1: \documentclass[12pt]{amsart} \usepackage{latexsym, amssymb}
2: \usepackage[T1]{fontenc}
3: \usepackage[latin1]{inputenc}
4: \usepackage{pstricks}
5:
6: % theorem environments
7:
8: \newtheoremstyle{slthm}% name
9: {9pt}% Space above, empty = `usual value'
10: {5pt}% Space below
11: {\slshape}% Body font
12: {}% Indent amount (empty = no indent, \parindent = para indent)
13: {\bfseries}% Thm head font
14: {.}% Punctuation after thm head
15: {.5em}% Space after thm head: " " = normal interword space;
16: % \newline = linebreak
17: {\thmname{#1}\thmnumber{ #2}\thmnote{ (#3)}}% Thm head spec
18:
19: \newtheoremstyle{prcl}% name
20: {9pt}% Space above, empty = `usual value'
21: {5pt}% Space below
22: {\slshape}% Body font
23: {}% Indent amount (empty = no indent, \parindent = para indent)
24: {\bfseries}% Thm head font
25: {.}% Punctuation after thm head
26: {.5em}% Space after thm head: " " = normal interword space;
27: % \newline = linebreak
28: {\thmname{#3}\thmnumber{ #2}}% Thm head spec
29:
30: \newtheoremstyle{prblm}% name
31: {9pt}% Space above, empty = `usual value'
32: {5pt}% Space below
33: {\rm}% Body font
34: {}% Indent amount (empty = no indent, \parindent = para indent)
35: {\bfseries}% Thm head font
36: {.}% Punctuation after thm head
37: {.5em}% Space after thm head: " " = normal interword space;
38: % \newline = linebreak
39: {\thmname{#3}\thmnumber{ #2}}% Thm head spec
40:
41: \theoremstyle{slthm}
42: \newtheorem{thm}{Theorem}[section]
43: \newtheorem{lemma}[thm]{Lemma}
44: \newtheorem{prop}[thm]{Proposition}
45: \newtheorem{cor}[thm]{Corollary}
46: \newtheorem{fact}[thm]{Fact}
47: \newtheorem{facts}[thm]{Facts}
48:
49: \theoremstyle{definition}
50: \newtheorem{exr}[thm]{Exercise}
51: \newtheorem{df}[thm]{Definition}
52: \newtheorem{dfs}[thm]{Definitions}
53: \newtheorem{rmkdf}[thm]{Remark and Definition}
54: \newtheorem{nrmk}[thm]{Remark}
55: \newtheorem{nrmks}[thm]{Remarks}
56: \newtheorem{expl}[thm]{Example}
57: \newtheorem{expls}[thm]{Examples}
58: \newtheorem{pgf}[thm]{}
59: \newtheorem{asmpt}[thm]{Assumption}
60: \newtheorem*{claim}{Claim}
61: \newtheorem*{term}{Terminology}
62: \newtheorem*{notation}{Notation}
63:
64: \theoremstyle{remark}
65: \newtheorem*{rmk}{Remark}
66: \newtheorem*{rmks}{Remarks}
67: \newtheorem*{ntn}{Notation}
68:
69: \theoremstyle{prcl}
70: \newtheorem*{prclaim}{Proclaim}
71: \newtheorem{nprclaim}[thm]{}
72:
73: \theoremstyle{prblm}
74: \newtheorem*{problem}{Problem}
75:
76: % change default numbering for enumerate environment to be in parentheses
77:
78: \renewcommand{\labelenumi}{\rm (\theenumi)}
79:
80: % enumerate environment with roman numbering
81:
82: \newenvironment{renumerate}
83: {\renewcommand{\labelenumi}{\rm (\roman{enumi})}
84: \begin{enumerate}}
85: {\end{enumerate}}
86:
87: %enumerate environment with first number chosen
88: \newcounter{flexnummark}
89: \newenvironment{flexnum}[1]
90: {\begin{list}{(\arabic{flexnummark})}{\usecounter{flexnummark}
91: \setcounter{flexnummark}{#1}
92: \addtocounter{flexnummark}{-1}}} {\end{list}}
93:
94:
95: %% various math operators (for closure, boundary, etc.)
96:
97: \DeclareMathOperator{\length}{length}
98: \DeclareMathOperator{\cl}{cl}
99: \DeclareMathOperator{\fr}{fr}
100: \DeclareMathOperator{\bd}{bd}
101: \DeclareMathOperator{\ir}{int}
102: \DeclareMathOperator{\dom}{dom}
103: \DeclareMathOperator{\codim}{codim}
104: \DeclareMathOperator{\Reg}{Reg}
105: \DeclareMathOperator{\mr}{MR}
106: \DeclareMathOperator{\tp}{tp}
107: \DeclareMathOperator{\id}{id}
108: \DeclareMathOperator{\rk}{rk}
109: \DeclareMathOperator{\fracf}{F}
110: \DeclareMathOperator{\puis}{P}
111: \DeclareMathOperator{\im}{Im}
112: \DeclareMathOperator{\re}{Re}
113: \DeclareMathOperator{\supp}{supp}
114: \DeclareMathOperator{\ord}{ord}
115: \DeclareMathOperator{\sord}{sord}
116: \DeclareMathOperator{\sign}{sgn}
117: \DeclareMathOperator{\ima}{im}
118: \DeclareMathOperator{\Span}{span}
119: \DeclareMathOperator{\bh}{b}
120: \DeclareMathOperator\Un{Un}
121: \DeclareMathOperator\An{An}
122: \DeclareMathOperator\cut{cut}
123: \DeclareMathOperator\val{val}
124: \DeclareMathOperator\Pos{Pos}
125: \DeclareMathOperator\st{st}
126: \DeclareMathOperator\gr{gr}
127: \DeclareMathOperator\bl{b}
128: \DeclareMathOperator\Bl{B}
129: \DeclareMathOperator\blh{bh}
130: \DeclareMathOperator\ex{e}
131: \DeclareMathOperator\tr{t}
132: \DeclareMathOperator\h{h}
133: \DeclareMathOperator\sh{sh}
134: \DeclareMathOperator\rh{rh}
135: \DeclareMathOperator\ph{ph}
136: \DeclareMathOperator\wh{wh}
137: \DeclareMathOperator\mh{mh}
138: \DeclareMathOperator\dih{dih}
139: \DeclareMathOperator\true{\bf T}
140: \DeclareMathOperator\false{\bf F}
141: \DeclareMathOperator\inv{i}
142: \DeclareMathOperator\dist{dist}
143: \DeclareMathOperator\Arg{Arg}
144:
145: \newcommand{\grad}{\nabla}
146: \newcommand{\linf}{\Lambda^{\infty}}
147: \newcommand{\tinf}{\text{\rm T}^{\infty}}
148: \newcommand{\fsigma}{\text{\rm F}_{\sigma}}
149:
150: % my spacing
151:
152: \newcommand{\mdots}{\dots}
153:
154: % logic NOT
155:
156: \newcommand{\negsim}{\sim\!}
157:
158: % \nsubseteq
159:
160: %\DeclareMathSymbol{\ssneq}{\mathalpha}{AMSb}{"20}
161: %\renewcommand{\nsubseteq}{\ \ssneq\ }
162:
163: % the infamous restriction sign!
164:
165: \newcommand{\rest}[1]{|_{#1}}
166:
167: % long arrow for functions: f:A \into B
168:
169: \newcommand{\into}{\longrightarrow}
170:
171: % changes \hat, \tilde and \bar to \widehat, \widetilde and \overline
172: % (looks better!)
173:
174: \renewcommand{\hat}{\widehat}
175: \renewcommand{\tilde}{\widetilde}
176: \renewcommand{\bar}{\overline}
177: \newcommand{\mdot}{\ \!\cdot\ \!}
178:
179: % Onshuus's Independence Defs. (\ind produces the anchor, \nind the
180: % same with a slash through it.)
181:
182: \def\Ind#1#2{#1\setbox0=\hbox{$#1x$}\kern\wd0\hbox to 0pt{\hss$#1\mid$\hss}
183: \lower.9\ht0\hbox to 0pt{\hss$#1\smile$\hss}\kern\wd0}
184: \def\ind{\mathop{\mathpalette\Ind{}}}
185: \def\Notind#1#2{#1\setbox0=\hbox{$#1x$}\kern\wd0\hbox to 0pt{\mathchardef
186: \nn=12854\hss$#1\nn$\kern1.4\wd0\hss}\hbox to
187: 0pt{\hss$#1\mid$\hss}\lower.9\ht0 \hbox to
188: 0pt{\hss$#1\smile$\hss}\kern\wd0}
189: \def\nind{\mathop{\mathpalette\Notind{}}}
190:
191: % norm and set
192:
193: \newcommand{\norm}[1]{\left\|#1\right\|}
194: \newcommand{\set}[1]{\left\{#1\right\}}
195: \newcommand{\lst}[1]{\left(#1\right)}
196:
197: % blackboard letters for naturals, rationals, reals etc.
198:
199: \newcommand{\NN}{\mathbb{N}}
200: \newcommand{\ZZ}{\mathbb{Z}}
201: \newcommand{\PP}{\mathbb{P}}
202: \newcommand{\QQ}{\mathbb{Q}}
203: \newcommand{\RR}{\mathbb{R}}
204: \newcommand{\CC}{\mathbb{C}}
205: \renewcommand{\AA}{\mathbb{A}}
206:
207: % various "curly" letters
208:
209: \newcommand{\curly}[1]{\mathcal{#1}}
210: \newcommand{\A}{\curly{A}}
211: \newcommand{\B}{\curly{B}}
212: \newcommand{\C}{\curly{C}}
213: \newcommand{\D}{\curly{D}}
214: \newcommand{\E}{\curly{E}}
215: \newcommand{\F}{\curly{F}}
216: \newcommand{\G}{\curly{G}}
217: \renewcommand{\H}{\curly{H}}
218: \newcommand{\I}{\curly{I}}
219: \newcommand{\J}{\curly{J}}
220: \newcommand{\K}{\curly{K}}
221: \newcommand{\Le}{\curly{L}}
222: \newcommand{\M}{\curly{M}}
223: \newcommand{\N}{\curly{N}}
224: \renewcommand{\O}{\curly{O}}
225: \renewcommand{\P}{\curly{P}}
226: \newcommand{\Q}{\curly{Q}}
227: \newcommand{\R}{\curly{R}}
228: \renewcommand{\S}{\curly{S}}
229: \newcommand{\T}{\curly{T}}
230: \newcommand{\U}{\curly{U}}
231: \newcommand{\V}{\curly{V}}
232: \newcommand{\W}{\curly{W}}
233: \newcommand{\X}{\curly{X}}
234: \newcommand{\Y}{\curly{Y}}
235: \newcommand{\Z}{\curly{Z}}
236:
237: % various fraktur letters
238:
239: \renewcommand{\a}{\mathbf{a}}
240: \renewcommand{\b}{\mathbf{b}}
241: \newcommand{\f}{\mathbf{f}}
242: \newcommand{\g}{\mathbf{g}}
243: \renewcommand{\k}{\mathbf k}
244: \renewcommand{\l}{\mathbf l}
245: \newcommand{\m}{\mathbf{m}}
246: \newcommand{\p}{\mathbf{p}}
247: \newcommand{\q}{\mathbf{q}}
248: \renewcommand{\r}{\mathbf{r}}
249: \newcommand{\s}{\mathbf{s}}
250: \renewcommand{\t}{\mathbf{t}}
251: \newcommand{\frB}{\mathbf{B}}
252: \newcommand{\frL}{\mathbf{L}}
253:
254:
255: % Borel and Laplace transforms
256:
257: \newcommand{\bo}{\curly{B}}
258: \newcommand{\la}{\curly{L}}
259:
260: % various expansions of the field of reals
261:
262: \newcommand{\Rbar}{\overline{\RR}}
263: \newcommand{\Rtilde}{\,\tilde\RR\,}
264: \newcommand{\pfaff}{\P\big(\Rtilde\big)}
265: \newcommand{\Ibar}{\bar{I}\,}
266: \newcommand{\Ran}{\RR_{\text{an}}}
267: \newcommand{\Ranexp}{\RR_{\text{an,exp}}}
268: \newcommand{\Rf}{\RR_{\F}}
269: \newcommand{\Rg}{\RR_{\G}}
270: \newcommand{\Rq}{\RR_{\Q}}
271: \newcommand{\Lg}{L_{\G}}
272: \newcommand{\Tg}{T_{\G}}
273: \newcommand{\Ranstar}{\RR_{\text{an}^*}}
274: \newcommand{\Ranomega}{\RR_{\text{an}^{\omega}}}
275: \newcommand{\Rb}{\RR_{\curly{B}}}
276: \newcommand{\Ra}{\R\langle a \rangle}
277:
278: % general power series ring
279:
280: \newcommand{\Rgamma}{\RR\left(\left(t^{\Gamma}\right)\right)}
281:
282: % power series with blackboard font coefficients.
283: % these commands have two arguments: the first is for the ring of coefficients,
284: % which will appear in blackboard font, the second is for the indeterminate(s).
285: % Use as follows: \Ps{R}{X} gives the usual power series with real
286: % coefficients,
287: % and \Pc{R}{X,Y} gives the ring of convergent power series with real
288: % coefficients
289: % and indeterminates X and Y.
290:
291: \newcommand{\Ps}[2]{\mathbb{#1}[\![#2]\!]}
292: \newcommand{\Pc}[2]{\mathbb{#1}\{#2\}}
293:
294: % power series with normal font coefficients.
295: % same as above, except that the ring of coefficients will be in normal font.
296: % (Used for example over a general ring A.)
297:
298: \newcommand{\As}[2]{#1[\![#2]\!]}
299: \newcommand{\Ac}[2]{#1\{#2\}}
300:
301: % fractions with \pi in the numerator.
302: % example: \piover{k} gives the fraction \pi/k.
303:
304: \newcommand{\piover}[1]{\frac {\pi}{#1}}
305:
306: % closed sector
307:
308: \newcommand{\Sbar}{\overline{S}}
309:
310: % regular part of cell decomposition
311:
312: \newcommand{\Creg}{\C_{\text{reg}}}
313: \newcommand{\Cregtan}{\C_{\text{reg}}^{\text{tan}}}
314:
315: % page settings
316:
317: %\voffset = -0.8cm
318: %\topmargin = 0in
319: %\evensidemargin = 0in
320: %\oddsidemargin = 0in
321: %\textheight = 9in
322: %\headheight = 12pt
323: %\textwidth = 6.5in
324: %\evensidemargin = 0.5in
325: %\oddsidemargin = 0.5in
326: %\textwidth = 5.5in
327: %\headheight = 12pt
328:
329: % equation numbering
330:
331: \numberwithin{equation}{section}
332:
333: % display breaks
334:
335: \allowdisplaybreaks[2]
336:
337:
338: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%% User specified LaTeX commands.
339: %\usepackage{hyperref}
340: %\usepackage{pstricks}
341:
342: %\usepackage{babel}
343: %\makeatother
344:
345:
346: \begin{document}
347:
348: \title{Transition maps at non-resonant hyperbolic
349: singularities are o-minimal}
350:
351:
352: \author{T. Kaiser, J.-P. Rolin and P. Speissegger}
353:
354: \address{Universit\"at Regensburg \\ NWF I-Mathematik \\ 93040
355: Regensburg, \linebreak Germany}
356: \email{tobias.kaiser@mathematik.uni-regensburg.de}
357:
358: \address{Universit\'e de Bourgogne \\ UFR Sciences et Techniques \\ 9
359: avenue Alain Savary - B.P. 47870 \\ 21078 Dijon Cedex \\ France}
360: \email{Jean-Philippe.Rolin@u-bourgogne.fr}
361:
362: \address{McMaster University \\ Department of Mathematics \&
363: Statistics, \linebreak 1280 Main Street West \\ Hamilton, Ontario L8S
364: 4K1 \\ Canada} \email{speisseg@math.mcmaster.ca}
365:
366: \subjclass {37C27, 37E35, 03C64}
367:
368: \keywords {Vector fields, transition maps, o-minimal structures}
369:
370: \thanks {Supported by DFG, CNRS and NSERC}
371:
372: \date{\today; Preprint.}
373:
374: \begin{abstract}
375: We construct a model complete and o-minimal expansion $\Rq$ of the
376: field of real numbers such that, for any planar analytic vector
377: field $\xi$ and any isolated, non-resonant hyperbolic singularity
378: $p$ of $\xi$, a transition map for $\xi$ at $p$ is definable in
379: $\Rq$. This expansion also defines all convergent generalized power
380: series with natural support and is polynomially bounded.
381: \end{abstract}
382:
383: \maketitle
384: \markboth{T. Kaiser, J.-P. Rolin and P. Speissegger}{Transition maps
385: at non-resonant hyperbolic singularities}
386: %\tableofcontents{}
387:
388: \section*{Introduction}
389:
390: One of the motivations for this paper is the following: let $\xi$ be a
391: (real) analytic vector field on $\RR^2$ such that $\xi^{-1}(0) =
392: \{p\}$ is an isolated singularity of $\xi$. We assume here that the
393: flow of $\xi$ near $p$ is as pictured in Figure 1 below:
394: there are two trajectories of $\xi$ at $p$, one incoming to $p$,
395: called $\gamma^-$, and the other outgoing from $p$, called $\gamma^+$.
396: To describe the flow of $\xi$ near these trajectories, we fix two
397: small segments $\Lambda^-$ and $\Lambda^+$ transverse to $\xi$ and
398: equipped with analytic charts $x$ and $y$ such that $x = 0$ is the
399: intersection point of $\gamma^-$ with $\Lambda^-$ and $x>0$ to the
400: right of $\gamma^-$, and similarly $y=0$ is the intersection point of
401: $\gamma^+$ with $\Lambda^+$ and $y>0$ above $\gamma^+$. Then for all
402: sufficiently small $x>0$, the trajectory of $\xi$ crossing $\Lambda^-$
403: in the point $x$ later crosses $\Lambda^+$ in the point $y = g(x)$ of
404: $\Lambda^+$.
405:
406: \begin{figure}[htbp]
407: \begin{center}
408: \input{transitionmap.pst}
409: \end{center}
410: \caption{\label{cap:Transition-map}Transition map $g$ at the
411: hyperbolic singular point $p$}
412: \end{figure}
413:
414: For any sufficiently small $\epsilon > 0$, the map $g:(0,\epsilon)
415: \into (0,\infty)$ defined in this way is called a \textbf{transition
416: map} of $\xi$ at $p$. The study of transition maps is at the heart
417: of Ilyashenko's solution of Dulac's Problem \cite{ily:dulac}.
418: Somewhat more precisely, Ilyashenko proves that any finite composition
419: of such transition maps has only finitely many isolated fixed points.
420: (Independently, Ecalle \cite{eca:dulac} proves that these maps are
421: \textit{analyzable} and deduces his own proof of Dulac's Problem.)
422:
423: Ilyashenko's analysis of transition maps suggests to us the following
424: \medskip
425:
426: \noindent\textbf{Question:} are the transition maps of $\xi$ near
427: $p$ definable in some \textit{fixed} \hbox{o-minimal} expansion $\R$
428: of the real field? \medskip
429:
430: If the answer to this question is positive, it would follow that some
431: \textit{Poincar\'e return map near any polycycle of $\xi$} (see
432: Section \ref{family} for details) is also definable in $\R$, because
433: the family of functions definable in $\R$ is closed under composition.
434: It would then follow from Dulac's arguments \cite{dul:cycles} that
435: $\xi$ has at most finitely many limit cycles.
436:
437: In this paper, we give a positive answer to the above question under
438: some restrictions on $\xi$. First, we assume that the singularity $p$
439: of $\xi$ is \textbf{hyperbolic}, that is, the linear part of $\xi$ at
440: $p$ has two nonzero real eigenvalues of opposite signs. In this
441: situation, Dulac proves in \cite[chapters 23 and 35]{dul:cycles} that
442: for a transition map $g$ as above, there exist a choice of charts $x$
443: and $y$, a $p_0>0$, real polynomials $p_j$ in one variable for $j = 1,
444: 2, \dots$, and real numbers $0 < \nu_0 < \nu_1 < \cdots$, such that
445: $\lim_j \nu_j = +\infty$ and for every $n \in \NN$ the following
446: asymptotic relation holds:
447: \begin{itemize}
448: \item[(D)] $g(x) - p_0x^{\nu_0} - \sum_{j=1}^n p_j(\log x) x^{\nu_j} =
449: o\left(x^{\nu_n}\right)$ as $x \to 0$.
450: \end{itemize}
451: Moreover, Ilyashenko obtains the following strengthening in Chapter 1
452: of \cite{ily:dulac}: a set $W \subseteq \CC$ is a \textbf{standard
453: quadratic domain} if there are constant $c \in \RR$ and $C > 0$ such
454: that $$W = \set{z \in \CC:\ \re z < c - C \sqrt{|\im z|}}.$$ Then for
455: $g$ as above,
456: \begin{itemize}
457: \item[(I1)] there exists a standard quadratic domain $W \subseteq \CC$
458: such that $g \circ \exp$ extends to a holomorphic mapping $G:W \into
459: \CC$;
460: \item[(I2)] for all $n \in \NN$, we have the asymptotic relation
461: \begin{equation*}
462: G(z) - p_0 e^{\nu_0 z} - \sum_{j=1}^n p_j(z) e^{\nu_j z} =
463: o\left(e^{\nu_n \re z}\right) \text{ as } |z| \to +\infty \text{
464: in } W.
465: \end{equation*}
466: \end{itemize}
467: Slightly abusing notations (to be clarified in Section
468: \ref{asymptotic_section} below), we summarize here conditions (I1) and
469: (I2) by saying that there is a standard quadratic domain $W \subseteq
470: \CC$ such that $g \sim_W p_0 x^{\nu_0} + \sum_{n=1}^\infty p_j(\log x)
471: x^{\nu_n}$. A Phragmen-Lindel\"of argument \cite[p. 23]{ily:dulac}
472: shows that these conditions suffice to conclude that $g$ has at most
473: finitely many isolated fixed points.
474:
475: The main body of Ilyashenko's proof consists in extending this
476: Phrag\-men-Lindel\"of argument to \textit{finite compositions} of
477: transition maps (not just in the hyperbolic case). In contrast, our
478: approach is to try to prove that all transition maps generate an
479: o-minimal expansion of the real field. Since finite compositions of
480: functions definable in an \hbox{o-minimal} structure are again
481: definable in that same structure, it would then follow that all finite
482: compositions of transition maps have finitely many isolated fixed
483: points.
484:
485: In this paper, we carry out our approach under the additional
486: hypotheses of hyperbolicity and
487: \begin{itemize}
488: \item[(NR)] the singularity $p$ is \textbf{non-resonant}, that is, the
489: ratio of the two eigenvalues of the linear part of $\xi$ at $p$ is
490: an irrational number.
491: \end{itemize}
492: It follows from Dulac's argument that under the assumption (NR), the
493: polynomials $p_j$ in the asymptotic series of $g$ above are all
494: constant.
495:
496: Thus, we let $\Ps{R}{X^*}^\omega$ be the set of all formal power
497: series $F = \sum_{\alpha \ge 0} a_\alpha X^\alpha$ such that $a_\alpha
498: \in \RR$ for each $\alpha \ge 0$ and the support
499: \begin{equation*}
500: \supp(F) := \set{\alpha \ge 0:\ a_\alpha \ne 0}
501: \end{equation*}
502: of $F$ is such that $\supp(F) \cap [0,R]$ is finite for every $R>0$.
503: Note that $\Ps{R}{X^*}^\omega$ is a subset of the set $\Ps{R}{X^*}$ of
504: all generalized power series defined by Van den Dries and Speissegger
505: in \cite{vdd-spe:genpower}. (In the latter, o-minimality is
506: established for the expansion of the real field by all
507: \textit{convergent} generalized power series; in contrast, the
508: generalized power series studied here are in general not convergent.)
509: Since we do not use the larger class $\Ps{R}{X^*}$ here, we shall
510: routinely omit the superscript $\omega$.
511:
512: Next, for every $\epsilon > 0$, we let $\Q_\epsilon$ be the set of all
513: functions $f:[0,\epsilon] \into \RR$ for which there exist an $F \in
514: \Ps{R}{X^*}$ and a standard quadratic domain $W \subseteq \CC$ such
515: that $(0,\epsilon] \subseteq \exp(W)$ and $f \sim_W F$. (Thus by
516: definition, for any $f \in \Q_\epsilon$ the only point in
517: $[0,\epsilon]$ where $f$ is not necessarily analytic is $0$.) Our
518: goal here is to prove that the expansion of the real field by all
519: functions in $\Q_1$ is o-minimal; to do so, we follow the method
520: developed in \cite{vdd-spe:genpower}.
521:
522: Roughly speaking, the method in \cite{vdd-spe:genpower} goes as
523: follows: starting from a quasianalytic class $\C$ of functions in
524: several variables, we consider ``mixed'' functions that behave like
525: functions in $\C$ for some of the variables and are analytic in the
526: remaining variables \cite[section 5]{vdd-spe:genpower}. We show that
527: the algebra of these mixed functions possesses certain closure
528: properties \cite[section 6]{vdd-spe:genpower}, most notably closure
529: under blow-up substitutions (which correspond to the charts of certain
530: blowings-up) and under Weierstrass preparation with respect to the
531: analytic variables. These closure properties allow us \cite[section 7
532: and Proposition 8.4]{vdd-spe:genpower} to use resolution of
533: singularities to describe \textit{$\C$-sets}, which are defined by
534: equations and inequalities among functions from $\C$; in particular,
535: we show that $\C$-sets have finitely many connected components. We
536: then adapt in \cite[section 8]{vdd-spe:genpower} Gabrielov's fiber
537: cutting argument to conclude that the complement of a projection of
538: a $\C$-set is again the projection of a $\C$-set. The o-minimality
539: of the expansion of the real field by the functions in $\C$ follows as
540: outlined in \cite[section 2]{vdd-spe:genpower}.
541:
542: This means in particular that we need to define classes $\Q_\rho$,
543: where $\rho \in (0,\infty)^m$ is a polyradius, of functions
544: $f:[0,\rho_1] \times \cdots \times [0,\rho_m] \into \RR$ with
545: analytic-extension and asymptotic properties in several variables
546: corresponding to (I1) and (I2) above. It turns out, however, that the
547: most natural definition of these $\Q_\rho$ is insufficient to obtain
548: all the necessary closure properties, and we refer the reader to
549: Section \ref{division} for the correct definition of these classes.
550:
551: Once the classes $\Q_\rho$ are introduced we define, for each $m \in
552: \NN$ and $m$-variable function $f \in \Q_{1, \dots, 1}$, a total
553: function $\tilde{f}:\RR^m \into \RR$ given by
554: \begin{equation*}
555: \tilde{f}(x) := \begin{cases} f(x) &\text{if } x \in [0,1]^m, \\ 0
556: &\text{otherwise.} \end{cases}
557: \end{equation*}
558: We let $\Rq := \left(\RR, <, 0, 1, +, -, \mdot, \left(\tilde{f}:\ f
559: \in \Q_{1, \dots, 1}, m \in \NN\right)\right)$; note that in
560: particular, every function in $\Q_\epsilon$, for any $\epsilon > 0$,
561: is definable in $\Rq$. Our partial answer to the question above is
562: the following:
563:
564: \begin{prclaim}[Theorem A]
565: The structure $\Rq$ is model complete, o-minimal and admits analytic
566: cell decomposition.
567: \end{prclaim}
568:
569: Moreover, our method of proving Theorem A also gives a kind of
570: ``Puiseux theorem'' for the one-variable functions definable in $\Rq$:
571:
572: \begin{prclaim}[Theorem B]
573: Let $\epsilon > 0$ and $f:(0,\epsilon) \into \RR$ be definable in
574: $\Rq$. Then there are a function $g \in \Q_\delta$ for some $\delta
575: \in (0,\epsilon)$ and an $r \in \RR$ such that $g(0) \ne 0$ and
576: $f(x) = x^r g(x)$ for all $x \in (0,\delta)$.
