal{ W}_2 } \{C\}:$] For every
9792: closed subset $C' \subset C$ there is a smooth vector field $ p
9793: \mapsto v(p)$ directed by the field
9794: of cones $p\mapsto {\sf C }_p^{ \mathcal{ W}_1, \mathcal{ W
9795: }_2}$ such that there exists a simple $\mathcal{ C}^{2,
9796: \alpha}$-smooth curve $\gamma': [-1,1] \to E$ whose range
9797: $\gamma'([-1,1])$ is contained in a single integral curve of $p
9798: \mapsto v (p)$ with $\gamma'(-1) \not \in C'$, $\gamma' (0) \in C'$
9799: and $\gamma' (1) \not \in C'$, there exists a local $(n -
9800: 1)$-dimensional transversal $R \subset E$ to $\gamma'$ passing through
9801: $\gamma' (0)$ and there exists a thin open neighborhood $V$ of
9802: $\gamma' ( [-1, 1])$ in $E$ such that if $\pi: V \to R$ denotes the
9803: semi-local projection parallel to the flow lines of $v$, then
9804: $\gamma'(0)$ lies on the boundary, relatively to the topology of $R$,
9805: of $\pi (C' \cap V)$.
9806: \end{itemize}
9807: \smallskip
9808: 
9809: The proper application of Theorem~1.2' is the following.
9810: Its proof follows by a direct examination of the
9811: proof of Theorem~1.2'.
9812: 
9813: \def\thecorollary{13.13}\begin{corollary}
9814: Let $E \subset \C^n$ $(n \geq 2)$ be a real analytic maximally real
9815: submanifold, let $(\mathcal{ W }_1, \mathcal{ W }_2)$ a double wedge
9816: attached to $E$. Let $C$ be a proper closed subset of $E$ satisfying
9817: the nontransversality property $\mathcal{ F}_{ \mathcal{ W }_1,
9818: \mathcal{ W}_2} \{C\}$ above. Let two meromorphic functions $f_j\in
9819: \mathcal{ M} ( \mathcal{ W}_j)$ for $j=1,2$ such that both $\overline{
9820: \Gamma_{f_1 }} \cap [(E \backslash C) \times P_1( \C)]$ and
9821: $\overline{ \Gamma_{ f_2 }} \cap [( E \backslash C)\times P_1( \C)]$
9822: coincide with the graph of a {\rm (}single{\rm )} continuous mapping
9823: from $E \backslash C$ to $ P_1 (\C)$, Then there exists a neighborhood
9824: $\mathcal{ D}$ of $E$ in $\C^2$ which depends
9825: only on $(\mathcal{ W}_1,\mathcal{ W}_2)$ and a meromorphic function
9826: \def\theequation{13.14}\begin{equation}
9827: f\in \mathcal{ M}\left(\mathcal{ W}_1 \cup
9828: \mathcal{ D} \cup \mathcal{ W}_2\right), 
9829: \end{equation}
9830: extending the $f_j$, namely such that $f|_{ \mathcal{ W }_j} =f_j$,
9831: for $j = 1,2$. 
9832: \end{corollary}
9833: 
9834: \subsection*{13.15.~Further applications}
9835: We now conclude this section by suggesting two applications of
9836: Theorem~1.2' in higher dimensions, in the case where $M^1$ is not
9837: everywhere totally real. However, we must mention that we did not try
9838: to generalize the results of E.~Bishop to understand the local
9839: geometry of complex tangencies of generic submanifolds of CR dimension
9840: $1$ in $\C^n$, for $n\geq 3$. Consequently, our formulations
9841: should be considered as mild generalizations of
9842: Theorem~1.1 and~1.3.
9843: 
9844: Thus, let $M$ be a $\mathcal{ C }^{2, \alpha}$-smooth generic
9845: submanifold of $\C^n$ {\rm (}$n \geq 2${\rm )} of CR dimension $1$,
9846: let $M^1$ be a codimension one submanifold of $M$ which is maximally
9847: real except at every point of some proper closed subset $E \subset
9848: M^1$. Let $C$ be a proper closed subset of $M^1$. For simplicity, we
9849: assume that $M$ is locally minimal at every point, an assumption which
9850: insures that for every closed subset $\widetilde{ C}$ of $M$, 
9851: both $M$ and $M \backslash \widetilde{ C}$ are globally
9852: minimal.
9853: 
9854: Firstly, applying Theorem~1.2' to remove
9855: $C \cap (M^1 \backslash E)$ and then Theorem~1.1 of~\cite{ mp2}
9856: to remove $C \cap E$, we deduce the following. 
9857: 
9858: \def\thecorollary{13.16}\begin{corollary}
9859: Assume that $E$ is of vanishing $(n - 1)$-dimensional Hausdorff
9860: content, and that the nontransversality condition $\mathcal{ F}_{M^1
9861: \backslash E}^c \{C \cap (M^1 \backslash E)\}$ holds. Then $C$ is
9862: CR-, $\mathcal{ W}$- and $L^{ \sf p}$-removable.
9863: \end{corollary}
9864: 
9865: Secondly, we may generalize the notion of hyperbolic tree and assume
9866: that $E$ consists of finitely many compact submanifolds of codimension
9867: $2$ in $M^1$ joined by a collection of finitely many codimension one
9868: submanifolds of $M^1$ with boundaries in $E$ which are foliated by
9869: characteristic curves. Under some easily found assumptions, one could
9870: formulate a second corollary analogous to Theorem~1.3.
9871: 
9872: \section*{\S14.~An example of a nonremovable three-dimensional torus}
9873: 
9874: This final section is devoted to exhibit a crucial example of a closed
9875: subset $C$ violating the main nontransversality condition $\mathcal{
9876: F}_{M^1}^c\{ C\}$ of Theorem~1.2' such that $C$ is truly nonremovable.
9877: In addition, we may require that $M$ and $M^1$ have the simplest
9878: possible topology.
9879: 
9880: \def\thelemma{14.1}\begin{lemma}
9881: There exists a triple $(M, M^1, C)$, where
9882: \begin{itemize}
9883: \item[{\bf (i)}]
9884: $M$ is a $\mathcal{ C}^\infty$-smooth generic submanifold in
9885: $\C^3$ of CR dimension $1$, diffeomorphic to a real $4$-ball{\rm ;}
9886: \item[{\bf (ii)}]
9887: $M^1$ is a $\mathcal{ C}^\infty$-smooth one-codimensional
9888: submanifold of $M$ which is maximally real in $\C^n$ and diffeomorphic
9889: to a real $3$-ball{\rm ;}
9890: \item[{\bf (iii)}]
9891: $C$ is a compact subset of $M^1$
9892: diffeomorphic to a real three-dimensional torus which is everywhere
9893: transversal to the characteristic foliation $\mathcal{ F}_{M^1}^c$, 
9894: hence the nontransversality condition $\mathcal{ F}_{M^1}^c\{C\}$ of
9895: Theorem~1.2' clearly does not hold{\rm ;}
9896: \item[{\bf (iv)}]
9897: $M$ of finite type $4$ in the sense of
9898: T.~Bloom and I.~Graham at every point, hence globally minimal,
9899: \end{itemize} 
9900: such that $C$ is neither CR- nor 
9901: $\mathcal{ W}$- nor $L^{\sf p}$-removable with respect to $M$.