577: \end{prclaim}
578:
579: Our discussion above of transition maps now implies:
580:
581: \begin{prclaim}[Corollary]
582: Assume that $p$ is a non-resonant hyperbolic singularity of $\xi$,
583: and let $g:(0,\epsilon) \into (0,\infty)$ be a transition map of
584: $\xi$ at $p$, expressed in the charts $x$ and $y$ such that (D)
585: holds. Then for every $\delta \in (0,\epsilon)$, the function
586: $g\rest{(0,\delta)}$ is definable in $\Rq$. \qed
587: \end{prclaim}
588:
589: The transition maps discussed here are not the only functions of
590: interest definable in $\Rq$: in two forthcoming papers, the first
591: author shows that both the Riemann maps and the solutions of
592: Dirichlet's Problem on certain subanalytic domains with non-analytic
593: boundary are definable in $\Rq$.
594:
595: Moreover, we construct in Section \ref{family} of this paper an
596: \textit{analytic unfolding $\xi_\mu$} of $\xi$ in a neighbourhood of a
597: polycycle $\Gamma$ of $\xi$, in the spirit of Roussarie
598: \cite{rou:bifurcations}. The unfolding $\xi_\mu$ is such that each
599: $\xi_\mu$ has the same set of singularities as $\xi$ and each of these
600: singularities is non-resonant hyperbolic. Thus we obtain an analytic
601: family of transition maps at each of these singularities; in fact, we
602: show that each such family of transition maps is definable in $\Rq$.
603: It follows that there is a definable family of functions $P_\mu$ such
604: that for every $\mu$, the function $P_\mu$ is a Poincar\'e return map
605: for $\xi_\mu$ near $\Gamma$. The uniform finiteness property of
606: o-minimal structures \cite[chapter 3, section 3]{vdd:vol1} then
607: implies that there is a uniform bound on the number of fixed points of
608: these functions $P_\mu$. The question whether such uniform bounds
609: exist is related to Hilbert's 16th problem and remains open for
610: general $\xi$ and $\Gamma$. (See \cite{ily:history} for a survey on
611: Hilbert's 16th problem and \cite{rou:bifurcations} for the
612: relationship between Hilbert's 16th problem and analytic unfoldings.)
613:
614: Except for Section \ref{family}, the content of this paper is entirely
615: focussed on the construction of the o-minimal structure $\Rq$. There
616: are various related questions that we do not know how to answer at
617: this point, such as:
618: \begin{enumerate}
619: \item Is $\Rg$, the o-minimal structure generated by all functions
620: that are multisummable in the positive real direction
621: \cite{vdd-spe:multisummable}, a reduct of $\Rq$? Or is $\Rq$ a
622: reduct of the Pfaffian closure of $\Rg$ (see \cite{spe:pfaffian})?
623: \item Are transition maps near \textit{resonant} hyperbolic
624: singularities of $\xi$ definable in the Pfaffian closure of $\Rq$?
625: \end{enumerate}
626:
627: This paper is organized as follows: in Section \ref{natural}, we
628: define the class of generalized power series with natural support, and
629: we establish some truncation properties needed later. Our first and
630: most straightforward attempt at defining the classes $\Q_\rho$ is made
631: in Section \ref{asymptotic_section}, where we also establish some
632: useful criteria for functions to belong to these classes. The reasons
633: why this first attempt is insufficient are explained in Section
634: \ref{tdte}, which lead us to a correct definition in Section
635: \ref{division}, where we also establish the first closure properties
636: needed to apply the method of \cite{vdd-spe:genpower}. In fact, in
637: order to apply this method, we need to introduce corresponding
638: ``mixed'' functions that are, roughly speaking, of asymptotic type
639: (I1) and (I2) in some variables and analytic in the other variables.
640: One of the reasons for studying these mixed classes is discussed
641: in Section \ref{weierstrass}, where we show that they admit
642: Weierstrass preparation with respect to the analytic variables. To
643: reduce the study of sets defined by functions in the $\Q_\rho$ to that
644: of sets defined by functions in the mixed classes, we use the
645: blow-up substitutions introduced in Section \ref{blowups}. Having
646: established all the properties necessary to apply the method of
647: \cite{vdd-spe:genpower}, we obtain Theorems A and B in Section
648: \ref{o-minimal}.
649:
650:
651: \section{Generalized power series with natural
652: support} \label{natural}
653:
654: Let $m \in \NN$, and let $X = (X_1, \dots, X_m)$ be a tuple of
655: indeterminates. For $\alpha= (\alpha_1, \dots, \alpha_m) \in
656: [0,\infty)^m$, we write $X^\alpha := X_1^{\alpha_1} \cdots
657: X_m^{\alpha_m}$, and we let $X^*$ be the multiplicative monoid
658: consisting of all such $X^\alpha$, multiplied according to $X^\alpha
659: \cdot X^\beta = X^{\alpha+\beta}$. The identity element of $X^*$ is
660: $X^0 = 1$, where $0 = (0, \dots, 0)$.
661:
662: Let $A$ be a commutative ring with $1 \ne 0$. We let $\As{A}{X^*}$
663: denote the set of all formal power series $f(X) = \sum_{\alpha \ge 0}
664: a_\alpha X^\alpha$ such that $a_\alpha \in A$ for each $\alpha \ge 0$
665: and the \textbf{support} $\supp(f) := \set{\alpha \ge 0:\ a_\alpha \ne
666: 0}$ of $f$ is well-ordered, as defined in Section 4.4 of
667: \cite{vdd-spe:genpower}. The elements of $\As{A}{X^*}$ are called
668: \textbf{generalized power series}.
669:
670: \begin{df}
671: \label{df:natural}
672: A set $S \subseteq [0,\infty)^m$ is \textbf{natural} if
673: $\Pi_{X_i}(S) \cap [0,R]$ is finite for every $R>0$ and each $i=1,
674: \dots, m$, where $\Pi_{X_i}:\RR^m \into \RR$ is the projection on
675: the coordinate $X_i$.
676:
677: We denote by $\As{A}{X^*}^\omega$ the set of all generalized power
678: series in $X$ whose support is a natural subset of $[0,\infty)^m$.
679: \end{df}
680:
681: \subsection*{Convention}
682: All results established in Section 4 of \cite{vdd-spe:genpower} go
683: through literally with $\As{A}{X^*}^\omega$ in place of $\As{A}{X^*}$,
684: as already pointed out in the second concluding remark of that paper.
685: To simplify notations, we shall from now on omit the superscript
686: $\omega$. Thus, throughout this paper, every series in $\As{A}{X^*}$
687: is assumed to have natural support, and all results in
688: \cite{vdd-spe:genpower} referenced here are interpreted in this
689: context. \medskip
690:
691: Let $Y = (Y_1, \dots, Y_n)$ be another tuple of indeterminates. For
692: $a,b \in \RR^k$, we write $a\le b$ (resp. $a < b$) if and only if $a_i
693: \le b_i$ (resp. $a_i < b_i$) for $i=1, \dots, k$.
694:
695: \begin{df}
696: \label{truncation_division}
697: Let $S \subseteq [0,\infty)^{m} \times \NN^n$ and $F = \sum
698: a_{(\alpha,\beta)} X^{\alpha} Y^\beta \in \As{A}{X^*,Y}$. We define
699: $\inf S = (a,b) = (a_1, \dots, a_m, b_1, \dots, b_n)$, where $a_i
700: := \inf(\Pi_{X_i}(S))$ for $i=1, \dots, m$ and $b_i :=
701: \min(\Pi_{Y_i}(S))$ for $i=1, \dots, n$, and we put
702: \begin{equation*}
703: F_{S} := \sum_{(\alpha,\beta) \in S}
704: a_{(\alpha,\beta)} X^{\alpha-a} Y^{\beta-b},
705: \quad\text{an element of } \As{A}{X^*,Y}.
706: \end{equation*}
707: For $\gamma \in [0,\infty)^m \times \NN^n$, we write $F_\gamma$ in
708: place of $F_{\{\alpha:\ \alpha \ge \gamma\}}$.
709: \end{df}
710:
711: \begin{rmk}
712: % \label{truncation_properties}
713: Let $F,G \in \As{A}{X^*,Y}$ and $\gamma \in [0,\infty)^m \times
714: \NN^n$. Then $(F+G)_\gamma = F_\gamma + G_\gamma$.
715: \end{rmk}
716:
717: In the remainder of this section, we study how the operation $F
718: \mapsto F_\gamma$ behaves with respect to various other operations on
719: natural power series.
720:
721: \subsection*{Differentiation}
722:
723: Let $F = \sum a_{(\alpha,\beta)} X^\alpha Y^\beta \in \As{A}{X^*,Y}$.
724: For $\beta \in \NN^n$ and $j=1, \dots, n$, we put $\beta^j :=
725: (\beta_1, \dots, \beta_{j-1}, \beta_j-1, \beta_{j+1}, \dots,
726: \beta_n)$. We define
727: \begin{equation*}
728: \partial_i F := \sum \alpha_i \cdot a_{(\alpha,\beta)} X^\alpha
729: Y^\beta \quad\text{for } i=1, \dots, m,
730: \end{equation*}
731: and
732: \begin{equation*}
733: \frac{\partial F}{\partial Y_j} := \sum \beta_j \cdot
734: a_{(\alpha,\beta)} X^{\alpha} Y^{\beta^j}
735: \quad\text{for } j = 1, \dots, n.
736: \end{equation*}
737: Note that each $\partial_i F$ and each $\partial F/\partial Y_j$
738: belongs to $\As{A}{X^*,Y}$. Since each $\partial_i$ is a derivation
739: on $\As{A}{X^*,Y}$, we obtain:
740:
741: \begin{lemma}
742: \label{derivation_truncation}
743: Let $F \in \As{A}{X^*,Y}$ and $\gamma \in [0,\infty)^m$. Then:
744: \begin{enumerate}
745: \item $(\partial F/\partial Y_j)_{(\gamma,0)}
746: = \partial (F_{(\gamma,0)})/\partial Y_j$ for each $j = 1, \dots, n$.
747: % \item If $i \in K$ and $\gamma_i > 0$, then $X^{\gamma^i}
748: % (\partial_i F)_{\gamma^i} = \partial_i\left(X^\gamma \cdot
749: % F_\gamma\right)$.
750: \item $(\partial_i F)_{(\gamma,0)} = \gamma_i \cdot F_{(\gamma,0)} +
751: \partial_i(F_{(\gamma,0)})$ for each $i=1, \dots, m$; in
752: particular, if $\gamma_i = 0$, then $(\partial_i F)_{(\gamma,0)}
753: = \partial_i(F_{(\gamma,0)})$.
754: \end{enumerate}
755: \end{lemma}
756:
757: \begin{proof}
758: Parts (1) straightforward. Since $X^\gamma \cdot G_\gamma$ is just
759: the truncation of $G$ at $\gamma$ for any $G \in \As{A}{X^*}$, we have
760: $X^\gamma \cdot (\partial_i F)_{(\gamma,0)} = \partial_i \left(X^\gamma
761: \cdot F_{(\gamma,0)}\right)$. Part (2) follows from the latter, because
762: $\partial_i$ is a derivation.
763: \end{proof}
764:
765: Before continuing with the behavior of the operation $F \mapsto
766: F_\gamma$, we make some crucial observations.
767:
768: \subsection*{Representation}
769: Let $I \subset \{1,\dots,m\}$. Below we write $X_I := (X_{i})_{i\in
770: I}$, and if $x \in \RR^{m}$ we write $x_I:=(x_{i})_{i\in I}$. We
771: let $\Pi_I:\RR^{m+n} \into \RR^{|I|}$ be the projection defined by
772: $\Pi_I(x,y) := x_I$. We write $\bar{I} := \{1, \dots, m\} \setminus
773: I$. For $F \in \As{A}{X^*,Y}$ and $\gamma \in [0,\infty)^m$, we put
774: $B_{\gamma,\emptyset} := \{0\}$, and for nonempty $I \subset
775: \{1,\dots,m\}$, we put $$B_{\gamma,I}=B_{\gamma,I}(F) :=
776: \set{\alpha\in \Pi_I(\supp F):\ \alpha < \gamma_I}.$$
777:
778: \begin{lemma}
779: \label{lemma:representation}
780: Let $F \in \As{A}{X^*,Y}$ and $\gamma \in [0,\infty)^m$. Then for
781: each $I \subseteq \{1, \dots, m\}$, the set $B_{\gamma,I}$ is
782: finite, and for each $\alpha \in B_{\gamma,I}$, there is a unique
783: $F_{\gamma, I, \alpha} \in \As{A}{X_{\bar I}^*,Y}$ such that
784: \begin{equation}
785: \label{eq:representation}
786: F(X,Y) = \sum_{I \subset \{1,\dots,m\}} X_{\bar I}^{\gamma_{\bar I}}
787: \left(\sum_{\alpha \in B_{\gamma,I}} X_I^{\alpha}
788: F_{\gamma,I,\alpha}(X_{\bar I},Y)\right).
789: \end{equation}
790: \end{lemma}
791:
792: \begin{proof}
793: For each $(\alpha,\beta) \in \supp F$, there is a unique $I
794: \subseteq \{1, \dots, m\}$ such that for all $i=1, \dots, m$, we
795: have $\alpha_i < \gamma_i$ iff $i \in I$. Moreover, $\alpha_I \in
796: B_{\gamma,I}$ and $X^\alpha / \left(X_{\bar I}^{\gamma_{\bar I}}
797: X_I^{\alpha_I}\right)$ is a monomial in $X_{\bar I}$.
798: \end{proof}
799:
800: \begin{df}
801: \label{df:representation}
802: Let $F \in \As{A}{X^*,Y}$ and $\gamma \in [0,\infty)^m$. We call
803: the right-hand side of equation \eqref{eq:representation} the
804: \textbf{$\gamma$-representation} of $F$. For each $I \subseteq \{1,
805: \dots, m\}$, we put
806: \begin{equation*}
807: F_{\gamma,I}(X,Y) := \sum_{\alpha \in B_{\gamma,I}} X_I^{\alpha}
808: F_{\gamma,I,\alpha}(X_{\bar I},Y),
809: \end{equation*}
810: so that $F(X,Y) = \sum_{I \subseteq \{1, \dots, m\}} X_{\bar
811: I}^{\gamma_{\bar I}} F_{\gamma,I}(X,Y)$.
812: \end{df}
813:
814: \begin{nrmk}
815: \label{rmk:elementary}
816: In the situation of Definition \ref{df:representation}, we have
817: $F_{(\gamma,0)} = F_{\gamma,\emptyset,0}$. More generally, for each
818: $\gamma \in [0,\infty)^m$, each $I \subseteq \{1, \dots, m\}$ and
819: each $\alpha \in B_{\gamma,I}$, the series
820: $F_{\gamma,I,\alpha}(X_{\bar I},Y)$, considered as an element of
821: $\As{A}{X^*,Y}$, is equal to $F_S$ for some $S \subseteq
822: [0,\infty)^m \times \NN^n$ of the following form:
823: \end{nrmk}
824:
825: \begin{df}
826: \label{elementary}
827: Recall that a box in $\RR^k$ is a set of the form $B = \set{\alpha
828: \in \RR^k:\ a_i \ast_{i,1} \alpha_i \ast_{i,2} b_i \text{ for each
829: } i}$, where $a,b \in (\RR \cup \{\infty\})^k$ and $\ast_{i,1},
830: \ast_{i,2} \in \{<,\le\}$. A set $E \subseteq \RR^k$ is
831: \textbf{elementary} if $E$ is a finite Boolean combination of boxes.
832: \end{df}
833:
834: \begin{nrmk}
835: \label{boolean}
836: If $B \subseteq [0,\infty)^k$ is a box, then $[0,\infty)^k \setminus
837: B$ is a finite union of pairwise disjoint boxes; also, any finite
838: intersection of boxes is a box. Therefore, every elementary set is
839: a finite union of pairwise disjoint boxes.
840: \end{nrmk}
841:
842: \begin{expl}
843: \label{elementary_example}
844: \begin{enumerate}
845: \item For each $i \in \{1, \dots, k\}$ and every $a>0$, the set
846: $\{\alpha \in [0,\infty)^k:\alpha_i = a\}$ is elementary.
847: \item For all $a_1, \dots, a_k,b \ge 0$ and every natural set $S
848: \subseteq [0,\infty)^k$, there is an elementary set $E \subseteq
849: \RR^k$ such that $$S \cap \set{\alpha \in
850: [0,\infty)^k:\sum_{i=1}^k a_i \cdot \alpha_i \ge b} = S \cap
851: E.$$
852: \end{enumerate}
853: \end{expl}
854:
855: Intersections of natural sets and boxes can be further
856: simplified:
857:
858: \begin{lemma}
859: \label{natural_elementary}
860: Let $B \subseteq [0,\infty)^k$ be a box and $S \subseteq
861: [0,\infty)^k$ be a natural set. Then there exist $\gamma, \delta^1,
862: \dots, \delta^k \in [0,\infty)^k$ such that $\gamma \le \delta^j$
863: for each $j$ and $S \cap B = S \cap \{\alpha \ge \gamma\} \setminus
864: \bigcup_{j=1}^k S \cap \{\alpha \ge \delta^j\}$.
865: \end{lemma}
866:
867: \begin{proof}
868: Say $B = \set{\alpha \in \RR^k:\ a_i \ast_{i,1} \alpha_i \ast_{i,2}
869: b_i \text{ for each } i}$, where $a,b \in (\RR \cup \{\infty\})^k$
870: and $\ast_{i,1}, \ast_{i,2} \in \{<,\le\}$. For each $i \in \{1,
871: \dots, k\}$, we define $\gamma_i := \min\set{r \in \Pi_{X_i}(S): a_i
872: \ast_{i,1} r}$ and $\delta_i := \max\set{r \in \Pi_{X_i}(S): r
873: \ast_{i,2} b_i}$. Then $S \cap B = S \cap \set{\gamma \le \alpha
874: \le \delta}$, so the lemma follows with $\delta^j$ defined by
875: $\delta^j_i := \gamma_i$ if $i \ne j$ and $\delta^j_j :=
876: \min\{r \in \Pi_{X_j}(S):\ r > \delta_j\}$.
877: \end{proof}
878:
879: We now return to the study of the operation $F \mapsto F_\gamma$.
880: Since the observations below are rather technical and not clearly
881: motivated at this point, the reader may want to skip the rest of this
882: section and come back to it later as needed (while reading Section
883: \ref{division}, say).
884:
885: \subsection*{Blow-up substitutions}
886:
887: Let $i,j \in \{1, \dots, m\}$ be such that $i \ne j$, and let $\rho >
888: 0$ and $\lambda \ge 0$. Using the binomial expansion
889: \begin{equation*}
890: (\lambda + X_j)^\beta := \sum_{p \in \NN} \begin{pmatrix} \beta \\
891: p \end{pmatrix} \lambda^{\beta-p} X_j^p \quad\text{if } \lambda >
892: 0 \text{ and } \beta \ge 0,
893: \end{equation*}
894: we let $\frB^{\rho,\lambda}_{ij}:\As{A}{X^*} \into \As{A}{X^*}$ be the
895: unique $A$-algebra homomorphism satisfying
896: \begin{equation*}
897: \frB^{\rho,\lambda}_{ij}(X_k) =
898: \begin{cases}
899: X_k &\text{if } k \ne i, \\ X_j^\rho (\lambda + X_i) &\text{if } k
900: = i.
901: \end{cases}
902: \end{equation*}
903: The homomorphism $\frB^{\rho,\lambda}_{ij}$ is called a \textbf{blow-up
904: substitution}; we call $\frB^{\rho,0}_{ij}$ \textbf{singular}, and if
905: $\lambda > 0$, we call $\frB^{\rho,\lambda}_{ij}$ \textbf{regular}. We
906: shall often write $\frB^{\rho,\lambda}_{ij} F$ in place of
907: $\frB^{\rho,\lambda}_{ij}(F)$, for $F \in \As{A}{X^*}$.
908:
909: \begin{nrmk}
910: \label{singular_expl}
911: The substitution $\frB^{\rho,0}_{ij}$ is the $A$-algebra
912: homomorphism $s^\rho_{ij}$ defined in Section 4.13 of
913: \cite{vdd-spe:genpower}.
914: \end{nrmk}
915:
916: \begin{prop}
917: \label{blowup_truncation}
918: Let $F \in \As{A}{X^*}$ and put $X' := (X_1, \dots, X_{m-1})$.
919: \begin{enumerate}
920: \item For $\lambda > 0$, we have $\frB^{\rho,\lambda}_{m,m-1} F \in
921: \As{A}{(X')^*,X_m}$.
922: \item Let $\gamma \in [0,\infty)^m$, and assume that $\gamma_m = 0$
923: if $\lambda > 0$. Then there are $\alpha_1, \dots, \alpha_k \in
924: [0,\infty)^m$ and elementary sets $E_1, \dots, E_k \subseteq
925: [0,\infty)^m$ such that
926: \begin{equation*}
927: \left(\frB^{\rho,\lambda}_{m,m-1} F\right)_\gamma = \sum_{l=1}^k
928: X^{\alpha_l} \cdot \frB^{\rho,\lambda}_{m,m-1} F_{E_l}.
929: \end{equation*}
930: \end{enumerate}
931: \end{prop}
932:
933: \begin{proof}
934: Part (1) follows from the binomial expansion. Next, for $k=1,
935: \dots, m$, we put
936: \begin{equation*}
937: \delta_k :=
938: \begin{cases}
939: \gamma_k &\text{if } k \ne m, \\ \max\set{\gamma_m,
940: \frac{\gamma_{m-1}}{\rho}} &\text{if } k=m,
941: \end{cases}
942: \end{equation*}
943: and $\delta := (\delta_1, \dots, \delta_m)$. Let
944: \begin{equation*}
945: F(X) = \sum_{I \subset \{1, \dots, m\}} X_{\bar I}^{\delta_{\bar
946: I}} \left( \sum_{\alpha \in B_{\delta,I}} X_I^\alpha
947: F_{\delta,I,\alpha}(X_{\bar I}) \right)
948: \end{equation*}
949: be the $\delta$-representation of $F$. Put $B'_{\delta,I} :=
950: B_{\delta,I}$ if $I = \emptyset$ or $I = \{m-1\}$,
951: $B'_{\delta,\{m\}} := \set{\alpha \in B_{\delta,I}:\ \alpha_m \ge
952: \gamma_m}$ and $$B'_{\delta,\{m-1,m\}} := \set{\alpha \in
953: B_{\delta,I}:\ \alpha_m \ge \gamma_m,\,\alpha_{m-1} + \rho
954: \alpha_m \ge \gamma_{m-1}}.$$ Then by the hypothesis on $\gamma$,
955: \begin{multline*}
956: X^\gamma \cdot \left(\frB^{\rho,\lambda}_{m,m-1}F(X)\right)_\gamma
957: \\ = \sum_{I \subseteq \{m-1,m\}} \sum_{\alpha \in B'_{\delta,I}}
958: \frB^{\rho,\lambda}_{m,m-1} \left(X_{\bar I}^{\delta_{\bar I}}
959: X_I^\alpha\right) \cdot \frB^{\rho,\lambda}_{m,m-1}
960: (F_{\delta,I,\alpha}(X_{\bar I})).
961: \end{multline*}
962: Since each term $\frB^{\rho,\lambda}_{m,m-1} \left(X_{\bar
963: I}^{\delta_{\bar I}} X_I^\alpha\right)$ on the right-hand side
964: is divisible by $X^\gamma$, part (2) follows.
965: \end{proof}
966:
967: \subsection*{Composition}
968:
969: Let $F \in \As{A}{X^*,Y}$. For the next lemma, we also let $\q =
970: (\q_1, \dots, \q_n) \in \NN^n$ and put $\k := |\q|$. We let $Z =
971: (Z_1, \dots, Z_\k)$ be a tuple of indeterminates and define
972: \begin{equation*}
973: F_{\q}(X,Z) := F\left(X, Z_1 + \cdots +
974: Z_{\q_1}, \dots,
975: Z_{\q_1 + \cdots + \q_{n-1}+1} + \cdots + Z_\k\right).
976: \end{equation*}
977: Note that $F_\q \in \As{A}{X^*,Z}$.
978:
979: \begin{prop}
980: \label{comp_representation}
981: Let $G = (G_1, \dots, G_n) \in (\As{A}{X^*,Y})^n$ be such that $G(0)
982: = 0$. Then the series $F(X,G(X,Y))$ belongs to $\As{A}{X^*,Y}$, and
983: for each $\gamma \in [0,\infty)^m$, there are
984: \begin{renumerate}
985: \item a $p \in \NN$ and a tuple $\q \in \NN^n$ and, with
986: $\k:= |\q|$,
987: \item elementary $E_1, \dots, E_p \subseteq
988: [0,\infty)^m \times \NN^\k$ and $B_{i,j} \subseteq [0,\infty)^{m}
989: \times \NN^n$ for each pair $(i,j)$ satisfying $i \in \{1, \dots,
990: n\}$ and $j \in \{1, \dots, \q_i\}$,
991: \end{renumerate}
992: such that, with $G_B := ((G_1)_{B_{1,1}}, \dots,
993: (G_1)_{B_{1,\q_1}}, (G_2)_{B_{2,1}}, \dots,
994: (G_n)_{B_{n,\q_n}})$, we have $G_B(0) = 0$, each term
995: $(X,G_B(X,Y))^{\inf E_j}$ is divisible by $X^\gamma$ in
996: $\Ps{A}{X^*,Y}$ and $$F(X,G(X,Y))_{(\gamma,0)} = \sum_{j=1}^p \frac
997: {(X,G_B(X,Y))^{\inf E_j}}{X^\gamma} \cdot
998: (F_\q)_{E_j}(X,G_B(X,Y)).$$
999: \end{prop}
1000:
1001: \begin{proof}
1002: It is standard to check that $F(Y,G(X,Y)) \in \As{A}{X^*,Y}$; we
1003: leave the details to the reader. Let $\gamma \in [0,\infty)^m$, and
1004: for each $j \in \{1, \dots, n\}$, we let
1005: \begin{equation*}
1006: G_j = \sum_{I \subseteq
1007: \{1,\dots,m\}} X_{\bar
1008: I}^{\gamma_{\bar I}} \left(\sum_{\alpha \in
1009: B_{I}(G_j)} X_I^\alpha \cdot
1010: (G_j)_{I,\alpha}(X_{\bar I},Y)\right)
1011: \end{equation*}
1012: be the $\gamma$-representation of $G_j$. (We omit the subscript
1013: $\gamma$ in these notations for the duration of this proof.) We let
1014: $\J$ be the set of all triples $(j,I,\alpha)$ such that $j \in \{1,
1015: \dots, n\}$, $I \subseteq \{1, \dots, m\}$ and $\alpha \in
1016: B_{I}(G_j)$, and we put $\kappa:= |\J|$ and fix a bijection
1017: $\sigma:\{1, \dots, \kappa\} \into \J$. Below, we write
1018: $\sigma(\lambda) = (j(\lambda),I(\lambda),\alpha(\lambda))$ for
1019: $\lambda = 1, \dots, \kappa$; without loss of generality, we may
1020: assume that $j(\lambda) \le j(\lambda')$ whenever $\lambda \le
1021: \lambda'$.
1022:
1023: For each $\lambda = 1, \dots, \kappa$, we let $Z_\lambda$
1024: be a new indeterminate and put
1025: \begin{equation*}
1026: G_\lambda(X,Y) := X_{\bar
1027: I}^{\gamma_{\bar I}} X_I^\alpha \cdot
1028: (G_j)_{I,\alpha}(X_{\bar I},Y)
1029: \end{equation*}
1030: with $(j,I,\alpha) = (j(\lambda),I(\lambda),\alpha(\lambda))$. We
1031: write $G_B := (G_1, \dots, G_\kappa)$; note that $G_B(0) = 0$. We
1032: also put
1033: \begin{equation*}
1034: H(X,Z) := F\left(X, \sum_{j(\lambda)=1} Z_\lambda\,,
1035: \dots, \sum_{j(\lambda)=n} Z_\lambda\right),
1036: \end{equation*}
1037: and we write $H(X,Z) = \sum c_{\mu,\nu} X^\mu Z^\nu$, where $\mu$
1038: ranges over $[0,\infty)^m$ and $\nu$ ranges over $\NN^\kappa$. Note
1039: that $F(X,G(X,Y)) = H(X,G_B(X,Y))$ and $H = F_{\q}$ for some
1040: $\q \in \NN^n$ with $|\q| = \kappa$.
1041:
1042: For $Q \subseteq \{1, \dots, \kappa\}$, we put $I_Q :=
1043: \bigcap_{\lambda \in Q} I(\lambda)$. Let $\mu \in [0,\infty)^{m}$
1044: and $\nu \in \NN^\kappa$, and put $Q(\nu) := \set{\lambda:\
1045: \nu_\lambda \ne 0}$. Then
1046: \begin{multline}
1047: \label{term_truncation}
1048: X^\gamma \cdot
1049: \left(X^\mu G_B(X,Y)^\nu\right)_{(\gamma,0)} = \\
1050: \begin{cases}
1051: X^\mu G_B(X,Y)^\nu &\text{if } \mu_i + \sum_{\lambda=1}^\kappa
1052: \nu_\lambda\cdot
1053: \alpha(\lambda)_i \ge \gamma_i \text{ for each } i \in I_{Q(\nu)}, \\
1054: 0 &\text{otherwise.}
1055: \end{cases}
1056: \end{multline}
1057: Therefore, for each $Q \subseteq \{1, \dots, \kappa\}$, we let $S_Q
1058: \subseteq [0,\infty)^{m} \times \NN^k$ be the set defined by
1059: \begin{multline*}
1060: S_Q := \Big\{(\mu,\nu) \in \supp(H):\ \nu_\lambda = 0 \text{ iff }
1061: \lambda \notin Q,\\ \quad \mu_i + \sum_{\lambda \in Q}
1062: \nu_\lambda\cdot \alpha(\lambda)_i \ge \gamma_i \text{ for each }
1063: i \in I_Q \Big\}.
1064: \end{multline*}
1065: $S_Q$ is in turn the disjoint union of the following sets: for each
1066: map $\eta:I_Q \into Q$, we define
1067: \begin{multline*}
1068: S_{Q, \eta} := \Big\{ (\mu,\nu) \in S_Q:\ \mu_i +
1069: \sum_{\lambda \in Q,\,\lambda < \eta(i)} \nu_\lambda \cdot \alpha(\lambda)_i
1070: < \gamma_i \text{ and } \\ \mu_i + \sum_{\lambda \in Q,\,\lambda \le
1071: \eta(i)} \nu_\lambda \cdot \alpha(\lambda)_i \ge \gamma_i \text{ for
1072: each } i \in I_Q\Big\}.
1073: \end{multline*}
1074: Then $S_Q = S_{Q,0} \cup \bigcup_{\eta} S_{Q,\eta}$, where $S_{Q,0}
1075: := \set{(\mu,\nu) \in S_Q:\ \mu_{I_Q} \ge \gamma_{I_Q}}$. Moreover
1076: for each $\eta:I_Q \into Q$, we write $\max \eta := \max_{i \in I_Q}
1077: \eta(i)$ and $J_\eta := \{\lambda < \max\eta:\ \alpha(\lambda)_i >
1078: 0$ for some $i \in I_Q\}$. Then the set $C_{Q,\eta} :=
1079: \set{(\mu_{I_Q},\nu_{J_\eta}):\ (\mu,\nu) \in S_{Q,\eta}}$ is finite,
1080: and $S_{Q,\eta}$ is the disjoint union of the sets $S_{Q,\eta} \cap
1081: E_{Q,\eta,\beta}$ as $\beta$ ranges over $C_{Q,\eta}$, where
1082: \begin{multline*}
1083: E_{Q,\eta,\beta} := \Big\{ (\mu,\nu):\
1084: (\mu_{I_Q},\nu_{J_\eta}) = \beta, \\
1085: \nu_{\max\eta} \cdot \alpha(\max\eta)_i \ge \gamma_i - \mu_i -
1086: \sum_{\lambda \in Q,\,\lambda < \max\eta} \nu_\lambda \cdot
1087: \alpha(\lambda)_i \text{ for each } i \in I_Q\Big\}.
1088: \end{multline*}
1089:
1090: By Example \ref{elementary_example}, each set $E_{Q,\eta,\beta}$ is
1091: elementary. It follows from \eqref{term_truncation} above that
1092: \begin{multline*}
1093: X^\gamma \cdot F(X,G(X,Y))_{(\gamma,0)} = \\
1094: \sum_{Q,\eta,\beta} (X,G_B(X,Y))^{\inf E_{Q,\eta,\beta}} \cdot
1095: H_{E_{Q,\eta,\beta}}(X,G_B(X,Y)),
1096: \end{multline*}
1097: and it follows from the definition of each $E_{Q,\eta,\beta}$ that
1098: $X^\gamma$ divides the factor $(X,G_B(X,Y))^{\inf E_{Q,\eta,\beta}}$
1099: in $\Ps{A}{X^*,Y}$, as required.
1100: \end{proof}
1101:
1102:
1103:
1104: \section{Natural asymptotic expansions} \label{asymptotic_section}
1105:
1106: Throughout this paper, we denote by $\|z\|$ the Euclidean norm of $z
1107: \in \CC^n$, and we put $|z|:= z_1 + \cdots + z_n$ for such $z$.
1108: We let $\frL = \set{(r,\varphi):\ r>0,\, \varphi \in \RR}$ be the
1109: Riemann surface of the logarithm. We fix an arbitrary
1110: $m\in\mathbb{N}$ and write $$x = (x_1, \dots, x_m) = ((r_1,\varphi_1),
1111: \dots, (r_m, \varphi_m))$$ for elements of $\frL^{m}$. For such $x$,
1112: we put $\|x\| := \|(r_1, \dots, r_m)\|$, $\arg(x) := (\varphi_1,
1113: \dots, \varphi_m)$ and $$\log_m x := (\log r_1+i\varphi_1, \dots, \log
1114: r_m + i\varphi_m) \in \CC^m;$$ we omit the subscript $m$ whenever it
1115: is clear from context. Note that $\log:\frL^m \into \CC^m$ is an
1116: analytic isomorphism. Below, we let $z = (z_1, \dots, z_m)$ range
1117: over $\CC^m$.
1118:
1119: Recall that for an open set $U \subseteq \frL$, a function $f:U \into
1120: \CC$ is holomorphic if the function $f\circ\log^{-1}: \log(U) \into
1121: \CC$ is holomorphic. The set of holomorphic functions on $U$ is
1122: denoted by $\O(U)$.
1123:
1124: \begin{df}
1125: \label{df:holomorphic}
1126: Let $U \subseteq \frL^m$ be open. For $f \in \O(U)$ and $i \in \{1,
1127: \dots, m\}$, we define $\partial_i f:U \into \CC$ by $$\partial_i
1128: f(x) := \frac{\partial (f \circ \log^{-1})}{\partial z_i} (\log
1129: x).$$
1130:
1131: Note that $\partial_i f \in \O(U)$.
1132: \end{df}
1133:
1134: \begin{expl}
1135: \label{expl:powers}
1136: Let $\alpha = (\alpha_{1},\dots,\alpha_{m}) \in \CC^{m}$. We put
1137: $x^\alpha := x_{1}^{\alpha_{1}} \cdots x_{m}^{\alpha_{m}}$, where
1138: $x_i^{\alpha_i} := \exp(\alpha_i \log(x_i))$ for each $i$. The
1139: function $(\cdot)^{\alpha} : \frL^{m} \longrightarrow \CC$ is
1140: holomorphic and $\partial_i(x^\alpha) = \alpha_i x^\alpha$ for each
1141: $i$.
1142: \end{expl}
1143:
1144: For $R>0$, we write $B_\frL(R):= \set{x \in \frL:\ \|x\| < R}$.
1145:
1146: \begin{df}
1147: \label{quadratic}
1148: Let $W \subseteq \frL$. The set $W$ is a \textbf{standard quadratic
1149: domain} if there are constants $c,C>0$ such that
1150: \begin{equation*}
1151: W = \set{(r,\varphi) \in \frL:\ 0 < r < c \exp\left( -C \sqrt{|\varphi|}
1152: \right)}.
1153: \end{equation*}
1154: \end{df}
1155:
1156: Below, we put $\frL_0 := \frL \cup \{0\}$, and we extend the topology
1157: on $\frL$ to a topology on $\frL_0$ by taking the set $B_\frL(R) \cup
1158: \{0\}$, for $R>0$, as a basis of open neighborhoods of $0$ in
1159: $\frL_0$. For a subset $W$ of $\frL^m$, we shall write $\cl_0(W)$ for
1160: the topological closure of $W$ in $\frL_0^m$.
1161:
1162: \begin{rmk}
1163: Note that if $W \subseteq \frL$ is a standard quadratic domain, then
1164: $W \cup \{0\}$ is not an open neighborhood of $0$ in $\frL_0$.
1165: In particular, $\ir(\cl_0(W)) = W$.
1166: \end{rmk}
1167:
1168: \begin{df}
1169: \label{quadratic2}
1170: A set $W \subseteq \frL$ is a \textbf{quadratic domain} if $W$
1171: contains a standard quadratic domain and $\ir(\cl_0(W)) = W$.
1172:
1173: Let $U \subseteq \frL^m$ and $k \le m$. We say that $U$ is
1174: \textbf{$k$-quadratic} if
1175: \begin{renumerate}
1176: \item there is a quadratic domain $W \subseteq \frL$ and an $R>0$
1177: such that $W^k \times B_\frL(R)^{m-k} \subseteq U$;
1178: \item $0 \notin \ir(\cl_0(\Pi_{X_i} U))$ for each $i=1, \dots, k$.
1179: \end{renumerate}
1180: \end{df}
1181:
1182: \begin{nrmks}
1183: \label{rmk:quadratic}
1184: Let $k \le m$.
1185: \begin{enumerate}
1186: \item Let $R>0$ and $W \subseteq \frL$ be a quadratic domain such
1187: that $W \subseteq B_\frL(R)$. If $l \ge k$, then $W^l \times
1188: B_\frL(R)^{m-l} \subseteq W^k \times B_\frL(R)^{m-k}$; therefore, every
1189: $k$-quadratic domain in $\frL^m$ contains an $l$-quadratic domain.
1190: \item Let $U \subseteq \frL^m$ be a $k$-quadratic domain, and let
1191: $r \in (0,\infty)^m$. Then the set $V := \log^{-1}(\log(U) - r)$
1192: is a $k$-quadratic domain.
1193: \end{enumerate}
1194: \end{nrmks}
1195:
1196: We now fix an $m$-quadratic domain $U \subseteq \frL^m$.
1197:
1198: \begin{df}
1199: \label{asymptotic}
1200: Let $f \in \O(U)$ and $F = \sum a_\alpha X^\alpha \in \Ps{C}{X^*}$. We say
1201: that $f$ has \textbf{asymptotic expansion} $F$ on $U$ and write $f
1202: \sim_U F$, if for each $\a>0$ there is an $m$-quadratic domain $U_{\a}
1203: \subseteq U$ such that
1204: \begin{equation*}
1205: f(x) - \sum_{|\alpha| \le \a} a_{\alpha} x^{\alpha} =
1206: o\left(\|x\|^{\a}\right) \quad\text{as } \|x\| \to 0 \text{ on }
1207: U_\a.
1208: \end{equation*}
1209: Note that in this situation, $F$ is the unique series $G \in
1210: \Ps{C}{X^*}$ with $f \sim_U G$; we therefore also write $Tf := F$,
1211: and we put $f(0) := \lim_{\|x\| \to 0,\,x \in U_\a} f(x)$ for any
1212: $\a > 0$. We let $\A(U)$ be the set of all $f \in \O(U)$ for which
1213: there is an $F \in \Ps{C}{X^*}$ such that $f \sim_U F$.
1214: \end{df}
1215:
1216: \begin{nrmk}
1217: \label{O-rmk}
1218: \begin{enumerate}
1219: \item If $f \in \A(U)$ and $n \ge m$, then we consider $f$ as an
1220: holomorphic function $f:U \times \frL^{n-m}$ in the obvious way;
1221: under this identification, we get $\A(U) \subseteq \A(U
1222: \times \frL^{n-m})$.
1223: \item Let $f \in \O(U)$ and $V \subseteq U$. Then $f\rest{V} \in
1224: \A(V)$ if and only if $f \in \A(U)$.
1225: \item The set $\A(U)$ is a $\CC$-algebra, and the map $T:\A(U) \into
1226: \Ps{C}{X^*}$ given by $T(f) := Tf$ is a $\CC$-algebra homomorphism
1227: satisfying $f(0) = (Tf)(0)$.
1228: \end{enumerate}
1229: \end{nrmk}
1230:
1231: For $R>0$, we put
1232: \begin{equation*}
1233: (0,R)_\frL := \set{x \in \frL:\ \|x\| < R \text{ and } \arg(x) = 0}.
1234: \end{equation*}
1235:
1236: \begin{prop}
1237: \label{asymptotic_to_0}
1238: The map $T:\A(U) \into \Ps{C}{X^*}$ is injective.
1239: \end{prop}
1240:
1241: \begin{proof}
1242: Let $f \in \A(U)$, and assume that $f \sim_U 0$; it suffices to show
1243: that $f = 0$. Let $c,C>0$ and $$W := \set{(r,\varphi) \in \frL:\
1244: 0<r<c \exp\left(-C\sqrt{|\varphi|}\right)}$$ be such that $W^m
1245: \subseteq U$. For $s=(s_{1},\dots,s_{m}) \in (0,1]^{m}$, we define
1246: $f_{s}: W \into \CC$ by $f_s(r,\varphi) := f((s_{1}r,\varphi),
1247: \dots, (s_{m}r,\varphi))$. Then $f_s \in \O(W)$, and our hypothesis
1248: implies that $f_s \sim_W 0$. This means that
1249: \begin{equation*}
1250: \|f_s \circ \log^{-1}(z)\| = o(e^{-n\re z}) \quad\text{as } \re
1251: z \to -\infty \text{ in } \log(W)
1252: \end{equation*}
1253: for every $n \in \NN$; hence $f_s = 0$ by Theorem 2 on p. 23 of
1254: \cite{ily:dulac}. In particular, $f((s_1c,0), \dots, (s_mc,0))
1255: = 0$ for all $s \in (0,1)^m$, that is, $f \rest{(0,c)_\frL^{m}} = 0$.
1256: Therefore, the holomorphic map $h := f \circ \log^{-1}$ vanishes on
1257: $(-\infty,\log c)^{m}$. Since $\log(U) \subseteq \CC^{m}$ is
1258: connected, it follows that $h=0$ and hence that $f=0$.
1259: \end{proof}
1260:
1261: \begin{prop}
1262: \label{prop:bad_derivatives}
1263: Let $f \in \A(U)$ and $i \in \{1, \dots, m\}$. Then $\partial_i f$
1264: belongs to $\A(U)$ and satisfies $T(\partial_i f) = \partial_i(Tf)$.
1265: \end{prop}
1266:
1267: \begin{proof}
1268: Let $\a>0$, and assume that $U$ is an $m$-quadratic domain and
1269: $\|f(x)\| = o(\|x\|^\a)$ as $\|x\| \to 0$ with $x \in U$. By Remark
1270: \ref{O-rmk}(2) and Example \ref{expl:powers}, it suffices to find an
1271: $m$-quadratic domain $V \subseteq U$ such that $\|\partial_i f(x)\|
1272: = o(\|x\|^\a)$ as $\|x\| \to 0$ with $x \in V$.
1273:
1274: We claim that $V := \log^{-1}(\log(U)-(1, \dots, 1))$ works; by
1275: Remark \ref{rmk:quadratic}(2), $V$ is an $m$-quadratic domain
1276: contained in $U$. To see the claim, for each $r>0$, we put
1277: \begin{equation*}
1278: M_r := \max\set{\|f(x)\|:\ x \in U \text{ and } \|x\| \le r}
1279: \end{equation*}
1280: and
1281: \begin{equation*}
1282: N_r := \max\set{\|\partial_i f(x)\|:\ x \in V \text{ and } \|x\| \le
1283: r}.
1284: \end{equation*}
1285: By assumption, we have $M_r / r^\a \to 0$ as $r \to 0$; we need to
1286: show that $N_r / r^\a \to 0$ as $r \to 0$. To do so, it suffices to
1287: show that $N_r \le M_{r \cdot e}$ for each $r>0$, where $e :=
1288: \exp(1)$. By the Cauchy estimates, we have for all $x \in V$ with
1289: $\|x\| \le r$ that
1290: \begin{equation*}
1291: \|\partial_if(x)\| = \left\|\frac{\partial (f \circ \log^{-1})} {\partial
1292: z_i} (\log x)\right\| \le M_{r \cdot e},
1293: \end{equation*}
1294: because $M_{r \cdot e}$ is the maximum of all $\|(f \circ
1295: \log^{-1})(z)\|$ such that $z \in \log(U)$ and $\re z_j \le \log r
1296: - 1$ for each $j$. This finishes the proof of the proposition.
1297: \end{proof}
1298:
1299: We now also fix a $k \le m$.
1300:
1301: \begin{df}
1302: \label{hol_proj}
1303: We let $\A^m_k(U)$ be the set of all $f \in \A(U)$ such that $Tf \in
1304: \Ps{C}{X_{\{1, \dots, k\}}^*, X_{\{k+1, \dots, m\}}}$. We also let
1305: $\pi^m_k:\frL_0^m \into \frL_0^k \times \CC^{m-k}$ be the map
1306: defined by $\pi^m_k(x) = (y_1, \dots, y_m)$, where
1307: \begin{equation*}
1308: y_i :=
1309: \begin{cases}
1310: x_i &\text{if } i \le k, \\ x_i^1 &\text{if } i>k \text{
1311: and } x_i \ne 0, \\ 0 &\text{otherwise.}
1312: \end{cases}
1313: \end{equation*}
1314: \end{df}
1315: We also denote by $\cl_0$ the topological closure in $\frL_0^k \times
1316: \CC^{m-k}$. As usual, we shall omit the superscript $m$ if clear from
1317: context.
1318:
1319: Finally, for each $i = 1, \dots, m$, we let $p_i:\frL^m \into \frL^m$
1320: be the map defined by $p_i(x) := y$ with $y_j:= x_j$ if $j \ne i$,
1321: $\|y_i\| := \|x_i\|$ and $\arg(y_i):= \arg(x_i)+2\pi$.
1322:
1323: \begin{prop}
1324: \label{prop:characterization}
1325: Let $f \in \A(U)$. The following are equivalent:
1326: \begin{enumerate}
1327: \item $f \in \A_k(U)$.
1328: \item There are a $k$-quadratic domain $V \subseteq \frL^m$ and a
1329: holomorphic $f^\sharp : \ir(\cl_0(\pi_k(V))) \into \CC$ such that
1330: $f(x) = f^\sharp(\pi_k(x))$ for all $x \in U \cap V$.
1331: \item For every $i = k+1, \dots, m$ and every $x \in U$ satisfying
1332: $p_i(x) \in U$, we have $f(x) = f(p_i(x))$.
1333: \end{enumerate}
1334: \end{prop}
1335:
1336: \begin{proof}
1337: (2) $\Rightarrow$ (3): straightforward, since $U$ is connected.
1338: \smallskip
1339:
1340: \noindent (3) $\Rightarrow$ (2): without loss of generality, we may
1341: assume that $U = W^m$ for some quadratic domain $W \subseteq \frL$.
1342: Let $R>0$ be such that $B(0,R) \subseteq \ir(\cl(\pi^1_0(W)))$, and
1343: put $V := W^k \times B_\frL(R)^{m-k}$. Then the assumption on $f$ and
1344: Riemann's theorem on removable singularities imply that there is a
1345: holomorphic $g : \ir(\cl_0(\pi_k(V))) \into \CC$ such that $f(x) =
1346: g(\pi_k(x))$ for all $x \in U \cap V$, which proves (2). \smallskip
1347:
1348: \noindent (3) $\Rightarrow$ (1): Assume (3) and let $i \in \{k+1,
1349: \dots, m\}$. Write $Tf(X) = \sum a_\alpha X^\alpha$, and let
1350: $\alpha \in [0,\infty)^m$ be such that $\alpha_i \notin \NN$; we
1351: need to show that $a_\alpha = 0$. Permuting coordinates if
1352: necessary, we may assume that $i=m$. Let $R>0$ be such that the set
1353: $V := (0,R)_\frL^m \cup p_m((0,R)_\frL^m)$ is contained in $U$.
1354:
1355: Put $\a:= |\alpha|>0$, and define $g:V \into \CC$ by $g(x) := f(x) -
1356: \sum_{|\beta| \le \a} a_\beta x^\beta$. Since $f \in \A(U)$, we
1357: have $g(x) = o\left(\|x\|^\a\right)$ as $\|z\| \to 0$ in $V$; in
1358: particular, for $s \in (0,R)$ and any $t \in (0,1)^m$, we have
1359: \begin{equation}
1360: \label{eq:diff_asym}
1361: g(st_1, \dots, st_m) - g(p_m(st_1, \dots, st_m)) = o(s^\a)
1362: \quad\text{as } s \to 0.
1363: \end{equation}
1364: On the other hand,
1365: \begin{equation*}
1366: g(st_1, \dots, st_m) - g(p_m(st_1, \dots, st_m)) =
1367: \sum_{\substack{|\beta| \le \a \\ \beta_m \notin \NN}} a_\beta
1368: \left(e^{2\pi i \cdot
1369: \beta_m}-1\right) t^\beta s^{|\beta|}.
1370: \end{equation*}
1371: It follows from \eqref{eq:diff_asym} that $\sum_{\substack{|\beta| \le
1372: \a \\ \beta_m \notin \NN}} a_\beta \left(e^{2\pi i \cdot
1373: \beta_m}-1\right) t^\beta = 0$. Since $t \in (0,1)^m$ was
1374: arbitrary, we obtain that $a_\beta = 0$ for every $\beta \in
1375: [0,\infty)^m$ satisfying $|\beta|\le \a$ and $\beta_m \notin \NN$;
1376: in particular, $a_\alpha = 0$. \smallskip
1377:
1378: \noindent (1) $\Rightarrow$ (3): assume that $Tf \in \Ps{C}{X_{\{1,
1379: \dots, k\}}^*, X_{\{k+1, \dots, m\}}}$, and let $i \in \{k+1,
1380: \dots, m\}$. Then there is a quadratic domain $V \subseteq U$ such
1381: that $p_i(V) \subseteq U$, and we define $g:V \into \CC$ by $g(x) :=
1382: f(x) - f(p_i(x))$. Then $g \in \O(V)$, and since $Tf \in
1383: \Ps{C}{X_{\{1, \dots, k\}}^*, X_{\{k+1, \dots, m\}}}$, we have $g
1384: \sim_V 0$. Thus $g = 0$ by Proposition \ref{asymptotic_to_0}, which
1385: proves (3).
1386: \end{proof}
1387:
1388: Assume that $U$ is a $k$-quadratic domain and let $f \in \A_k(U)$.
1389: Let $V$ and $f^\sharp$ be as in Proposition
1390: \ref{prop:characterization}(2); by analytic continuation, we may
1391: assume that $U = V$. We extend $f$ to a function on
1392: $\pi_k^{-1}(\ir(\cl_0(\pi_k(U))))$ by putting $f(x) :=
1393: f^\sharp(\pi_k(x))$. For example, we let $W \subseteq \frL$ be a
1394: quadratic domain and $R>0$ be such that $W^k \times B_\frL(R)^{m-k}
1395: \subseteq U$. Then the value $f(x',0)$ is well defined for all
1396: $x' \in W^k \times B_\frL(R)^{m-k-1}$, and we have:
1397:
1398: \begin{cor}
1399: \label{value_at_0}
1400: The function $g:W^k \times B_\frL(R)^{m-k-1} \into \CC$ defined by $g(x')
1401: := f(x',0)$ belongs to $\A_k^{m-1}(W^k \times B_\frL(R)^{m-k-1})$ and
1402: satisfies $Tg(X') = Tf(X',0)$. \qed
1403: \end{cor}
1404:
1405: Moreover, we let $i \in \{k+1, \dots, m\}$. Then the partial
1406: derivative $\partial f^\sharp/\partial z_i:\pi_k(U) \into \CC$ is
1407: defined as usual; using this, we define the partial derivative
1408: $\partial f/\partial x_i:U \into \CC$ by
1409: \begin{equation*}
1410: \frac {\partial f}{\partial x_i} (x) := \frac {\partial
1411: f^\sharp}{\partial z_i}(\pi_k(x)).
1412: \end{equation*}
1413:
1414: Using the Cauchy estimates (similarly as in the proof of Proposition
1415: \ref{prop:bad_derivatives}) and Proposition \ref{asymptotic_to_0}, we
1416: obtain:
1417:
1418: \begin{cor}
1419: \label{good_derivatives}
1420: Assume that $U$ is a $k$-quadratic domain, and let $f \in \A_k(U)$.
1421: Then for every $i=k+1, \dots, m$, the partial derivative $\partial
1422: f/\partial x_i$ belongs to $\A_k(U)$ and satisfies $T(\partial
1423: f/\partial x_i) = \partial (Tf)/\partial X_i$; in particular, $x_i
1424: \cdot (\partial f/\partial x_i)(x) = \partial_i f(x)$ for all $x \in
1425: U$. \qed
1426: \end{cor}
1427:
1428: Finally, we establish some criteria for membership in $\A_k(U)$: we
1429: fix an $l \in \{k, \dots, m\}$. For $x \in \frL^m$ we write $Y :=
1430: (X_1, \dots, X_{l})$, $y := (x_1, \dots, x_{l})$, $Z:= (X_{l+1},
1431: \dots, X_m)$ and $z := (x_{l+1}, \dots, x_m)$. We assume that $U =
1432: W^k \times B_\frL(R)^{m-k}$ for some quadratic domain $W \subseteq \frL$
1433: and some $R>0$, and we let $f \in \A_k(U)$. By Proposition
1434: \ref{prop:characterization}, for each $x = (y,z) \in U$ we have a
1435: convergent power series representation
1436: \begin{equation*}
1437: f(x) = \sum_{p \in \NN^{m-l}} a_p(y) z^p.
1438: \end{equation*}
1439: Writing $Tf = \sum_{\substack{p \in \NN^{m-l} \\ \alpha \in [0,\infty)^k
1440: \times \NN^{l-k}}} a_{p,\alpha} \cdot Y^\alpha Z^p$, we put
1441: \begin{equation*}
1442: A_p(Y) := \sum_{\alpha \in [0,\infty)^k \times \NN^{l-k}} a_{p,\alpha} \cdot
1443: Y^\alpha, \quad\text{for each } p \in \NN^{m-l}.
1444: \end{equation*}
1445: Shrinking $U$ if necessary, there is an $M>0$ such that $\|f(x)\| \le
1446: M$ for all $x \in U$. Finally, we put $U' := W^k \times B_\frL(R)^{l-k}$.
1447: Since $a_p(y) = \frac1{p!} \cdot \partial^p f/\partial (z)^p(y,0)$ for
1448: all $p \in \NN^{m-l}$ and all $y \in U'$, we obtain from the Cauchy
1449: estimates and Corollaries \ref{value_at_0} and \ref{good_derivatives}:
1450:
1451: \begin{prop}
1452: \label{coefficients}
1453: For each $p \in \NN^{m-l}$, the function $a_p:U' \into \CC$ belongs
1454: to $\A^l_k(U')$ and satisfies $Ta_p = A_p$ and $\|a_p(y)\| \le
1455: M/R^{|p|}$ for all $y \in U'$. \qed
1456: \end{prop}
1457:
1458: The following converse to Proposition \ref{coefficients} is our
1459: principal test for membership in $\A_k(U)$:
1460:
1461: \begin{prop}
1462: \label{O-membership}
1463: Let $S \subseteq [0,\infty)^k \times \NN^{l-k}$ be natural, let
1464: $A>0$, and for each $p \in \NN^{m-l}$ let $b_p \in \A^l_k(U')$ be
1465: such that
1466: \begin{renumerate}
1467: \item $\supp(Tb_p) \subseteq S$;
1468: \item $||b_p(y)\| \le A/R^{|p|}$ for all $y \in U'$.
1469: \end{renumerate}
1470: Then the function $g:U \into \CC$ defined by $g(x) := \sum_{p \in
1471: \NN^{m-l}} b_p(y) \cdot z^p$ belongs to $\A_k(U)$ and satisfies
1472: $Tf = \sum_{p \in \NN^{m-l}} Tb_p(Y) \cdot Z^p$.
1473: \end{prop}
1474:
1475: \begin{proof}
1476: It follows from the assumptions that $g$ is holomorphic on $U$; it
1477: remains to show that $g \sim_U \sum_{p \in \NN^{m-l}} Tb_p(Y) Z^p$.