9902: \end{lemma}
9903: 
9904: By type 4 at a point $p\in M$, we mean of course that
9905: the Lie brackets of the complex tangent bundle $T^cM$ up to length $4$
9906: generate $T_pM$.
9907: 
9908: \subsection*{14.2.~The geometric recipe}
9909: We first construct the $3$-torus $C$, then construct the maximally
9910: real $M^1$ and finally define $M$ as a certain thickening of
9911: $M^1$. The argument for insuring global minimality of $M$ involves
9912: computations with Lie brackets and is postponed to the
9913: end.
9914: 
9915: Firstly, in $\R^3 =\R^3\oplus i \{0\}\subset \C^3$ equipped with the
9916: coordinates $(x_1,x_2,x_3)$, where $x_j={\rm Re}\, z_j$ for $j=1,2,3$,
9917: pick the ``standard'' $2$-dimensional torus $T^2$ of Cartesian
9918: equation
9919: \def\theequation{14.3}\begin{equation}
9920: \left(\sqrt{x_1^2+x_2^2}-2\right)^2+x_3^2=1.
9921: \end{equation}
9922: This torus is stable under the rotations directed by the $x_3$-axis;
9923: its intersection with the $(x_1,x_3)$-plane consists of two circles of
9924: radius $1$ centered at the points $x_1=2$ and $x_1=-2$; it bounds a
9925: three-dimensional open ``full'' torus $T^3$; both $T^2$ and $T^3$ are
9926: contained in the ball $B^3$ of radius $5$ centered at the origin.
9927: 
9928: It is better to drop the square root: one checks that the equations
9929: of $T^2$ and $T^3$ are equally given by $T^2:=\{\rho =0\}$ and
9930: $T^3:= \{\rho<0\}$, by means of the {\it polynomial}\, defining 
9931: function
9932: \def\theequation{14.4}\begin{equation}
9933: \rho(x_1,x_2,x_3):=
9934: (x_1^2+x_2^2+x_3^2+3)^2-16\, (x_1^2+x_2^2),
9935: \end{equation}
9936: which has nonvanishing differential at every point of $T^2$.
9937: Consequently, the extrinsinc complexification of $T^2$, namely the
9938: complex hypersurface defined by
9939: \def\theequation{14.5}\begin{equation}
9940: \Sigma:= \{(z_1,z_2,z_3)\in \C^3: \
9941: \rho(z_1,z_2,z_3)=0\}
9942: \end{equation}
9943: cuts $\R^3$ along $T^2$ with the transversality property $T_x \R^3
9944: \cap T_x\Sigma= T_x T^2$ for every point $x\in T^2$.
9945: 
9946: Secondly, according to G.~Reeb ({\it see}\, \cite{ cln}, pp.~25--27;
9947: {\it see}\, also the figures there), by considering the space
9948: $\R^3\equiv S^3 \backslash \{ \infty\}$ as a punctured
9949: three-dimensional sphere $S^3$, one may glue a second
9950: three-dimensional full torus $\widetilde{ T}^3$ to $T^3$ along $T^2$
9951: with $\infty\in \widetilde{ T}^3$ and then construct a foliation of
9952: $S^3$ by $2$-dimensional surfaces all of whose leaves, except one, are
9953: diffeomorphic to $\R^2$, are contained in either $T^3$ or in
9954: $\widetilde{ T}^3$ and are accumulating on $T^2$, and finally, whose
9955: single compact leaf is the above $2$-torus $T^2$. This yields the
9956: so-called {\it Reeb foliation} of $S^3$, which is $\mathcal{
9957: C}^\infty$-smooth and orientable. Consequently, there exists a
9958: $\mathcal{ C}^\infty$ smooth vector field $L= a_1(x)\, \partial_{x_1}+
9959: a_2(x) \, \partial_{x_2} + a_3(x) \, \partial_{x_3}$ of norm $1$,
9960: namely $a_1(x)^2+ a_2(x)^2+ a_3(x)^2=1$ for every $x\in \R^3$, which
9961: is everywhere orthogonal (with respect to the standard Euclidean
9962: structure) to the leaves of the Reeb foliation. Geometrically, the
9963: integral curves of $L$ accumulate asymptotically on the two nodal
9964: (central) circles of $T^3$ and of $\widetilde{ T}^3$.
9965: 
9966: The open ball $B^3\subset \R^3$ of radius $5$ centered at the origin
9967: will be our maximally real submanifold $M^1$. The two-dimensionaly
9968: torus $T^2$ will be our nonremovable closed set $C$. The integral
9969: curves of the vector field $L$ will be our characteristic lines.
9970: Since $L$ is orthogonal to $T^2$, these characteristic lines will of
9971: course be everywhere transverse to $C$, so that $\mathcal{
9972: F}_{M^1}^c\{C\}$ clearly does not hold.
9973: 
9974: Thirdly, it remains to construct the generic submanifold $M$ of
9975: CR dimension $1$ containing $M^1$ and to check that $C$ will be
9976: nonremovable.
9977: 
9978: First of all, we notice that $L$ provides the characteristic
9979: directions of $M^1$ if and only if $T_x M= T_p\R^3\oplus \R \, J\,
9980: L(x)$ for every point $x\in M^1\equiv B^3$. Consequently, all
9981: submanifolds $M\subset \C^3$ obtained by slightly thickening $M^1$ in
9982: the direction of $J\, L(x)$ will be convenient; in other words, only
9983: the first jet of $M$ along $M^1$ is prescribed by our choice of the
9984: characterisctic vector field $L$. Notice that all such thin strips
9985: $M$ along $M^1$ will be diffeomorphic to a real $4$-ball.
9986: 
9987: The fact that $C$ is nonremovable for all such generic submanifolds
9988: $M$ is now clear: the hypersurface $\Sigma=\{z\in \C^3 : \, \rho(z)
9989: =0\}$ satisfying $T_x\Sigma= T_xT^2\oplus \R \, J\, T_xT^2$ for all
9990: $x\in T^2$ and $L$ being transversal to $T^2$, we easily deduce the
9991: transversality property $T_x \Sigma + T_x M= T_x\C^3$ for all $x\in
9992: T^2$, a geometric property which insures that the holomorphic function
9993: $1/\rho(z)$, which is CR on $M\backslash C$, does not extend
9994: holomorphically to any wedge of edge $M$ at any point of $C$.
9995: Intuitively, $T_x\Sigma/T_xM$ absorbs all the normal space 
9996: $T_x\C^3/T_x M$ at every point $x\in T^2$, leaving no room
9997: for any open cone.
9998: 
9999: Finally, to fulfill all the hypotheses of Theorem~1.2' (except of
10000: course $\mathcal{ F}_{M^1}^c\{C\}$), we have to insure that $M$ is
10001: globally minimal. We claim that by bending strongly the second and the
10002: fourth order jet of $M$ along $M^1$ (without modifying the first order
10003: jet which must be prescribed by $J\, L$), one may insure that $M$ is
10004: of type 4 in the sense of T.~Bloom and I.~Graham at every point of
10005: $M^1$; since being of finite type is an open property, it follows that
10006: $M$ is finite type at every point provided that, as a strip, $M$ is
10007: sufficiently thin along $M^1$. As is known, finite-typeness at every
10008: point implies local minimality at every point which in turn implies
10009: global minimality. This completes the recipe.
10010: 
10011: We would like to mention that by following a similar recipe, one may
10012: construct an elementary example of a non-removable compact subset of a
10013: generic submanifold of codimension one diffeomorphic to a $4$-ball
10014: lying in a globally minimal hypersurface in $\C^3$ which is (also)
10015: diffeomorphic to a $5$-ball ({\it cf.} \cite{ js}).
10016: 
10017: \subsection*{14.6.~Finite-typisation}
10018: Thus, it remains to construct a generic submanifold $M\subset \C^3$ of
10019: CR dimension $1$ satisfying $T_x M = T_xM^1 \oplus \R \, J \, L(x)$
10020: for every $x\in M^1$, which is of {\it type $4$ at every point $x\in
10021: M^1$}.
10022: 
10023: First of all, let us denote by $L=a_1(x)\, \partial_{x_1}+ a_2(x) \,
10024: \partial_{x_2}+ a_3(x) \, \partial_{x_3}$ the unit vector field which
10025: was constructed as a field orthogonal to the Reeb foliation: it is
10026: defined over $\R^3$ and has $\mathcal{ C}^\infty$-smooth coefficients
10027: satisfying $a_1(x)^2+ a_2(x)^2+ a_3(x)^2= 1$ for all $x\in\R^2$. The
10028: two-dimensional quotient vector bundle $T\R^3 / (\R L)$ with
10029: contractible base being necessarily trivial, it follows that we can
10030: complete $L$ by two other $\mathcal{ C }^\infty$-smooth unit
10031: vector fields $K^1$ and $K^2$ defined over $\R^3$ such that the triple
10032: $(L(x), K^1(x), K^2(x))$ forms a direct orthonormal frame at every
10033: point $x\in \R^3$. Let us denote the coefficients of $K^1$ and of
10034: $K^2$ by
10035: \def\theequation{14.7}\begin{equation}
10036: \aligned
10037: K^1 
10038: & \
10039: = \rho_1 \, \partial_{x_1}+ 
10040: \rho_2\, \partial_{x_2} + 
10041: \rho_3 \, \partial_{x_3},\\
10042: K^2 
10043: & \
10044: =
10045: r_1 \, \partial_{x_1}+ 
10046: r_2 \, \partial_{x_2}+ 
10047: r_3 \, \partial_{x_3},
10048: \endaligned
10049: \end{equation}
10050: where $\rho_j$ and $r_j$ for $j=1,2,3$ are $\mathcal{
10051: C}^\infty$-smooth functions of $x\in \R^3$ satisfying $\rho_1^2+
10052: \rho_2^2+ \rho_3^2=1$ and $r_1^2+r_2^2+r_3^2=1$. In our case, $K^1$
10053: and $K^2$ may even be constructed directly by means of a
10054: trivialization of the bundle tangent to the Reeb foliation.
10055: 
10056: Let $P>0$ be a constant, which will be chosen later to be large. Since
10057: by construction we have the two orthogonality relations $a_1 \rho_1 +
10058: a_2 \rho_2 + a_3 \rho_3= 0$ and $a_1 r_1 + a_2 r_2 + a_3 r_3= 0$, it
10059: follows that every generic submanifold $M_P\subset 
10060: \C^3$ defined by the two
10061: Cartesian equations
10062: \def\theequation{14.8}\begin{equation}
10063: \aligned
10064: 0 
10065: & \
10066: = \rho 
10067: = 
10068: y_1 \, \rho_1(x) + 
10069: y_2 \, \rho_2(x) +
10070: y_3 \, \rho_3(x) + 
10071: P 
10072: \left[
10073: y_1^2+ y_2^2+y_3^2
10074: \right], \\
10075: 0 
10076: & \
10077: = r 
10078: = 
10079: y_1\, r_1(x) + 
10080: y_2 \, r_2(x) + 
10081: y_3 \, r_3(x)+
10082: P^3
10083: \left[
10084: y_1^4 + y_2^4 + y_3^4
10085: \right]
10086: \endaligned
10087: \end{equation}
10088: enjoys the property that the vector field $J L(x)= a_1(x) \,
10089: \partial_{y_1}+ a_2(x) \, \partial_{x_2}+ a_3(x) \, \partial_{x_3}$ is
10090: tangent to $M_P$ at every $x\in \R^3$. As desired, we deduce that
10091: $T_x^c M = \R L(x) \oplus J \R L(x)$ for every $x\in \R^3$, a property
10092: which insures that $\R L(x)$ is the characteristic direction of $M^1$
10093: in $M_P$, independently of $P$.
10094: 
10095: To complete the final minimalization argument for the construction of
10096: a nonremovable compact set $C:= T^2 \subset M^1\subset M$ which
10097: appears in the Introduction, it suffices now to apply the following
10098: lemma with $R= 5$. Though calculatory, its proof is totally
10099: elementary.
10100: 
10101: \def\thelemma{14.9}\begin{lemma}
10102: For every $R>0$, there exist $P>0$ sufficiently large
10103: such that $M_P$ is of type $4$ at every point
10104: $x\in \R^3$ with $x_1^2+ x_2^2+ x_3^2\leq R^2$. 
10105: \end{lemma}
10106: 
10107: \proof
10108: As above, let $M_P = \{z\in \C^3: \, \rho =r =0\}$. By writing the
10109: tangency condition, one checks immediately that the one-dimensional
10110: complex vector bundle $T^{1, 0}M_P$ is generated over $\C$ by the
10111: vector field $\LL := A_1 \, \partial_{z_1}+ A_2 \, \partial_{ z_2}+
10112: A_3\, \partial_{ z_3}$, with the explicit expressions
10113: \def\theequation{14.10}\begin{equation}
10114: \aligned
10115: A_1:= 
10116: & \ 
10117: 4\rho_{z_3}r_{z_2}- 4 \rho_{z_2} r_{z_3}, \\
10118: A_2:= 
10119: & \ 
10120: 4\rho_{z_1}r_{z_3}- 4 \rho_{z_3} r_{z_1}, \\
10121: A_3:= 
10122: & \ 
10123: 4\rho_{z_2}r_{z_1}- 4 \rho_{z_1} r_{z_2}. \\
10124: \endaligned
10125: \end{equation}
10126: Using the expressions~\thetag{ 14.8} for $\rho$ and $r$, we see that
10127: these three components restrict on $\{y=0\}$ as the Pl\"ucker
10128: coordinates of the bivector $(K^1,K^2)$, namely
10129: \def\theequation{14.11}\begin{equation}
10130: \aligned
10131: A_1 \vert_{y=0}=
10132: & \
10133: \rho_2 r_3- \rho_3 r_2=:
10134: \Delta_{2,3}, \\
10135: A_2\vert_{y=0}=
10136: & \
10137: \rho_3 r_1- \rho_1 r_3
10138: =:\Delta_{3,1}, \\
10139: A_3 \vert_{y=0}=
10140: & \
10141: \rho_1 r_2- \rho_2 r_1
10142: =:\Delta_{1,2}.