1478: Let $\a>0$; for each $p \in \NN^{m-l}$, we write $Tb_p = \sum
1479: b_{p,\alpha} Y^\alpha$ and define $\epsilon_p: U' \into \CC$ by
1480: \begin{equation*}
1481: \epsilon_p(y) := b_p(y) - \sum_{|\alpha| \le \a-|p|} b_{p,\alpha}
1482: y^\alpha.
1483: \end{equation*}
1484: After shrinking $W$ if necessary, there is a constant $C>0$ such
1485: that $|\epsilon_p(y)| \le C\|y\|^{\a-|p|}$ whenever $|p| \le \a$ and
1486: $y \in U'$. For every $x \in U$, we have
1487: \begin{equation*}
1488: g(x) - \sum_{|(\alpha,p)| \le \a} b_{p,\alpha} y^\alpha z^p
1489: = \sum_{|p| \le \a} \epsilon_p(y) z^p + \sum_{|p| > \a}
1490: b_p(y) z^p.
1491: \end{equation*}
1492: By the above, $\sum_{|p| \le \a} \epsilon_p(y) z^p =
1493: o\left(\|x\|^\a\right)$ as $\|x\| \to 0$ in $U$; so it suffices to
1494: show, after shrinking $U$ again if necessary, that $\sum_{|p| > \a}
1495: b_p(y) z^p = o\left(\|x\|^\a\right)$ as $\|x\| \to 0$ in $U$. Let
1496: $\rho:= \inf\set{|p|-\a:\ |p| > \a} > 0$. By assumption (ii), we have
1497: for $x \in U$ satisfying $\|z\| \le \frac \rho 2$ that
1498: \begin{equation*}
1499: \left|\sum_{|p| > \a} b_p(y) z^p \right| \le \frac{A}{R^{\a+\rho}}
1500: \left(\sum_{|p| \ge \a + \rho} 2^{\a+\rho-|p|}\right) \cdot
1501: \|z\|^{\a+\rho} = o\left(\|x\|^\a\right)
1502: \end{equation*}
1503: as $\|x\| \to 0$, as required.
1504: \end{proof}
1505:
1506: The following criterion for membership in $\A_k(U)$ will be
1507: useful in Section \ref{family}:
1508:
1509: \begin{cor}
1510: \label{uniform_criterion}
1511: Let $f:U \into \CC$ be holomorphic, and let $W \subseteq \frL$ be a
1512: quadratic domain and $R>0$ be such that $W^k \times B_\frL(R)^{m-k}
1513: \subseteq U$. Let $S \subseteq [0,\infty)^k$ be a natural set, and
1514: assume that for each $\alpha \in S$, there is a holomorphic function
1515: $a_\alpha:B_\frL(R)^{m-k} \into \CC$ such that
1516: \begin{renumerate}
1517: \item for every $z \in B_\frL(R)^{m-k}$, the function $f_z:W^k \into \CC$
1518: defined by $f_z(y):= f(y,z)$ belongs to $\A_k(W^k)$ with $Tf_z =
1519: \sum_{\alpha \in S} a_\alpha(z) y^\alpha$;
1520: \item for each $\nu>0$, there are constants $K_\nu, \epsilon_\nu >
1521: 0$ and a quadratic domain $W_\nu \subseteq W$ such that $\left\|
1522: f(y,z) - \sum_{|\alpha| \le \nu} a_\alpha(z) y^\alpha \right\|
1523: \le K_\nu \|y\|^{\nu+\epsilon_\nu}$ for all $(y,z) \in W_\nu^k
1524: \times B_\frL(R)^{m-k}$.
1525: \end{renumerate}
1526: Then $f \in \A_k\left(W^k \times B_\frL(R)^{m-k}\right)$, and if
1527: $Ta_\alpha = \sum a_{\alpha,p} z^p$ for each $\alpha \in S$, then
1528: $Tf = \sum a_{\alpha,p} y^\alpha z^p$.
1529: \end{cor}
1530:
1531: \begin{proof}
1532: By assumption (i) and Proposition \ref{O-membership}, it suffices to
1533: show that the function $f^{(l)} : W^k \into \CC$ defined by
1534: $f^{(l)}(y) := (\partial^l f/\partial y^l)(y,0)$ belongs to
1535: $\A_k(W^k)$ for each $l \in \NN^{m-k}$. But for any such $l$, any
1536: $\nu>0$ and any $y \in W_\nu$, we get from the Cauchy estimates that
1537: \begin{align*}
1538: \left\| f^{(l)}(y) - \sum_{|\alpha| \le \nu} a_\alpha^{(l)}(0)
1539: y^\alpha \right\| &= \left\| \frac{\partial^l}{\partial y^l}\left(
1540: f(y,z) - \sum_{|\alpha| \le \nu} a_\alpha(z) y^\alpha \right)
1541: (y,0)\right\| \\ &\le \frac {K_\nu}{R^{|l|}} \|y\|^{\nu+\epsilon_\nu},
1542: \end{align*}
1543: as required.
1544: \end{proof}
1545:
1546:
1547: \section{Truncation-division, Taylor expansion and composition in the holomorphic variables}
1548: \label{tdte}
1549:
1550: It will be convenient from now on to more explicitely separate the
1551: holomorphic variables from the non-holomorphic ones. Thus, we let
1552: $m,n \in \NN$, and let $U \subseteq \frL^{m+n}$ be an $m$-quadratic
1553: domain. Below, we let $y = (y_1, \dots, y_n)$ range over $\frL^n$ and
1554: $Y = (Y_1, \dots, Y_n)$ be indeterminates.
1555:
1556: \begin{nrmk}
1557: \label{permutations}
1558: Let $\sigma$ be a permutation of $\{1, \dots, m\}$ and $\tau$ be a
1559: permutation of $\{1, \dots, n\}$. We associate to $\sigma$ and
1560: $\tau$ the substitution automorphisms $\sigma,\tau:\Ps{C}{X^*,Y}
1561: \into \Ps{C}{X^*,Y}$ defined by $\sigma(X,Y) := (X_{\sigma(1)},
1562: \dots, X_{\sigma(m)},Y)$ and $\tau(X,Y):= (X,Y_{\tau(1)}, \dots,
1563: Y_{\tau(n)})$ and the maps $\sigma, \tau:\frL^{m+n} \into
1564: \frL^{m+n}$ defined by $\sigma(x,y) := (x_{\sigma(1)}, \dots,
1565: x_{\sigma(m)},y)$ and $\tau(x,y) := (x,y_{\tau(1)}, \dots,
1566: y_{\tau(n)})$. Then for every $f \in \A_m(U)$, we have $f \circ
1567: \sigma \in \A_m(\sigma^{-1}(U))$ with $T(f \circ \sigma) =
1568: \sigma(Tf)$ and $f \circ \tau \in \A_m(\tau^{-1}(U))$ with $T(f
1569: \circ \tau) = \tau(Tf)$.
1570: \end{nrmk}
1571:
1572: \subsection*{Truncation-division} First, we show that the natural
1573: operations of truncation and division by monomials in the $Y$
1574: variables of $Tf$, where $f \in \A_m(U)$, lead to new functions in
1575: $\A_m(U)$.
1576:
1577: \begin{prop}
1578: \label{holo_division}
1579: Let $f \in \A_m(U)$ and $\delta \in \NN^n$.
1580: \begin{enumerate}
1581: \item Assume that $Tf = Y^\delta \cdot G$ for some $G \in
1582: \Ps{C}{X^*,Y}$. Then there is a $g \in \A_m(U)$ such that $Tg =
1583: G$.
1584: \item There is an $f_{(0,\delta)} \in \A_m(U)$ such that
1585: $Tf_{(0,\delta)} = (Tf)_{(0,\delta)}$.
1586: \end{enumerate}
1587: \end{prop}
1588:
1589: \begin{proof}
1590: Below, we write $\hat{y}_j := (y_1, \dots, y_{j-1}, 0, y_{j+1},
1591: \dots, y_n)$ for each $j = 1, \dots, n$.
1592:
1593: (1) Working by induction on $|\delta|$ and using Remark
1594: \ref{permutations}, we may assume that $\delta = (0, \dots, 0, 1)$.
1595: Given $h \in \A_m(U)$, the function $I(h):U \into \CC$ defined by
1596: \begin{equation*}
1597: I(h)(x,y) := \int_0^1 h(x,y_1, \dots, y_{n-1}, ty_n) dt
1598: \end{equation*}
1599: belongs to $\A_m(U)$, where $t \cdot (r,\varphi) := (tr,\varphi)$
1600: for $t>0$ and $(r,\varphi) \in \frL$. Since $f(x,\hat{y}_n) = 0$,
1601: we get from the fundamental theorem of calculus that $f(x,y) = y_n
1602: \cdot I(\partial f/\partial y_n)(x,y)$ for all $(x,y) \in U$, so
1603: part (1) follows from Corollary \ref{good_derivatives}.
1604:
1605: (2) We define $h:U \into \CC$ by
1606: \begin{equation*}
1607: h(x,y) := f(x,y) - \sum_{j=1}^n \sum_{p=0}^{\delta_j-1} \frac1{p!}
1608: \frac {\partial^p f}{\partial y_j^p}(x,\hat{y}_j) \cdot y_j^p.
1609: \end{equation*}
1610: By Corollaries \ref{value_at_0} and \ref{good_derivatives}, the
1611: function $h$ belongs to $\A_m(U)$ and $Th = Y^\delta \cdot G$ for
1612: some $G \in \Ps{C}{X^*,Y}$. Part (2) now follows from part (1).
1613: \end{proof}
1614:
1615: \begin{nrmk}
1616: \label{truncation_remark}
1617: We do not know whether, in the situation of Proposition
1618: \ref{holo_division}, a corresponding statements holds for all
1619: variables, that is, whether
1620: \begin{itemize}
1621: \item[$(\ast)_f$] for every elementary set $S \subseteq [0,\infty)^m
1622: \times \NN^n$, there are an $m$-quadratic $V \subseteq U$ and an
1623: $f_S \in \A_m(V)$ such that $T f_S = (Tf)_S$.
1624: \end{itemize}
1625: This is the first of two reasons to eventually restrict our
1626: attention to a subclass of $\A_m(U)$ introduced in Section
1627: \ref{division}.
1628: \end{nrmk}
1629:
1630: \subsection*{Taylor expansion}
1631: Next, we establish a Taylor expansion result with respect to the $y$
1632: variables. Let $V \subseteq \frL^p$ be open; recall that a map $\f:V
1633: \into \frL^p$ is holomorphic if the function $\log_p \circ \f \circ
1634: \log_p^{-1}:\log_p(V) \into \CC^p$ is holomorphic. Note that if $W
1635: \subseteq \frL^q$ and $\g:W \into \frL^p$ is holomorphic such that
1636: $\g(W) \subseteq V$, then the composition $\f \circ g:W \into \frL^p$
1637: is holomorphic.
1638:
1639: \begin{df}[Translation]
1640: \label{shift}
1641: Let $\lambda \in \CC$ be nonzero, and denote by $\Arg\lambda \in
1642: (-\frac\pi2, \frac\pi2]$ the standard argument of $\lambda$. We
1643: put $$\CC_\lambda := \set{z \in \CC:\ \Arg\lambda - \frac\pi2 < \arg
1644: z < \Arg\lambda + \frac\pi2},$$ and we let $\l_\lambda:\CC_\lambda
1645: \into \frL$ be defined by $\l_\lambda(z) := (|z|,\arg z)$ and put
1646: $D(\lambda) := \l_\lambda(B(\lambda,|\lambda|))$. We let
1647: $\t_\lambda:B_\frL(\lambda) \into D(\lambda)$ be the holomorphic map
1648: defined by $\t_\lambda(x) := \l_\lambda\left(\lambda +
1649: (x)^1\right)$.
1650:
1651: For completeness' sake, we also define $\t_0:\frL \into \frL$ by
1652: $\t_0(x) := x$.
1653: \end{df}
1654:
1655: For $\lambda \in \CC^p$ and $w \in \frL^p$, we write $\t_\lambda(w) :=
1656: (\t_{\lambda_1}(w_1), \dots, \t_{\lambda_p}(w_p))$. Abusing
1657: notation, we identify the set $\big\{x \in \frL^p_0:\ -\pi < \arg
1658: x_i \le \pi$ for each $i \big\}$ with $\CC^p$.
1659:
1660: \begin{nrmk}
1661: \label{translate_derivative}
1662: Let $\lambda \in \CC^{m+n} \cap \cl_0(U)$ and $f \in \O(U)$. An
1663: elementary calculation shows that for each $i \in \{1, \dots, m\}$,
1664: we have
1665: \begin{equation*}
1666: \partial_i(f \circ \t_\lambda) = \Big((\partial_i f) \circ
1667: \t_\lambda\Big) \cdot \frac {x_i^1}{\lambda_i + x_i^1}.
1668: \end{equation*}
1669: \end{nrmk}
1670:
1671: We now let $l \in \{1, \dots, n\}$, and we write $y' := (y_1, \dots,
1672: y_{n-l})$, \linebreak $Y' := (Y_1, \dots, Y_{n-l})$, $z = (z_1, \dots,
1673: z_l):= (y_{n-l+1}, \dots, y_n)$ and $Z = (Z_1, \dots, Z_l) :=
1674: (Y_{n-l+1}, \dots, Y_n)$. We assume that $U = W^m \times B_\frL(R)^n$ for
1675: some quadratic domain $W \subseteq \frL$ and some $R>0$. We let $f
1676: \in \A_m(U)$ and $\lambda \in B(0,R)^l$, and we put $U':= W^m \times
1677: B_\frL(R)^{n-l}$.
1678:
1679: \begin{lemma}
1680: \label{real_constants}
1681: Assume $(\ast)_f$ holds. Then the formal sum $Tf(X,Y',\lambda)$
1682: gives a series in $\Ps{C}{X^*,Y'}$. Moreover, the function $(x,y')
1683: \mapsto f(x,y',\lambda):U' \into \CC$, denoted simply by
1684: $f(x,y',\lambda)$, satisfies $(\ast)_{f(x,y',\lambda)}$ with
1685: $$T(f(x,y',\lambda)) = (Tf)(X,Y',\lambda).$$
1686: \end{lemma}
1687:
1688: \begin{proof}
1689: We assume that $l=1$; the general case follows by induction on
1690: $l$. Throughout this proof, we let $\gamma$ and $p$ range over
1691: $[0,\infty)^m \times \NN^{n-1}$ and $\NN$, respectively. We write
1692: $f(x,y) = \sum a_p(x,y') z^p$ as a convergent power series in $z =
1693: y_n$ and $Tf = \sum a_{(\gamma,p)} (X,Y')^\gamma Z^p$. From
1694: $(\ast)_f$ and Propositions \ref{coefficients} and
1695: \ref{asymptotic_to_0}, we see that $(\ast)_{a_p}$ holds (with $n-1$
1696: in place of $n$) for each $p \in \NN$. Hence
1697: \begin{renumerate}
1698: \item $f_{(\gamma,0)}(x,y) = \sum (a_p)_\gamma(x,y') z^p$ for all
1699: $\gamma$ and all sufficiently small $(x,y) \in U$;
1700: \item $a_{(\gamma,p)} = (a_p)_\gamma(0,0)$ for all $\gamma$ and $p$.
1701: \end{renumerate}
1702: It follows from (ii) and Proposition \ref{coefficients} that
1703: $$Tf(X,Y',\lambda) = \sum_\gamma
1704: \left(\sum_p a_{(\gamma,p)} \lambda^p\right) (X,Y')^\gamma$$ belongs
1705: to $\Ps{C}{X^*,Y'}$, which proves the first assertion.
1706:
1707: Next, we let $\a>0$ and consider the finite set $$S_\a :=
1708: \set{\gamma \in \Pi_{m+n-1}(\supp Tf):\ |\gamma| \le \a}.$$ Note
1709: that, by (ii) above and Proposition \ref{coefficients}, for every
1710: $\gamma \in S_\a$ we have $f_{\{\gamma\} \times \NN}(x,y) = \sum_p
1711: a_{(\gamma,p)} z^p$ for all sufficiently small $(x,y) \in U$.
1712: Therefore, after shrinking $U$ if necessary, the function $$g(x,y)
1713: := f(x,y) - \sum_{\gamma \in S_\a} f_{\{\gamma\} \times \NN}(x,y)
1714: \cdot (x,y')^\gamma$$ belongs to $\A_m(U)$ and satisfies $Tg = Tf -
1715: \sum_{\gamma \in S_\a} (Tf)_{\{\gamma\} \times \NN} \cdot
1716: (X,Y')^\gamma$. By the above, it follows that $$Tg(X,Y) = Tf(X,Y) -
1717: \sum_{\gamma \in S_\a} \sum_p a_{(\gamma,p)} \cdot (X,Y')^\gamma
1718: Z^p\ ;$$ in particular, $f(x,y',\lambda) - \sum_{|\gamma| \le \a}
1719: \left(\sum_p a_{(\gamma,p)} \lambda^p \right) \cdot (x,y')^\gamma =
1720: o(\|(x,y')\|^\a)$ as $\|(x,y')\| \to 0$ in $U'$. Hence
1721: $f(x,y',\lambda) \in \A_m(U')$ with $T(f(x,y',\lambda)) =
1722: (Tf)(X,Y',\lambda)$.
1723:
1724: Finally, given an elementary set $S \subseteq [0,\infty)^m \times
1725: \NN^{n-1}$ and arguing as above with $f_{S \times \{0\}}$ in place
1726: of $f$, we see that $(\ast)_{f(x,y',\lambda)}$ holds.
1727: \end{proof}
1728:
1729: We put $R':= \min_{i=1, \dots, l} (R - |\lambda_i|)$ and $V := W^m
1730: \times B_\frL(R)^{n-l} \times B_\frL(R')^l$. From the Taylor expansion theorem
1731: for holomorphic functions, Propositions \ref{coefficients} and
1732: \ref{O-membership} and Lemma \ref{real_constants}, we obtain:
1733:
1734: \begin{cor}
1735: \label{taylor}
1736: Assume $(\ast)_f$ holds. Then the function $g:V \into \CC$ defined
1737: by $g(x,y) := f(x,y',\t_\lambda(z))$ satisfies $(\ast)_g$ with
1738: \begin{equation*}
1739: Tg(X,Y) = \sum_{p \in \NN^l} \frac1{p!} \frac
1740: {\partial^p (Tf)}{\partial Z^p}(X,Y',\lambda) \cdot
1741: Z^p. \qed
1742: \end{equation*}
1743: \end{cor}
1744:
1745: In the situation of Corollary \ref{taylor}, we also write
1746: $\t_{(0,\lambda)} f$ and $T_{(0,\lambda)} f$ for the function $g$ and
1747: the series $Tg$, respectively.
1748:
1749: \begin{nrmk}
1750: \label{taylor_remark}
1751: We are not aware of a statement corresponding to Corollary
1752: \ref{taylor} for translation in the $x$ variables. This is the
1753: second reason to restrict our attention to a subclass of $\A_m(U)$,
1754: done in Section \ref{division}.
1755: \end{nrmk}
1756:
1757: \subsection*{Composition} Let $f \in \A_m(U)$. Let $V \subseteq
1758: \frL^{m+n}$ be an $m$-quadratic domain and let $g = (g_1, \dots, g_n)
1759: \in \A_m(V)^n$. Abusing notation, for $(x,y) \in V$ we write
1760: $(x,g(x,y)) \in U$ to mean $(x, g(x,y)) \in \ir(\cl_0(\pi_m(U)))$, and
1761: if the latter is the case, we also write $f(x,g(x,y))$ in place of
1762: $f^\sharp(x,g(x,y))$.
1763:
1764: For the next lemma, we assume that $g(0,0) = 0$ and $(x, g(x,y))
1765: \in U$ for all $(x,y) \in V$, and we define the holomorphic function
1766: $h:V \into \CC$ by $ h(x,y) := f(x,g(x,y)). $
1767:
1768: \begin{prop}
1769: \label{A-composition}
1770: The function $h$ belongs to $\A_m(V)$ and $Th(X,Y) =
1771: Tf(X,Tg(X,Y))$.
1772: \end{prop}
1773:
1774: We will deduce this proposition from the following two special cases:
1775:
1776: \begin{lemma}
1777: \label{comp1}
1778: Assume that $Tg(X,0) = Tg(X,Y)$. Then $h \in \A_m(V)$ and $Th(X,Y)
1779: = Tf(X,Tg(X,Y))$.
1780: \end{lemma}
1781:
1782: \begin{proof}
1783: We write $Tf(X,Y) = \sum a_{\gamma,p} X^\gamma Y^p$ and $Tg_j(X) =
1784: \sum b_{j,\delta} X^\delta$; note that $b_{j,0} = 0$ for each $j$.
1785: Then $(Tf)(X,(Tg)(X,Y)) = \sum c_\alpha X^\alpha$, where for each
1786: $\alpha \in [0,\infty)^m$,
1787: \begin{multline*}
1788: \Sigma(\alpha) := \big\{(\gamma,p,\delta):\ \gamma \in
1789: \Pi_m(\supp(Tf)),\, p \in \NN^n, \text{ and } \\ \delta =
1790: (\delta^1, \dots, \delta^{|p|}) \in \left(\bigcup
1791: \supp(g_j)\right)^{|p|} \text{ with } \gamma + \delta^1 + \cdots
1792: \delta^{|p|} = \alpha\big\}
1793: \end{multline*}
1794: and
1795: \begin{equation*}
1796: c_\alpha := \sum_{(\gamma,p,\delta) \in \Sigma(\alpha)}
1797: a_{\gamma,p} \cdot \prod_{j=1}^n \prod_{l=p_1 + \cdots + p_{j-1} +
1798: 1}^{p_1 + \cdots + p_l} b_{j,\delta^l}.
1799: \end{equation*}
1800: Since each $b_{j,0} = 0$, each set $\Sigma(\alpha)$ is finite; in
1801: fact, with
1802: \begin{equation*}
1803: q(r) := \sum_{i=1}^m \left| \Pi_{X_i}\left(\set{\beta \in
1804: \bigcup\supp(g_j):\ |\beta| \le r}\right) \right|,
1805: \end{equation*}
1806: we have $|p| \le q(|\alpha|)$ for all $(\gamma,p,\delta) \in
1807: \Sigma(\alpha)$.
1808:
1809: Let now $\a > 1$, and for all suitable $(x,y) \in \frL^{m+n}$, we
1810: define
1811: \begin{equation*}
1812: f^\a(x,y) := f(x,y) - \sum_{|\gamma| + |p| \le \a + q(\a)}
1813: a_{\gamma,p} x^\gamma y^p
1814: \end{equation*}
1815: and
1816: \begin{equation*}
1817: g_j^\a(x) := g_j(x,0) - \sum_{|\delta| \le \a} b_{j,\delta}
1818: x^\delta, \quad\text{for } j=1, \dots, n.
1819: \end{equation*}
1820: Then $f^\a(x,y) = o\left(\|(x,y)\|^{\a+q(\a)}\right)$ as $\|(x,y)\| \to
1821: 0$ in some $m$-quadratic domain, and $g_j^\a(x) = o(\|x\|^\a)$ for
1822: each $j$ as $\|x\| \to 0$ in some quadratic domain. Thus,
1823: $f^\a(x,g^a(x)) = o(\|x\|^\a)$, and there is a polynomial $P(x) = \sum
1824: d_\beta X^\beta$ such that $|\beta| > \a$ whenever $d_\beta \ne 0$
1825: and
1826: \begin{equation*}
1827: P(x) = f(x,g(x,0)) - f^\a(x,g^\a(x)) - \sum_{|\alpha| \le \a}
1828: c_\alpha x^\alpha
1829: \end{equation*}
1830: for all sufficiently small $x$. In particular, $f(x,g(x,0)) -
1831: \sum_{|\alpha| \le \a} c_\alpha x^\alpha = o(\|x\|^\a)$, which proves
1832: the lemma.
1833: \end{proof}
1834:
1835: \begin{lemma}
1836: \label{comp2}
1837: Assume that $Tg(X,0) = 0$. Then $h \in \A_m(V)$ and $Th(X,Y)
1838: = Tf(X,Tg(X,Y))$.
1839: \end{lemma}
1840:
1841: \begin{proof}
1842: After shrinking $U$ and $V$ if necessary, we may assume that $f$ is
1843: bounded on $U$. By Proposition \ref{coefficients}, we can write
1844: $f(x) = \sum a_p(x) y^p$ for all $(x,y) \in U$ and $g_j(x,y) = \sum
1845: b_{j,p}(x) y^p$ for all $(x,y) \in V$ and $j=1, \dots, n$, and there
1846: are an $m$-quadratic domain $W \subseteq \frL^m$ and constants
1847: $A,B>0$ such that $a_p, b_{j,p} \in \A_m(W^m)$ and $\|a_p(x)\|,
1848: \|b_{j,p}(x)\| \le AB^{|p|}$ for all $x \in W^m$, $p \in \NN^n$ and
1849: $j=1, \dots, n$. Our assumption on $g$ implies that $b_{j,0} = 0$
1850: for each $j$. Thus, after shrinking $W$ if necessary, there is an
1851: $R>0$ such that for all $(x,y) \in W \times B_\frL(R)^n$,
1852: \begin{equation*}
1853: h(x,y) = f(x,g(x,y)) = \sum_{p \in \NN^n} a_p(x) g(x,y)^p = \sum_{r
1854: \in \NN^n} c_r(x) y^r,
1855: \end{equation*}
1856: where, for $r \in \NN^n$ and $x \in W^m$, we put
1857: \begin{equation*}
1858: c_r(x) := \sum_{(p,q) \in \Sigma(r)} a_p(x) \cdot \prod_{i=1}^n
1859: \prod_{j=p_1 + \cdots + p_{i-1}+1}^{p_1 + \cdots + p_i} b_{i,q^j}(x)
1860: \end{equation*}
1861: with
1862: \begin{multline*}
1863: \Sigma(r) := \big\{(p,q):\ p \in \NN^n \text{ with } |p| \le |r|,
1864: \text{ and } \\ q = (q^1, \dots, q^{|p|}) \in (\NN^n
1865: \setminus\{0\})^{|p|} \text{ with } q^1 + \cdots + q^{|p|} =
1866: r\big\}.
1867: \end{multline*}
1868: Note that each $\Sigma(r)$ is finite, because $|p| \le |r|$ for each
1869: $(p,q) \in \Sigma(r)$; we only need to consider such $p$, because
1870: each $b_{j,0} = 0$.) Since $h$ is bounded and holomorphic on $W^m
1871: \times B_\frL(R)^n$, there are $C,D>0$ such that $\|c_r(x)\| \le
1872: CD^{|r|}$ for all $x \in W^m$ and $r \in \NN^n$. Finally, writing
1873: $Tf(X,Y) = \sum_{p \in \NN^n} A_p(X) Y^p$ and $Tg_j(X,Y) = \sum_{q
1874: \in \NN^n} B_{j,p}(X) Y^q$ for $j=1, \dots, n$, it follows from
1875: Remark \ref{O-rmk}(3) that each $c_r$ belongs to $\A_m(W^m)$ and
1876: satisfies
1877: \begin{equation*}
1878: T c_r(X) = \sum_{(p,q) \in \Sigma(r)} A_p(X) \cdot \prod_{i=1}^n
1879: \prod_{j=p_1 + \cdots + p_{i-1}+1}^{p_1 + \cdots + p_i} B_{i,q^j}(x).
1880: \end{equation*}
1881: The claim now follows from Proposition \ref{O-membership}, because
1882: $\supp(Tc_r) \subseteq \Pi_m\big(\supp Tf(X,Tg(X,Y))\big)$ for
1883: each $r \in \NN^n$ and the latter is a natural set by Proposition
1884: \ref{comp_representation}.
1885: \end{proof}
1886:
1887: \begin{proof}[Proof of Proposition \ref{A-composition}]
1888: We define $f'(x,z,y) := f(x,z+y)$, $g^0(x) := g(x,0)$ and $g'(x,y)
1889: := g(x,y) - g(x,0)$ for all suitable $x \in \frL^m$ and $y,z \in
1890: \frL^n$. By Lemma \ref{comp2}, there is an $m$-quadratic $U'
1891: \subseteq \frL^{m+2n}$ such that $f' \in \A_m(U')$. Note that
1892: $T(g^0)(X,0) = T(g^0)(X,Y)$ and $T(g')(X,0) = 0$. Hence by Lemmas
1893: \ref{real_constants} and \ref{comp1}, there is an $m$-quadratic $V'
1894: \subseteq \frL^{m+n}$ such that $f'(x,g^0(x,y),y) \in \A_m(V')$, and
1895: by Lemma \ref{comp2}, there is an $m$-quadratic $V'' \subseteq
1896: \frL^{m+n}$ such that $f'(x,g^0(x,y),g'(x,y)) \in \A_m(V'')$. Since
1897: $f(x,g(x,y)) = f'(x,g^0(x),g'(x,y))$ for all suitable $(x,y) \in
1898: \frL^{m+n}$, the proposition follows from Remark \ref{O-rmk}(2).
1899: \end{proof}
1900:
1901:
1902: \section{Blow-up substitutions in the non-holomorphic variables}
1903: \label{blowups}
1904:
1905: We continue to work with $m,n \in \NN$ and an $m$-quadratic domain $U
1906: \subseteq \frL^{m+n}$. For each real $\rho > 0$, the map
1907: $\p^\rho:\frL \into \frL$ defined by
1908: \begin{equation*}
1909: \p^\rho(r,\varphi) := (r^\rho,\rho\varphi)
1910: \end{equation*}
1911: is holomorphic, and the map $\m:\frL^2 \into \frL$ defined by
1912: \begin{equation*}
1913: \m((r_1,\varphi_1),(r_2,\varphi_2)) := (r_1r_2, \varphi_1+\varphi_2)
1914: \end{equation*}
1915: is holomorphic. Note that for all $x,x_1,x_2 \in \frL$, we have
1916: $(\p^\rho(x))^1 = x^\rho$ for each $\rho > 0$ and $(\m(x_1,x_2))^1 =
1917: (x_1)^1 \cdot (x_2)^1$.
1918:
1919: If $m \ge 2$ and $\rho \in (0,\infty)^m$, we define the holomorphic
1920: map $\p^\rho:\frL^m \into \frL$ by induction on $m$:
1921: \begin{equation*}
1922: \p^\rho(x) := \m\left(\p^{\rho'}(x'), \p^{\rho_m}(x_m)\right),
1923: \end{equation*}
1924: where $x' := (x_1, \dots, x_{m-1})$ and $\rho' := (\rho_1, \dots,
1925: \rho_{m-1})$.
1926:
1927: \begin{df}
1928: \label{singular_blowup}
1929: Let $m \ge 2$ and $i,j \in \{1, \dots, m\}$ be such that $i \ne j$,
1930: and let $\rho > 0$. The \textbf{singular blowing-up} $\s^\rho_{ij}
1931: : \frL^{m+n} \into \frL^{m+n}$ is defined as $\s^\rho_{ij} (x,y) =
1932: (z,y)$, where
1933: \begin{equation*}
1934: z_k :=
1935: \begin{cases}
1936: x_k &\text{if } k \ne i, \\ \m(\p^\rho(x_j),x_i) &\text{if } k =
1937: i.
1938: \end{cases}
1939: \end{equation*}
1940: \end{df}
1941:
1942: \begin{prop}
1943: \label{singular_prop}
1944: Let $ m \ge 2$ and $f \in \A_m(U)$. Then the there is an
1945: $m$-quadratic $V \subseteq \frL^{m+n}$ such that $\s^\rho_{ij}(V)
1946: \subseteq U$ and the function $f \circ \s^\rho_{ij}$ belongs to
1947: $\A_m(V)$ and satisfies $T\left(f \circ \s^\rho_{ij}\right) =
1948: \frB^{\rho,0}_{ij}(Tf)$.
1949: \end{prop}
1950:
1951: \begin{proof}
1952: Without loss of generality, we may assume that $i = m$ and $j =
1953: m-1$. Below, we write $\s$ and $\frB$ in place of $\s^\rho_{ij}$
1954: and $\frB^{\rho,0}_{ij}$. Let $W \subseteq \frL$ be quadratic and
1955: $1 > R>0$ be such that $f \in \A_m(W^m \times B_\frL(R)^n)$; we may
1956: assume that
1957: \begin{equation*}
1958: W = \set{(r,\varphi) \in \frL:\ 0 < r <
1959: c\exp\left(-C\sqrt{|\varphi|}\right)}
1960: \end{equation*}
1961: for some $c,C>0$ satisfying $c<R$. We let $D :=
1962: C/\min\{\sqrt{\rho},1\}$ and put
1963: \begin{equation*}
1964: W' := \set{(r,\varphi) \in \frL:\ 0 < r <
1965: c\exp\left(-D\sqrt{|\varphi|}\right)} \subseteq W
1966: \end{equation*}
1967: and $V := (W')^m \times B_\frL(R)^n$; we claim that $\s(V) \subseteq U$.
1968: To see this, we write $x_k = (r_k,\varphi_k)$ for $k=1, \dots, m$.
1969: Then
1970: \begin{align*}
1971: \left\|\left(\s(x)\right)_m\right\| &\le c^{\rho+1}
1972: \exp\left( -D\left( \rho\sqrt{|\varphi_{m-1}|} +
1973: \sqrt{|\varphi_m|} \right)\right) \\ &\le c \exp\left(
1974: -D\left( \rho\sqrt{|\varphi_{m-1}|} + \sqrt{|\varphi_m|}
1975: \right)\right);
1976: \end{align*}
1977: since
1978: \begin{align*}
1979: C \sqrt{|\rho\varphi_{m-1} + \varphi_m|} &\le C\left( \sqrt\rho
1980: \sqrt{|\varphi_{m-1}|} + \sqrt{|\varphi_m|}\right) \\ &\le D
1981: \left(\rho \sqrt{|\varphi_{m-1}|} + \sqrt{|\varphi_m|}\right),
1982: \end{align*}
1983: the claim follows.
1984:
1985: Since $f
1986: \circ \s$ is holomorphic on $V$, for each $\beta \in \NN^n$ the function
1987: $a_\beta:(W')^m \into \CC$ defined by
1988: \begin{equation*}
1989: a_\beta(x') := \frac1{\beta!} \left(\frac {\partial^\beta
1990: f}{\partial y^\beta} \circ \s\right) (x,0)
1991: \end{equation*}
1992: is holomorphic. By Proposition \ref{O-membership}, it suffices to
1993: show that $a_\beta \in \A((W')^m)$ for each $\beta$. We
1994: fix $\beta \in \NN^n$ and write $T\left(\frac {\partial^\beta
1995: f}{\partial y^\beta}(X,0)\right) = \sum a_{\alpha} X^\alpha$,
1996: and we let $\a>0$. Shrinking $W$ and $W'$ if necessary, we may
1997: assume by Lemma \ref{real_constants} that
1998: \begin{equation}
1999: \label{preblow_asymptotic}
2000: \left\|\frac {\partial^\beta f}{\partial y^\beta}(x,0) -
2001: \sum_{|\alpha| \le \a} a_{\alpha}
2002: x^\alpha \right\|
2003: = o\left(\|x\|^\a\right) \quad\text{as } \|x\| \to 0
2004: \text{ in } W^m.
2005: \end{equation}
2006: We now define $\rho:[0,\infty)^m \into [0,\infty)^m$
2007: by $$\rho(\alpha) := (\alpha_1, \dots, \alpha_{m-2},
2008: \alpha_{m-1}+\rho\alpha_m, \alpha_m).$$ Note that
2009: $\frB(T(\partial^\beta f/\partial Y^\beta))(X,0) = \sum a_{\alpha}
2010: X^{\rho(\alpha)}$ and $\s(x)^{\alpha} = x^{\rho(\alpha)}$ for all $x
2011: \in (W')^m$ and $\alpha \in [0,\infty)^m$. Since $W \subseteq
2012: B_\frL(1)$, we have $\|\s(x)\| \le \|x\|$, so it follows from
2013: \eqref{preblow_asymptotic} that
2014: \begin{equation}
2015: \label{blow_asymptotic}
2016: \left\| \left(\frac {\partial^\beta
2017: f}{\partial y^\beta} \circ \s\right)(x,0) - \sum_{|\alpha|
2018: \le \a} a_{\alpha} x^{\rho(\alpha)} \right\| =
2019: o\left(\|x\|^\a\right)
2020: \end{equation}
2021: as $\|x\| \to 0$ in $(W')^m$. Finally, for $x \in W^m$ we have
2022: \begin{multline*}
2023: \left\|\left(\frac {\partial^\beta f}{\partial y^\beta} \circ
2024: \s\right) (x,0) - \sum_{|\rho(\alpha)| \le \a} a_{\alpha}
2025: x^{\rho(\alpha)} \right\| \\ \le \left\| \left(\frac
2026: {\partial^\beta f}{\partial y^\beta} \circ \s\right)(x,0) -
2027: \sum_{|\alpha| \le \a} a_{\alpha}
2028: x^{\rho(\alpha)} \right\| +
2029: \left\|\sum_{|\alpha| \le \a < |\rho(\alpha)|} a_{\alpha}
2030: x^{\rho(\alpha)} \right\|.
2031: \end{multline*}
2032: The right-hand side above is $o\left(\|x\|^\a\right)$ as $\|x\| \to
2033: 0$ in $(W')^m$, by \eqref{blow_asymptotic} and because the sum in
2034: the second summand is finite and each of its summands has an
2035: exponent $\gamma$ satisfying $|\gamma| > \a$. This proves the
2036: proposition.
2037: \end{proof}
2038:
2039: \begin{df}
2040: \label{regular_blowup}
2041: Let $m \ge 2$ and $\lambda > 0$. The \textbf{regular blowing-up}
2042: $\r^{\rho,\lambda}: \frL^{m-1} \times B_\frL(\lambda) \times \frL^n \into
2043: \frL^{m+n}$ is defined as $\r^{\rho,\lambda}(x,y) = (z,y)$, where
2044: \begin{equation*}
2045: z_k :=
2046: \begin{cases}
2047: x_k &\text{if } k < m, \\ \m\left( \p^\rho(x_{m-1}),
2048: \t_\lambda(x_m) \right) &\text{if } k=m.
2049: \end{cases}
2050: \end{equation*}
2051: \end{df}
2052:
2053: \begin{prop}
2054: \label{regular_prop}
2055: Let $m \ge 2$, $\lambda > 0$ and $f \in \A_m(U)$. Then there is an
2056: $(m-1)$-quadratic $V \subseteq \frL^{m-1} \times B_\frL(\lambda) \times
2057: \frL^n$ such that $\r^{\rho,\lambda}(V) \subseteq U$ and the
2058: function $f \circ \r^{\rho,\lambda}:V \into \CC$ belongs to
2059: $\A_{m-1}(V)$ and satisfies $T\left(f \circ \r^{\rho,\lambda}\right)
2060: = \frB^{\rho,\lambda}_{m,m-1}(Tf)$.
2061: \end{prop}
2062:
2063: \begin{proof}
2064: Below, we write $\r$ and $\frB$ in place of $\r^{\rho,\lambda}$ and
2065: $\frB^{\rho,\lambda}_{m,m-1}$. Let $W' \subseteq \frL$ be quadratic
2066: and $\min\{1,\lambda\} > R>0$ be such that $U':= (W')^m \times
2067: B_\frL(R)^n \subseteq U$; we may assume that
2068: \begin{equation*}
2069: W' = \set{(r,\varphi) \in \frL:\ 0 < r <
2070: c\exp\left(-C\sqrt{|\varphi|}\right)}
2071: \end{equation*}
2072: for some $c,C>0$ satisfying $c<R$. We let
2073: \begin{equation*}
2074: D:= \frac C{\min\{\sqrt\rho,1\}} \quad\text{and}\quad d :=
2075: \min\set{c, \left(
2076: \frac c{2\lambda \exp\left(D \sqrt{\pi/2}\right)} \right)^{1/\rho}},
2077: \end{equation*}
2078: and we put
2079: \begin{equation*}
2080: W := \set{(r,\varphi) \in \frL:\ 0 < r <
2081: d\exp\left(-D\sqrt{|\varphi|}\right)} \subseteq W'
2082: \end{equation*}
2083: and $V := W^{m-1} \times B_\frL(R)^{n+1}$; we claim that $\r(V) \subseteq
2084: U$. To see this, we write $x_k = (r_k,\varphi_k)$ for $k=1, \dots,
2085: m$. Then
2086: \begin{align*}
2087: \left\|\left(\r(x)\right)_m\right\| &\le d^\rho \exp\left( -D
2088: \rho\sqrt{|\varphi_{m-1}|}\right) \cdot 2\lambda \\ &\le c
2089: \exp\left( -D\left( \rho\sqrt{|\varphi_{m-1}|} +
2090: \sqrt{\pi/2} \right)\right).
2091: \end{align*}
2092: Since $|\arg(\t_\lambda(w))| \le \pi/2$ for all $w \in \frL$, we
2093: also get
2094: \begin{align*}
2095: C \sqrt{|\rho\varphi_{m-1} + \arg(\t_\lambda(x_m))|} &\le C\left(
2096: \sqrt\rho \sqrt{|\varphi_{m-1}|} + \sqrt{\pi/2}\right) \\
2097: &\le D \left(\rho \sqrt{|\varphi_{m-1}|} +
2098: \sqrt{\pi/2}\right).
2099: \end{align*}
2100: The claim follows.
2101:
2102:
2103: We write $x' := (x_1, \dots, x_{m-1})$ for $x \in \frL^m$. Since $f
2104: \circ \r$ is holomorphic on $V$, for each $p \in \NN$ and each
2105: $\beta \in \NN^n$ the function $a_{(p,\beta)}:(W')^{m-1} \into \CC$
2106: defined by
2107: \begin{equation*}
2108: a_{(p,\beta)}(x') := \frac1{p! \beta!} \frac {\partial^p((\partial
2109: f/\partial y^\beta) \circ \r)}{\partial
2110: x_m^p} (x',\lambda,0)
2111: \end{equation*}
2112: is holomorphic. Moreover, we put $X' := (X_1, \dots, X_{m-1})$, and
2113: $\alpha' := (\alpha_1, \dots, \alpha_{m-1})$, and we fix $p \in \NN$
2114: and $\beta \in \NN^n$ and write $T\left(\frac {\partial^\beta
2115: f}{\partial y^\beta}(X,0)\right) = \sum a_{\alpha} X^\alpha$.
2116: By the above and Proposition \ref{O-membership}, it now suffices to
2117: show that $a_{(p,\beta)} \in \A((W')^{m-1})$ with $Ta_{(p,\beta)} =
2118: \frac1{\beta!} A_{(p,\beta)}$,
2119: where $$A_{(p,\beta)}(X') := \sum_{\alpha} \begin{pmatrix} \alpha_m \\
2120: p \end{pmatrix} \lambda^{\alpha_m - p} a_{\alpha} (X')^{\alpha'}
2121: X_{m-1}^{\rho\alpha_m}.$$ Let $\a>0$, and choose $\a' >
2122: \a/\min\{\rho,1\}$. Shrinking $W$ and $W'$ if necessary, we may
2123: assume by Lemma \ref{real_constants} that
2124: \begin{equation*}
2125: \left\|\frac {\partial^\beta f}{\partial y^\beta}(x,0) -
2126: \sum_{|\alpha| \le \a'} a_{\alpha}
2127: x^\alpha \right\|
2128: = o\left(\|x\|^{\a'}\right) \quad\text{as } \|x\| \to 0
2129: \text{ in } W^m.
2130: \end{equation*}
2131: Therefore,
2132: \begin{equation}
2133: \label{blow_asymptotic2}
2134: \left\| \left(\frac {\partial^\beta
2135: f}{\partial y^\beta} \circ \r\right)(x,0) -
2136: \sum_{|\alpha| \le \a'} a_{\alpha} \cdot
2137: (\r(x))^{\alpha} \right\| =
2138: o\left(\|\r(x)\|^{\a'}\right)
2139: \end{equation}
2140: as $\|x\| \to 0$ in $(W')^m$. We now define $\rho:[0,\infty)^m
2141: \into [0,\infty)^{m-1}$ by $\rho(\alpha) := (\alpha_1, \dots,
2142: \alpha_{m-2}, \alpha_{m-1}+\rho\alpha_m)$. Note that formally
2143: $A_{(p,\beta)}(X') = \sum_{\alpha} \begin{pmatrix} \alpha_m \\
2144: p \end{pmatrix} \lambda^{\alpha_m - p} a_{\alpha}
2145: (X')^{\rho(\alpha)}$, and for all $x \in (W')^m$ and $\alpha
2146: \in [0,\infty)^m$ that $\r(x)^{\alpha} =
2147: (x')^{\rho(\alpha)} \cdot \sum_{q \in \NN} \begin{pmatrix} \alpha_m
2148: \\ q \end{pmatrix} \lambda^{\alpha_m-q} x_m^q$. Differentiating
2149: \eqref{blow_asymptotic2}, it follows from the Cauchy estimates and
2150: our choice of $\a'$ that
2151: \begin{equation*}
2152: \left\| \beta! \cdot a_{(p,\beta)}(x') - \sum_{|\alpha| \le
2153: \a'} \begin{pmatrix} \alpha_m \\
2154: p \end{pmatrix} \lambda^{\alpha_m-p} a_{\alpha}
2155: (x')^{\rho(\alpha)} \right\| =
2156: o\left(\|x'\|^\a\right)
2157: \end{equation*}
2158: as $\|x'\| \to 0$ in $(W')^{m-1}$. Finally, for $x' \in (W')^{m-1}$ we
2159: have
2160: \begin{multline*}
2161: \left\| \beta! \cdot a_{(p,\beta)}(x') - \sum_{|\rho(\alpha)|
2162: \le
2163: \a} \begin{pmatrix} \alpha_m \\
2164: p \end{pmatrix} \lambda^{\alpha_m-p} a_{\alpha}
2165: (x')^{\rho(\alpha)} \right\| \\ \le \left\| \beta! \cdot
2166: a_{(p,\beta)}(x') - \sum_{|\alpha| \le
2167: \a'} \begin{pmatrix} \alpha_m \\
2168: p \end{pmatrix} \lambda^{\alpha_m-p} a_{\alpha}
2169: (x')^{\rho(\alpha)} \right\| \\ +
2170: \left\|\sum_{\substack{|\alpha| \le
2171: \a' \\ |\rho(\alpha)| > \a}} \begin{pmatrix} \alpha_m \\
2172: p \end{pmatrix} \lambda^{\alpha_m-p} a_{\alpha}
2173: (x')^{\rho(\alpha)} \right\|.
2174: \end{multline*}
2175: The right-hand side above is $o\left(\|x'\|^\a\right)$ as $\|x'\|
2176: \to 0$ in $(W')^{m-1}$, by \eqref{blow_asymptotic2} and because the
2177: sum in the second summand is finite and each of its summands has an
2178: exponent $\beta$ satisfying $|\beta| > \a$, as $|\rho(\alpha)| \ge
2179: |\alpha| \min\{\rho,1\}$ for all $\alpha$. This proves the
2180: proposition.
2181: \end{proof}
2182:
2183:
2184: \section{The class $\Q$}
2185: \label{division}
2186:
2187: Let $m,n \in \NN$, and let $U \subseteq \frL^{m+n}$ be an
2188: $m$-quadratic domain. Below, we let $y = (y_1, \dots, y_n)$ range
2189: over $\frL^n$ and $Y = (Y_1, \dots, Y_n)$ be indeterminates.
2190:
2191: With Remarks \ref{truncation_remark} and \ref{taylor_remark} in mind,
2192: we now restrict our attention to a subclass of $\A_m(U)$.
2193: Abusing notation, we identify $[0,\infty)$ with the set $\{0\} \cup
2194: (0,\infty)_\frL \subseteq \frL_0$.
2195:
2196: \begin{df}
2197: \label{Q}
2198: We define the class $\Q^{m+n}_m(U)$ to be the set of all $f \in
2199: \A_m(U)$ such that for every $\gamma \in [0,\infty)^m$,
2200: \begin{itemize}
2201: \item[(TD)] there are an $m$-quadratic $V = V(f,\gamma)
2202: \subseteq U$ and an $f_{(\gamma,0)} \in \A_m(V)$ such that $T
2203: f_{(\gamma,0)} = (Tf)_{(\gamma,0)}$;
2204: \item[(TE)]for every $\kappa \in [0,\infty)^m$ with $(\kappa,0) \in
2205: \cl_0(U)$, there is an $m$-quadratic $W= W(f_{(\gamma,0)},\kappa)
2206: \subseteq \frL^{m+n}$ such that $(\t_\kappa(x),y) \in V$ for all
2207: $(x,y) \in W$ and the function $\t_{(\kappa,0)} f_{(\gamma,0)} : W
2208: \into \CC$ defined by $(\t_{(\kappa,0)} f_{(\gamma,0)})(x,y) :=
2209: f_{(\gamma,0)}(\t_\kappa(x),y)$ belongs to $\A_m(W)$.
2210: \end{itemize}
2211: \end{df}
2212:
2213: We shall omit the superscript $m+n$ whenever clear from context.
2214:
2215: \begin{nrmks}
2216: \label{quasirmks}
2217: \begin{enumerate}
2218: \item By Proposition \ref{asymptotic_to_0}, for each $f \in \Q_m(U)$
2219: and each $\gamma \in [0,\infty)^m$, the function $f_{(\gamma,0)}$
2220: in the definition above is unique in $\A_m(V(f,\gamma))$.
2221: \item Let $f:U \into \CC$ be holomorphic, and let $V \subseteq U$ be
2222: an $m$-quadratic domain. Then by Remark \ref{O-rmk}(2) and the
2223: above definition, $f\rest{V} \in \Q_m(V)$ iff $f \in \Q_m(U)$.
2224: \item By definition, the collection $\Q_1(U)$ is equal to the set of
2225: all $f \in \A_1(U)$ such that (TD) holds for every $\gamma \in
2226: [0,\infty)^m$; in particular, $\Q^1_1(U) = \A^1_1(U)$. Moreover,
2227: $\A_0(U) = \Q_0(U)$ by Proposition \ref{holo_division} and
2228: Corollary \ref{taylor}.
2229: \item Let $\sigma$ be a permutation of $\{1, \dots, m\}$ and $\tau$
2230: be a permutation of $\{1, \dots, n\}$, and let $f \in \Q_m(U)$.
2231: Then $f \circ \sigma$ and $f \circ \tau$ belong to $\Q_m(U)$.
2232: \end{enumerate}
2233: \end{nrmks}
2234:
2235: For $p \in \NN$ and $q \in \{1, \dots, p\}$, we say that $\rho \in
2236: [0,\infty)^p$ is \textbf{$q$-zero} if $\rho_1 = \cdots = \rho_q =
2237: 0$ and $\rho_{q+1}, \dots, \rho_p > 0$. From Proposition
2238: \ref{prop:characterization}, we obtain:
2239:
2240: \begin{cor}
2241: \label{strongQ}
2242: Let $f \in \Q_m(U)$, and let $\gamma \in [0,\infty)^m$, $k \in \{1,
2243: \dots, m\}$, $\kappa \in [0,\infty)^m$ be $k$-zero such that
2244: $(\kappa,0) \in \cl_0(U)$ and $\sigma$ a permutation of $\{1, \dots,
2245: m\}$. Then there is a $k$-quadratic domain $W \subseteq \frL^{m+n}$
2246: such that $\sigma(\t_\kappa(x),y) \in V(f,\gamma)$ for all $(x,y)
2247: \in W$ and the function $\t_{(\kappa,0)}(f_{(\gamma,0)} \circ
2248: \sigma)$ belongs to $\A_k(W)$. \qed
2249: \end{cor}
2250:
2251: \begin{df}
2252: \label{germs}
2253: Let $\E^{m+n}_m$ be the union of all $\Q_m(U)$ as $U$ ranges over
2254: the $m$-quadratic domains in $\frL^{m+n}$. We define an equivalence
2255: relation $\equiv$ on $\E^{m+n}_m$ as follows: $f \equiv g$ if and
2256: only of there is an $m$-quadratic domain $U \subseteq \frL^{m+n}$
2257: such that $f \rest{U} = g\rest{U}$. We let $\Q^{m+n}_m$ be the set
2258: of all $\equiv$-equivalence classes.
2259: \end{df}
2260:
2261: We shall omit the superscript $m+n$ whenever it is clear from context.
2262: We will not distinguish between $f \in \Q_m(U)$ and its equivalence
2263: class in $\Q_m$, which we also denote by $f$. With this
2264: identification, whenever $U \subseteq \frL^{m+n}$ is an $m$-quadratic
2265: domain, we have $\Q_m(U) \subseteq \Q_m$. Moreover, for every $f,g
2266: \in \E_m$ such that $f \equiv g$, we have $Tf = Tg$; hence, the map $f
2267: \mapsto Tf:\E_m \into \Ps{C}{X^*,Y}$ induces a map $f \mapsto Tf:\Q_m
2268: \into \Ps{C}{X^*,Y}$. Finally, for $r \ge 0$ we simply write $x^r$
2269: for the germ of the function $x \mapsto x^r:\frL \into \CC$.
2270:
2271: \begin{lemma}
2272: \label{Q-addition}
2273: \begin{enumerate}
2274: \item Let $f,g \in \Q_m$ and $a \in \CC$. Then $f+g \in \Q_m$
2275: and $af \in \Q_m$.
2276: \item If $m+n \ge l \ge m$, then $\Q_m \subseteq \Q_l$.
2277: \item Let $f \in \Q_m$ and $(\lambda,\mu) \in (0,\infty)^{m+n}$.
2278: Then the function $$f(\m(\lambda_1,x_1), \dots, \m(\lambda_m,x_m),
2279: \m(\mu_1,y_1), \dots, \m(\mu_n,y_n))$$ belongs to $\Q_m$.
2280: \end{enumerate}
2281: \end{lemma}
2282:
2283: \begin{proof}
2284: (1) Let $\gamma \in [0,\infty)^m$; then $(Tf)_{(\gamma,0)} +
2285: (Tg)_{(\gamma,0)} = (T(f+g))_{(\gamma,0)}$ and $a (Tf)_{(\gamma,0)}
2286: = T(af)_{(\gamma,0)}$. Let also $\kappa \in [0,\infty)^m$ be
2287: sufficiently small. Then $f \circ \t_{(\kappa,0)} + g \circ
2288: \t_{(\kappa,0)} = (f+g) \circ \t_{(\kappa,0)}$ and $a (f \circ
2289: \t_{(\kappa,0)}) = (af) \circ \t_{(\kappa,0)}$, so we define
2290: $\t_{(\kappa,0)}(f+g) := \t_{(\kappa,0)}f + \t_{(\kappa,0)}g$ and
2291: $\t_{(\kappa,0)}(af) := a \t_{(\kappa,0)} f$.
2292:
2293: (2) Let $m+n \ge l \ge m$, and let $f \in \Q_m$ and $\rho \in
2294: [0,\infty)^l$. Let also $\rho'$ be the least $\tau \in [0,\infty)^m
2295: \times \NN^{l-m}$ such that $\tau \ge \rho$. Then we can take
2296: $f_{(\rho,0)} := (x,y)^{\rho'-\rho} \cdot f_{(\rho',0)}$. It
2297: follows easily that $f \in \Q_l$.
2298:
2299: (3) Writing $\m((\lambda,\mu),(x,y)):= (\m(\lambda_1,x_1), \dots,
2300: \m(\mu_n,y_n))$, and writing $(\lambda,\mu) \cdot (X,Y) :=
2301: (\lambda_1 X_1, \dots, \mu_n Y_n)$, we see that
2302: \begin{equation*}
2303: Tf((\lambda,\mu) \cdot (X,Y))_{(\gamma,0)} =
2304: (Tf)_{(\gamma,0)}((\lambda,\mu) \cdot (X,Y))
2305: \end{equation*}
2306: for all $\gamma \in [0,\infty)^m$, and that
2307: \begin{equation*}
2308: \t_{(\kappa,0)}(f(\m((\lambda,\mu),(x,y)))) =
2309: (\t_{(\lambda\kappa,0)} f)(\m((\lambda,\mu),(x,y)))
2310: \end{equation*}
2311: for all sufficiently small $(x,y) \in \frL^{m+n}$, where
2312: $\lambda\kappa := (\lambda_1\kappa_1, \dots, \lambda_m\kappa_m)$.