10143: \endaligned
10144: \end{equation}
10145: As $K^1$ and $K^2$ are of norm $1$ and orthogonal at every point, it
10146: follows by direct computation that $\Delta_{2,3}^2 + \Delta_{3,1}^2+
10147: \Delta_{ 1,2}^2 = 1$ and that the vector of coordinates $(\Delta_{
10148: 2,3}, \Delta_{3,1}, \Delta_{1,2})$ is orthogonal to both $K^1$ and
10149: $K^2$. Moreover, as the orthonormal trihedron $(L(x), K^1(x), K^2(x))$
10150: is direct at every point, we deduce that necessarily
10151: \def\theequation{14.12}\begin{equation}
10152: \Delta_{2,3} \equiv a_1, \ \ \ \ \
10153: \Delta_{3,1} \equiv a_2, \ \ \ \ \
10154: \Delta_{1,2} \equiv a_3.
10155: \end{equation}
10156: 
10157: Next, we compute in length $A_1$, $A_2$ and $A_3$ using~\thetag{
10158: 14.8}. As their complete explicit development will not be crucial for
10159: the sequel and as we shall perform with them differentiations and
10160: linear combinations yielding relatively complicated expressions, let
10161: us adopt the following notation: by $\mathcal{ R}^0$, we denote
10162: various expressions which are polynomials in the jets of the functions
10163: $\rho_1,\rho_2,\rho_3$ and $r_1,r_2,r_3$. Similarly, by $\mathcal{
10164: R}^{I}$, by $\mathcal{ R}^{II}$, by $\mathcal{ R}^{III}$ and by
10165: $\mathcal{ R}^{IV}$, we denote polynomials in the transverse variables
10166: $(y_1,y_2,y_3)$ which are homogeneous of degree $1$, $2$, $3$ and $4$
10167: and have as coefficients various expressions $\mathcal{ R}^0$.
10168: 
10169: Importantly, we make the convention that such expressions $\mathcal{
10170: R}^0$, $\mathcal{ R}^I$, $\mathcal{ R}^{II}$, $\mathcal{ R}^{III}$ and
10171: $\mathcal{ R}^{IV}$ should be totally independent of the constant
10172: $P$. Consequently, if $P$ appears somehow, we shall write it as a
10173: factor, as for instance in $P \, \mathcal{ R}^I$ or in $P^3 \,
10174: \mathcal{ R}^{III}$.
10175: 
10176: With this convention at hand, we may develope~\thetag{ 14.10} using the
10177: expressions~\thetag{ 14.8} by writing out only the terms which will be
10178: useful in the sequel and by treating the rest as controlled
10179: remainders. Let us detail the computation of $A_1$:
10180: \def\theequation{14.13}\begin{equation}
10181: \aligned
10182: A_1 
10183: & \
10184: =
10185: 4 \, 
10186: \left[
10187: -\frac{ i}{2} \rho_3 - iP y_3 + \mathcal{ R}^I
10188: \right] 
10189: \left[
10190: -\frac{ i}{2} r_2 - 2i P^3 y_2^3 + 
10191: \mathcal{ R}^I
10192: \right]- \\
10193: & \ \ \ \ \ \ \ 
10194: - 
10195: 4\, 
10196: \left[
10197: -\frac{ i}{2} \rho_2 - iP y_2 + \mathcal{ R}^I
10198: \right] 
10199: \left[
10200: - \frac{ i}{2} r_3 - 2i P^3 y_3^3 + 
10201: \mathcal{ R}^I
10202: \right] \\
10203: & 
10204: \
10205: =
10206: -\rho_3 r_2 -4 P^3 \rho_3y_2^3+ \mathcal{ R}^I - 
10207: 2P r_2y_3+ P^4 \mathcal{ R}^{IV}+ P \mathcal{ R}^I + 
10208: \mathcal{ R}^I + P^3\mathcal{ R}^{IV}+ 
10209: \mathcal{ R}^{II} \\
10210: & \ \ \ \ \
10211: +
10212: \rho_2 r_3 + 4 P^3 \rho_2 y_3^3+
10213: \mathcal{ R}^I +
10214: 2P r_3y_2+ P^4 \mathcal{ R}^{IV}+ P \mathcal{ R}^I + 
10215: \mathcal{ R}^I + P^3\mathcal{ R}^{IV}+ 
10216: \mathcal{ R}^{II} \\
10217: & \
10218: =
10219: \rho_2 r_3 - \rho_3 r_2 + 
10220: 2P r_3 y_2 - 2P r_2 y_3 + 
10221: 4P^3 \rho_2 y_3^3 - 4 P^3 \rho_3 y_2^3 + \\
10222: & \ \ \ \ \
10223: +
10224: \mathcal{ R}^I + \mathcal{ R}^{II}+
10225: P \mathcal{ R}^{II}+
10226: P^3 \mathcal{ R}^{IV}+
10227: P^4 \mathcal{ R}^{IV}.
10228: \endaligned
10229: \end{equation}
10230: In the development, before simplification, we firstly write out in 
10231: lines 3 and 4 all
10232: the $9\times 2$ terms of the two product: for instance, the third term
10233: of the first product, namely 
10234: $4(-\frac{i}{ 2} \rho_3 )(\mathcal{ R}^I)$, yields
10235: a term $\mathcal{ R}^I$ whereas the fifth term $4(-iPy_3)(-2iP^3
10236: y_2^3)$ yields a term $P^4 \mathcal{ R}^{IV}$; secondly, we simplify
10237: the obtained sum: by our convention, $\mathcal{ R}^{I}+ \mathcal{
10238: R}^I=\mathcal{ R}^I$, whereas $\mathcal{ R}^I+ P\mathcal{ R}^I$ cannot
10239: be simplified, since the large constant $P$ will be chosen later.
10240: With these technical explanations at hand, we shall not provide any
10241: intermediate detail for the further computations, whose rules are
10242: totally analogous. For $A_1$, $A_2$ and $A_3$, we obtain
10243: \def\theequation{14.14}\begin{equation}
10244: \left\{
10245: \aligned
10246: A_1 
10247: & \
10248: =
10249: \rho_2 r_3 - \rho_3 r_2 + 
10250: 2P r_3 y_2 - 2P r_2 y_3 + 
10251: 4P^3 \rho_2 y_3^3 - 4 P^3 
10252: \rho_3 y_2^3 + \\
10253: & \ \ \ \ \
10254: +
10255: \mathcal{ R}^I + \mathcal{ R}^{II}+
10256: P \mathcal{ R}^{II}+
10257: P^3 \mathcal{ R}^{IV}+
10258: P^4 \mathcal{ R}^{IV}, \\
10259: A_2
10260: & \
10261: =
10262: \rho_3 r_1 - \rho_1 r_3 + 
10263: 2P r_1 y_3 - 2P r_3 y_1 + 
10264: 4P^3 \rho_3 y_1^3 - 4 P^3 
10265: \rho_1 y_3^3 + \\
10266: & \ \ \ \ \
10267: +
10268: \mathcal{ R}^I + \mathcal{ R}^{II}+
10269: P \mathcal{ R}^{II}+
10270: P^3 \mathcal{ R}^{IV}+
10271: P^4 \mathcal{ R}^{IV}, \\
10272: A_3
10273: & \
10274: =
10275: \rho_1 r_2 - \rho_2 r_1 + 
10276: 2P r_2 y_1 - 2P r_1 y_2 + 
10277: 4P^3 \rho_1 y_2^3 - 4 P^3 
10278: \rho_2 y_1^3 + \\
10279: & \ \ \ \ \
10280: +
10281: \mathcal{ R}^I + \mathcal{ R}^{II}+
10282: P \mathcal{ R}^{II}+
10283: P^3 \mathcal{ R}^{IV}+
10284: P^4 \mathcal{ R}^{IV}.