2313: Part (3) follows.
2314: \end{proof}
2315:
2316: \begin{prop}
2317: \label{Q-basics}
2318: Let $f \in \Q_m$.
2319: \begin{enumerate}
2320: \item For every elementary set $S \subseteq [0,\infty)^m \times
2321: \NN^n$, there is a unique $f_S \in \Q_m$ such that $T(f_S) =
2322: (Tf)_S$.
2323: \item For every $k \in \{1, \dots, m\}$, every sufficiently small
2324: $k$-zero $(\kappa,\lambda) \in [0,\infty)^m \times \RR^n$ and
2325: every permutation $\sigma$ of $\{1, \dots, m\}$, the function
2326: $\t^\sigma_{(\kappa,\lambda)} f := f \circ \sigma \circ
2327: \t_{(\kappa,\lambda)}$ belongs to $\Q_k$.
2328: \end{enumerate}
2329: \end{prop}
2330:
2331: \begin{proof}
2332: (1) By Remark \ref{boolean} and Lemmas \ref{natural_elementary} and
2333: \ref{Q-addition}, it suffices to consider $S = \set{((\alpha,\beta)
2334: \in [0,\infty)^m \times \NN^n:\ (\alpha,\beta) \ge
2335: (\gamma,\delta)}$ for some $(\gamma,\delta) \in [0,\infty)^m
2336: \times \NN^n$. By Lemma \ref{Q-addition} and Proposition
2337: \ref{asymptotic_to_0}, we may even assume that either $\gamma = 0$
2338: or $\delta = 0$. We assume first that $\delta = 0$ and let
2339: $f_{(\gamma,0)}$ be as in (TD); we need to show that $f_{(\gamma,0)}
2340: \in \Q_m$. So let $\gamma' \in [0,\infty)^m$ and $\kappa \in
2341: [0,\infty)^m$ be sufficiently small. Since
2342: $(Tf)_{(\gamma+\gamma',0)} = (Tf_{(\gamma,0)})_{(\gamma',0)}$, we
2343: can take $(f_{(\gamma,0)})_{(\gamma',0)} := f_{(\gamma+\gamma',0)}$
2344: and $\t_{(\kappa,0)}\big((f_{(\gamma,0)})_{(\gamma',0)}\big)
2345: := \t_{(\kappa,0)} f_{(\gamma+\gamma',0)}$. Second, the case
2346: $\gamma = 0$ follows from Proposition \ref{holo_division} and
2347: Corollary \ref{taylor}.
2348:
2349: (2) Let $k \in \{1, \dots, m\}$ and $(\kappa,\lambda) \in
2350: [0,\infty)^m \times \RR^n$ be sufficiently small and $k$-zero; by
2351: Remark \ref{quasirmks}(4), it suffices to prove that
2352: $\t_{(\kappa,\lambda)} f$ belongs to $\Q_k$. Since
2353: $\t_{(\kappa,\lambda)} f = \t_{(\kappa,0)}(\t_{(0,\lambda)} f)$, we
2354: may assume by Corollary \ref{taylor} that $\lambda = 0$. Let
2355: $\gamma \in [0,\infty)^k \times \{0\}^{m-k}$. By (1), there is for
2356: each $I \subseteq \{1, \dots, m\}$ and each $\alpha \in B_{\gamma,I}
2357: = B_{\gamma,I}(Tf)$ a unique $f_{\gamma,I, \alpha} \in \Q_m$ such
2358: that
2359: \begin{equation*}
2360: f = \sum_{I \subseteq \{1, \dots, m\}} x_{\bar
2361: I}^{\gamma_{\bar I}} \left( \sum_{\alpha \in B_{\gamma,I}}
2362: x_I^\alpha \cdot f_{\gamma,I,\alpha} \right)
2363: \end{equation*}
2364: and each $f_{\gamma,I,\alpha}$ depends only on the variables
2365: $x_{\bar I}$ and $y$. Since $\gamma_{k+1} = \cdots = \gamma_m = 0$,
2366: we have $B_{\gamma,I} = \emptyset$ whenever $I \nsubseteq \{1,
2367: \dots, k\}$. Therefore,
2368: \begin{equation*}
2369: \t_{(\kappa,0)} f = \sum_{I \subseteq \{1, \dots, m\}} x_{\bar
2370: I}^{\gamma_{\bar I}} \left( \sum_{\alpha \in B_{\gamma,I}}
2371: x_I^\alpha \cdot \t_{(\kappa,0)} f_{\gamma,I,\alpha} \right),
2372: \end{equation*}
2373: and hence $T(\t_{(\kappa,0)} f) = \sum_I X_{\bar I}^{\gamma_{\bar
2374: I}} \left( \sum_\alpha X_I^\alpha \cdot T(\t_{(\kappa,0)}
2375: f_{\gamma,I,\alpha}) \right)$ is the unique
2376: $\gamma$-representation of $T(\t_{(\kappa,0)} f)$. Since
2377: $f_{\gamma,\emptyset,0} = f_{(\gamma,0)}$, it follows that we can
2378: take $(\t_{(\kappa,0)} f)_{(\gamma,0)} := \t_{(\kappa,0)}
2379: f_{(\gamma,0)}$. Moreover, if $\kappa' \in [0,\infty)^k \times
2380: \{0\}^{m-k}$ is sufficiently small, then
2381: $$\t_{(\kappa',0)}((\t_{(\kappa,0)} f)_{(\gamma,0)}) =
2382: \t_{(\kappa',0)}(\t_{(\kappa,0)} f_{(\gamma,0)}) =
2383: \t_{(\kappa+\kappa',0)} f_{(\gamma,0)},$$ so part (2) follows.
2384: \end{proof}
2385:
2386: From Proposition \ref{Q-basics} and Lemma \ref{lemma:representation},
2387: we obtain:
2388:
2389: \begin{cor}
2390: \label{cor:Q-basics}
2391: Let $f \in \Q_m$ and $\gamma \in [0,\infty)^m$. Then for each
2392: $I \subset \{1, \dots, m\}$ and each $\alpha \in B_{\gamma,I} =
2393: B_{\gamma,I}(Tf)$, there is a unique $f_{\gamma,I,\alpha} \in \Q_m$
2394: such that
2395: \begin{equation*}
2396: f = \sum_{I \subset \{1, \dots, m\}} x_{\bar I}^{\gamma_{\bar
2397: I}} \left( \sum_{\alpha \in B_{\gamma,I}} x_I^\alpha \cdot
2398: f_{\gamma,I,\alpha} \right)
2399: \end{equation*}
2400: and each $f_{\gamma,I,\alpha}$ depends only on the variables
2401: $x_{\bar I}$ and $y$. \qed
2402: \end{cor}
2403:
2404: \begin{prop}
2405: \label{Q-derivatives}
2406: Let $f \in \Q_m$.
2407: \begin{enumerate}
2408: \item For each $i=1, \dots, m$, the function $\partial_i f$ belongs to
2409: $\Q_m$ and satisfies $T(\partial_i f) = \partial_i(Tf)$.
2410: \item For each $j = 1, \dots, n$, the function $\partial f/\partial
2411: y_j$ belongs to $\Q_m$ and satisfies $T(\partial f/\partial y_j)
2412: = \partial (Tf)/\partial Y_j$.
2413: \item The function $g := f(x,y_1, \dots, y_{n-1},0)$ belongs to
2414: $\Q^{m+n-1}_m$.
2415: \end{enumerate}
2416: \end{prop}
2417:
2418: \begin{proof}
2419: Let $\gamma \in [0,\infty)^m$. It follows from Propositions
2420: \ref{prop:bad_derivatives} and \ref{Q-basics} and Lemmas
2421: \ref{derivation_truncation} and \ref{Q-addition} that $(\partial_i
2422: f)_{(\gamma,0)} := \gamma_i \cdot f_{(\gamma,0)}
2423: + \partial_i(f_{(\gamma,0)})$ belongs to $\A_m(V)$ for a suitable $V
2424: \subseteq \frL^{m+n}$ and satisfies $T((\partial_i f)_{(\gamma,0)})
2425: = \partial_i((Tf)_{(\gamma,0)})$. Moreover, we let $k \in \{1,
2426: \dots, m\}$ and a $k$-zero $\kappa \in [0,\infty)^m$ be sufficiently
2427: small. If $i \le k$, then $\t_{(\kappa,0)}(\partial_i f_{(\gamma,0)})
2428: = \partial_i(\t_{(\kappa,0)} f_{(\gamma,0)})$ by Remark
2429: \ref{translate_derivative}. If $i > k$, then $\t_{(\kappa,0)}
2430: f_{(\gamma,0)}$ belongs to $\Q_k$ by Proposition \ref{Q-basics}(2),
2431: and it follows from Corollary \ref{good_derivatives} and Remark
2432: \ref{translate_derivative} that
2433: \begin{equation*}
2434: x_i^1 \cdot \frac {\partial}{\partial x_i} (\t_{(\kappa,0)}
2435: f_{(\gamma,0)}) = \partial_i (\t_{(\kappa,0)} f_{(\gamma,0)}) =
2436: \frac {x_i^1} {\kappa_i + x_i^1} \cdot \t_{(\kappa,0)} (\partial_i
2437: f_{(\gamma,0)}).
2438: \end{equation*}
2439: Therefore, $\t_{(\kappa,0)} (\partial_i f_{(\gamma,0)})$ belongs to
2440: $\A_m(V)$ for some suitable $V$ with $T(\t_{(\kappa,0)} (\partial_i
2441: f_{(\gamma,0)})) = (\kappa_i + X_i) \cdot (\partial/\partial
2442: X_i)(T(\t_{(\kappa,0)} f_{(\gamma,0)}))$, which proves part (1).
2443:
2444: Part (2) is more straightforward and follows from Corollary
2445: \ref{good_derivatives}, Proposition \ref{Q-basics} and Lemmas
2446: \ref{derivation_truncation} and \ref{Q-addition}. For (3), note
2447: that for all $\gamma \in [0,\infty)^m$, we can take $g_{(\gamma,0)}
2448: := f_{(\gamma,0)}(x,y_1, \dots, y_{n-1},0)$. Then for every
2449: sufficiently small $\kappa \in [0,\infty)^m$, we have
2450: $$\t_{(\kappa,0)} g_{(\gamma,0)} = (\t_{(\kappa,0)}
2451: f_{(\gamma,0)})(x,y_1, \dots, y_{n-1},0),$$ and part (3) follows.
2452: \end{proof}
2453:
2454:
2455: \subsection*{Composition}
2456:
2457: Let $f \in \Q_m(U)$. For the next lemma, we let $\q = (\q_1, \dots,
2458: \q_n) \in \NN^n$ and put $\k := |\q|$. We let $z = (z_1, \dots,
2459: z_\k)$ range over $\frL^\k$ and $U' \subseteq \frL^{m+\k}$ be an
2460: $m$-quadratic domain such that $(x, z_1 + \cdots + z_{\q_1}, \dots,
2461: z_{\q_1 + \cdots + \q_{n-1}+1} + \cdots + z_\k) \in U$ for all $(x,z)
2462: \in U'$. In this situation, we define the holomorphic function $f_\q:
2463: U' \into \CC$ by $$f_\q(x,z) := f(x, z_1 + \cdots + z_{\q_1}, \dots,
2464: z_{\q_1 + \cdots + \q_{n-1}+1} + \cdots + z_\k).$$
2465:
2466: \begin{lemma}
2467: \label{sum_composition}
2468: We have $f_\q \in \Q_m(U')$ and $T(f_\q) = (Tf)_\q$.
2469: \end{lemma}
2470:
2471: \begin{proof}
2472: We first show that $f_\q \in \A_m(U')$ and $T(f_\q) = (Tf)_\q$.
2473: Arguing by induction on $\k$ (simultaneously for all $m$) and
2474: permuting the last $n$ coordinates if necessary, it suffices to
2475: consider the case where $n = 1$ and $\k = \q_1 = 2$. In this
2476: situation, by Proposition \ref{coefficients} and after shrinking $U$
2477: if necessary, we can write $f(x,y) = \sum_{p \in \NN} a_p(x) y^p$
2478: for all $(x,y) \in U$, and there are a quadratic domain $W \subseteq
2479: \frL$ and constants $A,B>0$ such that $a_p \in \A_m(W^m)$ and
2480: $\|a_p(x)\| \le AB^p$ for all $x \in W^m$ and each $p \in \NN$.
2481: Hence
2482: \begin{equation*}
2483: f_\q(x,z) = \sum_{p \in \NN} a_p(x)
2484: (z_1+z_2)^p = \sum_{p,q \in
2485: \NN} b_{p,q}(x) z_1^p z_2^q
2486: \end{equation*}
2487: for all sufficiently small $(x,z) \in U'$, where $b_{p,q} :=
2488: \begin{pmatrix} p+q \\ p \end{pmatrix} a_{p+q}$ for all $p,q \in
2489: \NN$. Since $\|b_{p,q}\| \le A (2B)^{p+q}$, it follows from
2490: Propositions \ref{coefficients} and \ref{O-membership} that $f_\q
2491: \in \A_m(U')$, as required.
2492:
2493: Next, for every $\gamma \in [0,\infty)^m$, we have
2494: $T(f_{(\gamma,0)})_\q) = (T f_\q)_{(\gamma,0)}$, so we can take
2495: $(f_\q)_{(\gamma,0)} := (f_{(\gamma,0)})_\q$. Moreover, for every
2496: sufficiently small $\kappa \in [0,\infty)^m$, the previous paragraph
2497: and Proposition \ref{O-membership} now also show that
2498: $\t_{(\kappa,0)} (f_\q)_{(\gamma,0)}$ belongs to $\A_m(V')$ for some
2499: appropriate $V'$.
2500: \end{proof}
2501:
2502: For the next proposition, we let $g = (g_1, \dots, g_n) \in \Q_m(V)^n$
2503: be such that $g(0) = 0$ and $(x, g(x,y)) \in U$ for all $(x,y) \in V$,
2504: and we define the holomorphic function $h:V \into \CC$ by $ h(x,y) :=
2505: f(x,g(x,y)). $
2506:
2507: \begin{prop}
2508: \label{Q-composition}
2509: The function $h$ belongs to $\Q_m(V)$ and $Th(X,Y) =
2510: Tf(X,Tg(X,Y))$.
2511: \end{prop}
2512:
2513: \begin{proof}
2514: First, let $\kappa \in [0,\infty)^m$ be such that $(\kappa,0) \in
2515: \cl_0(V)$. Then $\t_{(\kappa,0)} h =
2516: (\t_{(\kappa,0)}f)(x,\t_{(\kappa,0)} g)$, so Corollary \ref{taylor}
2517: (with $\lambda$ there equal to $\t_{(\kappa,0)}g(0,0)$) and
2518: Proposition \ref{A-composition} show that $\t_{(\kappa,0)}h \in
2519: \A_m(W)$ for some appropriate $W$.
2520:
2521: Second, let $\gamma \in [0,\infty)^m$; we need to find an
2522: $m$-quadratic domain $V' \subseteq V$ and an $h' \in \A_m(V')$ such
2523: that $T(h') = (Th)_{(\gamma,0)}$. By Proposition
2524: \ref{comp_representation}, there are $p \in \NN$, a tuple $\q \in
2525: \NN^n$ and, with $\k:= |\q|$, elementary sets $E_1, \dots, E_p
2526: \subseteq [0,\infty)^m \times \NN^\k$ and $B_{i,j} \subseteq
2527: [0,\infty)^m \times \NN^n$ for each pair $(i,j)$ satisfying $i \in
2528: \{1, \dots, n\}$ and $j \in \{1, \dots, \q_i\}$, such that
2529: \begin{equation*}
2530: Tf(X,Tg)_{(\gamma,0)} = \sum_{q=1}^p \frac {(X,(Tg)_B)^{\inf
2531: E_q}}{X^\gamma} \cdot ((Tf)_\q)_{E_q}(X,(Tg)_B)
2532: \end{equation*}
2533: with $(Tg)_B := ((Tg_1)_{B_{1,1}}, \dots, (Tg_n)_{B_{n,\q_n}})$ and
2534: each $(X,(Tg)_B)^{\inf E_q}$ divisible by $X^\gamma$. After
2535: shrinking $V$ if necessary and writing $g_B := ((g_1)_{B_{1,1}},
2536: \dots, (g_n)_{B_{n,\q_n}})$, we get from Lemma
2537: \ref{sum_composition}, Proposition \ref{Q-basics} and the above
2538: that, for each $q =1 , \dots, p$, the function $$h_q := x^{-\gamma}
2539: \cdot (x,g_B)^{\inf E_q} \cdot (f_\q)_{E_q}(x,g_B)$$ belongs to
2540: $\A_m(V)$ and satisfies $$T h_\q = X^{-\gamma} \cdot
2541: (X,(Tg)_B)^{\inf E_q} \cdot ((Tf)_\q)_{E_q}(X,(Tg)_B).$$ Hence by
2542: Lemma \ref{Q-addition}, we can take $h' := h_1 + \cdots + h_p$.
2543:
2544: Finally, it follows from the last paragraph and the first
2545: observation above that $h \in \Q_m$.
2546: \end{proof}
2547:
2548: Here are some immediate applications of Proposition \ref{Q-composition}:
2549:
2550: \begin{prop}
2551: \label{Q-algebra}
2552: The set $\Q_m$ is a $\CC$-algebra, and the map $f \mapsto Tf: \Q_m
2553: \longrightarrow \Ps{C}{X^*,Y}$ is an injective $\CC$-algebra
2554: homomorphism such that $f(0) = (Tf)(0)$ for all $f \in \Q_m$.
2555: \end{prop}
2556:
2557: \begin{proof}
2558: Let $f,g \in \Q_m$; we need to show that $fg \in \Q_m$. Put $f_1 :=
2559: f - f(0)$ and $g_1 := g - g(0)$; then $f_1, g_1 \in \Q_m$ by Lemma
2560: \ref{Q-addition}, and $fg = P(f_1,g_1)$ with $P(Y_1,Y_2) := (f(0) +
2561: Y_1) (g(0) + Y_2)$. Hence $fg \in \Q_m$ by Proposition
2562: \ref{Q-composition}.
2563: \end{proof}
2564:
2565: \begin{prop}
2566: \label{Q-units}
2567: Let $f \in \Q_m$. Then
2568: \begin{enumerate}
2569: \item $f$ is a unit in $\Q_m$ if and only if $f(0) \ne 0$;
2570: \item if $f(0) = 0$, then there are $(\gamma,\delta) \in
2571: (0,\infty)^m \times (\NN \setminus \{0\})^n$ and $f_1, \dots,
2572: f_{m+n} \in \Q_m$ such that $f = X_1^{\gamma_1} f_1 + \cdots +
2573: Y_n^{\delta_n} f_{m+n}$.
2574: \end{enumerate}
2575: \end{prop}
2576:
2577: \begin{proof}
2578: (1) Assume first that $f$ is a unit in $\Q_m$, and let $g \in Q_m$
2579: be such that $f\cdot g = 1$. Then $Tf \cdot Th = 1$ and hence $f(0)
2580: = Tf(0) \ne 0$. Conversely, assume that $f(0) \ne 0$; we may assume
2581: that $f(0) = 1$, and we put $f_1:= 1 - f \in \Q_m$. Let $\a>0$ and
2582: $U_\a \subseteq \frL^{m+n}$ be an $m$-quadratic domain such that
2583: $f_1(x,y) = o(\|(x,y)\|^{\a})$ as $\|(x,y)\| \to 0$ in $U_{\a}$.
2584: Thus, there is an $m$-quadratic domain $U \subseteq U_{\a}$ such
2585: that $\|f_1(x,y)\| \leq \frac{1}{2}$ for all $(x,y) \in U$. Let
2586: $\phi: B(0,1) \into \CC$ be the holomorphic function defined by
2587: $\phi(z) := \frac{1}{1-z}$, and define $g:U \into \CC$ by $g(x,y) :=
2588: \phi(f_1(x,y))$. Then $f \cdot g = 1$, and $g \in \Q_m(U)$ by
2589: Proposition \ref{Q-composition}.
2590:
2591: (2) follows from Lemma 4.8 of \cite{vdd-spe:genpower} and
2592: Proposition \ref{Q-basics}.
2593: \end{proof}
2594:
2595: Finally, Proposition \ref{Q-composition} allows us to make sense of
2596: certain substitutions in the $x$-variables:
2597:
2598: \begin{df}
2599: \label{x-substitution}
2600: Let $W \subseteq \frL$ be a quadratic domain and $R>0$, and let $f
2601: \in \Q_m(W^m \times B_\frL(R))$. Let also $V \subseteq \frL^{m+n}$ be
2602: $m$-quadratic and $g = (g_1, \dots, g_m) \in \Q_m(V)^m$ be such that
2603: $\lambda := g(0,0) \in W^m \cap (0,\infty)^m$. Then $g(x,y) =
2604: \lambda + h(x,y)$ with $h \in \Q_m(V)^m$ satisfying $h(0) = 0$, and
2605: we define $f(g(x,y),y) := (\t_{(\lambda,0)} f) (h(x,y),y)$.
2606: \end{df}
2607:
2608: \begin{cor}
2609: \label{bad_substitution_1}
2610: The function $f(g(x,y),y)$ in Definition \ref{x-substitution}
2611: belongs to $\Q_m$. \qed
2612: \end{cor}
2613:
2614: Some of the substitutions not covered by the previous corollary are
2615: the blow-up substitutions:
2616:
2617: \begin{prop}
2618: \label{blowup_homom}
2619: Let $\rho,\lambda > 0$ and $i,j \in \{1, \dots, m\}$ be distinct.
2620: \begin{enumerate}
2621: \item The function $f \circ \s^\rho_{i,j}$ belongs to $\Q_m$ for
2622: every $f \in \Q_m$, and the map $\s^\rho_{ij}:\Q_m \into \Q_m$
2623: defined by $\s^\rho_{ij}(f) := f \circ \s^\rho_{ij}$ is a
2624: $\CC$-algebra homomorphism such that $T \circ \s^\rho_{ij} =
2625: \frB^{\rho,0}_{ij} \circ T$.
2626: \item The function $f \circ \r^{\rho,\lambda}$ belong to $\Q_{m-1}$
2627: for every $f \in \Q_m$, and the map $\r^{\rho,\lambda}:\Q_m \into
2628: \Q_{m-1}$ defined by $\r^{\rho,\lambda}(f) := f \circ
2629: \r^{\rho,\lambda}$ is a $\CC$-algebra homomorphism such that $T
2630: \circ \r^{\rho,\lambda} = \frB^{\rho,\lambda}_{m,m-1} \circ T$.
2631: \end{enumerate}
2632: \end{prop}
2633:
2634: Whenever convenient, we shall write $\s^\rho_{ij} f$ and
2635: $\r^{\rho,\lambda} f$ in place of $\s^\rho_{ij}(f)$ and
2636: $\r^{\rho,\lambda}(f)$.
2637:
2638: \begin{proof}
2639: The proofs for parts (1) and (2) are similar; we prove (1) here and
2640: leave (2) to the reader. We may assume that $i=m$ and $j=m-1$, and
2641: we write $\s$ and $\B$ in place of $\s^\rho_{m,m-1}$ and
2642: $\B^\rho_{m,m-1}$. Let $f \in \Q_m$; if suffices to prove that $f
2643: \circ \s \in \Q_m$.
2644:
2645: To do so, we let $W \subseteq \frL$ be quadratic and $1 > R>0$ be
2646: such that $f \in \A_m(W^m \times B_\frL(R)^n)$, and we let $W'$ and $V$
2647: be as in the proof of Proposition \ref{singular_prop}. We also let
2648: $\kappa \in [0,\infty)^m$ be nonzero such that $(\kappa,0) \in
2649: \cl_0(V)$, and let $W_\kappa \subseteq W'$ be quadratic such that
2650: $\t_{(\kappa,0)}(x,y) \in V$ for all $(x,y) \in V_\kappa :=
2651: (W_\kappa)^m \times B_\frL(R)^n$. By Propositions
2652: \ref{blowup_truncation} and \ref{Q-basics}, it remains to prove that
2653: \begin{itemize}
2654: \item[$(\ast)$] $\t_{(\kappa,0)}(f \circ \s)$ belongs to
2655: $\A_m(V_\kappa)$.
2656: \end{itemize}
2657: Writing $\kappa'\:= (\kappa_1, \dots, \kappa_{m-2})$ and $\kappa'':=
2658: (\kappa_{m-1}, \kappa_m)$, we see that $\t_{(\kappa,0)}(f \circ \s)
2659: = \t_{(0,\kappa'',0)}( \t_{(\kappa',0,0)}(f \circ \s))$; since
2660: $\t_{(\kappa',0,0)}(f \circ \s) = (\t_{(\kappa',0,0)} f) \circ \s$,
2661: we may even assume that $\kappa_1 = \cdots = \kappa_{m-2} = 0$. We
2662: now distinguish three cases: \medskip
2663:
2664: \noindent\textbf{Case 1:} both $\kappa_{m-1}$ and $\kappa_m$ are
2665: nonzero. Then
2666: \begin{equation*}
2667: \t_{(\kappa,0)}(f \circ \s)(x,y) =
2668: (\t_{(0,\kappa_{m-1},\kappa_{m-1}^\rho \kappa_m,0)}
2669: f)(x',g(x_{m-1}, x_m),y),
2670: \end{equation*}
2671: where $x' := (x_1, \dots, x_{m-1})$ and $g$ is an analytic function
2672: satisfying $g(0) = 0$. Since $\t_{(0,\kappa_{m-1},\kappa_{m-1}^\rho
2673: \kappa_m,0)} f$ belongs to $\Q_{m-2}$ by Proposition
2674: \ref{Q-basics}, $(\ast)$ follows from Proposition
2675: \ref{Q-composition} in this case. \medskip
2676:
2677: \noindent\textbf{Case 2:} $\kappa_{m-1} = 0$ and $\kappa_m > 0$.
2678: Then $\t_{(\kappa,0)} (f \circ \s) = f \circ \r^{\rho,\kappa_m}$,
2679: so $(\ast)$ follows from Proposition \ref{regular_prop} in this
2680: case.
2681: \medskip
2682:
2683: \noindent\textbf{Case 3:} $\kappa_{m-1} \ne 0$ and $\kappa_m = 0$.
2684: We define $\phi(z_1, \dots, z_{m+1},y) := \t_{(\kappa,0)} f(z_1,
2685: \dots, z_{m-1},z_{m+1},y)$. Then $\phi \in \Q_{m+1}$, and there is
2686: an analytic one-variable function $g$ with $g(0) = 0$ such that
2687: \begin{equation*}
2688: \t_{(\kappa,0)} (f \circ \s)(x,y) = \left(\phi \circ
2689: \r^{1,\kappa_{m-1}^\rho}\right) (x, g(x_{m-1}),y).
2690: \end{equation*}
2691: Thus, $(\ast)$ follows from Propositions \ref{regular_prop} and
2692: \ref{A-composition} in this case.
2693: \end{proof}
2694:
2695: As a consequence of Proposition \ref{blowup_homom}, we extend
2696: Corollary \ref{bad_substitution_1} to certain functions with zero
2697: constant coefficient:
2698:
2699: \begin{df}
2700: \label{x-substitution_2}
2701: Let $m \ge 1$, let $W \subseteq \frL$ be a quadratic domain and
2702: $R>0$, and let $f \in \Q_m(W^m \times B_\frL(R)^n)$. Let also $V
2703: \subseteq \frL$ be a quadratic domain and $g \in Q_1(V)$ be such
2704: that $g(t) \in W \cap (0,\infty)$ for all $t \in V \cap (0,\infty)$.
2705: Then $g(t) = t^\rho(\lambda + h(t))$ for some $\rho, \lambda > 0$
2706: and some $h \in \Q_1(V)$ with $h(0) = 0$. We write $x' := (x_1,
2707: \dots, x_{m-1})$ and let $\tilde{f} \in \Q_{m+1}(W^{m+1} \times
2708: B_\frL(R)^n)$ be the function defined by $\tilde{f}(x',u,v,y) :=
2709: f(x',v,y)$. Then $\r^{\rho,\lambda} \tilde{f} \in \Q_m^{m+n+1}$,
2710: and we define
2711: $$f(x',g(t),y) := \left(\r^{\rho,\lambda} \tilde{f}\right)
2712: (x',t,h(t),y).$$
2713: \end{df}
2714:
2715: \begin{cor}
2716: \label{bad_substitution_2}
2717: The function $f(x',g(t),y)$ in Definition \ref{x-substitution_2}
2718: belongs to $\Q_m^{m+n}$. \qed
2719: \end{cor}
2720:
2721:
2722: \section{Weierstrass Preparation} \label{weierstrass}
2723:
2724: We continue to work in the setting of the previous section. In this
2725: section, we establish a Weierstrass Preparation Theorem for the
2726: classes $\Q_m$. We follow Brieskorn and Kn\"orrer's exposition in
2727: Section 8.2 of \cite{bri-kno:plane}; to do so, we need to first
2728: establish an implicit function theorem and a theorem on symmetric
2729: functions. We thank Lou van den Dries for his helpfull suggestions on
2730: this section, especially the proof of Corollary \ref{multi_implicit}
2731: below.
2732:
2733: We start with a single implicit variable and write $y' :=
2734: (y_1, \dots, y_{n-1})$ and $Y' := (Y_1, \dots, Y_{n-1})$.
2735:
2736: \begin{prop}
2737: \label{implicit}
2738: Let $f \in \Q_m^{m+n}$, and assume that $f(0) = 0$ and $\partial
2739: f/\partial y_n(0) \ne 0$. Then there is an $h \in \Q_m^{m+n-1}$ such
2740: that $h(0) = 0$ and $f(x,y',h) = 0$.
2741: \end{prop}
2742:
2743: \begin{proof}
2744: We let $U = W^m \times B_\frL(R)^n$ for some quadratic $W \subseteq \frL$
2745: and $R>0$ be such that $f \in \Q_m(U)$, and we put $V := W^m \times
2746: B_\frL(R)^{n-1}$. During the proof below, we may have to shrink $W$ and
2747: $R$ (and all related quantities introduced below) on various
2748: occasions; we will not explicitely mention this. By Proposition
2749: \ref{Q-derivatives}, the function $(\partial f/\partial
2750: y_n)(x,y',0)$ belongs to $\Q_m(V)$. Hence by hypothesis, there is a
2751: constant $c>0$ such that $|(\partial f/\partial y_n)(x,y',0)| \ge c$
2752: for all $(x,y') \in V$. On the other hand, by Proposition
2753: \ref{coefficients}, we can write $f(x,y) = \sum_{p \in \NN} a_{p}(x,y')
2754: y_n^p$ with each $a_{p} \in \O_m(V)$.
2755:
2756: Define $g:U \into \CC$ by $g(x,y):= f(x,y) - a_0(x,y')$; note that
2757: $g \in \O_m(U)$. By the above and the usual arguments for the
2758: inverse function theorem, there is a $\rho>0$ such that for every
2759: $(x,y') \in V$, we have $\|a_0(x,y')\| \le \rho/2$ and the function
2760: $g_{x,y'}:B_\frL(R) \into \CC$ defined by $g_{x,y'}(y_n) := g(x,y)$ is
2761: injective and satisfies $g_{x,y'}(0) = 0$ and $B(0,\rho) \subseteq
2762: g_{x,y'}(B_\frL(R))$, and such that its compositional inverse
2763: $g_{x,y'}^{-1}:B_\frL(\rho) \into \CC$ is given by a convergent power
2764: series
2765: \begin{equation*}
2766: g^{-1}_{x,y'}(z) = \sum_{p \in \NN} b_p(x,y') y_n^p.
2767: \end{equation*}
2768:
2769: We claim that the function $H:V \times B_\frL(\rho/2) \into \CC$ defined
2770: by $H(x,y) := g_{x,y'}^{-1}(y_n)$ belongs to $\Q_m(V \times
2771: B_\frL(\rho/2))$. The proposition follows from this claim by defining
2772: $h:V \into \CC$ as $h(x,y') := H(x,y',y_n-a_0(x,y'))\rest{y_n=0}$.
2773:
2774: To see the claim, we note first from the Lagrange inversion formula
2775: (see for instance Whittaker and Watson
2776: \cite[p. 133]{whi-wat:analysis}) that for all $p \in \NN$,
2777: \begin{equation*}
2778: b_p(x,y') = \frac1{p!} \left[ \frac{\partial^{p-1}}{\partial y_n^{p-1}}
2779: \left( \frac {y_n}{g(x,y)} \right)^p \right]_{y_n=0}.
2780: \end{equation*}
2781: Since $a_1(0) \ne 0$ by hypothesis, it follows from Propositions
2782: \ref{Q-units}, \ref{Q-algebra} and \ref{Q-derivatives} and Remark
2783: \ref{quasirmks}(2) that $b_p \in \Q_m(V)$ for each $p$.
2784:
2785: Second, we let $\gamma \in [0,\infty)^m$, and we claim that
2786: $H_{(\gamma,0)} \in \A_m(V \times B_\frL(\rho/2))$. It suffices, by
2787: Propositions \ref{O-membership} and \ref{holo_division}, to find
2788: constants $A,B > 0$ such that $\|(b_p)_{(\gamma,0)}(x,y')\| \le A
2789: B^p$ for all $p \in \NN$ and $(x,y') \in V$. To do so, we shall
2790: assume that $U \subseteq B_\frL(1)^{m+n}$, and we put $\g(x,y) :=
2791: y_n/g(x,y) \in \Q_m(U)$. We let $\J$ be the set of all ordered
2792: pairs $(I,\alpha)$ such that $I \subseteq \{1, \dots, m\}$ and
2793: $\alpha \in B_I = B_{\gamma,I}(T\mathfrak{g})$, where the latter is
2794: defined as in Lemma \ref{lemma:representation}. By Corollary
2795: \ref{cor:Q-basics}, there is a constant $C>0$ and for each
2796: $(I,\alpha) \in \J$ a function $\g_{I,\alpha} \in \Q_m(U)$,
2797: depending only on the variables $x_{\bar I}$ and $y$, such that
2798: \begin{equation*}
2799: \g(x,y) = \sum_{(I,\alpha) \in \J} x_{\bar I}^{\gamma_{\bar I}}
2800: x_I^\alpha \g_{I,\alpha}(x,y)
2801: \end{equation*}
2802: and $\|\g_{I,\alpha}(x,y)\| \le C$ for all $(x,y) \in U$. We fix $p
2803: \in \NN$ and write $\g^p(x,y) := (\g(x,y))^p$ for all $(x,y) \in U$.
2804: Since
2805: \begin{multline*}
2806: \g^p(x,y) = \\ \sum_{((I_1,\alpha_1), \dots, (I_p, \alpha_p))
2807: \in \J^p} x_{I_1}^{\alpha_1} \cdots x_{I_p}^{\alpha_p} \cdot
2808: x_{\bar{I_1}}^{\gamma_{\bar{I_1}}} \cdots
2809: x_{\bar{I_p}}^{\gamma_{\bar{I_p}}} \cdot
2810: \g_{I_1,\alpha_1}(x,y) \cdots
2811: \g_{I_p,\alpha_p}(x,y),
2812: \end{multline*}
2813: and since each $\mathfrak{g}_{I,\alpha}$ only depends on the
2814: variables $x_{\bar I}$ and $y$, we get that
2815: \begin{multline*}
2816: (\g^p)_{(\gamma,0)}(x,y) = \\ \sum_{((I_1,\alpha_1), \dots, (I_p,
2817: \alpha_p)) \in \J_p} \frac {x_{I_1}^{\alpha_1} \cdots
2818: x_{I_p}^{\alpha_p} \cdot x_{\bar{I_1}}^{\gamma_{\bar{I_1}}}
2819: \cdots x_{\bar{I_p}}^{\gamma_{\bar{I_p}}}} {x^\gamma} \cdot
2820: \g_{I_1,\alpha_1}(x,y) \cdots
2821: \g_{I_p,\alpha_p}(x,y),
2822: \end{multline*}
2823: where $\J_p$ is the set of all $((I_1, \alpha_1), \dots,
2824: (I_p,\alpha_p)) \in \J^p$ such that the monomial $x_{I_1}^{\alpha_1}
2825: \cdots x_{I_p}^{\alpha_p} \cdot x_{\bar{I_1}}^{\gamma_{\bar{I_1}}}
2826: \cdots x_{\bar{I_p}}^{\gamma_{\bar{I_p}}}$ is divisible by
2827: $x^\gamma$. As $U \subseteq B_\frL(1)^{m+n}$, it follows for all $x
2828: \in U$ that
2829: \begin{align*}
2830: \left\|(\g^p)_{(\gamma,0)}(x,y)\right\| &\le
2831: \sum_{((I_1,\alpha_1), \dots, (I_p, \alpha_p)) \in \J^p}
2832: \|\g_{I_1,\alpha_1}(x,y)\| \cdots
2833: \|\g_{I_p,\alpha_p}(x,y)\| \\ &= \left(\sum_{(I,\alpha) \in \J}
2834: \|\g_{I,\alpha}(x,y)\|\right)^p \\ &\le \left(|\J|
2835: C\right)^p.
2836: \end{align*}
2837: It follows from Lemma \ref{derivation_truncation} and the Cauchy
2838: estimates that
2839: \begin{equation*}
2840: \|(b_p)_{(\gamma,0)}(x,y')\| = \frac 1{p!} \left\| \frac {\partial^{p-1}
2841: (\g^p)_{(\gamma,0)}} {\partial y_n^{p-1}} (x,y',0) \right\| \le \frac
2842: {\left(|\J|C\right)^p} {\rho^{p-1}}
2843: \end{equation*}
2844: for all $(x,y') \in V$; so we can take $A := \rho$ and $B :=
2845: |\J|C/\rho$.
2846:
2847: Finally, let $\kappa \in [0,\infty)^m$ be such that $(\kappa,0) \in
2848: \cl_0(V \times B_\frL(\rho/2))$. Then $\t_{(\kappa,0)} (b_p)_{(\gamma,0)}
2849: \in \A_m(V')$ for some appropriate $V'$ independent of $p$ by Remark
2850: \ref{O-rmk}(2). Since $\|\t_{(\kappa,0)} (b_p)_{(\gamma,0)}(x,y')\|
2851: \le AB^p$ for all $(x,y') \in V'$ by the above, it follows that
2852: $\t_{(\kappa,0)} H_{(\gamma,0)}$ belongs to $\A_m(V' \times
2853: B_\frL(\rho/2))$, and the proposition is proved.
2854: \end{proof}
2855:
2856: The case of several implicit variables can be reduced to that of one
2857: implicit variable: below, we let $l \in \{1, \dots, n\}$, and we write
2858: $y' := (y_1, \dots, y_{n-l})$, $z = (z_1, \dots, z_l):= (y_{n-l+1},
2859: \dots, y_n)$, $Y' := (Y_1, \dots, Y_{n-l})$ and $Z = (Z_1, \dots, Z_l)
2860: := (Y_{n-l+1}, \dots, Y_n)$.
2861:
2862: \begin{cor}[Implicit Function Theorem]
2863: \label{multi_implicit}
2864: Let $f \in (\Q_m)^l$ such that $f(0) = 0$ and $\partial
2865: f/\partial z(0) \ne 0$. Then there is an $h \in (\Q_m^{m+n-l})^l$
2866: such that $h(0) = 0$ and $f(x,y',h) = 0$.
2867: \end{cor}
2868:
2869: \begin{proof}
2870: By induction on $l$; the case $l=1$ corresponds to Proposition
2871: \ref{implicit}, so we assume that $n \ge l>1$ and the corollary
2872: holds for lower values of $l$. After permuting the component
2873: functions of $f$, we may assume that $\partial f_l/\partial y_n(0)
2874: \ne 0$. Hence, writing $z' := (y_{n-l+1}, \dots, y_{n-1})$, we
2875: obtain from Proposition \ref{implicit} a function $w \in
2876: \Q_m^{m+n-1}$ such that $f_l(x,y', z', w) = 0$. Moreover, there are
2877: constants $c_1, \dots, c_{l-1} \in \CC$ such that for each $i=1,
2878: \dots, l-1$, defining $f'_i := f_i - c_i f_l$ gives $(\partial
2879: f'_i/\partial y_n)(0) = 0$. By the hypothesis of the corollary, the
2880: map $g \in (\Q_m^{m+n-1})^{l-1}$ defined by
2881: \begin{equation*}
2882: g_i(x,y',z') := f'_i(x,y',z',w(x,y',z')) \quad\text{for } i=1,
2883: \dots, l-1
2884: \end{equation*}
2885: satisfies $g(0) = 0$ and $(\partial g / \partial z')(0) \ne 0$.
2886: Hence by the inductive hypothesis, there is an $h' \in
2887: (\Q_m^{m+n-l})^{l-1}$ such that $g(x,y',h') = 0$. The corollary
2888: follows with $h \in (\Q_m^{m+n-l})^l$ defined by $h_i := h'_i$ if $i
2889: = 1, \dots, l-1$ and $h_l(x,y') := w(x,y',h'(x,y'))$.
2890: \end{proof}
2891:
2892: For the next proposition, we let $\sigma = (\sigma_1, \dots,
2893: \sigma_l)$ be the elementary symmetric functions in the variables $z$.
2894: Recall that $f \in \Q_m$ is \textbf{symmetric in the variables $z$} if
2895: $f(x,y',z) = f(x,y', \lambda(z))$ for every permutation $\lambda$ of
2896: $\{1, \dots, l\}$.
2897:
2898: \begin{prop}[Symmetric Function Theorem]
2899: \label{symmetric}
2900: Let $f \in \Q_m$ be symmetric in the variables $z$. Then there is a
2901: $g \in \Q_m$ such that $f(x,y',z) = g(x,y',\sigma)$.
2902: \end{prop}
2903:
2904: \begin{proof}
2905: First, let $\gamma \in [0,\infty)^m$ and assume there is a $G \in
2906: \Ps{C}{X^*,Y',Z}$ such that $Tf(X,Y',Z) = G(X,Y',\sigma_1(Z), \dots,
2907: \sigma_l(Z))$. Then $(Tf)_{(\gamma,0,0)}$ is symmetric in $Z$
2908: and $$(Tf)_{(\gamma,0,0)}(X,Y',Z) =
2909: G_{(\gamma,0,0)}(X,Y',\sigma_1(Z), \dots, \sigma_l(Z)).$$ Moreover,
2910: if $g \in \A_m(V)$ is such that $f(x,y',z) = g(x,y',\sigma)$, and if
2911: $\kappa \in [0,\infty)^m$ is sufficiently small, then
2912: $\t_{(\kappa,0,0)} f$ is symmetric in $z$ and $\t_{(\kappa,0,0)}
2913: f(x,y',z) = \t_{(\kappa,0,0)} g(x,y',\sigma)$.
2914:
2915: Therefore, we assume that $f \in \A_m(U)$ for some $m$-quadratic $U
2916: \subseteq \frL^{m+n}$ and we need to find, after shrinking $U$ if
2917: necessary, a $g \in \A_m(U)$ such that $f(x,y',z) = g(x,y',\sigma_1,
2918: \dots, \sigma_l)$ for all $(x,y',z) \in U$. Without loss of
2919: generality, we also assume that $U = W^m \times B_\frL(R)^n$ for some
2920: quadratic domain $W \subseteq \frL$ and some $R>0$, and we put $U'
2921: := W^m \times B_\frL(R)^{n-l}$.
2922:
2923: By Proposition \ref{coefficients}, there are $a_q \in
2924: \A_m^{m+n-l}(U')$, for $q \in \NN^l$, and constants $B,C>0$ such
2925: that $\|a_q(x,y')\| \le B C^{|q|}$ for each $q$ and $f(x,y',z) =
2926: \sum_{q \in \NN^l} a_q(x,y') z^q$. Let also $\sim$ be the
2927: equivalence relation on $\NN^l$ defined by $p \sim q$ if and only if
2928: there is a permutation $\lambda$ of $\{1, \dots, l\}$ such that $p =
2929: (q_{\sigma(1)}, \dots, q_{\sigma(l)})$, and let $E_1, E_2, \dots$ be
2930: an enumeration of all equivalence classes of $\sim$. Since $f$ is
2931: symmetric in $z$, we get for all $j \in \NN$ that $a_p = a_q$ for
2932: all $p,q \in E_j$. Thus, for each $j \in \NN$, we define $b_j :=
2933: a_p$ for some $p \in E_j$; then
2934: \begin{equation*}
2935: \sum_{p \in E_j} a_p(x,y') z^p = b_j(x,y') \cdot \sum_{p \in E_j}
2936: z^p \quad\text{for all } j \in \NN.
2937: \end{equation*}
2938: Let $j \in \NN$, and note that the sum $\sum_{p \in E_j} z^p$ is a
2939: symmetric polynomial in $z$ that is homogeneous of degree $d_j :=
2940: |p|$ for any $p \in E_j$. By the main theorem on symmetric
2941: polynomials (see for instance Van der Waerden \cite{vdw:algebra}),
2942: there is a unique polynomial $S_j \in \CC[Z]$ of weighted degree
2943: $d_j$ such that $\sum_{p \in E_j} z^p = S_j(\sigma_1(z), \dots,
2944: \sigma_l(z))$. (Here ``of weighted degree $d_j$'' means that any
2945: term $c z^p$ occurring in $S_j$ satisfies $p_1 + 2p_2 + \cdots +
2946: lp_l = d_j$.)
2947:
2948: We now define $S \in \Ps{C}{Z}$ by $S(Z) := \sum_{j \in \NN}
2949: S_j(Z)$; we claim that $S$ is convergent. To see this, note that
2950: the product $\Pi_{i=1}^l (1-Z_i)$ is a symmetric polynomial; hence,
2951: there is a polynomial $P \in \CC[Z]$ such that $P(\sigma_1(Z),
2952: \dots, \sigma_l(Z)) = \Pi_{i=1}^l (1-Z_i)$. Since $P(0) \ne 0$, we
2953: have that $1/P \in \Ps{C}{Z}$ converges; but $(1/P)(\sigma_1(Z),
2954: \dots, \sigma_l(Z)) = \sum_{q \in \NN^l} z^q$, and the claim
2955: follows.
2956:
2957: We now claim that the sum $g(x,y',z) := \sum_{j \in \NN} b_j(x,y')
2958: S_j(z)$ defines a function $g \in \A_m(U)$. Assuming the claim
2959: holds, we necessarily have $Tg(X,Y',Z) = G(X,Y',Z) := \sum_{j \in \NN}
2960: Tb_j(X,Y') S_j(Z)$, so $g$ has the required properties by the
2961: injectivity of the map $T$ and because $Tf(X,Y',Z) = G(X,Y',
2962: \sigma_1(Z), \dots, \sigma_l(Z))$.
2963:
2964: To see the claim, we first need to rewrite the sum: for each $q \in
2965: \NN^l$, we put $D_q := \{j \in \NN:\ q \in \supp S_j\}$ and $c_q :=
2966: \sum_{j \in D_q} b_j$, so that $g(x,y',z) = \sum_{q \in \NN^l}
2967: c_q(x,y') z^q$ for all $(x,y',z) \in U$. Note that for all $q \in
2968: \NN^l$, $j \in D_q$ and $p \in E_j$, we have $|p| = q_1 + 2q_2 +
2969: \cdots + lq_l \le l|q|$. Hence, for each $q \in \NN^l$, there is a
2970: set $C_q \subseteq \set{r \in \NN^l:\ |r| \le l|q|}$ such that $c_q
2971: = \sum_{r \in C_q} a_r$. Since $|C_q| \le (l|q|)^l$, it follows
2972: that there are constants $\a,\b > 0$ such that $\|c_q(x,y')\| \le
2973: \a\b^{|q|}$ for every $q \in \NN^l$ and all $(x,y') \in U'$. Since
2974: $c_q \in \A_m^{m+n-l}(U')$ for every $q \in \NN^l$, the claim
2975: follows from Proposition \ref{O-membership}.
2976: \end{proof}
2977:
2978: For the remainder of this section, we write again $Y' = (Y_1, \dots,
2979: Y_{n-1})$. We recall from Definition 4.16 of \cite{vdd-spe:genpower}
2980: that $F \in \Ps{C}{X^*,Y}$ is regular in $Y_n$ of order $d \in \NN$ if
2981: $F(0,0,Y_n) = c Y_n^d + $ terms of higher order in $Y_n$.
2982:
2983: \begin{prop}[Weierstrass Preparation]
2984: \label{wp_theorem}
2985: Assume that $n>0$, and let $f \in \Q_m$ be such that $Tf$ is regular
2986: in $Y_n$ of order $d \in \NN$.
2987: \begin{enumerate}
2988: \item For every $g \in \Q_m$, there are a unique $q \in \Q_m$ and a
2989: unique $r \in \Q_m^{m+n-1}[Y_n]$ such that $g = qf + r$ and
2990: $\deg_{Y_n}(r) < d$.
2991: \item There are a unique unit $u \in \Q_m$ and a unique $w \in
2992: \Q_m^{m+n-1}[Y_n]$ such that $f = uw$ and $w$ is monic of degree
2993: $d$ in $Y_n$.
2994: \end{enumerate}
2995: \end{prop}
2996:
2997: \begin{proof}
2998: The proof of Theorems 1 and 2 on p. 338 of \cite{bri-kno:plane} goes
2999: through almost literally, using the properties established for the
3000: classes $\Q_m$ in the previous sections as well as the Implicit
3001: Function Theorem and the Symmetric Function Theorem above, except
3002: for the following trivial changes: the variable $t$ and $z_1, \dots,
3003: z_n$ there correspond to $y_n$ and $x_1, \dots, x_m, y_1, \dots,
3004: y_{n-1}$ here, and the roles of $f$ and $g$ are exchanged. (Note
3005: that the uniqueness also follows directly from Proposition 4.17 in
3006: \cite{vdd-spe:genpower} and the injectivity of the map $T:\Q_m \into
3007: \Ps{C}{X^*,Y}$.)
3008: \end{proof}
3009:
3010:
3011: \section{$\Q$-semianalytic sets and model completeness}
3012: \label{o-minimal}
3013:
3014: In this section we prove model completeness and o-minimality of $\Rq$.
3015: We also show that $\Rq$ admits analytic cell decomposition. (An
3016: o-minimal expansion $\Rtilde$ of the ordered field of real numbers is
3017: said to admit \textbf{analytic cell decomposition} if for any $A_{1},
3018: \dots, A_{k} \subseteq \RR^{m}$ definable in $\Rtilde$, there is a
3019: decomposition of $\RR^{m}$ into analytic cells definable in $\Rtilde$
3020: and compatible with each $A_{i}$.)
3021:
3022: We let $m,n \in \NN$ and $\rho \in (0,\infty)^{m+n}$, and we put
3023: \begin{equation*}
3024: I_{m,n;\rho} := [0,\rho_1] \times \cdots \times [0,\rho_m] \times
3025: [-\rho_{m+1}, \rho_{m+1}] \times \cdots \times [-\rho_{m+n},
3026: \rho_{m+n}],
3027: \end{equation*}
3028: a subset of $\RR^{m+n}$. We also write $I_{m,n;\epsilon}$ instead of
3029: $I_{m,n;(\epsilon, \dots, \epsilon)}$, for $\epsilon > 0$, and we put
3030: $I_{m,n;\infty} := [0,\infty)^m \times \RR^n$. Abusing notation, we
3031: identify $I_{m,n;\rho}$ with the set
3032: \begin{multline*}
3033: \{(x,y) \in [0,\infty)_{\frL_0}^m \times \RR^n:\ 0 \le \|x_i\| \le
3034: \rho_i \text{ and } \\ -\rho_{m+j} \le y_j \le \rho_{m+j} \
3035: \text{for } i=1, \dots, m \text{ and } j=1, \dots, n\}.
3036: \end{multline*}
3037: Given an $m$-quadratic $U \subseteq \frL^{m+n}$ such that
3038: $I_{m,n;\rho} \subseteq \ir(\cl(\pi^{m+n}_m(U))$, and given an $f \in
3039: \Q_m(U)$, we write $f\rest{I_{m,n;\rho}}$ for the function
3040: $f^\sharp\rest{I_{m,n;\rho}}$.
3041:
3042: \begin{df}
3043: \label{real_Q}
3044: We let $\Q_{m,n;\rho}$ be the set of all functions $f:I_{m,n;\rho}
3045: \into \RR$ for which there exist an $m$-quadratic domain $U
3046: \subseteq \frL^{m+n}$ and a $g \in \Q_m(U)$ such that $I_{m,n;\rho}
3047: \subseteq \ir(\cl(\pi^{m+n}_m(U))$ and $f = g \rest{I_{m,n;\rho}}$.
3048: \end{df}
3049:
3050: \begin{nrmk}
3051: \label{restricted}
3052: For every $f \in \Q_{m,n;\rho}$, there are $\rho' > \rho$ and $g \in
3053: \Q_{m,n,\rho'}$ such that $f = g\rest{I_{m,n;\rho}}$.
3054: \end{nrmk}
3055:
3056: \begin{prop}
3057: \label{specialize}
3058: Let $U \subseteq \frL^{m+n}$ be $m$-quadratic such that $I_{m,n;\rho}
3059: \subseteq \ir(\cl(\pi^{m+n}_m(U))$, and let $f \in \Q_m(U)$. Then
3060: $f\rest{I_{m,n;\rho}} \in \Q_{m,n;\rho}$ if and only if $Tf \in
3061: \Ps{R}{X^*,Y}$.
3062: \end{prop}
3063:
3064: \begin{proof}
3065: The necessity is clear from Definition \ref{asymptotic}, so we
3066: assume that $Tf \in \Ps{R}{X^*,Y}$. Define $\bar{(r,\varphi)} :=
3067: (r,-\varphi)$ for $(r,\varphi) \in \frL$ and $\bar{(x,y)} :=
3068: (\bar{x_1}, \dots, \bar{y_n})$ for $(x,y) \in \frL^{m+n}$. Then the
3069: function $\bar{f}:U \into \CC$ defined by $\bar{f}(x) :=
3070: \bar{f(\bar{x})}$ belongs to $\A(U)$ and satisfies
3071: $T\left(\bar{f}\right) = Tf$. Hence by Proposition
3072: \ref{asymptotic_to_0}, we have $\bar{f} = f$, which proves the
3073: proposition.
3074: \end{proof}
3075:
3076: Correspondingly, we put $\Pc{R}{X^*,Y}_{\Q,\rho} := \set{Tf:\ f \in
3077: \Q_{m,n;\rho}}$. If $\epsilon > 0$, we write
3078: $\Pc{R}{X^*,Y}_{\Q,\epsilon}$ and $\Q_{m,n;\epsilon}$ instead of
3079: $\Pc{R}{X^*,Y}_{\Q,(\epsilon,\mdots, \epsilon)}$ and
3080: $\Q_{m,n;(\epsilon,\mdots, \epsilon)}$. Next, we
3081: put $$\Pc{R}{X^*,Y}_{\Q} := \bigcup_{\rho \in (0, \infty)^{m+n}}
3082: \Pc{R}{X^*,Y}_{\Q,\rho}.$$ For $n=0$ we just write
3083: $\Pc{R}{X^*}_{\Q,\rho}$ instead of $\Pc{R}{X^*,Y}_{\Q,\rho}$.
3084:
3085: The properties described Sections \ref{division}, \ref{weierstrass}
3086: and \ref{blowups} of the algebras $\Q_m(U)$ are easily seen to imply
3087: corresponding properties of the algebras $\Q_{m,n;\rho}$ and
3088: $\Pc{R}{X^*,Y}_{\Q,\rho}$. Due to Proposition \ref{Q-algebra}, we
3089: need no longer formally distinguish between $f \in \Q_{m,n;\rho}$ and
3090: $Tf \in \Pc{R}{X^*,Y}_{\Q,\rho}$; in particular, the notations in
3091: Sections 7, 8 and 9 of \cite{vdd-spe:genpower} make sense in our
3092: setting.
3093:
3094: \begin{df}
3095: \label{basicset}
3096: A set $A \subseteq I_{m,n;\rho}$ is called a \textbf{basic
3097: $\Q_{m,n;\rho}$-set} if there are $f,g_1,\mdots, g_k \in
3098: \Q_{m,n;\rho}$ such that $$A = \set{z \in I_{m,n;\rho} :\ f(z) = 0,\
3099: g_1(z) > 0,\mdots,\ g_k(z) > 0}.$$ A \textbf{$\Q_{m,n;\rho}$-set}
3100: is a finite union of basic $\Q_{m,n;\rho}$-sets. Note that the
3101: $\Q_{m,n;\rho}$-sets form a boolean algebra of subsets of
3102: $I_{m,n;\rho}$.