10285: \endaligned\right.
10286: \end{equation}
10287: 
10288: Now that we have written the complex vector field $\LL$ and its
10289: coefficients $A_1$, $A_2$ and $A_3$, in order to establish Lemma~14.9,
10290: it suffices to choose $P>0$ sufficiently large in order that the four
10291: complex vector fields
10292: \def\theequation{14.15}\begin{equation}
10293: \overline{ \LL}\, \vert_{y=0} , 
10294: \ \ \ \ \
10295: \LL \vert_{y=0}, 
10296: \ \ \ \ \
10297: \left[
10298: \overline{ \LL}, \LL
10299: \right]\, \vert_{y=0}, 
10300: \ \ \ \ \
10301: \left[
10302: \overline{\LL}, 
10303: \left[
10304: \overline{\LL}, 
10305: \left[
10306: \overline{ \LL}, \LL
10307: \right]\right]\right]
10308: \vert_{y=0}
10309: \end{equation}
10310: are linearly independent at every point $x \in \R^3$ with $x_1^2 +
10311: x_2^2 + x_3^2 \leq R^2$. At the end of the proof, we shall explain why
10312: we cannot insure type $3$ at every point, namely why the consideration
10313: of $\left[ \overline{\LL}, \left[ \overline{ \LL}, \LL
10314: \right]\right]\vert_{y=0}$ instead of the length four last Lie bracket
10315: in~\thetag{ 14.15} would fail.
10316: 
10317: As promised, we shall now summarize all the subsequent
10318: computations. As we aim to restrict the last Lie bracket to $\{y=0\}$
10319: which is of length four and whose coefficients involve derivatives of
10320: order at most three of the coefficients $A_1$, $A_2$ and $A_3$, we can
10321: already neglect the last two remainders $P^3\mathcal{ R}^{IV}$ and
10322: $P^4 \mathcal{ R}^{IV}$ in~\thetag{ 14.14}. In other words, we can
10323: consider $A^1$, $A^2$ and $A^3 \ {\rm mod} (IV)$. Similarly, in the
10324: computation of the Lie bracket
10325: \def\theequation{14.16}\begin{equation}
10326: \left[\overline{\LL},
10327: \LL\right]=:
10328: C_1 \, \partial_{z_1}+
10329: C_2 \, \partial_{z_2}+
10330: C_3 \, \partial_{z_3}- 
10331: \overline{C_1} \, \partial_{\bar z_1}- 
10332: \overline{C_2}\, \partial_{\bar z_2} - 
10333: \overline{C_3} \, \partial_{\bar z_3},
10334: \end{equation}
10335: before restriction to $\{y=0\}$, we can restrict our task to
10336: developing the coefficients
10337: \def\theequation{14.17}\begin{equation}
10338: \aligned
10339: C_1 
10340: & \
10341: :=
10342: \overline{A_1} A_{1,\bar z_1}+
10343: \overline{A_2} A_{1,\bar z_2}+ 
10344: \overline{A_3} A_{1,\bar z_3}, \\
10345: C_2
10346: & \
10347: :=
10348: \overline{A_1} A_{2,\bar z_1}+
10349: \overline{A_2} A_{2,\bar z_2}+ 
10350: \overline{A_3} A_{2,\bar z_3}, \\
10351: C_3
10352: & \
10353: :=
10354: \overline{A_1} A_{3,\bar z_1}+
10355: \overline{A_2} A_{3,\bar z_2}+ 
10356: \overline{A_3} A_{3,\bar z_3} \\
10357: \endaligned
10358: \end{equation}
10359: only modulo order $(III)$, which yields by means of
10360: the expressions~\thetag{ 14.14}
10361: \def\theequation{14.18}\begin{equation}
10362: \aligned
10363: C_1 \ {\rm mod} \, 
10364: (III) 
10365: & \
10366: \equiv 
10367: -iP \rho_1 +
10368: 6i P^3 a_3 \rho_2 y_3^2 -
10369: 6i P^3 a_2 \rho_3 y_2^2 +
10370: \mathcal{ R}^0 +
10371: \mathcal{ R}^I+ \\
10372: & \ \ \ \ \ \ 
10373: +
10374: P\mathcal{ R}^I+
10375: P^2 \mathcal{ R}^I
10376: +\mathcal{ R}^{II}+
10377: P\mathcal{ R}^{II} +
10378: P^2 \mathcal{ R}^{II}, \\ 
10379: C_2 \ {\rm mod} \, 
10380: (III) 
10381: & \
10382: \equiv 
10383: -iP \rho_2 +
10384: 6i P^3 a_1 \rho_3 y_1^2 -
10385: 6i P^3 a_3 \rho_1 y_3^2 +
10386: \mathcal{ R}^0 +
10387: \mathcal{ R}^I+ \\
10388: & \ \ \ \ \ \ 
10389: +
10390: P\mathcal{ R}^I+
10391: P^2 \mathcal{ R}^I
10392: +\mathcal{ R}^{II}+
10393: P\mathcal{ R}^{II} +
10394: P^2 \mathcal{ R}^{II}, \\
10395: C_3 \ {\rm mod} \, 
10396: (III) 
10397: & \
10398: \equiv 
10399: -iP \rho_3 +
10400: 6i P^3 a_2 \rho_1 y_2^2 -
10401: 6i P^3 a_1 \rho_2 y_1^2 +
10402: \mathcal{ R}^0 +
10403: \mathcal{ R}^I+ \\
10404: & \ \ \ \ \ \ 
10405: +
10406: P\mathcal{ R}^I+
10407: P^2 \mathcal{ R}^I
10408: +\mathcal{ R}^{II}+
10409: P\mathcal{ R}^{II} +
10410: P^2 \mathcal{ R}^{II}. 
10411: \endaligned
10412: \end{equation} 
10413: We must mention the use of natural rule hold for computing the partial
10414: derivatives $A_{j, \bar z_k}$: we have for instance $\partial_{\bar
10415: z_k} \left( \mathcal{ R}^{II} \right)= \mathcal{ R}^I+ \mathcal{
10416: R}^{II}$. Also, we have used the hypothesis that $(L(x), K^1(x),
10417: K^2(x))$ provides a direct orthonormal frame at every $x\in \R^3$,
10418: which yields in particular the three relations
10419: \def\theequation{14.19}\begin{equation}
10420: a_2 r_3 - a_3 r_2=-\rho_1, \ \ \ \ \ 
10421: a_3 r_1-a_1 r_3= -\rho_2, \ \ \ \ \ 
10422: a_1r_2-a_2r_1= -\rho_3.
10423: \end{equation}
10424: After mild computation, the coefficients 
10425: $F_1$, $F_2$ and $F_3$ of the length four Lie bracket
10426: \def\theequation{14.20}\begin{equation}
10427: \left[
10428: \overline{\LL}, 
10429: \left[
10430: \overline{\LL}, 
10431: \left[
10432: \overline{\LL}, \LL
10433: \right]
10434: \right]
10435: \right]=
10436: F_1 \, \partial_{z_1}+
10437: F_2 \, \partial_{z_2}+
10438: F_3 \, \partial_{z_3}+
10439: G_1 \, \partial_{\bar z_1}+
10440: G_2 \, \partial_{\bar z_2}+
10441: G_3 \, \partial_{\bar z_3}
10442: \end{equation}
10443: are given, after restriction to $\{y=0\}$, by 
10444: \def\theequation{14.21}\begin{equation}
10445: \aligned
10446: F_1\vert_{y=0} 
10447: & \
10448: =
10449: 3i P^3 a_2^3\rho_3 -3i P^3 a_3^3 \rho_2 +
10450: \mathcal{ R}^0+ 
10451: P\mathcal{R}^0 +
10452: P^2\mathcal{ R}^0, \\
10453: F_2\vert_{y=0} 
10454: & \
10455: =
10456: 3i P^3 a_3^3\rho_1 -3i P^3 a_1^3 \rho_3 +
10457: \mathcal{ R}^0+ 
10458: P\mathcal{R}^0 +
10459: P^2\mathcal{ R}^0, \\
10460: F_3\vert_{y=0} 
10461: & \
10462: =
10463: 3i P^3 a_1^3\rho_2 -3i P^3 a_2^3 \rho_1 +
10464: \mathcal{ R}^0+ 
10465: P\mathcal{R}^0 +
10466: P^2\mathcal{ R}^0, \\
10467: \endaligned
10468: \end{equation}
10469: 
10470: We can now complete the proof of Lemma~14.9. In the basis
10471: $(\partial_{z_1}, \partial_{z_2}, \partial_{z_3}, \partial_{\bar z_1},
10472: \partial_{\bar z_2}, \partial_{\bar z_3})$, the $4\times 6$ matrix
10473: associated with the four vector fields~\thetag{ 14.15} (without
10474: mentioning $\vert_{y=0}$)
10475: \def\theequation{14.22}\begin{equation}
10476: \left(
10477: \begin{array}{cccccc}
10478: 0 & 0 & 0 & a_1 & a_2 & a_3 \\
10479: a_1 & a_2 & a_3 & 0 & 0 & 0 \\
10480: C_1 & C_2 & C_3 & -\overline{C_1} 
10481: & - \overline{C_2} & -\overline{C_3} \\
10482: F_1 & F_2 & F_3 & G_1 & G_2 & G_3
10483: \end{array}
10484: \right)
10485: \end{equation}
10486: has rank four at a point $x\in\R^3$ if and only if the $3\times 3$
10487: determinant in the left low corner is nonvanishing, namely if and only
10488: if the developped expression
10489: \def\theequation{14.23}\begin{equation}
10490: \aligned
10491: {}
10492: & \
10493: \left\vert
10494: \begin{array}{ccc}
10495: a_1 & a_2 & a_3 \\
10496: -iP\rho_1+\mathcal{ R}^0 &
10497: -iP\rho_2+\mathcal{ R}^0 &
10498: -iP\rho_3+\mathcal{ R}^0 \\
10499: 3iP^3a_2^3\rho_3-3iP^3a_3^3\rho_2+ &
10500: 3iP^3a_3^3\rho_1-3iP^3a_1^3\rho_3+ &
10501: 3iP^3a_1^3\rho_2-3iP^3a_2^3\rho_1+ \\ 
10502: +\mathcal{ R}^0+P\mathcal{ R}^0+P^2\mathcal{ R}^0 &
10503: +\mathcal{ R}^0+P\mathcal{ R}^0+P^2\mathcal{ R}^0 &
10504: +\mathcal{ R}^0+P\mathcal{ R}^0+P^2\mathcal{ R}^0 
10505: \end{array}
10506: \right\vert \\
10507: & \
10508: =
10509: 3P^4 \left(
10510: r_3 [a_1^3 \rho_2 -a_2^3 \rho_1]+
10511: r_2[a_3^3\rho_1 -a_1^3 \rho_3]+
10512: r_1 [a_2^3 \rho_3 -a_3^3 \rho_2]
10513: \right)+ \\
10514: & \ \ \ \ \ \
10515: +
10516: \mathcal{ R}^0+
10517: P\mathcal{ R}^0+
10518: P^2 \mathcal{ R}^0+
10519: P^3\mathcal{ R}^0 +
10520: P^4\mathcal{ R}^0 \\
10521: & \
10522: =
10523: 3P^4 \left(a_1^4+a_2^4+a_3^4\right)+
10524: \mathcal{ R}^0+
10525: P\mathcal{ R}^0+
10526: P^2 \mathcal{ R}^0+
10527: P^3\mathcal{ R}^0 +
10528: P^4\mathcal{ R}^0
10529: \endaligned
10530: \end{equation}
10531: is nonvanishing. 
10532: 
10533: At this point, the conclusion of the lemma is now an immediate
10534: consequence of the following trivial assertion: {\it Let $a_1$, $a_2$
10535: and $a_3$ be $\mathcal{ C }^\infty$-smooth functions on $\R^3$
10536: satisfying $a_1 (x)^2+ a_2 (x)^2+a_3(x)^2=1$ for all $x \in \R^3$ and
10537: let $\mathcal{ R }_0^0$, $\mathcal{ R}_1^0$, $\mathcal{ R }_2^0$,
10538: $\mathcal{ R }_3^0$ and $\mathcal{ R}_4^0$ be $\mathcal{ C
10539: }^\infty$-smooth functions on $\R^3$. For every $R>0$, there exists a
10540: constant $P>0$ large enough so that the function
10541: \def\theequation{14.24}\begin{equation}
10542: 3P^4 \left(
10543: a_1^4+a_2^4+a_3^4\right)+
10544: \mathcal{ R}_0^0+P\mathcal{ R}_1^0+
10545: P^2\mathcal{ R}_2^0 +
10546: P^3 \mathcal{ R}_3^0+
10547: P^4\mathcal{ R}_4^0 
10548: \end{equation}
10549: is positive at every $x\in \R^3$ with $x_1^2 + x_2^2+ x_3^2 \leq R^2$}.
10550: 
10551: If we had put $y_1^3+ y_2^3+ y_3^3$ instead of $y_1^4+ y_2^4+ y_3^4$
10552: in the second equation~\thetag{14.8}, we would have considered the
10553: length three Lie bracket $\left[ \overline{\LL}, \left[
10554: \overline{\LL}, \LL \right] \right] \vert_{y=0}$ instead of the length
10555: four Lie bracket in~\thetag{14.15}, and hence instead of the quartic
10556: $a_1^4+ a_2^4+ a_3^4$ in~\thetag{ 14.24}, we would have obtained the
10557: cubic $a_1^3+a_2^3+a_3^3$, a function which (unfortunately) vanishes,
10558: for instance if $a_1(x) =\frac{ 1}{\sqrt{ 2}}$, $a_2(x)= -\frac{ 1}{
10559: \sqrt{ 2}}$ and $a_3(x) =0$. We notice that in our example, this value
10560: of $(a_1,a_2,a_3)$ is indeed attained at the point $x \in T^2$ of
10561: coordinates $(\frac{ 3}{\sqrt{2}}, -\frac{ 3}{ \sqrt{ 2}}, 0)$, whence
10562: the necessity of passing to type 4.
10563: The proof of Lemma~14.9 is complete.
10564: \endproof
10565: 
10566: \begin{thebibliography}{XL}
10567: 
10568: \bibitem[A]{a}
10569: {\sc Ayrapetian}, R.A.: 
10570: {\em Extending CR functions from piecewise smooth CR manifolds}. 
10571: Mat. Sbornik {\bf 134} (1987), 108--118. Trad. in english in Math. 
10572: {\sc Ussr} Sbornik {\bf 62} (1989), 1, 111--120.
10573: 
10574: \bibitem[BT]{bt} 
10575: {\sc Baouendi}, M.S.;
10576: {\sc Treves}, F: 
10577: {\em A property of the functions and distributions annihilated by a
10578: locally integrable system of complex vector fields},
10579: Ann. of Math.
10580: {\bf 113} (1981), no.2, 387--421.
10581: 
10582: \bibitem[BER]{ber} 
10583: {\sc Baouendi}, M.S.;
10584: {\sc Ebenfelt}, P.; {\sc Rothschild}, L.P.: 
10585: {\em Real submanifolds in complex space and their mappings}.
10586: Princeton Mathematical Series, {\bf 47}, Princeton University Press,
10587: Princeton, NJ, 1999, xii+404 pp.
10588: 
10589: \bibitem[BK]{bk}
10590: {\sc Bedford}, E.; {\sc Klingenberg}, W.:
10591: {\em On the envelope of holomorphy of a $2$-sphere in $\C^2$}, 
10592: J. Amer. Math. Soc. {\bf 4} (1991), 623--646.
10593: 
10594: \bibitem[Be]{be}
10595: {\sc Bennequin}, D.:
10596: {\em Entrelacements et \'equations de Pfaff}, Ast\'erisque 107/108, 
10597: 87--161, 1983.
10598: 
10599: \bibitem[B]{b}
10600: {\sc Bishop}, E.:
10601: {\em Differentiable manifolds in complex Euclidean space}, 
10602: Duke Math. J. {\bf 32} (1965), 1--22.
10603: 
10604: \bibitem[Bo]{bo} 
10605: {\sc Boggess, A.}:
10606: {\em CR manifolds and the tangential Cauchy-Riemann complex}. 
10607: Studies in Advanced Mathematics. CRC Press,
10608: Boca Raton, FL, 1991, xviii+364 pp. 
10609: 
10610: \bibitem[CLN]{cln}
10611: {\sc Camacho}, C.; {\sc Lins Neto}, A.:
10612: {\em Geometric theory of foliations}, 
10613: Birkh\"auser, Boston, 1985.
10614: 
10615: \bibitem[CS]{cs} 
10616: {\sc Chirka}, E.M.; {\sc Stout}, E.L.:
10617: {\em Removable singularities in the boundary}. 
10618: Contributions to complex analysis and analytic geometry, 43--104,
10619: Aspects Math., E26, Vieweg, Braunschweig, 1994.
10620: 
10621: \bibitem[DS]{ds} 
10622: {\sc Dinh, T.C.}; {\sc Sarkis, F.}: {\it Wedge removability of
10623: metrically thin sets and application to the CR meromorphic
10624: extension}. Math. Z. {\bf 238} (2001), no.3, 639--653.
10625: 
10626: \bibitem[D]{d}
10627: {\sc Duval}, J.: 
10628: {\em Surfaces convexes dans un bord pseudoconvexe}, 
10629: Colloque d'Analyse Complexe et G\'eom\'etrie, Marseille, 1992~;
10630: Ast\'erisque {\bf 217}, Soc. Math. France, 
10631: Montrouge, 1993, 6, 103--118.
10632: 
10633: \bibitem[E]{e}
10634: {\sc Eliashberg}, Y.:
10635: {\em Classification of overtwisted contact structures
10636: on 3-manifolds}, 
10637: Invent. Math. {\bf 98} (1989), no.~3, 623--637.
10638: 
10639: \bibitem[FM]{fm}
10640: {\sc Forn{\ae}ss}, J.E.; {\sc Ma}, D.:
10641: {\em A $2$-sphere in $\C^2$ that cannot be filled
10642: in with analytic discs}, International Math. Res. 
10643: Notices, 1995, no.~1, 17--22.
10644: 
10645: \bibitem[FS]{fs}
10646: {\sc Forstneri$\check{\text{\sc c}}$}, F.; {\sc Stout}, E.L.:
10647: {\em A new class of polynomially convex sets}, Ark. Mat. {\bf 29}
10648: (1991), 52--62.
10649: 
10650: \bibitem[HT]{ht} 
10651: {\sc Hanges}, N; {\sc Treves}, F.:
10652: {\em Propagation of holomorphic extendability of CR functions},
10653: Math. Ann. {\bf 263} (1983), 157--177.
10654: 
10655: \bibitem[Ha]{ha}
10656: {\sc Hartman, P.}:
10657: {\em Ordinary Differential Equations}.
10658: Birkh\"auser, Boston 1982.
10659: 
10660: \bibitem[HL]{hl} 
10661: {\sc Harvey, R.}; {\sc Lawson, B.}:
10662: {\em On boundaries of complex analytic varieties}. Ann. of Math., I:
10663: {\bf 102} (1975), no.2, 233--290; II: {\bf 106} (1977), no.2, 213--238.
10664: 
10665: \bibitem[I]{i} 
10666: {\sc Ivashkovitch}, S.M.; 
10667: {\em The Hartogs-type extension theorem for meromorphic maps 
10668: into compact K\"ahler manifolds}, 
10669: Invent. Math., {\bf 109} (1992), no.1, 47--54.
10670: 
10671: \bibitem[J1]{j1}
10672: {\sc J\"oricke}, B.:
10673: {\em Removable singularities of CR-functions}, 
10674: Ark. Mat. {\bf 26} (1988), 117--143.
10675: 
10676: \bibitem[J2]{j2}
10677: {\sc J\"oricke, B.}: 
10678: {\em Deformation of CR-manifolds, minimal points and 
10679: CR-manifolds with the microlocal analytic extension property},
10680: J. Geom. Anal. {\bf 6} (1996), no.4, 555--611.
10681: 
10682: \bibitem[J3]{j3}
10683: {\sc J\"oricke}, B.: 
10684: {\em Local polynomial hulls of discs near isolated parabolic points},
10685: Indiana Univ. Math. J. {\bf 46} (1997), no.3, 789--826. 
10686: 
10687: \bibitem[J4]{j4}
10688: {\sc J\"oricke}, B.: 
10689: {\em Boundaries of singularity sets, removable
10690: singularities, and CR-invariant subsets of CR-manifolds},
10691: J. Geom. Anal. {\bf 9} (1999), no.2, 257--300.
10692: 
10693: \bibitem[J5]{j5}
10694: {\sc J\"oricke, B.}: 
10695: {\em Removable singularities of $L^{\sf p}$ 
10696: CR functions on hypersurfaces},
10697: J. Geom. Anal. {\bf 9} (1999), no.3, 429--456.
10698: 
10699: \bibitem[JS]{js}
10700: {\sc J\"oricke}, B.; {\sc Shcherbina}, N.:
10701: {\em A nonremovable generic 4-ball in the unit sphere of $\C^3$},
10702: Duke Math. J. {\bf 102} (2000), no.1, 87--100.
10703: 
10704: \bibitem[JP]{jp}
10705: {\sc J\"oricke, B.}; {\sc Porten}, E.:
10706: {\em Hulls and analytic extension from nonpseudoconvex boundaries}, 
10707: U.U.M.D. Report 2002:19, Uppsala University, 38~pp., 2002.
10708: 
10709: \bibitem[KR]{kr}
10710: {\sc Kytmanov}, A.M.; {\sc Rea}, C.:
10711: {\em Elimination of $L^1$ singularities on H\"older peak 
10712: sets for CR functions}, Ann. Scuola Norm. Sup. Pisa, Classe 
10713: di Scienze, {\bf 22} (1995), 211--226.
10714: 
10715: \bibitem[L]{l}
10716: {\sc Laurent-Thi\'ebaut}, C.:
10717: {\em Sur l'extension des fonctions CR dans une vari\'et\'e
10718: de Stein}, Ann. Mat. Pura Appl. {\bf 150} (1988), 141--151.
10719: 
10720: \bibitem[LP]{lp}
10721: {\sc Laurent-Thi\'ebaut}, C.; {\sc Porten}, E.:
10722: {\em Analytic extension from nonpseudoconvex boundaries and 
10723: $A(D)$-convexity}, Ann. Inst. Fourier (Grenoble), 2003, to appear.
10724: 
10725: \bibitem[Lu]{lu}
10726: {\sc Lupacciolu}, G.:
10727: {\em Characterization of removable sets in strongly pseudoconvex
10728: boundaries}, Ark. Mat. {\bf 32} (1994), 455--473.
10729: 
10730: \bibitem[M1]{m1} 
10731: {\sc Merker}, J.:
10732: {\em Global minimality of generic manifolds and holomorphic
10733: extendibility of CR functions}. Internat. Math. Res. Notices 1994,
10734: no.8, 329--342.
10735: 
10736: \bibitem[M2]{m2} 
10737: {\sc Merker}, J.:
10738: {\em On removable singularities for CR functions 
10739: in higher codimension}, 
10740: Internat. Math. Res. Notices 1997, no.1, 21--56.
10741: 
10742: \bibitem[MP1]{mp1}
10743: {\sc Merker, J.}; {\sc Porten, P.}:
10744: {\em On removable singularities for integrable CR functions},
10745: Indiana Univ. Math. J. {\bf 48} (1999), no.3, 805--856. 
10746: 
10747: \bibitem[MP2]{mp2}
10748: {\sc Merker, J.}; {\sc Porten, P.}:
10749: {\em On the local meromorphic extension of CR meromorphic functions}.
10750: Complex analysis and applications (Warsaw,
10751: 1997). Ann. Polon. Math. {\bf 70} (1998), 163--193.
10752: 
10753: \bibitem[MP3]{mp3}
10754: {\sc Merker}, J; {\sc Porten}, E.:
10755: {\em On wedge extendability of CR meromorphic functions}, 
10756: Math. Z. {\bf 241} (2002) 485--512.
10757: 
10758: \bibitem[P]{p}
10759: {\sc Pinchuk}, S.: 
10760: {\em A boundary uniqueness theorem for holomorphic functions of
10761: several complex variables}, 
10762: Mat. Zametki {\bf 15} (1974), 205--212.
10763: 
10764: \bibitem[P1]{p1}
10765: {\sc Porten}, P.:
10766: {\em Analytic extension and removable singularities of the integrable
10767: CR-functions}, 
10768: Preprint, Humboldt-Universit\"at zu Berlin, 2000, no.~3, 18~pp.
10769: 
10770: \bibitem[P2]{p2}
10771: {\sc Porten}, P.:
10772: {\em Totally real discs in nonpseudoconvex boundaries}, 
10773: Ark. Mat. {\bf 41} (2003), no.1, 133--150.
10774: 
10775: \bibitem[P3]{p3}
10776: {\sc Porten}, P.:
10777: {\em Habilitationsschrift} (in preparation).
10778: 
10779: \bibitem[Sa]{sa} 
10780: {\sc Sarkis, F.}:
10781: {\em CR-meromorphic extension and the non embeddability of the
10782: Andreotti-Rossi CR structure in the projective space},
10783: Int. J. Math. {\bf 10} (1999), no.7, 897--915.
10784: 
10785: \bibitem[Sl]{sl}
10786: {\sc Slapar}, M:
10787: {\em On Stein neighborhood basis of real surfaces},
10788: e-print {\tt arXiv:math.CV/0309218}, september 2003.
10789: 
10790: \bibitem[St]{st}
10791: {\sc Stout}, E.L.:
10792: {\em Removable singularities for the boundary values of holomorphic
10793: functions}. Several Complex Variables (Stockholm, 1987/1988),
10794: 600--629, {\it Math. Notes}, {\bf 38}, Princeton Univ. Press,
10795: Princeton, NJ, 1993.
10796: 
10797: \bibitem[Su]{su} 
10798: {\sc Sussmann}, H.J.: 
10799: {\em Orbits of families of vector fields and integrability of 
10800: distributions},
10801: Trans. Amer. Math. Soc. {\bf 180} (1973), 171--188.
10802: 
10803: \bibitem[Tr1]{tr1}
10804: {\sc Tr\'epreau}, J.-M.: 
10805: {\em Sur le prolongement holomorphe des fonctions CR d\'efinies sur une
10806: hypersurface r\'eelle de classe ${\bf C}^{2}$}, Invent. Math.
10807: {\bf 83} (1986), 583--592.
10808: 
10809: \bibitem[Tr2]{tr2}
10810: {\sc Tr\'epreau}, J.-M.:
10811: {\em Sur la propagation des singularit\'es dans les vari\'et\'es CR},
10812: Bull. Soc. Math. Fr. {\bf 118} (1990), no.4, 403--450.
10813: 
10814: \bibitem[Trv]{trv} 
10815: {\sc Treves}, F.: 
10816: {\em Approximation and
10817: representation of functions and distributions annihilated by a system
10818: of complex vector fields},
10819: Palaiseau, \'Ecole
10820: Polytechnique, Centre de Math\'ematiques, 1981.
10821: 
10822: \bibitem[Tu1]{tu1}
10823: {\sc Tumanov}, A.E.: 
10824: {\em Extending CR-functions into a wedge},
10825: Mat. Sbornik {\bf 181} (1990), 951--964. Trad. in English in Math.
10826: {\sc Ussr} Sbornik {\bf 70} (1991), 2, 385--398.
10827: 
10828: \bibitem[Tu2]{tu2}
10829: {\sc Tumanov}, A.E.:
10830: {\em Connections and propagation of analyticity for CR functions},
10831: Duke Math. J. {\bf 73} (1994), no.1, 1--24.
10832: 
10833: \bibitem[Tu3]{tu3}
10834: {\sc Tumanov, A.E.}:
10835: {\em On the propagation of extendibility of CR functions}. Complex
10836: analysis and geometry (Trento, 1993), 479--498, Lecture Notes in Pure
10837: and Appl. Math., {\bf 173},
10838: Dekker, New York, 1996.
10839: 
10840: \end{thebibliography}
10841: 
10842: \vfill
10843: \end{document}
10844: 
10845: \endproof
10846: 
10847: \endproof