3103: \end{df}
3104:
3105: Given a point $a = (a_1,\mdots, a_{m+n}) \in \RR^{m+n}$ and a choice
3106: of signs $\sigma \in \{-1, 1\}^m$, we let $h_{a,\sigma} :\ \RR^{m+n}
3107: \longrightarrow \RR^{m+n}$ be the bijection given by $$h_{a,\sigma}(z)
3108: := \left(a_1 + \sigma_1 z_1, \mdots, a_{m+n} + z_{m+n}\right). $$ Note
3109: that the maps $h_{a,\sigma}$ (with $a \in \RR^{m+n}$ and $\sigma \in
3110: \{-1, 1\}^m$) form a group of permutations of $\RR^{m+n}$.
3111:
3112: \begin{df}
3113: \label{gensemianalytic}
3114: A set $X \subseteq \RR^{m+n}$ is \textbf{$\Q_{m,n}$-semianalytic at
3115: $a \in \RR^{m+n}$} if there is an $\epsilon > 0$ such that for
3116: each $\sigma \in \{-1, 1\}^m$ there is a $\Q_{m,n;\epsilon}$-set
3117: $A_{\sigma} \subseteq I_{m,n;\epsilon}$ with $$X \cap
3118: h_{a,\sigma}(I_{m,n;\epsilon}) = h_{a,\sigma}(A_{\sigma}).$$ A set
3119: $X \subseteq \RR^{m+n}$ is \textbf{$\Q_{m,n}$-semianalytic} if it is
3120: $\Q_{m,n}$-semianalytic at every point $a \in \RR^{m+n}$. For
3121: convenience, if $X \subseteq \RR^m$ is $\Q_{m,0}$-semianalytic we
3122: also simply say that $X$ is $\Q_m$-semianalytic.
3123: \end{df}
3124:
3125: \begin{nrmk}
3126: \label{gensemrmk}
3127: \begin{enumerate}
3128: \item If $X, Y \subseteq \RR^{m+n}$ are $\Q_{m,n}$-semianalytic at
3129: $a$, then so are $X \cup Y$, $X \cap Y$ and $X \setminus Y$.
3130: \item Let $X \subseteq \RR^{m+n}$ be $\Q_{m,n}$-semianalytic, $a \in
3131: \RR^{m+n}$ and $\sigma \in \{-1, 1\}^m$. Then the set
3132: $h_{a,\sigma}(X)$ is $\Q_{m,n}$-semianalytic. Moreover by Lemma
3133: \ref{Q-addition}(3), for each $\lambda \in (0, \infty)^{m+n}$ the
3134: set $E_{\lambda}(X)$ is $\Q_{m,n}$-semianalytic, where
3135: $E_{\lambda} : \RR^{m+n} \longrightarrow \RR^{m+n}$ is defined by
3136: $E_{\lambda}(z) := (\lambda_1 z_1, \dots, \lambda_{m+n} z_{m+n})$.
3137: \item If $X \subseteq \RR^n$ is semianalytic, then $X$ is
3138: $\Q_{0,n}$-semianalytic.
3139: \end{enumerate}
3140: \end{nrmk}
3141:
3142: Below we write $0$ for the point $(0,\mdots, 0) \in \RR^{m+n}$. The
3143: following lemma is now proved just as in \cite[Section
3144: 7]{vdd-spe:genpower} with obvious changes: ``$\R_{\mdots}$-set'' is
3145: replaced by ``$\Q_{\dots}$-set'', $\Pc{R}{X^*,Y}$ by
3146: $\Pc{R}{X^*,Y}_{\Q}$ and the algebras $\R_{m,n,\dots}$ by
3147: $\Q_{m,n;\dots}$. Also, the results from Sections 4, 5 and 6 there
3148: need to be replaced by the corresponding results of Sections
3149: \ref{division}, \ref{weierstrass} and \ref{blowups} here. (For
3150: example, we use Proposition \ref{Q-basics}(2) here in place of
3151: Corollary 6.7 there; the other replacements are more straightforward.)
3152:
3153: \begin{lemma} \label{gensembasics}
3154: \begin{enumerate}
3155: \item Let $A \subseteq \RR^{m+n}$ be $\Q_{m,n}$-semianalytic at $0$
3156: and let $\sigma$ be a permutation of $\set{1,\mdots, m}$. Then
3157: $\sigma(A)$ is $\Q_{m,n}$-semianalytic at $0$.
3158: \item If $n \geq 1$, then each $\Q_{m,n}$-semianalytic subset of
3159: $\,\RR^{m+n}$ is also $\Q_{m+1,n-1}$-semianalytic.
3160: \item Every $\Q_{m,n;\rho}$-set $A \subseteq I_{m,n;\rho}$ is
3161: $\Q_{m,n}$-semianalytic. \qed
3162: \end{enumerate}
3163: \end{lemma}
3164:
3165: Note that Remark \ref{gensemrmk}(3) and Lemma \ref{gensembasics}(2)
3166: imply in particular that every semianalytic subset of $\RR^{m+n}$ is
3167: $\Q_{m,n}$-semianalytic. Also, since every $f \in \R_{m,n,\rho}$
3168: extends to a holomorphic function $g:B_\frL(\rho') \into \CC$ for some
3169: $\rho'>\rho$, we see that $\R_{m,n,\rho}^\omega \subseteq
3170: \Q_{m,n;\rho}$, where $\R_{m,n,\rho}^\omega$ consists of all $f \in
3171: \R_{m,n,\rho}$ with natural support. In particular, every
3172: $\R_{m,n}^\omega$-semianalytic subset of $\RR^{m+n}$ is
3173: $\Q_{m,n}$-semianalytic.
3174:
3175: We now consider the system $\Lambda = (\Lambda_p)_{p \in \NN}$, where
3176: $$\Lambda_p := \set{A \subseteq I^p:\ A \text{ is }
3177: \Q_{p} \text{-semianalytic}}.$$ Note that if $A \subseteq I^p$ is
3178: $\Q_{m,n}$-semianalytic with $m+n=p$, then $A$ is also
3179: $\Q_{p}$-semianalytic by Lemma \ref{gensembasics}(2), so $A \in
3180: \Lambda_p$. A set $A \subseteq \RR^m$ is called a
3181: \textbf{$\Lambda$-set} if $A \in \Lambda_n$, and $B \subseteq \RR^m$
3182: is called a \textbf{sub-$\Lambda$-set} if there exist $n \in \NN$ and
3183: a $\Lambda$-set $A \subseteq \RR^{m+n}$ such that $B = \Pi_m(A)$.
3184:
3185: \begin{prop} \label{mc+om1}
3186: Let $A \subseteq [-1,1]^m$ be a sub-$\Lambda$-set. Then $[-1,1]^m
3187: \setminus A$ is also a sub-$\Lambda$-set.
3188: \end{prop}
3189:
3190: \begin{proof}[Sketch of proof]
3191: By Theorem 2.7 of \cite{vdd-spe:genpower}, we need to establish
3192: Axioms (I)-(IV) listed in \cite[Section 2]{vdd-spe:genpower}; the
3193: first three are straightforward. For Axiom (IV), the proof proceeds
3194: almost literally as for \cite[Corollary 8.15]{vdd-spe:genpower},
3195: with the obvious changes indicated earlier, as well as the
3196: following: $\Pc{R}{X^*, Y}_{\dots}$ there is replaced by
3197: $\Pc{R}{X^*,Y}_{\Q,\dots}$ here, and the facts of Section
3198: \ref{o-minimal} here are used in place of the corresponding facts
3199: from Section 7 there. Moreover, note that Lemma 6.1 there goes
3200: through unchanged here.
3201: \end{proof}
3202:
3203: Recall that by the remarks after Definition \ref{real_Q}, $$\Rq =
3204: \left(\RR, <, 0, 1, +, -, \mdot, \left(\tilde{f}:\ f \in
3205: \Q_{m,0;1}\right)\right).$$ It is clear from Remark
3206: \ref{gensemrmk}(2) that every \textit{bounded} $\ \Q_p$-semianalytic
3207: set, for $p \in \NN$, is quantifier-free definable in $\Rq$. We are
3208: now ready to prove Theorem A.
3209:
3210: \begin{thm}
3211: The expansion $\Rq$ is model complete, o-minimal and admits analytic
3212: cell decomposition.
3213: \end{thm}
3214:
3215: \begin{proof}
3216: The theorem follows from the previous remark in view of Propositions
3217: \ref{mc+om1} above and \cite[Corollary 2.9]{vdd-spe:genpower}. For
3218: analytic cell decomposition, we proceed exactly as in the proof of
3219: Corollary 6.10 of \cite{vdd-spe:multisummable}.
3220: \end{proof}
3221:
3222: For Theorem B, we proceed as in Section 9 of \cite{vdd-spe:genpower},
3223: with the following changes: we do not need 9.3 there, and use
3224: Corollaries \ref{bad_substitution_1} and \ref{bad_substitution_2} here
3225: in place of Lemma 9.4 there. Moreover, we do not need 9.7 and Lemma
3226: 9.8 there, and we give a much simpler proof of Lemma 9.9 there:
3227:
3228: \begin{lemma}
3229: \label{comp_inverse}
3230: Let $0 < f \in \Pc{R}{T^*}_\Q$ with $f(0) = 0$. Then there exists
3231: $g \in \Pc{R}{T^*}_\Q$ such that $g > 0$, $g(0) = 0$ and $f(g(T)) =
3232: T$.
3233: \end{lemma}
3234:
3235: \begin{proof}
3236: By the hypotheses, there are $\lambda,\alpha > 0$ and $h \in
3237: \Pc{R}{T^*}_\Q$ such that $f(t) = t^\alpha(\lambda + h(t))$ and
3238: $h(0) = 0$. We put $\rho := 1/\alpha$ and let $F,H \in
3239: \Pc{R}{T^*,Y^*}_\Q$ be defined by $F(T,Y) := f(Y)$ and $H(T,Y):=
3240: h(Y)$. By Proposition \ref{blowup_homom}, the functions
3241: $\r^{\rho,\lambda} F(t,y)$ and $\r^{\rho,\lambda} H(t,y)$ belong to
3242: $\Pc{R}{T^*,Y}_\Q$. We define
3243: \begin{equation*}
3244: \phi(t,y) := (\lambda+y)^\alpha \left(\lambda + \r^{\rho,\lambda}
3245: H(t,y)\right);
3246: \end{equation*}
3247: then $\r^{\rho,\lambda} F(t,y) = t \cdot \phi(t,y)$, so $\phi \in
3248: \Pc{R}{T^*,Y}_\Q$ by Proposition \ref{Q-basics}(1). Moreover, we
3249: have $\phi(0,0) = \lambda^{\alpha+1}$ and $\frac {\partial
3250: \phi}{\partial y}(0,0) = \alpha\lambda^\alpha > 0$; hence by the
3251: implicit function theorem, there is a $\psi \in \Pc{R}{T^*}_\Q$ such
3252: that $\phi(t,\psi(t)) = 1$. Therefore, we have $\r^{\rho,\lambda} F
3253: (t,\psi(t)) = t$, so we take $g(t) := t^{1/\alpha} (\lambda +
3254: \psi(t))$.
3255: \end{proof}
3256:
3257: Using this lemma in place of Lemma 9.9 in \cite{vdd-spe:genpower}, we
3258: finish the proof of Theorem B as it is done there, and we obtain
3259: corresponding corollaries for $1$-dimensional sets definable in
3260: $\Rq$.
3261:
3262:
3263: \section{Example of a definable family of transition maps}
3264: \label{family}
3265:
3266: Let $\xi$ be an analytic vector field in $\RR^2$. A
3267: \textbf{polycycle} $\Gamma$ of $\xi$ is a cyclically ordered finite
3268: set of singular points $p_0 = p_k, p_{1},\dots, p_{k}, p_{k+1} = p_1$
3269: (with possible repetitions), called \textbf{vertices}, and
3270: trajectories $\gamma_{1},\dots, \gamma_{k}$, called
3271: \textbf{separatrices}, connecting the vertices in the order following
3272: the flow of $\xi$, as in Figure 2. We assume here that each $p_i$ is
3273: a non-resonant hyperbolic singularity of $\xi$. For each $i$, we fix
3274: two segments $\Lambda_i^-$ and $\Lambda_i^+$ transverse to $\xi$ and
3275: intersecting the separatrices $\gamma_{i-1}$ and $\gamma_i$,
3276: respectively, close to $p_i$.
3277:
3278: \begin{figure}[htbp]
3279: \label{cap:poly}
3280: \begin{center}
3281: \input{polycycle.pst}
3282: \end{center}
3283: \caption{The polycycle $\Gamma$ of $\xi = \xi_0$}
3284: \end{figure}
3285:
3286: For each $i$, we fix analytic charts $x_i:(-1,1) \into \Lambda_i^-$
3287: and $y_i:(-1,1) \into \Lambda_i^+$ such that $x_i(0)$ and $y_i(0)$ are
3288: the points of intersection of $\Lambda_i^-$ with $\gamma_{i-1}$ and of
3289: $\Lambda_i^+$ with $\gamma_i$, respectively, and such that $x_i(t)$
3290: and $y_i(t)$ lie inside the region circumscribed by $\Gamma$ for all
3291: $t \in (0,1)$. We denote by $g_i:(0,1) \into (0,1)$ the corresponding
3292: transition map in the coordinates $x_i$ and $y_i$; we extend $g_i$ to
3293: all of $(-1,1)$ by putting $g_i(t) := 0$ for $t \in (-1,0]$. After an
3294: analytic change of coordinates if necessary, it follows from the
3295: general theory of analytic differential equations that there are
3296: analytic functions $f_i:(-1,1) \into (-1,1)$, for $i=1, \dots, k$,
3297: representing the flow of $\xi$ from $\Lambda_{i-1}^+$ to $\Lambda_i^-$
3298: in the charts $y_{i-1}$ and $x_i$. In fact, these functions $f_i$ are
3299: \textit{restricted analytic}, that is, they extend analytically to a
3300: neighbourhood of $[-1,1]$; in particular, they are definable in $\Rq$.
3301: The restriction of $P := f_k \circ g_k \circ f_{k-1} \circ \cdots
3302: \circ f_1 \circ g_1$ to $(0,1)$ represents the Poincar\'e first return
3303: map of $\xi$ at $p_1$ in the chart $x_1$.
3304:
3305: By the corollary and the explanations in the introduction each $g_i$,
3306: and hence $P$, is definable in $\Rq$. Our goal in this section is to
3307: show that for certain analytic unfoldings $\xi_\mu$ of the vector
3308: field $\xi$, the corresponding Poincar\'e return map \textit{with
3309: parameter} $\mu$ is also definable in $\Rq$.
3310:
3311: More precisely, we let $\xi_{\mu}$ be an analytic unfolding of $\xi$,
3312: with $\mu \in \RR^p$ and $\xi_0 = \xi$, defined in a neighborhood $U$
3313: of $\Gamma$ containing each $\Lambda_i^-$ and $\Lambda_i^+$, with the
3314: same singular points inside $U$ and \textit{with the same linear part
3315: at each of these singular points as $\xi$}. We assume that the
3316: unfolding is small, in the sense that for each $\mu \in \RR^p$ and
3317: each $i \in \{1, \dots, k\}$, both $\Lambda_i^-$ and $\Lambda_i^+$
3318: remain transverse to $\xi_\mu$, the transition map of $\xi_\mu$ at
3319: $p_i$ is given by a function $g_{\mu,i}:(0,1) \into (0,1)$ in the
3320: charts $x_i$ and $y_i$ (with the latter as above, independent of
3321: $\mu$), and there are analytic functions $f_{\mu,i}:(-1,1) \into
3322: (-1,1)$ representing the flow of $\xi_\mu$ from $\Lambda_{i-1}^+$ to
3323: $\Lambda_i^-$ in the charts $y_{i-1}$ and $x_i$. We extend each
3324: $g_{\mu,i}$ to $(-1,1)$ by putting $g_{\mu,i}(t) := 0$ if $t \in
3325: (-1,0]$. Then the restriction of $P_\mu:= f_{\mu,k} \circ g_{\mu,k}
3326: \circ f_{\mu,k-1} \circ \cdots \circ f_{\mu,1} \circ g_{\mu,1}$ to
3327: $(0,1)$ represents the Poincar\'e first return map of $\xi_\mu$ at
3328: $p_1$ in the chart $x_1$.
3329:
3330: We define $g_i:(-1,1) \times \RR^p \into (-1,1)$, $f_i:(-1,1) \times
3331: \RR^p \into (-1,1)$ and $P:(-1,1) \times \RR^p \into (-1,1)$ by
3332: $g_i(t,\mu):= g_{\mu,i}(t)$, $f_i(t,\mu):= f_{\mu,i}(t)$ and
3333: $P(t,\mu):= P_\mu(t)$. Since the chosen unfolding $\xi_\mu$ is small,
3334: each $f_i$ is a restricted analytic map and hence definable in $\Rq$.
3335: Moreover, we have the following:
3336:
3337: \begin{prop}
3338: \label{finite_cyclicity}
3339: Each $g_i$ is definable in the structure $\Rq$. In particular, $P$
3340: is definable in $\Rq$, and the number of isolated fixed points of
3341: $P_\mu$ is uniformly bounded in $\mu$.
3342: \end{prop}
3343:
3344: For the proof of this proposition, we assume that $p_1 = 0 \in \RR^2$
3345: and show that $g_1$ is definable in $\Rq$. First, since the ratio
3346: $\lambda = \lambda_2/\lambda_1$ of the eigenvalues $\lambda_1$ (with
3347: respect to $x_1$) and $\lambda_2$ (with respect to $y_1)$ of $\xi_\mu$
3348: at $0$ is irrational and independent of $\mu$, we may assume, after a
3349: change of coordinates that is analytic in both $(x,y)$ and $\mu$, that
3350: the incoming and outgoing separatrices of $\xi_\mu$ at $0$ are
3351: represented by the $x$-axis and the $y$-axis, respectively, for every
3352: $\mu$. In this situation, the normalization method in
3353: \cite[pp. 70--73]{dul:cycles} goes through uniformly in the parameter
3354: $\mu$ and yields:
3355:
3356: \begin{lemma}
3357: \label{dulac_reduction}
3358: Let $N \in \NN$ be positive. Then there exist analytic functions
3359: $\phi_N,A_N:\RR^{2+p} \into \RR$ such that $A_N(0,0,0) = 0$, the map
3360: $\Phi^N:\RR^{2+p} \into \RR^{2+p}$ defined by $(u,v,\mu) =
3361: \Phi^N(x,y,\mu) := (x,\phi_N(x,y,\mu),\mu)$ is a change of coodinates
3362: fixing $0$ and for each $\mu \in \RR^p$, the push-forward
3363: $\Phi^N_*\xi_\mu$ satisfies the equations
3364: \begin{equation}
3365: \begin{split}
3366: \label{normal_form}
3367: \dot{u} & = u\\
3368: \dot{v} & =
3369: v\left(\lambda+u^{N}v^{N}A_N\left(u,v,\mu\right)\right). \qed
3370: \end{split}
3371: \end{equation}
3372: \end{lemma}
3373:
3374: Second, we fix a segment $\Lambda^-:= (0,\epsilon) \times \{y_0\}$,
3375: parametrized by the $x$-coordinate along $\Lambda^-$, and a segment
3376: $\Lambda^+:= \{x_0\} \times (0,\epsilon)$, parametrized by the
3377: $y$-coordinate along $\Lambda^+$. We assume that $x_0$, $y_0$ and
3378: $\epsilon$ are small enough such that $\Lambda^-$ and $\Lambda^+$ are
3379: transverse to each $\Phi_*^N \xi_\mu$. We denote by
3380: $g^N_\mu:(0,\epsilon) \into (0,\epsilon)$ the corresponding transition
3381: map of $\Phi_*^N\xi_\mu$, and we define $g^N:(0,\epsilon) \times
3382: \RR^p \into (0,\epsilon)$ by $g^N(t,\mu):= g^N_\mu(t)$. In this
3383: situation, the estimates obtained on pp. 24--29 in \cite{ily:dulac} go
3384: through uniformly in $\mu$; in particular, the domain $\Omega$ defined
3385: by inequality $( \overset{*}{_{**}} )$ on p. 29 is independent of
3386: $\mu$. Therefore, we obtain:
3387:
3388: \begin{lemma}
3389: \label{dulac_parameter}
3390: Let $\nu>0$. Then there exist an integer $N = N(\nu) > 0$,
3391: constants $K = K(\nu) > 0$ and $\epsilon = \epsilon(\nu) > 0$ and a
3392: quadratic domain $\Omega = \Omega(\nu)$ such that
3393: \begin{enumerate}
3394: \item the map $g^N$ is analytic in the variable $\mu$ and admits an
3395: analytic extension to $\Omega \times \RR^p$;
3396: \item $\left|g^N(t,\mu) - t^{\lambda}\right| \le
3397: K|t|^{\nu+\epsilon}$ for all $(t,\mu) \in \Omega \times
3398: \RR^p$. \qed
3399: \end{enumerate}
3400: \end{lemma}
3401:
3402: Third, it follows from the theory of analytic differential equations
3403: that for each $N \in \NN$, there are analytic functions $h_N^-:(-1,1)
3404: \times \RR^p \into (-\epsilon,\epsilon) \times \RR^p$ and
3405: $h_N^+:(-\epsilon,\epsilon) \times \RR^p \into (-1,1) \times \RR^p$
3406: such that $h_N^-(0,\mu) = h_N^+(0,\mu) = 0$ for all $\mu$ and
3407: $g_1(t,\mu) = h_N^+(g^N(h_N^-(t,\mu),\mu),\mu)$ for all $(t,\mu)$.
3408:
3409: We write $\langle 1,\lambda \rangle$ for the additive submonoid of
3410: $\RR$ generated by $1$ and $\lambda$. By the binomial theorem, there
3411: is for each $\alpha \in \langle 1,\lambda \rangle$ and each $N \in
3412: \NN$ an analytic function $c_{\alpha,N}:\RR^p \into \RR$ such that for
3413: each $\mu \in \RR^p$,
3414: \begin{equation*}
3415: h_N^+\left( (h_N^-(t,\mu))^\lambda,\mu\right) = \sum_{\alpha \in
3416: \langle 1,\lambda \rangle} c_{\alpha,N}(\mu) \cdot t^\alpha.
3417: \end{equation*}
3418: On the other hand, given $\nu >0$, it follows from Lemma
3419: \ref{dulac_parameter} for each $\mu \in \RR^p$ and each $N \ge N(\nu)$
3420: that
3421: \begin{equation*}
3422: g_1(t,\mu) - \sum_{\alpha \le \nu} c_{\alpha,N}(\mu) \cdot
3423: t^\alpha = o\left(\|t\|^\nu\right) \quad\text{as } \|t\| \to 0;
3424: \end{equation*}
3425: in particular, $c_{\alpha,N} = c_{\alpha,N'}$ whenever $|\alpha| \le
3426: \nu$ and $N,N' \ge N(\nu)$. Thus, for each $\alpha \in \langle
3427: 1,\lambda \rangle$ we put $c_\alpha := c_{\alpha,N(|\alpha|)}$; then
3428: by Lemma \ref{dulac_parameter} again, we have for every $\nu > 0$ and
3429: all $(t,\mu) \in \Omega(N(\nu)) \times \RR^p$ that
3430: \begin{equation}
3431: \label{uni_estimate}
3432: \left\|g_1(t,\mu) - \sum_{\alpha \le \nu} c_\alpha(\mu) \cdot
3433: t^\alpha\right\| \le K(\nu) \cdot \|t\|^{\nu+\epsilon(\nu)}.
3434: \end{equation}
3435: It follows from Corollary \ref{uniform_criterion} that $g_1 \in
3436: \A_1(\Omega(0) \times \RR^p)$. Finally, given any $\nu > \gamma \ge
3437: 0$, we define $(g_1)_\gamma:\Omega(N(\gamma)) \times \RR^p \into \RR$
3438: by
3439: \begin{equation*}
3440: (g_1)_\gamma(t,\mu) := t^{-\gamma} \left(g_1(t,\mu) - \sum_{\alpha <
3441: \gamma}
3442: c_\alpha(\mu) t^\alpha\right).
3443: \end{equation*}
3444: Then by \eqref{uni_estimate} again, we have for all $(t,\mu) \in
3445: \Omega(N(\nu)) \times \RR^p$ that
3446: \begin{equation*}
3447: \left\|(g_1)_\gamma(t,\mu) - \sum_{\gamma \le \alpha \le \nu}
3448: c_\alpha(\mu) \cdot
3449: t^{\alpha-\gamma}\right\| \le K(\nu) \cdot
3450: \|t\|^{\nu+\epsilon(\nu)-\gamma}.
3451: \end{equation*}
3452: Hence by Corollary \ref{uniform_criterion}, each $(g_1)_\gamma$
3453: belongs to $\A_1(\Omega(N(\gamma)) \times \RR^p)$, that is, $g_1$
3454: satisfies condition (TE). It follows that $g_1$ belongs to
3455: $\Q_1(\Omega(0) \times \RR^p)$, which proves Proposition
3456: \ref{finite_cyclicity}.
3457:
3458:
3459:
3460: \begin{thebibliography}{10}
3461: \itemsep -0cm
3462:
3463: \bibitem{bri-kno:plane}
3464: {\em E.~Brieskorn and H.~Kn\"orrer}, Plane algebraic curves, Birkh\"auser
3465: Verlag, 1986.
3466:
3467: \bibitem{vdd:vol1}
3468: {\em L.~van~den Dries}, Tame Topology and O-minimal Structures, no.~248 in LMS
3469: Lecture Note Series, Cambridge University Press, 1998.
3470:
3471: \bibitem{vdd-spe:genpower}
3472: {\em L.~van~den Dries and P.~Speissegger}, The real field with convergent
3473: generalized power series is model complete and o-minimal, Trans. Amer. Math.
3474: Soc., {\bf 350} (1998), ~4377--4421.
3475:
3476: \bibitem{vdd-spe:multisummable}
3477: \leavevmode\vrule height 2pt depth -1.6pt width 23pt, The field of reals with
3478: multisummable series and the exponential function, Proc. London Math. Soc.
3479: (3), {\bf 81} (2000), ~513--565.
3480:
3481: \bibitem{dul:cycles}
3482: {\em H.~Dulac}, Sur les cycles limites, Bull. Soc. Math. France, {\bf 51}
3483: (1923), ~45--188.
3484:
3485: \bibitem{eca:dulac}
3486: {\em J.~Ecalle}, Introduction aux fonctions analysables et preuve constructive
3487: de la conjecture de Dulac, Hermann, Paris, 1992.
3488:
3489: \bibitem{ily:dulac}
3490: {\em Yu.~S. Ilyashenko}, Finiteness theorems for limit cycles, vol.~94 of
3491: Translations of Mathematical Monographs, American Mathematical Society, 1991.
3492:
3493: \bibitem{ily:history}
3494: \leavevmode\vrule height 2pt depth -1.6pt width 23pt, Centennial history of
3495: {H}ilbert's sixteenth problem, Bull. Amer. Math. Soc., {\bf 39} (2002),
3496: ~301--354.
3497:
3498: \bibitem{rou:bifurcations}
3499: {\em R.~Roussarie}, Bifurcations of planar vector fields and {H}ilbert's
3500: sixteenth problem, vol.~164 of Progress in Mathematics, Birkh\"auser Verlag,
3501: Basel, 1998.
3502:
3503: \bibitem{spe:pfaffian}
3504: {\em P.~Speissegger}, The {P}faffian closure of an o-minimal structure, J.
3505: Reine Angew. Math., {\bf 508} (1999), ~189--211.
3506:
3507: \bibitem{vdw:algebra}
3508: {\em B.L. van~der Waerden}, Algebra, Springer Verlag, 1967.
3509:
3510: \bibitem{whi-wat:analysis}
3511: {\em E.~Whittaker and G.~Watson}, A Course of Modern Analysis, Cambridge, 1927.
3512:
3513: \end{thebibliography}
3514:
3515:
3516: \end{document}
3517: