math0608309/p7.tex
1: \documentclass[]{amsart} 
2: 
3: \usepackage{amsmath} 
4: \usepackage{latexsym}
5: %\usepackage[active]{srcltx}
6: \usepackage{amssymb}
7: \ifx\pdftexversion\undefined
8:   \usepackage[dvips]{graphics}
9: \else
10:   \usepackage[pdftex]{graphics}
11: \fi
12: \newtheorem{Proposition}{Proposition}
13:   \newtheorem{Remark}{Remark}
14:   \newtheorem{Corollary}[Proposition]{Corollary}
15:   \newtheorem{Lemma}[Proposition]{Lemma}
16:    \newtheorem{Theorem}[Proposition] {Theorem}
17: \newtheorem{Condition}[Proposition]{Condition}
18: \newtheorem{Conditions}[Proposition]{Conditions}
19: \newtheorem{Definition}[Proposition]{Definition}
20: \newtheorem{Note}[Proposition]{Note}
21: \renewcommand{\theenumi}{\roman{enumi}}
22: 
23: %\newenvironment{proof}{\par\noindent {\sc Proof.}}{$\Box$}
24: \def\mb{\mathbf}
25: \def\z{\noindent}
26: \def\CC{\mathbb{C}}
27: \def\RR{\mathbb{R}}
28: \def\NN{\mathbb{N}}
29: \def\ZZ{\mathbb{Z}}
30: \def\lap{\mathcal{L}}
31: \def\bor{\mathcal{B}}
32: \def\phi{\varphi}
33: \def\bfk{\mathbf{k}}
34: \def\bflam{\boldsymbol{\lambda}}
35: \def\CC{\mathbb{C}}
36:  \def\RR{\mathbb{R}}
37:  \def\NN{\mathbb{N}}
38: \def\ZZ{\mathbb{Z}}
39: \def\bor{\mathcal{B}}
40: \def\lap{\mathcal{L}}
41: \def\sumb{\mathcal{LB}}
42: \def\bchi{\ensuremath{\mbox{\large $\chi$}}}
43: \def\bk{\mathbf{k}}
44: \def\bfC{\mathbf{C}}
45: \def\bj{\mathbf{j}}
46: \def\bfm{\mathbf{m}}
47: \def\bfe{\mathbf{e}}
48: \def\bfbet{\boldsymbol{\beta}}
49: \def\smallmo{\ensuremath{\,\rule{0.05em}{0.7em}\rule{0.5em}{0.05em}\rule{0.05em}{0.7em}}\,}
50: \def\largemo{\ensuremath{\,\rule{0.05em}{0.7em}\rule[0.7em]{0.5em}{0.05em}\rule{0.05em}{0.7em}}\,}
51: 
52: \def\anymo{\ensuremath{\,\rule{0.05em}{0.7em}\rule{0.55em}{0.05em}\hskip-0.55em
53: \rule[0.7em]{0.55em}{0.05em}\hskip-0.05em\rule{0.05em}{0.7em}}\,}
54: 
55: 
56: 
57: \def\smalltr{\,\ensuremath{\rule{0.05em}{0.7em}\rule{0.5em}{0.05em}\rule{0.05em}{0.7em}\rule{0.5em}{0.05em}\rule{0.05em}{0.7em}}\,}
58: 
59: \def\largetr{\,\ensuremath{\rule{0.05em}{0.7em}\rule[0.7em]{0.5em}{0.05em}\rule{0.05em}{0.7em}s
60: \rule[0.7em]{0.5em}{0.05em}\rule{0.05em}{0.7em}}\,}
61: 
62: \def\anytr{\,\ensuremath{\rule{0.05em}{0.7em}\rule{0.55em}{0.05em}\hskip-0.55em
63: \rule[0.7em]{0.55em}{0.05em}\hskip-0.05em\rule{0.05em}{0.7em}
64: \rule{0.55em}{0.05em}\hskip-0.55em
65: \rule[0.7em]{0.55em}{0.05em}\hskip-0.05em\rule{0.05em}{0.7em}}\,}
66: \def\spacetr{\stackrel{\approx}{\mathcal{T}}}
67: \def\erm{\mathrm{e}}
68: \def\mag{\,\mathrm{mag}}
69: \def\dom{\,\mathrm{dom}}
70: 
71: \def\drm{\mathrm{d}}
72: 
73: \begin{document}
74: 
75: 
76: \begin{abstract}
77:   
78:   Transseries in the sense of \'Ecalle are constructed using a
79:   topological approach. A general contractive mapping principle is
80:   formulated and proved, showing the closure of transseries under a
81:   wide class of operations.  
82:   
83:   In the second part we give an overview of results and methods
84:   reconstruction of actual functions and solutions of equations from
85:   transseries by generalized Borel summation with in ordinary and
86:   partial differential and difference equations.
87: 
88: \end{abstract}
89: 
90: \author{O.  Costin } 
91: \address{Department of Mathematics\\ Rutgers University
92: \\ Piscataway, NJ 08854-8019}
93: 
94: \title{Topological construction of transseries and introduction to
95:   generalized Borel summability}\gdef\shorttitle{Trannseries and
96:   summability}
97: 
98: \maketitle
99:  
100: 
101: %\today
102: \tableofcontents
103: \section{Introduction}
104: 
105: 
106: In the first part we give a rigorous and concise construction of a
107: space of transseries adequate for the study of a relatively large
108: class of ordinary differential and difference equations. The main
109: stages of the construction and the notations are similar to those in
110: \cite{EcalleNATO}, but the technical details rely on a generalized
111: contractivity principle which we prove for abstract multiseries
112: (Theorem~\ref{PFp}).  We then discuss how various classes of problems
113: are solved within transseries.  The last part of the paper is devoted
114: to a brief overview of generalized Borel summation, an isomorphism
115: used to associate actual functions to transseries, in the context of
116: ODEs, difference equations and the extension of some of these
117: techniques to PDEs.
118: 
119: 
120: Various spaces of transseries were introduced roughly at the same time
121: in logic (see \cite{VDD}), analysis \cite{Eca84}, \cite{EcalleNATO},
122: and the theory of surreal numbers \cite{Conway}; similar structures
123: were introduced independently by Berry and used as a powerful tool in
124: applied mathematics, see  \cite{Berry-hypB,BerryKeating}.
125: 
126: Informally, transseries are {\em asymptotic\footnote{An asymptotic
127:     expansion is one in which the terms are ordered decreasingly with
128:     respect to the order relation $\ll$, where $f_1(x)\ll f_2(x)$ if
129:     $f_1(x)=o(f_2(x))$}, finitely generated}\index{asymptotic
130:   expansion} combinations powers, logarithms and exponentials. It is a
131: remarkable fact that a wide class of functions can be asymptotically
132: described in terms of them. As Hardy noted \cite{Hardy} ``No function
133: has yet presented itself in analysis the laws of whose increase, in so
134: far as they can be stated at all, cannot be stated, so to say, in
135: logarithmic-exponential terms''. This universality of representation
136: can be seen as a byproduct of the fact that transseries are closed
137: under many common operations.
138: 
139: It is convenient to take $x^{-1}$, $x\to+\infty$, as the small
140: parameter in the transseries. When corresponding functions are
141: considered we change variables so that the limit is again $x\to
142: +\infty$.
143: 
144: A simple
145: example of a transseries generated by $x^{-1}$ and $e^{-x}$ is
146: 
147: $$\sum_{k,m=0}^{\infty}c_{km}e^{-kx}x^{-m}$$
148: where $c_{km}\in\CC$. This transseries is without logarithms or
149: log-free, and has {\em level }one since it contains no iterated
150: exponential; the simplest nontrivial transseries of level two is
151: $e^{e^x}$.  A more complicated example of a transseries
152: \index{transseries} of exponential level two, with level zero {\em
153:   generators} $x^{-1}$ and $x^{-\sqrt{2}}$, level one generator
154: $\exp(x)$, and level two generators $\exp(\sum_{k=0}^{\infty}c_k e^x
155: x^{-k})$ and $ \exp(-e^x)$ is
156: 
157: 
158: $$e^{\sum_{k=0}^{\infty}c_k e^x x^{-k}}+\sum_{k=0}^{\infty}
159: d_kx^{-k\sqrt{2}}+e^{-e^x}$$
160: Some examples of transseries-like expressions for $x\rightarrow
161: +\infty$, which are in fact {\em not} transseries, \index{transseries}
162: are $\sum_{k=0}^{\infty}x^k$ (it fails the asymptoticity condition)
163: and $\sum_{k=0}^{\!\infty}\!e^{-e^{nx}}$ (it does not have finitely
164: many generators, this property is described precisely in the sequel).
165: 
166: \subsection{Abstract multiseries}\index{transseries}
167: The underlying structure behind the condition of asymptoticity is that
168: of {\em well ordering}.  In order to formalize transseries
169: \index{transseries} and study their properties, it is useful to first
170: introduce and study more general abstract expansions, over a well
171: ordered set.
172: 
173: \subsubsection{Totally ordered sets; well ordered sets}
174: Let $A$ be an ordered set, with respect to $\le$. If $x\not\le y$ we write $x>y$ or $y<x$.
175:  $A$ is {\bf totally ordered} if any two elements are
176: comparable, i.e., if for any $x,y\in A$ we have $x\ge y$ or $y\ge
177: x$. If $A$ is not totally ordered, it is called
178: {\bf partially ordered}.
179: 
180: The set $A$ is {\bf well ordered} with respect to $>$ if every
181:   nonempty totally ordered subset ({\em chain}) of $A$ has a
182:   minimal element, i.e.
183: 
184: $$A'\subset A\implies \exists M\in A'\mbox{ such that } \forall
185: x\in A',\,\,M\le x$$
186: 
187: If any nonempty totally ordered subset of $A$ has a {\bf maximal} element, we say that
188: $A$ is well ordered with respect to $<$.
189: \subsubsection{Finite chain property} $A$ has the {\bf finite descending chain} property
190: if there is no {\em infinite strictly decreasing} sequence in $A$, in
191: other words if $f:\NN\mapsto A$ is decreasing, then $f$ is constant for
192: large $n$.
193: 
194: 
195: \begin{Proposition}\label{WOFC}
196: $A$ is well ordered with respect to ``$>$'' {\em iff} it has the finite
197: descending
198: chain property.
199: \end{Proposition}
200: 
201: \begin{proof}
202:   A strictly decreasing infinite sequence is obviously totally ordered
203:   and has no minimal element. For the converse, if there exists
204:   $A'\subset A$ such that $\forall\,x\in A'\,\exists\, y=:f(x)\in
205:   A',\,\,f(x)<x$ then for $x_0\in A'$, the sequence
206:   $\{f^{(n)}(x_0)\}_{n\in\NN}$ is an infinite descending chain in $A$.
207: \end{proof}
208: 
209: {\em Example: multi-indices.} $\NN$ is well ordered with respect to $>$,
210: and so is 
211: 
212: $$\NN^M-\bk_0:=\{\bk\in\ZZ^M:\bk\ge-\bk_0\}$$
213: with respect to the order relation $\bf m\ge n$ iff $m_i\ge
214: n_i\,\forall\,i\ge k$. Indeed an infinite descending sequence $\mathbf{n}_i$
215: would be infinitely descending on at least one component.
216: 
217: 
218: 
219: \begin{Proposition}\label{Va}
220: Let $\bk_0\in\ZZ^M$ be fixed. Any infinite set $A$ in $\NN^M-\bk_0$
221: contains a {\em strictly increasing} (infinite) sequence.
222: \end{Proposition}
223: 
224: \begin{proof}
225:   The set $A$ is unbounded, thus there must exist at least one component
226:   $i\le M$ so that the set $\{m_i:\mathbf{m}\in A\}$ is also unbounded;
227:   say $i=1$. We can then choose a sequence
228:   $S=\{\mathbf{m}_n\}_{n\in\NN}$ so that $(\mathbf{m}_n)_1$ is strictly
229:   increasing. If the set $\{(\mathbf{m}_n)_j: \mathbf{m}\in A;\,\,
230:   j>1\}$ is bounded, then there is a subsequence $S'$ of $S$ so that
231:   $(m_2,...,m_n)_{n'}$ is a constant vector. Then $S'$ is a strictly
232:   increasing sequence. Otherwise, one component, say
233:   $(\mathbf{m}_{n'})_2$ is unbounded, and we can choose a subsequence
234:   $S''$ so that $(m_1,m_2)_{n''}$ is increasing.  The argument continues
235:   in this fashion until in at most $M$ steps an increasing sequence is
236:   constructed.
237: \end{proof}
238: \begin{Corollary}
239:   \label{nononcomp} Any infinite set of multi-indices
240:   in $\NN^M$ contains at least two comparable elements.
241: \end{Corollary}
242: 
243: \begin{Corollary}
244:   \label{Lowerset} Let $A$ be a nonempty set of multi-indices in
245:   $\NN^M-{\bk_0}$.  There exists a unique and {\bf finite} minimizer set
246:   $\mathcal{M}_A$ such that none of its elements are comparable
247: and for any $a'\in A$ there is an $a\in \mathcal{M}_A$
248:   such that $a\le a'$.
249: \end{Corollary}
250: \begin{proof}
251:   Consider the set $C$ of all {\em maximal} totally ordered subsets of
252:   $A$ (every chain is contained in a maximal chain; also, in view of
253:   countability, Zorn's lemma is not needed). Let $\mathcal{M}_A$ be the set of the
254:   least elements of these chains, i.e. $\mathcal{M}_A=\{\min c:c\in C\}$.  Then
255:   $\mathcal{M}_A$ is finite. Indeed, otherwise, by Corollary~\ref{Lowerset} at
256:   least two elements in $\mathcal{M}_A$ such that $a'_1<a'_2$. But this
257:   contradicts the maximality of the chain whose least element was
258:   $a'_2$. It is clear that if $\mathcal{M}'_A$ is a minimizer then $\mathcal{M}'_A\supset
259:   \mathcal{M}_A$. Conversely if $m\in \mathcal{M}'_A\setminus \mathcal{M}_A$ then $m\nless a,\,\,\forall 
260: a\in \mathcal{M}_A$ contradicting the definition of $\mathcal{M}_A$.
261: 
262: 
263: \end{proof}
264: \subsubsection{Definition and properties of abstract multiseries}\label{SFst}
265: If $\mathcal{G}$ is a commutative group with an order relation, we
266: call it an {\bf abelian ordered group} if the order relation is
267: compatible with the group operation, i.e., $a\le A\mbox{ and }b\le
268: B\implies ab\le AB$ (e.g.  $\RR$ or $\ZZ^M$ with addition). Let
269: $\mathcal{G}$ be an abelian ordered group and let $\mu:\ZZ^M\mapsto
270: \mathcal{G}$ be a {\em decreasing group morphism}, i.e.,
271: 
272: 
273: (1) $\mu_{\mathbf{0}}=1$.
274: 
275: (2) $\mu_{\bk_1+\bk_2}=\mu_{\bk_1}\mu_{\bk_2}$.
276: 
277: (3) $\bk_1>\bk_2\implies \mu_{\bk_1}<\mu_{\bk_2}$.
278: Then $\mu(\ZZ)$ is the subgroup finitely generated by
279: $\mu_{\mathbf{e}_j};j=1,...,M$ where $\mathbf{e}_j$ is the
280: unit vector in the direction $j$ in $\ZZ^M$, and since
281: $\mathbf{e}_j>0$ it follows that
282: 
283: $$\mu_{\mathbf{e}_j}<1,\ j=1,...,M$$
284: 
285: In view of our final goal, a simple example to keep in mind is the
286: multiplicative group of {\em monomials} $\mathcal{G}_1$, generated by the functions $x^{-1/2}$, $x^{-1/3}$
287: and $e^{-x}$, for large positive $x$. The order relation on $\mathcal{G}_1$ is
288: $\mu_1<\mu_2$ if $|\mu_1(x)|<|\mu_2(x)|$ for all large $x$.  When $M=3$ 
289: we choose $\mu(\bk)=x^{-k_1/2-k_2/3}e^{-k_3 x}$.
290: 
291: \begin{Remark}\label{R0} The relation $\mu_{\bk_1}=\mu_{\bk_2}$
292:   induces an equivalence relation on $\ZZ^r$; we denote it by $\equiv$.
293: \end{Remark} 
294: For instance in $\mathcal{G}_1$, since $1/2$ and $1/3$ are not rationally
295: independent, there exist distinct $\bk',\bk$ so that
296: $\bk'\in\NN^3:\mu_{\bk'}=\mu_{\bk}$.
297: 
298: \begin{Remark}\label{finigen}
299: Clearly any choice of $\mu_i$ with $\mu_i<1$ for $i=1,...,M$ defines
300: an order preserving  morphism via $\mu(\bk)=\prod_{i=1}^M\mu_{i}^{k_i}$.
301: \end{Remark}
302: 
303: 
304: Ordered morphisms preserve well-ordering:
305: \begin{Proposition}\label{(4)}
306: Let $P\subset\ZZ^M$ be {\em well ordered} (an important
307: example for us is $P=\NN^M-\bk_0$)  and $\mu$ an order preserving
308: morphism. Then $\mu(P)$ is well ordered.
309: \end{Proposition}
310: 
311: \begin{proof}
312:   Assuming the contrary, let $J=\{\bk_n\}_{n\in\NN}$ be such that
313:   $\boldsymbol\mu_J=\{\boldsymbol\mu_{\bk_n}\}_{n\in\NN}$ is an infinite
314:   strictly {\em ascending} chain in $\mu(P)$. Then the index set $J$ is
315:   clearly infinite, and then, by Proposition~\ref{Va} it has a strictly
316:   increasing subsequence $J'$.  Then $\boldsymbol\mu_{J'}$ is a {\em
317:   descending} subsequence of $\boldsymbol\mu_J$, which is a
318:   contradiction.
319: \end{proof}
320: \begin{Corollary}\label{bilateral}
321: For any $\bk\in\ZZ^M$, the set $\{\bk'\in\NN^M-\bk_0:\mu_{\bk'}=\mu_\bk\}$
322: is finite. In particular, given $\bk''$, the set $\{\bk,\bk'\in\NN^M-\bk_0:\bk+\bk'=
323: \bk''$ is finite.
324: \end{Corollary}
325: \begin{proof}
326:   By Proposition~\ref{Va}, the contrary would imply the existence of
327:   an strictly increasing subsequence of $\bk'$, for which then
328:   $\{\mu_{\bk'}\}$ would be strictly decreasing, a contradiction. The
329:   last part follows if we take $\mu_{\bk}=\bk$.
330: \end{proof}
331: 
332: \begin{Proposition}\label{2.9}
333:   The space of formal series\footnote{i.e.  the space of real or complex
334:     {\em functions} on $\mu(\NN^M-\bk_0)$ with usual addition and
335:     convolution\index{convolution}  (\ref{innerprod1}).}
336:     $$\tilde{\mathcal{A}}(\mu_1,...,\mu_M)=
337: \tilde{\mathcal{A}}=\{\tilde{S}=\sum_{\bk\ge
338:     \bk_0}c_{\bk}\mu_{\bk}:
339:  \bk_0\in\ZZ^M,\,c_{\bk}\in\CC\}$$
340: is an algebra with respect to
341: componentwise multiplication by scalars, componentwise
342: addition, and the inner multiplication
343: \begin{equation}\label{innerprod}
344:   \tilde{S} \tilde{S}'=\sum_{\bk\ge \bk_0}\sum_{\bk'\ge
345:   \bk'_0}c_{\bk}c_{\bk'}\mu_{\bk+\bk'}=
346: \sum_{\bk''\ge \bk_0+\bk'_0}\mu_{\bk''}c_{\bk''}
347: \end{equation}
348: where
349: \begin{equation}\label{innerprod1}
350: c_{\bk''}=
351: \mathop{\sum_{\bk\ge \bk_0,\bk'\ge\bk_0}}
352: _{\bk+\bk'=\bk''}c_{\bk}c_{\bk'}
353: \end{equation}
354: The same is true for $\tilde{\mathcal{A}}(\mu_1,...,\mu_M)$ factored by
355: the equivalence relation
356: \begin{equation}
357:   \label{eq:defequi}
358:   \tilde{S}\equiv \tilde{S}'\iff \sum_{\mathbf{k}'\equiv\bk}(c_{\mathbf{k}'}-
359: c'_{\mathbf{k}'})=0\ \forall\,
360: \bk\ge \bk_0
361: \end{equation}
362:  replacing $\bk+\bk'=\bk''$ with $\bk+\bk'\equiv\bk''$ in
363: (\ref{innerprod}) (note that by Corollary~\ref{bilateral} the
364: equivalence classes have finitely many elements).
365: \end{Proposition}
366: \begin{proof}
367:  Straightforward.
368: \end{proof}
369: We define $\tilde{\mathcal{A}}_{\bk_0}$ the linear subspace 
370: of $\tilde{\mathcal{A}}$ for which $\bk_0$ is fixed.
371: \begin{Definition}
372: The sum
373: $$\tilde{S}_c=\sum_{\bk\ge \bk_0}c_{\bk}\mu_{\bk}$$
374:  is in {\em collected} form if, by definition, $c_\bk\ne 0\implies
375: \bk=\max\{\bk':\bk'\equiv\bk\}$, where the maximum is with respect to
376: the lexicographic order. (In other words the coefficients are collected
377: and assigned to the earliest $\mu$ in its equivalence class.)
378: 
379: It is then natural to represent the equivalence class (v.
380: (\ref{eq:defequi})) $\{\tilde{S}\}$ of $\tilde{S}$, in
381: $\tilde{\mathcal{A}}_{\bk_0}/\equiv$,  by $\tilde{S}_c$.
382: \end{Definition}
383: \begin{Corollary}\label{collected}
384: By Proposition \ref{2.9} every nonzero sum can be written in collected form.
385: \end{Corollary}
386: \subsection{Topology on multiseries}
387: A topology
388:  is introduced in the following way:
389: \begin{Definition}
390:   \label{Topology}
391:   {\rm The sequence $\tilde{S}^{(n)}$ in $\tilde{\mathcal{A}}_{\bk_0}$
392:     converges in the asymptotic topology \index{topology}if for any $\bk$,
393:     $c_{\bk}^{(n)}$ becomes  constant (with respect to $n$) eventually.} This
394:   induces a natural topology on $\tilde{\mathcal{A}}_{\bk_0}/\equiv$.
395: \end{Definition}
396: This topology \index{topology}is metrizable. Indeed, any bounded
397: $w:\NN^M\mapsto (0,\infty)$ such that $w(\bk)\rightarrow 0$
398: iff all $k_i\rightarrow \infty$ (e.g. $w(\bk)=\sum_{i\le
399: M}e^{-k_i}$) provides a translation-invariant distance
400: $$d(\tilde{S}^{(1)},\tilde{S}^{(2)})=\sup_{\bk\ge \bk_0}\theta(c_{\bk}^{(1)}-c_{\bk}^{(2)})w(\bk-\bk_0)$$ 
401: where $\theta(x)=0$ if $x= 0$
402: and is one otherwise.  A  Cauchy sequence in
403: $\tilde{\mathcal{A}}_{\bk_0}$ is clearly convergent, and in this sense
404: $\tilde{\mathcal{A}}_{\bk_0}$ is a complete topological algebra.
405:  From this point on, we assume $\mathcal{G}$ is
406: a {\bf totally ordered abelian group}. Let
407: $\tilde{S}\in\tilde{\mathcal{A}}$. 
408: \begin{Remark}
409:   A subgroup of $\mathcal{G}$ generated by $n$ elements $\mu_1<1,...,\mu_n<1$
410: is totally ordered and well ordered, and thus can be indexed by a set
411: of ordinals $\Omega$, in such a way that $\omega_1<\omega_2$
412: implies $\mu_{\omega_1}>\mu_{\omega_2}$. A sum
413: \begin{eqnarray}
414:   \label{eq:asympt-form}
415:   \tilde{S}=\sum_{\omega\in\Omega_{\tilde{S}}}c_{\omega}\mu_{\omega}
416: \end{eqnarray}
417: \z where we agree to omit from $\Omega_{\tilde{S}}$ all ordinals for which
418: $c_\omega=0$ is called the {\em asymptotic} form of $\tilde{S}$.
419: \end{Remark}
420: \begin{Definition} {\bf Dominant term.}
421:   Assume $\tilde{S}=\sum_{\bk\ge\bk_0} c_\bk
422:   \mu_\bk\in\mathcal{A}_{\bk_0}$ is presented in collected form.  The
423:   set $\mu_{\bk}:c_{\bk}\ne 0$ is then totally ordered and must have a
424:   maximal element $\mu_{\bk_1}$, by Proposition~\ref{(4)}.  We say
425:   that $c_{\bk_1}\mu_{\bk_1}=:\dom(\tilde{S})$ is the dominant term of
426:   $\tilde{S}$ and $\mu_{\bk_1}=:\mag(\tilde{S})$ is the {\em
427:     (dominant) magnitude} of $\tilde{S}$ (equivalently,
428:   $\mag(\tilde{S})=\mu_{\min\Omega_{\tilde{S}}}$). We allow for
429:   $\mag(\tilde{S})$ to be zero, {\bf iff} $\tilde{S}=0$.
430: \end{Definition}
431:  The following property is an immediate consequence of
432: Corollar\ref{collected}:
433: \begin{Remark}\label{R2.19}
434: For any nonzero $\tilde{S}$ we can write 
435: $$\tilde{S}=c_{\bk_1}\mag(\tilde{S})(1+\sum c'_{\bk'}\mu_{\bk'})
436: =\dom(\tilde{S})(1+\tilde{S}_1)$$
437: where all the terms in $\tilde{S}_1$ are less than one.
438: \end{Remark}
439:  \begin{Remark}\label{Rfinitude} The magnitude is continuous:
440:   if $\tilde{S}_{\bk}\in\tilde{\mathcal{A}}_{\bk_0}$  and
441:   $\tilde{S}_{\bk}\rightarrow\tilde{S}$ in the
442:   asymptotic topology, then
443: $\mag(\tilde{S}_{\bk})\rightarrow\mag(\tilde{S})$
444: (i.e.  $\exists\bk_1$ so that
445:   $\mag(\tilde{S})=\mag(\tilde{S}_{\bk}),\,\forall \bk\ge \bk_1$).
446: \end{Remark}
447: \begin{proof}
448: This follows immediately from the definition of the topology
449: and of $\mag(\cdot)$.
450: \end{proof}
451: The proposition below discusses the closure of
452: $\tilde{\mathcal{A}}_{\bk_0}$ under restricted {\em infinite sums}.
453: \begin{Proposition}
454:   \label{genclos}
455:  Let $\mathbf{j}_0,\mathbf{k}_0,\mathbf{l}_0\in\ZZ^M$ with
456: $\mathbf{k}_0+\mathbf{l}_0=\mathbf{j}_0$ and consider the sequence
457: in $\tilde{\mathcal{A}}_{\bk_0}$
458: $$\tilde{S}^{(\mathbf{m})}=\sum_{\bk\ge \bk_0}c_{\bk}^{(\mathbf{m})}\mu_{\bk}$$
459: and a fixed $T\in\tilde{\mathcal{A}}_{\mathbf{l}_0}$,
460: $$T=\sum_{\bk\ge \mathbf{l}_0}c'_{\bk}\mu_{\bk}$$
461: Then the sum (``blending'')
462: $$T(\tilde{S}):=\sum_{\bk\ge \mathbf{l}_0}c'_{\bk}\mu_{\bk}\cdot
463: \tilde{S}^{(\mathbf{k})}$$
464: obtained by replacing each $\mu_{\bk}$ in
465: $T$ by the product $\mu_{\bk}\cdot \tilde{S}^{(\mathbf{k})}$ is well
466: defined in $\tilde{\mathcal{A}}_{\bj_0}$ as the limit as
467: $\bk'\uparrow\infty$ of the convergent sequence of {\em truncates}
468: \begin{multline}
469:   T^{[\bk']}(\tilde{S})=\sum_{\mathbf{l}_0\le\bk\le\bk'}c'_{\bk}\mu_{\bk}\cdot
470: \tilde{S}^{(\mathbf{k})}=
471: \sum_{\mathbf{j}\ge\bj_0}\mu_{\mathbf{j}}\sum_{B^{(\bk')}_\mathbf{j}}c'_{\mathbf{m}_1}c_{\mathbf{m}_2}^{(\mathbf{m}_1)}=
472: \sum_{\mathbf{j}\ge\bj_0}d_{\mathbf{j}}^{(\bk')}\mu_{\mathbf{j}}
473: \end{multline}
474: where
475: $$B_{\mathbf{j}}^{(\bk')}=\{\mathbf{m}_1,\mathbf{m}_2:
476: \mathbf{m}_1+\mathbf{m}_2=\mathbf{j},
477: \mathbf{l}_0\le\mathbf{m}_1\le\bk', \ \mathbf{m}_2\ge\bk_0\}
478: $$
479: \end{Proposition}
480: \begin{proof}\label{p} 
481: Given $\mathbf{j}$, the coefficient $d_{\mathbf{j}}^{(\bk')}$ is
482: constant for large  $\bk'$. Indeed, in the expression of
483: $B_{\mathbf{j}}^{(\bk')}$ we have $\mathbf{l}_0\le\mathbf{m}_2=
484: \mathbf{j}-\mathbf{m}_1\le \mathbf{j}-\mathbf{k}_0$ and 
485: similarly $\bk_0\le\mathbf{m}_2\le \mathbf{j}-\mathbf{l}_0$ and
486: therefore there is a bound independent of $\bk'$ on the number of
487: elements in the set $B^{(\bk')}_{\mathbf{j}}$. On the other hand, we
488: obviously have $B^{(\bk')}_{\mathbf{j}}\subset
489: B^{(\bk'')}_{\mathbf{j}}$ if $\bk''>\bk'$. Thus the set
490: $B^{(\bk')}_{\mathbf{j}}$ is constant if $\bk'$ is large enough, and
491: thus $d_{\mathbf{j}}^{(\bk')}$ is constant for all large $\bk'$, which
492: means the sums $T^{[\bk']}(\tilde{S})$ are convergent in the
493: asymptotic topology.
494: \end{proof}
495: {\bf Note}. The condition that $\mag(\tilde{S})^{(m)}$
496:  decreases strictly in $m$ does {\bf not} suffice
497: for $\sum_{m\ge 0}c_m\tilde{S}^{(m)}$ to be well defined. 
498: Indeed, the terms $\tilde{S}^{(m)}=x^{-m}+e^{-x}\in \mathcal{G}_1$ 
499: (cf. \S~\ref{SFst}) have strictly decreasing
500: magnitudes and yet the formal expression $\sum_{m\ge 0}(x^{-m}+e^{-x})$
501: is meaningless.
502: \subsection{Contractive operators}\label{SubContr}
503: \begin{Definition}\label{DFp}
504: Let $J$ be a linear operator from $\tilde{\mathcal{A}}_{\bk_0}$ or from
505: one of its subspaces, to $\tilde{\mathcal{A}}_{\bk_0}$,
506: \begin{equation}
507:   \label{EJlinear}
508:   J\tilde{S} =J\sum_{\bk\ge\bk_0}c_{\bk}\mu_{\bk}=\sum_{\bk\ge\bk_0}c_{\bk}J\mu_{\bk}
509: \end{equation}
510:  Then $J$
511: is called asymptotically contractive\index{asymptotically contractive} on $\tilde{\mathcal{A}}_{\bk_0}$ if
512: \begin{equation}
513:   \label{Econtractformula}
514:  J\mu_{\mathbf{j}}
515: =
516: \sum_{\mathbf{p}>0}c_{\mathbf{j;p}}\mu_{\bj+\mathbf{p}}
517: \end{equation}
518: thus
519: \begin{equation}
520:   \label{Econtractformula1}
521:  J\tilde{S}=\sum_{\bk\ge\bk_0}\mu_{\bk}
522: \mathop{\sum_{\bj+\mathbf{p}\equiv\bk}}_{\mathbf{p}>0;\bj\ge\bk_0}c_{\bj;\mathbf{p}}c_{\bj}
523: \end{equation}
524: \end{Definition}
525: \z We note that by (\ref{Econtractformula}) and
526: Proposition~\ref{genclos}, $J$ is well defined.
527: \begin{Definition}
528:   \label{nonlinear.contrac}
529:   The linear or nonlinear operator $J$ is (asymptotically)
530: contractive\index{asymptotically contractive}
531:   in the set $A\subset\mathcal{A}_{\bk_0}$ if $J:A\mapsto A$ and the
532:   following condition holds. Let $f_1$ and $f_2$ in $A$ be arbitrary and let
533: \begin{eqnarray}
534:     \label{Econtr.notat}
535:  f_1-f_2=\sum_{\bk\ge \bk_0}c_{\bk}\mu_{\bk}
536:   \end{eqnarray}
537:  Then
538:   \begin{eqnarray}
539:     \label{Econtr.nonl}
540:     J(f_1)-J(f_2)=\sum_{\bk\ge \bk_0}c_{\bk}\mu_{\bk}\tilde{S}_{\bk}
541:   \end{eqnarray}
542:  where $\mag(\tilde{S}_{\bk})=\mu_{\mathbf{p_\bk}}$ for some $\mathbf{p_\bk}>0$.\end{Definition}
543: \begin{Remark}\label{Rlinearity}
544:  The sum of asymptotically contractive\index{asymptotically contractive} operators is contractive; the
545:   composition
546: of contractive operators, whenever defined, is contractive.
547: \end{Remark}
548: \begin{Theorem}\label{PFp}
549: (i) If $J$ is linear and contractive\index{asymptotically contractive} on $\tilde{\mathcal{A}}_{\bk_0}$
550:   then for any $\tilde{S}_0\in\tilde{\mathcal{A}}_{\bk_0}$ the fixed
551:   point equation $\tilde{S}=J\tilde{S}+\tilde{S}_0$ has a unique
552:   solution $\tilde{S}\in\tilde{\mathcal{A}}_{\bk_0}$. 
553:   
554:   (ii) In general, if $A\in\mathcal{A}_{\bk_0}$ is closed and
555:   $J:A\mapsto A$ is a (linear or nonlinear) contractive\index{asymptotically contractive} operator on $A$,
556:   then $\tilde{S}=J(\tilde{S})$ has a unique solution is $A$.
557: \end{Theorem}
558: \begin{proof} (i) 
559:   Uniqueness: if we had two solutions $\tilde{S}_1$ and $\tilde{S}_2$ we
560:   would get
561:   $\mag(\tilde{S}_1-\tilde{S}_2)=\mag(J(\tilde{S}_1-\tilde{S}_2))<\mag(\tilde{S}_1-\tilde{S}_2)$
562:   by (\ref{Econtractformula}). We show that $J^n\tilde{S}_0\rightarrow
563:   0$ in the asymptotic topology implying that $\sum_{n=0}^{\infty}J^n
564:   \tilde{S}_0$ is convergent. We have
565:   \begin{align}\label{EJn(delta)}
566:     J^n\tilde{S}=\sum_{\bk\ge\bk_0}\mu_{\bk}
567:  \mathop{\sum_{\mathbf{p_1}>0,...,\mathbf{p}_n>0;\bj\ge\bk_0}}_{\bj+\mathbf{p}_1+\cdots+\mathbf{p}_n\equiv\bk}const_{\mathbf{p_1};...,\mathbf{p_n};\mathbf{j}}
568:   \end{align}
569:   and since $\left|\mathbf{p}_1+\cdots+\mathbf{p}_n\right|>n$ and since the set
570:  $\{\bk':\bk'\equiv\bk\}$ is finite by Corollary~\ref{bilateral}, for
571:  each $\bk$ the condition $\bj+\mathbf{p}_1+\cdots+\mathbf{p}_n\equiv\bk
572:  $ becomes impossible if $n$ exceeds some $n_0$, and then the
573:  coefficient of $\mu_\bk$ in $J^n\tilde{S}$ is zero for $n>n_0$.
574:  
575: (ii) Uniqueness follows in the same way
576:  as in the linear case. For existence, prove the convergence of the
577:  recurrence $f_{n+1}=J(f_n)$.
578:  
579:  With
580:  $f_{n}-f_{n-1}=\delta_{n-1}=\sum_{\bk\ge\bk_0}c_\bk^{(n-1)}\mu_\bk$
581:  we have, for some coefficients $C_{\mathbf{m};\mathbf{m}_1}^{(n-1)}$
582: 
583: \begin{multline}
584:   \label{En->n+1}
585: \delta_n=J(f_{n-1}+\delta_{n-1})-J(f_{n-1})=
586: \sum_{\bk\ge\bk_0}c_\bk^{(n-1)}\mu_\bk\tilde{S}^{(n-1)}_\bk\\=
587: \sum_{\bk\ge\bk_0}\mu_\bk\mathop{\sum_{\mathbf{m}+\mathbf{p}=\bk}}_{\mathbf{m}\ge\bk_0;\mathbf{p}>0}c_{\mathbf{m}}^{(n-1)}
588: C_{\mathbf{m};\mathbf{p}}^{(n-1)}:=\sum_{\bk\ge\bk_0}c_\bk^{(n)}\mu_\bk
589: \end{multline}
590: 
591: \z and therefore $\delta_n$ has an expression similar
592: to (\ref{EJn(delta)}),
593: 
594: 
595: 
596: $$\delta_n=\sum_{\bk\ge\bk_0}\mu_{\bk}
597:  \mathop{\sum_{\mathbf{p_1}>0,...,\mathbf{p}_n>0;\bj\ge\bk_0}}_{\bj+\mathbf{p}_1+\cdots+\mathbf{p}_n\equiv\bk}const_{\mathbf{p_1};...,\mathbf{p_n};\mathbf{j}}$$
598: 
599: \z Consequently $\delta_n\rightarrow 0$ and, as before,
600: it follows that $\sum_n\delta_n$ converges. 
601: \end{proof}
602: \begin{Corollary}\label{compo_n}
603: Let $\tilde{S}\in\mathcal{A}_0$ be arbitrary and
604: $\tilde{S}_n=\sum_{\bk>0}c_{\bk;n}\mu_{\bk}\in\mathcal{A}_0$ for
605: $n\in\NN$. Then the operator defined by
606: 
607: $$J(y)= \tilde{S}+\sum_{n\ge 2}\tilde{S}_{n-2}y^{n}$$
608:   \z is contractive\index{asymptotically contractive} in the set $\{y:\mag(y)<1\}$.
609: \end{Corollary}
610: 
611: \begin{proof}
612:   We have 
613: 
614: $$J(y+\delta)-J(y)=\delta\sum_{n\ge 2}\tilde{S}_{n-2}\left(
615: \sum_{j=1}^{n-1}y^j\delta^{n-j}\right)$$
616: 
617: \end{proof}
618: 
619: \subsubsection{The field of finitely generated formal series}
620: 
621: Let $\mathcal{G}$ be a totally ordered abelian group. We now define the algebra:
622: 
623: 
624: \begin{equation}
625:   \label{eq:defSapprox}
626:    \stackrel{\approx}{S}=\mathop
627: {\mathop{\bigcup_{\bk_0\in\ZZ^M}}_{\mu_1<1,...,\mu_M<1}}_{M\in\NN}
628: \tilde{\mathcal{A}}_{\bk_0}(\mu_1,...,\mu_M)
629: \end{equation}
630: 
631: \z modulo the obvious inclusions, and with the induced topology
632: (convergence in $\stackrel{\approx}{S}$ means convergence in one of the
633: $\tilde{\mathcal{A}}_{\bk_0}(\mu_1,...,\mu_M)$).
634: 
635: 
636: 
637: {\em Product form}. 
638: \begin{Proposition}\label{Pstd}
639:   Any $\tilde{S}\in \stackrel{\approx}{S}$ can be written in the form
640:   \begin{equation}
641:     \label{Estd}
642:     c\,\mag(\tilde{S})\left(1+\sum_{\bk> 0}c_{\bk}\mu_{\bk}\right)
643:   \end{equation}
644:   
645:   \z i.e., 
646: \begin{equation}
647:     \label{Estd2}
648:     c\,\mag(\tilde{S})(1+\tilde{S}_1)
649:   \end{equation}
650:   
651: \z where $\tilde{S}_1\in\tilde{\mathcal{A}}_{\bk_0}$ for some
652: $\bk_0> 0$ (cf. also Remark \ref{R2.19}) and $\mag(\tilde{S}_1)<1$. 
653: \end{Proposition}
654: \begin{proof}
655:   We have, by Remark \ref{R2.19},
656:  \begin{align}\label{caseI}
657:   \tilde{S}=
658: c\,\mag(\tilde{S})\left(1+\sum_{\bk\ge \bk_0}c'_{\bk}\mu_1^{k_1}\cdots
659: \mu_M^{k_M}\mag(\tilde{S})^{-1}\right)
660: \end{align}
661: 
662: \z where all the elements in the last sum are less than one.
663: 
664:   Let $A$ be the set of multi-indices in the sum in (\ref{caseI}) for which {\em some} $k_i <0$ and let $A'$   be
665: its minimizer in the sense of Corollary~\ref{Lowerset}, a finite set. We now
666: consider the extended set of generators
667: 
668: $$\{\overline{\mu}_i:i\le M'\}:=\{\nu\mag(\tilde{S})^{-1}:\nu=\mu_i\mbox{ with }i\le M\mbox{ or
669:   }\nu=\mu_\bk\mbox{ with }\bk\in A'\}$$
670: 
671: We clearly have $\overline{\mu}_i<1$. By the definition of $A'$, for any
672: term in the sum in (\ref{caseI}) either $\bk>0$ or
673: else $\bk=\bk'+\bk''$ with $\bk'\in A'$ and $\bk''\ge 0$. In both cases we
674: have  $c_{\bk}\mu_{\bk}=c_{\bk'}\overline{\mu}_{\bk'}$ with
675: $\bk'\in\ZZ^{M'}$ and $\bk'>0$. Thus $\tilde{S}$ can be rewritten in the
676: form
677: 
678: $$c\,\mag(\tilde{S})\left(1+\sum_{\bk> 0}c_{\bk}\mu_{\bk}\right)\ \
679: (\bk\in\NN^{M'})$$
680: 
681: \z where the assumptions of the Proposition are satisfied.
682: 
683: 
684: \end{proof}
685: \begin{Proposition}
686:   \label{Field}
687:  $\stackrel{\approx}{S}$ is a field.
688: \end{Proposition}
689: 
690: \begin{proof}
691:   The only property that needs verification is the existence of a
692:   reciprocal for any nonzero $\tilde{S}$. Using Proposition~\ref{Pstd}
693:   we only need to consider the case when
694: $$\tilde{S}=1+\sum_{\bk>0}c_{\bk}\mu^\bk$$
695: 
696: \z Since multiplication by $t$ is manifestly contractive\index{asymptotically contractive} (see
697: \S~\ref{SubContr}), $\tilde{S}^{-1}$ is the solution (unique by
698: Proposition~\ref{PFp}) of
699: 
700: $$\tilde{S}^{-1}=1-t\tilde{S}^{-1}$$
701: 
702: \end{proof}
703: 
704: {\em Closure under infinite sums.} 
705: \begin{Corollary}
706:   \label{infisum} (i) Let $\bk_0>0$ and
707: $\tilde{S}\in\tilde{\mathcal{A}}_{\bk_0}$ and $\{c_n\}_{n\in\NN}\in\CC$
708: be any sequence. Then
709: 
710: $$\sum_{n=0}^{\infty} c_n \tilde{S}^n\in\tilde{\mathcal{A}}_{\bk_0}$$
711: 
712: (ii) More generally,
713: if $\tilde{S}_{01},...,\tilde{S}_{0M}$ are of the form $\tilde{S}_0$ and
714: $\{c_{\bk}\}_{\bk\in\NN^M}$ is a multi-sequence of constants, then
715: $\sum_{\bk\ge 0}c_\bk \tilde{S}_0^{\bk}=\sum_{\bk\ge 0}c_\bk
716: \tilde{S}_{01}^{k_1}\cdots \tilde{S}_{0M}^{k_M}$ is well defined.
717: \end{Corollary}
718: 
719: \begin{proof}
720: For some $C\in\CC$ we have $\tilde{S}=C\mag(\tilde{S})(1+t)=C(1+t)\mu$.
721: Since $\mu<1$ we have $\sum_{n=0}^{\infty} c_n
722: C^n\mu^n\in\tilde{\mathcal{A}}_0(\mu) $  and by Proposition~\ref{genclos}
723: $$\sum_{n=0}^{\infty} c_n \tilde{S}^n=
724: \sum_{n=0}^{\infty} c_n C^n\mu_{n\bk_1}(1+t)^n\in\stackrel{\approx}{S}$$
725: \end{proof}
726: 
727: {\em Formal series with real coefficients. Order relation}. Let
728: $\stackrel{\approx}{S}_\RR$ the subfield consisting in $\tilde{S}\in
729: \stackrel{\approx}{S}_\RR$ which have real
730: coefficients.  We say that $\stackrel{\approx}{S}_\RR\ni \tilde{S}>0$ if
731: $\dom(\tilde{S})/\mag(\tilde{S})>0$. Then every nonzero $\tilde{S}\in \stackrel{\approx}{S}_\RR$ is
732: either positive or else $-\tilde{S}$ is positive. This induces a {\em total}
733: order relation on $\stackrel{\approx}{S}_\RR$, by writing $\tilde{S}_1>\tilde{S}_2$ if
734: $\tilde{S}_1-\tilde{S}_2>0$.
735: 
736: 
737: \vfill\eject
738: \subsection{Inductive construction of logarithm-free transseries \index{transseries}}\label{Slogfree}
739: \subsubsection{Transseries}
740: 
741: Transseries are constructed as a special instance of abstract series
742: in which the ablelian ordered group is constructed inductively.
743: 
744: In constructing spaces of transseries, one aims at constructing
745: differential fields containing $x^{-1}$, closed under all operations
746: of importance for a certain class of problems operations. Smaller
747: closed spaces can be endowed with better overall properties.
748: 
749: The construction presented below differs in a number of technical
750: respects from the one of \'Ecalle, and the transseries space
751: constructed here is smaller than his. Still some of the construction
752: steps and the structure of the final object are similar enough to
753: \'Ecalle's, to justify using his terminology and notations.
754: 
755: \subsubsection{\'Ecalle's notation} 
756: 
757: \begin{itemize}
758: \item $\smallmo$ ---small transmonomial.
759: \item  $\largemo$ ---large transmonomial.
760: \item $\anymo$ ---any transmonomial, large or small.
761: \item
762: $\smalltr$ ---small transseries. \index{transseries}
763: \item
764: $\largetr$ ---large transseries. \index{transseries}
765: \item
766: $\anytr$
767: ---any transseries, \index{transseries} small or large.
768: 
769: \end{itemize}
770: 
771: 
772: \subsubsection{Level 0: power series}\label{Level0} Let $x$  be large and positive
773: and let $\mathcal{G}$ be the totally ordered multiplicative group
774: $(x^{\sigma},\,\cdot, \,\ll), \sigma\in\RR$, with $x^{\sigma_1}\ll
775: x^{\sigma_2}$ if $x^{\sigma_1} =o(x^{\sigma_2})$ as
776: $x\rightarrow\infty$, i.e., if $\sigma_1<\sigma_2$. The space of level
777: zero log-free transseries \index{transseries} is by definition
778: $\tilde{\mathcal{T}}^{[0]}=\stackrel{\approx}{S}(\mathcal{G})$.  By
779: Proposition~\ref{Field}, $\tilde{\mathcal{T}}^{[0]}$ is a field.
780: 
781: 
782: If  $\tilde{T}\in\tilde{\mathcal{T}}^{[0]}$, then $\tilde{T}=\anymo$
783: iff $\tilde{T}=x^{\sigma}$ for some $\sigma\ne 0$, $\tilde{T}=\largemo$
784: if $\sigma>0$ and $\tilde{T}=\smallmo$ if $\sigma<0$.
785: 
786: The general element of $\tilde{\mathcal{T}}^{[0]}$ is a level zero
787: transseries \index{transseries}, $\anytr^{[0]}$ or $\anytr$ in short.  We have
788: 
789: \begin{equation}
790:   \label{defanytr}
791:   \anytr=\sum_{\bk\ge \bk_0}c_{\bk}\smallmo^{\bk}
792: \end{equation}
793: 
794: 
795: There are  two order
796: relations: $<$ and $\ll$ on $\tilde{T}\in\tilde{\mathcal{T}}^{[0]}$. 
797: We have $\anytr_1\ll \anytr_2$ iff $\mag(\anytr_1)\ll \mag(\anytr_2)$
798: (the sign of the leading coefficient is immaterial)
799: and $\anytr> 0$ if ($\anytr\ne 0$ and) the real number 
800: $\dom(\anytr)/\mag(\anytr)$ is positive.
801: 
802: \begin{Definition}
803:   \label{smalllarge}
804: {\rm A transseries \index{transseries} is {\bf small}, i.e. $\anytr=\smalltr$ iff in
805: (\ref{defanytr}) we have $c_{\bk}=0$ whenever $\smallmo^{\bk}\not\ll 1$.
806: Correspondingly, transseries \index{transseries} is {\bf large}, i.e. $\anytr=\largetr$ iff
807: in (\ref{defanytr}) we have $c_{\bk}=0$ whenever $\smallmo^{\bk}\not\gg
808: 1$. We note that $\anytr=\smalltr$ iff $\mag(\anytr)\ll 1$ (there is an
809: asymmetry: the condition $\mag(\anytr)\gg 1$ does {\em not} imply
810: $\anytr=\largetr$, since it does not prevent the presence of small
811: terms in $\anytr$). } Any transseries \index{transseries} can then be written uniquely as
812: 
813: \begin{multline}
814:   \label{Edecsmalllarge}
815:   \anytr=\sum_{\bk\ge\bk_0}c_{\bk}\smallmo^{\bk}
816: =\sum_{\bk\ge\bk_0;\ \smallmo^{\bk}>1}c_{\bk}\smallmo^{\bk}
817: +const+\sum_{\bk\ge\bk_0;\ \smallmo^{\bk}<1}c_{\bk}\smallmo^{\bk}\\
818: =\largetr+const+\smalltr:=L(\anytr)+C(\anytr)+s(\anytr)
819: \end{multline}
820: \end{Definition}
821: 
822: \subsubsection{Level 1: Exponential power series}\label{Level1} The set $\mathcal{G}^{[1]}$
823: of {transmonomials of exponentiality one} consists by definition in the
824: formal expressions
825: 
826: $$\anymo^{[1]}=\anymo^{[0]}\exp({\largetr^{[0]}}),\ \
827: \anymo^{[0]},\largetr^{[0]}\in \tilde{\mathcal{T}}^{[0]}$$
828: 
829: \z where we allow for $\largetr^{[0]}=0$ and set $\exp(0)=1$. With
830: respect to the operation
831: 
832: $$\anymo_1^{[0]}\exp(\largetr_1^{[0]})
833: \anymo_2^{[0]}\exp(\largetr_2^{[0]})=
834: (\anymo_1^{[0]}\anymo_2^{[0]})\exp(\largetr_1^{[0]}+\largetr_2^{[0]})$$
835: \z we see that $\mathcal{G}^{[1]}$ is a commutative group. 
836: 
837: The order relations are introduced in the following way.
838: 
839: \begin{multline}\label{Eorderrel}
840:   \anymo_1\exp(\largetr_1)\gg
841:   \anymo_2\exp(\largetr_2)\\ \mbox{ iff }
842:   \left(\largetr_1>\largetr_2\right)\mbox{ or }
843:   \left(\largetr_1= \largetr_2\mbox{ and
844:       }\anymo_1\gg\anymo_2\right)
845: \end{multline}
846: 
847: \z In particular, if $\largetr_1^{[0]}$ is positive, then
848: $\anymo_1^{[0]}\exp(\largetr_1^{[0]})\gg 1$. 
849: 
850: 
851: The second order relation, $>$, is defined by 
852: 
853: $$\anymo_1^{[0]}\exp(\largetr_1^{[0]})>0\iff\anymo_1^{[0]}>0$$
854: 
855: It is straightforward to check that $(\mathcal{G}^{[1]},\cdot,\gg)$
856: is an abelian ordered group. The abelian ordered
857: group of zero level monomials,  $(\mathcal{G}^{[0]},\cdot,\gg)$,
858: is naturally identified with the set of transmonomials for which
859: $\largetr^{[0]}=0$.
860: 
861: The space $\tilde{\mathcal{T}}^{[1]}$ of level one transseries
862: \index{transseries} is by definition
863: $\stackrel{\approx}{S}(\mathcal{G}^{[1]})$.  By Proposition~\ref{Field},
864: $\tilde{\mathcal{T}}^{[1]}$ is a field.  By construction, the space
865: $\tilde{\mathcal{T}}^{[0]}$ is embedded in
866: $\tilde{\mathcal{T}}^{[1]}$. Formula (\ref{defanytr}) is the general
867: expression of a level one transseries \index{transseries}, where now
868: $\smallmo$ is a transmonomial of level one.  The two order relations on
869: transseries \index{transseries} are the ones induced by transmonomials,
870: namely
871: 
872: \begin{equation}
873:   \label{Eord2}
874:   \anytr\gg 1\iff \mag(\anytr)\gg 1  \ \mbox{ and }\ \anytr> 0 \iff \dom(\anytr)/\mag(\anytr)>0
875: \end{equation}
876: 
877: 
878: 
879: 
880: \subsubsection{Induction step: level $n$ transseries \index{transseries}} Assuming the
881: transseries \index{transseries} of level $\le n-1$ are constructed, transseries \index{transseries} of level $n$
882: together with the order relation, are constructed exactly as in
883: \S~\ref{Level1}, replacing $[0]$ by $[n-1]$ and $[1]$ by $[n]$. 
884: The group $\mathcal{G}^{[1]}$ of transmonomials of order at most $n$
885: consists in expressions of the form
886: 
887: 
888: \begin{equation}
889:   \label{Ecannon}
890:   \anymo^{[n]}=x^{\sigma}\exp(\largetr^{[n-1]})
891: \end{equation}
892: 
893: \z where $\largetr^{[n-1]}$ is either zero or a large transseries of
894: level $n-1$ with the multiplication:
895: 
896: \begin{equation}
897:   \label{multcannon}
898:   x^{\sigma_1}\exp(\largetr_1^{[n-1]})x^{\sigma_2}\exp(\largetr_2^{[n-1]})
899: =x^{\sigma_1+\sigma_2}\exp(\largetr_1^{[n-1]}+\largetr_2^{[n-1]})
900: \end{equation}
901: 
902: \z The order relation is given by 
903: 
904: 
905: \begin{align}
906:   \label{ineqcannon}
907:   x^{\sigma_1}\exp(\largetr_1^{[n-1]})\gg
908: x^{\sigma_2}\exp(\largetr_2^{[n-1]})
909: \iff \\
910: \Big(\largetr_1^{[n-1]}>\largetr_2^{[n-1]}\Big) \ \mbox{or} \ 
911: \Big(\largetr_1^{[n-1]}=\largetr_2^{[n-1]} \ \mbox{and}\ \sigma_1>\sigma_2\Big) 
912: \end{align}
913: 
914: 
915: \begin{equation}
916:   \label{defanytrn}
917:   \anytr^{[n]}=\sum_{\bk\ge \bk_0}c_{\bk}(\smallmo^{[n]})^{\bk}
918: \end{equation}
919: 
920:  As
921: in \S~\ref{Level1}, $\tilde{\mathcal{T}}^{[n-1]}$ is naturally embedded in
922: $\tilde{\mathcal{T}}^{[n]}$.
923: 
924: 
925: \subsubsection{General log-free transseries \index{transseries}, 
926:   $\stackrel{\approx}{\mathcal{T}}$} This is the space of arbitrary
927: level transseries, the inductive limit of the finite level spaces of
928: transseries \index{transseries}:
929: 
930: $$\stackrel{\approx}{\mathcal{T}}=\bigcup_{n=0}^{\infty}
931: \tilde{\mathcal{T}}^{[n]}$$
932: 
933: \z Clearly $\stackrel{\approx}{\mathcal{T}}$ is a field.  The order
934: relation is the one inherited from $\tilde{\mathcal{T}}^{[n]}$. The
935: topology \index{topology}is also that of an inductive limit, namely a sequence
936: converges iff it converges in $\tilde{\mathcal{T}}^{[n]}$ for {\em
937:   some} $n$.
938: 
939: \subsubsection{Further properties of transseries}
940: 
941: 
942: 
943: 
944: 
945: \z {\em Definition}. The level $l(\anytr)$ of $\anytr$ is $n$ if
946: $\anytr\in\tilde{\mathcal{T}}^{[n]}$ and
947: $\anytr\not\in\tilde{\mathcal{T}}^{[n-1]}$.
948: 
949: 
950: 
951: 
952: \begin{Proposition}
953:   \label{Rform0}
954:   
955:   If $n=l(\largetr_1)>l(\largetr_2)$  then
956:   $\largetr_1\gg\largetr_2$.
957: 
958: \end{Proposition}
959: \begin{proof}
960:  We may clearly take $n\ge 1$.  Since (by definition)
961: $\largetr\gg 1$ we must have, in
962:  particular,
963: $\dom(\largetr)=cx^{\sigma}\exp(\largetr')$ with 
964: $\largetr'\ge 0$. 
965: By induction, and the assumption $l(\largetr_1)=n$ we must have 
966: $\largetr'_1>0$ and 
967: $l(\largetr_1')=n-1$. The proposition follows since,
968: by again by the induction step, $\largetr_1'\gg\largetr_2'$.
969: \end{proof}
970: \begin{Remark}
971:   \label{Rform}
972: 
973: If $\anymo$ is of level no less than $1$, then either $\anymo$ is large, and
974: then 
975: $\anymo\gg x^\alpha,\,\forall\alpha\in\RR$ or else  $\anymo$ is small, and then
976: $\anymo\ll x^{-\alpha},\,\forall\alpha\in\RR$.
977: 
978: \end{Remark}
979: 
980: \begin{Remark}
981:   \label{Rfinige1}
982:   
983:   We can define  {\em generating monomials} of $0\ne \anytr\in\tilde{\mathcal{T}}^{[n]}$  a minimal
984:   subgroup $\mathcal{G}=\mathcal{G}(\anytr)$ of ${\mathcal{G}}^{[n]}$
985:   with the following properties: 
986: \begin{itemize} \item $\anytr\in
987:   \stackrel{\approx}{S}(\mathcal{G})$; \item $x^\sigma_1
988:   \exp(\largetr_1)\in \mathcal{G}$ implies $x^\sigma_1 \in \mathcal{G}$
989:   and, if $\largetr_1\ne 0$, then $\mathcal{G}\supset \mathcal{G}(\largetr_1)$.
990: \end{itemize}
991: 
992: \z By induction we see that $\mathcal{G}(\anytr)$ is finitely generated
993: for any $\anytr\in \mathcal{T}^{[n]}$.
994: 
995: \end{Remark}
996: 
997: \subsubsection{Closure of $\stackrel{\approx}{\mathcal{T}}$ 
998:   under composition and differentiation} 
999: \begin{Proposition}
1000:   \label{Pdifferentiation}
1001: $\stackrel{\approx}{\mathcal{T}}$ and $\mathcal{T}^{[n]};\ n\in\NN$ are differential fields.
1002: \end{Proposition}
1003: 
1004: \begin{proof}
1005: Differentiation
1006: ${\mathcal{D}}=\frac{\mathrm{d}}{\mathrm{d}x}$
1007: is introduced
1008: inductively on $\stackrel{\approx}{\mathcal{T}}$, as term by term
1009: differentiation, in the following way. Differentiation in
1010: $\tilde{\mathcal{T}}^{[0]}$ is defined as:
1011: 
1012: 
1013: \begin{equation}
1014:   \label{defdif0}
1015:   {\mathcal{D}}\anytr=\sum_{\bk\ge \bk_0}c_{\bk}{\mathcal{D}}\smallmo^{\bk}
1016: \end{equation}
1017: 
1018: \z where, as mentioned in \S\ref{Level0} we have
1019: $\smallmo=x^{-\sigma}$ for some $\sigma\in\RR^+$ and, in the natural
1020: way, we set ${\mathcal{D}} x^{-\sigma}=-\sigma x^{-\sigma-1}$. This makes
1021: ${\mathcal{D}}\anytr\in \tilde{\mathcal{T}}^{[0]}$, and the generating
1022: transmonomials of 
1023: ${\mathcal{D}} \anytr$ are those of $\anytr$ together with $x^{-1}$.
1024: 
1025: We assume by induction that differentiation
1026: ${\mathcal{D}}:\tilde{\mathcal{T}}^{[n-1]}\mapsto
1027: \tilde{\mathcal{T}}^{[n-1]}$ has been defined for all transseries \index{transseries} of
1028: level at most $n-1$. (In particular,
1029: ${\mathcal{D}}\anytr$ is finitely generated.) We define 
1030: 
1031: \begin{multline}
1032:   \label{def:deriv:n}
1033:   {\mathcal{D}}
1034:   \left(\anymo^{[n]}\right)={\mathcal{D}}\left(x^{\sigma}\exp(\largetr^{[n-1]})\right)=
1035: \sigma x^{\sigma-1}\exp(\largetr^{[n-1]})
1036: \\+x^{\sigma}{\mathcal{D}}\largetr^{[n-1]}\exp(\largetr^{[n-1]})
1037: \end{multline}
1038: 
1039: \z A level $n$ transseries \index{transseries} is 
1040: 
1041: \begin{equation}
1042:   \label{defanytrn2}
1043:   \anytr=\sum_{\bk\ge \bk_0}c_{\bk}\smallmo^{\bk}
1044: =\sum_{\bk\ge \bk_0}c_{\bk}\prod_{j=1}^{M}\smallmo_j^{k_j}
1045: \end{equation}
1046: 
1047: \z and we write in a natural way
1048: \begin{multline}
1049:   \label{defanytrnp}
1050:   {\mathcal{D}}\anytr^{[n]}
1051: =\sum_{\bk\ge \bk_0}c_{\bk} \sum_{m=1}^M k_m\smallmo_m^{k_m-1}
1052: {\mathcal{D}} \smallmo_m
1053: \prod_{m\ne j=1}^{M}\smallmo_j^{k_j}\\=
1054: \sum_{m=1}^M\smallmo_m^{-1}\sum_{\bk\ge \bk_0}c_{\bk}  k_m\smallmo_m^{k_m-1}
1055: {\mathcal{D}} \smallmo_m
1056: \prod_{m\ne j=1}^{M}\smallmo_j^{k_j}
1057: \end{multline}
1058: \z and the result follows from the induction hypothesis, since
1059: 
1060: \begin{equation}
1061:   \label{oneterm}
1062:   \sum_{\bk\ge \bk_0}c_{\bk}  k_m \smallmo^{\bk}
1063: \frac{{\mathcal{D}} \smallmo_m}{ \smallmo_m}=\frac{{\mathcal{D}} \smallmo_m}{\smallmo_m}\sum_{\bk\ge \bk_0}c_{\bk}  k_m \smallmo^{\bk}
1064: \in\tilde{\mathcal{T}}^{[n]}
1065: \end{equation}
1066: 
1067: \end{proof}
1068: \begin{Corollary}
1069:   \label{Cfinigen} If $\mathcal{G}_{\anytr}$ is the group (finitely) generated by
1070: all generators in any of the levels of $\anytr$, then ${\mathcal{D}}\anytr$
1071: is generated by the transmonomials of $\mathcal{G}_{\anytr}$ together possibly
1072: with $x^{-1}$. If $\anytr\ne \mathrm{Const}.$ then $l(\anytr)=l(\anytr')$.
1073: \end{Corollary}
1074: \begin{proof}
1075:   Immediate induction; cf. also the beginning of the proof of
1076:   Proposition
1077: ~\ref{Pdifferentiation}.
1078: \end{proof}
1079: The properties of differentiation are the usual ones:
1080: \begin{Proposition}
1081:   \label{Pdiffprop}
1082: ${\mathcal{D}}(fg)=g{\mathcal{D}} f+f{\mathcal{D}} g$, ${\mathcal{D}} const=0$
1083: and ${\mathcal{D}}(f\circ g)=({\mathcal{D}} f)\circ g{\mathcal{D}} g$
1084: (for composition, see \S~\ref{Scompo}).
1085: \end{Proposition}
1086: \begin{proof}
1087:   The proof is straightforward induction.
1088: \end{proof}
1089: 
1090: In the space of transseries, \index{transseries} differentiation is also
1091: compatible with the order relation, a property which is not true in
1092: general, in function spaces. 
1093: 
1094: \begin{Proposition}
1095:   \label{Porrder} For any $\largetr_i, i=1,2$ and $\smalltr_i, i=1,2$ 
1096: we have
1097:   $$\largetr_1\gg\largetr_2 \Leftrightarrow \largetr_1'\gg\largetr_2' $$
1098: $$\smalltr_1\gg\smalltr_2 \Leftrightarrow \smalltr_1'\gg\smalltr_2' $$
1099: $$\largetr_1'\gg\smalltr_1'$$ 
1100: 
1101: \end{Proposition}
1102: \begin{proof}
1103:   The proof is by induction. It is true for power series, which are the
1104:   level zero transseries \index{transseries}. Assume the property holds for transseries \index{transseries} of
1105:   level $\le n-1$, and first prove the result for {\em transmonomials}
1106:   of order $n$, i.e. for the case when for $i=1,2$
1107:   $\anytr_{i}=\anymo_{i}^{[n]}=x^{\sigma_{i}}\exp(\largetr_{i})$, where at least one of $\largetr_{i}$ has  level $n-1$.  
1108: 
1109:  We have to evaluate
1110: 
1111: $$\exp(\largetr_1-\largetr_2)x^{\sigma_1-\sigma_2}
1112: \frac{\largetr_1'+\sigma_1x^{-1}}{\largetr_2'+\sigma_2x^{-1}}$$
1113: and it is plain that we can assume without loss of generality
1114: that $\largetr_1>0$.
1115: 
1116: 
1117: (1) If $l(\largetr_1-\largetr_2)=n-1$ then $M\gg 1$ by
1118: Proposition~\ref{Rform0}. The remaining case is that for $i=1,2$ we have
1119: $l(\largetr_{i})=n-1$ but $\largetr_{i}$ are equal  or else $l(\largetr_1-\largetr_2)\le n-2$
1120: (which obviously requires $n\ge 2$). Let
1121: $\Delta=\largetr_1-\largetr_2$. We have
1122: $l(\Delta)<l(\largetr_1)$ and also $\largetr_1\gg x^{\alpha}$
1123: for some $\alpha>0$, thus by Proposition~\ref{Rform0} and the induction
1124: hypothesis we have
1125: $\largetr_1' \gg \Delta'+\sigma x^{-1}$
1126: $$\largetr_1'+\Delta'+\frac{\sigma_2}{x}=\largetr_1'(1+\smalltr)$$
1127: \z  thus
1128: 
1129: $$\exp(\largetr_1-\largetr_2)x^{\sigma_1-\sigma_2}
1130: \frac{\largetr_1'+\sigma_1x^{-1}}{\largetr_2'+\sigma_2x^{-1}}=
1131: \exp(\largetr_1-\largetr_2)x^{\sigma_1-\sigma_2}(1+\smalltr)\gg 1$$
1132: 
1133: (2) We now let $\anytr$ be arbitrary with the property
1134: $\mag(\anytr)\ne 1$ and  use  Proposition~\ref{Pstd} to write
1135: 
1136: $$\anytr=c\anymo_0+c\sum_{\bk\ge\bk_0}c_\bk\anymo_0\smallmo^\bk
1137: =c\anymo+c\sum_{\bk\ge\bk_0}c_\bk\anymo_\bk$$
1138: 
1139: \z where $\anymo_\bk\ll \anymo_0$ and thus, by step (1) we have
1140: 
1141: $$\anytr'
1142: =c\anymo'+c\sum_{\bk\ge\bk_0}c_\bk\anymo_\bk'=c\anymo'(1+\smalltr)$$
1143: 
1144: \z The rest of the proof is immediate.\end{proof}
1145: 
1146: \begin{Corollary}
1147:   We have ${\mathcal{D}}\anytr=0\iff\anytr=\mathrm{Const.}$
1148: \end{Corollary}
1149: \begin{proof}
1150:   We have to show that if $\anytr=\largetr+\smalltr\ne 0$ then
1151:   $\anytr'\ne 0$. If $\largetr\ne 0$ then (for instance) $\largetr+\smalltr\gg
1152:   x^{-1}=\smalltr$ and then $\largetr'+\smalltr'\gg x^{-2}\ne 0$. If
1153:   instead $\largetr=0$ then $(1/\anytr)=\largetr_1+\smalltr_1+c$ and
1154: we see that $(\largetr_1+\smalltr_1)'=0$ which, by the above, implies
1155: $\largetr_1=0$ which gives $1/\smalltr=\smalltr_1$, a contradiction.
1156: \end{proof}
1157: 
1158: 
1159: \begin{Proposition}
1160:   \label{P dom} Assume $\anytr=\largetr$ or $\anytr=\smalltr$. Then:
1161: 
1162:  
1163: \z (i) If 
1164:  $l(\mag(\anytr))\ge 1$ then
1165: $l(\mag(\anytr^{-1}\anytr'))<l(\mag(\anytr))$.
1166: 
1167: \z (ii) 
1168: $\dom(\anytr')= \dom(\anytr)'(1+\smalltr)$.
1169: 
1170: 
1171: 
1172: \end{Proposition}
1173: 
1174: \begin{proof}
1175:   Straightforward induction.
1176: \end{proof}
1177: 
1178: \subsubsection{Transseries with complex coefficients}\label{complextr}
1179: Complex transseries $\mathcal{T}_{\CC}$ are constructed in a similar
1180: way as real transseries, replacing everywhere $\largetr_1>\largetr_2$
1181: by $\Re\largetr_1>\Re\largetr_2$. Thus there is only one order
1182: relation in $\mathcal{T}_{\CC}$, $\gg$. Difficulties arise when
1183: exponentiating transseries whose dominant term is imaginary.
1184: Operations with complex transseries are then limited.  We will only
1185: use complex transseries in contexts that will prevent these
1186: difficulties.
1187: 
1188: \subsubsection{Differential systems in $\mathcal{T}$}
1189: 
1190: The theory of differential equations in $\mathcal{T}$ is similar to the
1191: corresponding theory for functions.
1192: 
1193: {\em Example}.  The general solution of the differential equation 
1194: 
1195: \begin{equation}
1196:   \label{eqEi1}
1197:   f'+f=1/x
1198: \end{equation}
1199: 
1200: \z in $\mathcal{T}$ (for $x\rightarrow +\infty$)
1201: is $\anytr(x;C)=\sum_{k=0}^{\infty}k!x^{-k}+Ce^{-x}=\anytr(x;0)+Ce^{-x}$.
1202: 
1203: Indeed, the fact that $\anytr(x;C)$ is a solution
1204: follows immediately from the definition of the operations
1205: in $\mathcal{T}$. To show uniqueness, assume $\anytr_{1}$
1206: satisfies  (\ref{eqEi1}). Then $\anytr_2=\anytr_{1}-\anytr(x;0)$
1207: is a solution of ${\mathcal{D}}\anytr+\anytr=0$. Then 
1208: $\anytr_2=e^{x}\anytr$ satisfies ${\mathcal{D}}\anytr_2=0$
1209: i.e., $\anytr_2=\mathrm{Const.}$
1210: 
1211: The particular solution $\anytr(x;0)$ is the unique solution of the
1212: equation $f=1/x-{\mathcal{D}} f$ which is manifestly
1213: contractive\index{asymptotically contractive} in the space of level
1214: zero transseries \index{transseries} (cf. \S~\ref{SubContr}). However
1215: this same equation is not contractive\index{asymptotically
1216:   contractive} for transseries \index{transseries} of positive level,
1217: (because e.g. ${\mathcal{D}} e^x=e^x$); this could also have been
1218: anticipated noting that the solution is not unique.
1219: 
1220: \subsubsection{Restricted composition}\label{Scompo}
1221: The right composition $\anytr_1\circ\anytr_2$ is defined on
1222: $\spacetr$, if $\mag(\anytr_2)\gg 1$ and $\dom(\anytr_2)>0$. The
1223: definition is inductive.  
1224: 
1225: We first define the power and the exponential of a transseries \index{transseries}. Assume
1226: powers and exponentials have been defined for all transseries \index{transseries} of level
1227: $\le n-1$. Let
1228: $\anytr=c\,\mag(\anytr)(1+\smalltr)\in\tilde{\mathcal{T}}^{[n]}$ be any
1229: transseries \index{transseries} such that $ c>0$, cf.  Proposition~\ref{Pstd}.  By the
1230: definition of $\mag(\cdot)$ and (\ref{defanytrn}), $\mag(\anytr)$ is a
1231: transmonomial, $\mag(\anytr)=\largemo^{[n-1]}\exp(\largetr^{[n-1]})$.
1232: We let
1233: 
1234: \begin{multline}
1235:   \label{Edefpower}
1236:   \anytr^\sigma=c^{\sigma}\left(\largemo^{[n-1]}\right)^{\sigma}
1237:   \exp(\sigma\largetr^{[n-1]}) (1+\smalltr)^{\sigma}\\=
1238:   c^{\sigma}\left(\largemo^{[n-1]}\right)^{\sigma}
1239:   \exp(\sigma\largetr^{[n-1]}) (1+\smalltr)^{\sigma}\\=
1240:   c^{\sigma}\left(\largemo^{[n-1]}\right)^{\sigma}
1241:   \exp(\sigma\largetr^{[n-1]})
1242:   \sum_{k=0}^{\infty}\binom{n}{\sigma}\smalltr^{n}
1243: \end{multline}
1244: \z  where   $\binom{n}{\sigma}$    are    the  generalized    binomial
1245: coefficients,  the    infinite     sum   is      well   defined,    by
1246: Proposition~\ref{infisum}, and thus  $\anytr^{\sigma}$ is well defined
1247: as well.  Then, if $\boldsymbol\sigma\in(\RR^+)^M$
1248: and if $\anytr^{[0]}=\sum_{\bk\ge\bk_0}c_\bk x^{-\boldsymbol\sigma\cdot\bk}$
1249: is a level zero transseries \index{transseries}, we write
1250: 
1251: $$\anytr^{[0]}\circ \anytr=
1252: \sum_{\bk\ge\bk_0}c_\bk (\anytr^{-1})^{\boldsymbol\sigma\cdot\bk}$$
1253: 
1254: \z which is well defined by Proposition~\ref{infisum} (ii) and
1255: Proposition~\ref{Pstd}. We note that, under our assumptions for
1256: $\anytr$, $\anytr^{[0]}\circ\anytr>0$ is positive iff $\anytr^{[0]}>0$
1257: 
1258: 
1259: 
1260: 
1261: Similarly, we write cf. Definition~\ref{smalllarge}
1262: 
1263: 
1264: \begin{multline}
1265:   \exp(\anytr)=e^{L(\anytr)+C(\anytr) +s(\anytr)}\\=
1266: e^{C(\anytr)}e^{L(\anytr)}\sum_{k=0}^{\infty}\frac{s(\anytr)^n}{n!}
1267: =C'\anymo^{[n+1]}\anytr^{[n]}
1268: \end{multline}
1269: 
1270: \z well defined by the definition of a transmonomial,
1271: Proposition~\ref{infisum} (ii) and Proposition~\ref{Pstd}. Now the
1272: definition of general composition is straightforward induction. We
1273: assume that composition is defined at all $\le n-1$ levels, and that
1274: in addition $\anytr^{[n-1]}\circ\anytr>0$ if $\anytr^{[n-1]}>0$. Then
1275: 
1276: \begin{multline}
1277:   \anytr^{[n]}\circ\anytr=\sum_{\bk\ge\bk_0}c_{\bk}(\smallmo^{[n]}
1278:   \circ\anytr)^{\bk}
1279:   \\=\sum_{\bk\ge\bk_0}c_{\bk}(\anymo^{[n-1]}\circ\anytr)^{\bk}
1280:   \exp(-\largetr^{[n-1]}\circ\anytr)\\
1281: =\sum_{\bk\ge\bk_0}c_{\bk}(\anymo^{[n-1]}\circ\anytr)^{\bk}
1282:   \left[\exp(-L(\largetr^{[n-1]}\circ\anytr)C\exp(-s(\largetr^{[n-1]}\circ\anytr)\right]\\
1283: =\sum_{\bk\ge\bk_0}c'_{\bk}(\anymo^{[n-1]}\circ\anytr)^{\bk}
1284:   \smallmo^{[n]}\exp(-s(\largetr^{[n-1]}\circ\anytr)
1285: \end{multline}
1286: 
1287: \z and the last sum exists by the induction hypothesis and
1288: Proposition~\ref{genclos}.
1289: 
1290: 
1291: \subsection{The space $\mathcal{T}$ of general transseries}
1292: 
1293: We define 
1294: 
1295: \begin{align}
1296:   L_n(x)=\mathop{\underbrace{\log\log ... \log(x)}}_{n\ times}\\
1297: E_n(x)=\mathop{\underbrace{\exp\exp... \exp(x)}}_{n\ times}\\
1298: \end{align}
1299: \z with the convention $E_0(x)=L_0(x)=x$.
1300: 
1301: \z We write $\exp(\ln x)=x$ and then any log-free transseries \index{transseries} can be
1302: written as $\anytr(x)=\anytr\circ E_n(L_n(x))$. This defines right
1303: composition with $L_n$ in this trivial case, as $\anytr_1\circ L_n(x))=(\anytr\circ
1304: E_n)\circ L_n(x):=\anytr(x)$. 
1305: 
1306: More generally, we define $\mathcal{T}$, the space of general transseries \index{transseries}, 
1307: as a set of formal compositions
1308: 
1309: $$\mathcal{T}=\{\anytr\circ  L_n:\anytr\in\spacetr\}$$
1310: 
1311: \z with the algebraic operations (symbolized below by $*$) inherited
1312: from $\spacetr$ by
1313: 
1314: \begin{align}
1315:   \label{defop}
1316: (\anytr_1\circ  L_n)*(\anytr_2\circ  L_{n+k})=
1317: \left[(\anytr_1\circ E_k) *\anytr_2\right]\circ L_{n+k}
1318: \end{align}
1319: 
1320: 
1321: \z and using (\ref{defop}), differentiation is defined by 
1322: $${\mathcal{D}}( \anytr\circ
1323: L_n)=\left[(\prod_{k=0}^{n-1}L_k)^{-1}\right]({\mathcal{D}}\anytr)\circ L_n$$
1324: 
1325: 
1326: \begin{Proposition}
1327:   \label{Pbigfield}
1328: $\mathcal{T}$ is an ordered differential field, closed under restricted composition.
1329: \end{Proposition}
1330: 
1331: \begin{proof}
1332:   The proof is straightforward, by substitution from the results in
1333:   \S~\ref{Slogfree}.
1334: \end{proof}
1335: 
1336: We will denote generically the elements of $\mathcal{T}$
1337: with the same symbols that we used for $\spacetr$.
1338: \begin{Proposition}
1339: \label{Pintegr}
1340: $\mathcal{T}$ is closed under  integration.
1341: 
1342: \end{Proposition}
1343: \begin{proof}
1344:  
1345: The idea behind the construction of ${\mathcal{D}}^{-1}$ is the following: we
1346: first find an  invertible operator $J$ which is {\em to leading order}
1347: $\mathcal{D}^{-1}$; then the equation for the correction will be
1348: contractive\index{asymptotically contractive}.
1349: Let $\anytr=\sum_{\bk\ge\bk_0}\smallmo^\bk\circ L_n$. To unify 
1350: the treatment, it is convenient to use the identity
1351: $$\int_x\anytr(s)ds=\int_{L_{n+2}(x)}\left(\anytr\circ
1352: E_{n+2}\right)(t)\prod_{j\le
1353:   n+1}E_{j}(t)\mathrm{d}t\,=\int_{L_{n+2}(x)}\anytr_1(t)\mathrm{d}t\,$$
1354: 
1355: \z where the last integrand, $\anytr_1(t)$ is a log-free transseries \index{transseries} and
1356: moreover 
1357: 
1358: $$\anytr_1(t)=\sum_{\bk\ge\bk_0}c_{\bk}\smallmo_1^{k_1}\cdots\smallmo_M^{k_M}=
1359: \sum_{\bk\ge\bk_0}c_{\bk}e^{-k_1\largetr_1-...-k_M\largetr_M} $$
1360: 
1361: \z The case $\bk=0$ is trivial and it thus suffices to find
1362: $\mathcal{D}^{-1} e^{-\largetr}$, where $n=l(\largetr)\ge 1$.  Then  $\largetr\gg x^m$ for any $m$ and thus also $\mathcal{D}\largetr\gg
1363: x^m$ for all $m$. Therefore, since
1364: $\mathcal{D}e^{-\largetr}=-(\mathcal{D}\largetr) e^{-\largetr}$ we expect that
1365: $\dom(\mathcal{D}^{-1}e^{-\largetr})=-(\mathcal{D}\largetr)^{-1}e^{-\largetr}$
1366: and we look for a $\Delta$ such that
1367: \begin{equation}
1368:   \label{defDelta}
1369: \mathcal{D}^{-1}e^{-\largetr}=-\frac{e^{-\largetr}}{\mathcal{D}\largetr}(1+\Delta)  
1370: \end{equation}
1371: 
1372: Then $\Delta$ satisfies the equation
1373: 
1374: 
1375: \begin{equation}
1376:   \label{eq:intContrac}
1377:   \Delta=-\frac{\mathcal{D}^2\largetr}{(\mathcal{D}\largetr)^2}-
1378: \frac{\mathcal{D}^2\largetr}{(\mathcal{D}\largetr)^2}\Delta+(\mathcal{D}\largetr)^{-1}\mathcal{D}\Delta
1379: \end{equation}
1380: 
1381: 
1382: \z By Propositions~\ref{Rform0},
1383: Corollary~\ref{Cfinigen} and Proposition~\ref{P dom}, (\ref{eq:intContrac})
1384: is contractive\index{asymptotically contractive} in $\mathcal{T}^{[n]}$. The proof now follows from  Proposition~\ref{genclos}.
1385: 
1386: \end{proof}
1387: 
1388: \z In the following we also use the notation ${\mathcal{D}}\anytr=\anytr'$
1389: and we write $\mathcal{P}$ for the antiderivative $\mathcal{D}^{-1}$
1390: constructed above.
1391: 
1392: 
1393: \begin{Proposition}
1394:   \label{noconst}
1395: $\mathcal{P}$ is an antiderivative without constant terms, i.e, 
1396: 
1397: 
1398: $$\mathcal{P}\anytr=\largetr+\smalltr$$
1399: \end{Proposition}
1400: 
1401: \begin{proof}
1402:   This follows from Proposition~\ref{Rform0}, together with the fact that 
1403: $\Delta,\largetr$ and $\mathcal{D}$ in (\ref{defDelta}) belong to $\mathcal{T}^{[n]}$.
1404: \end{proof}
1405: \begin{Proposition}
1406:   \label{Ppropint}
1407: We have 
1408: \begin{align}
1409:   \label{Epropint}
1410: \mathcal{P}(\anytr_1+\anytr_2)=\mathcal{P}\anytr_1+\mathcal{P}\anytr_2\nonumber\\
1411: (\mathcal{P}\anytr)'=\anytr;\ \mathcal{P}\anytr'=\anytr(0)\nonumber\\
1412: \mathcal{P}(\anytr_1\anytr_2')=\anytr_1\anytr_2-\mathcal{P}(\anytr_1'\anytr_2)\nonumber\\
1413: \anytr_1\gg\anytr_2\implies \mathcal{P}\anytr_1\gg \mathcal{P}\anytr_2\nonumber\\
1414: \anytr>0\implies\mathcal{P}\anytr>0
1415: \end{align}
1416: 
1417: \z where
1418: 
1419:  $$\anytr=\sum_{\bk\ge\bk_0}c_\bk\smallmo^{\bk}\implies\anytr(0)=\sum_{\bk\ge\bk_0;\bk\not =
1420:   0}c_\bk\smallmo^{\bk}$$
1421: 
1422: \end{Proposition}
1423: \begin{proof}
1424:   All the properties are straightforward; preservation of inequalities
1425: uses Proposition~\ref{Porrder}.
1426: \end{proof}
1427: 
1428: 
1429: \begin{Remark}
1430:   \label{Rexpcontrac}
1431: Let $\smalltr_0\in\mathcal{T}$. The operators defined by
1432: 
1433: \begin{align}
1434:   \label{EJ1}
1435: J_1(\anytr)=\mathcal{P}(e^{-x}(\mathrm{Const.}+\smalltr_0)\anytr(x))\\
1436: J_2(\anytr)=e^{\pm x}x^{\sigma}\mathcal{P}(x^{-2}x^{-\sigma}e^{\mp
1437:   x}(\mathrm{Const.}+\smalltr_0)\anytr(x))\label{EJ2}
1438: \end{align}
1439: 
1440: 
1441: \z are contractive\index{asymptotically contractive} on $\mathcal{T}$.
1442: \end{Remark}
1443: 
1444: \begin{proof} For (\ref{EJ1}) it is   enough to show contractivity\index{asymptotically contractive}
1445:   of $\mathcal{P}(e^{-x}\cdot)$. This is a straightforward calculation
1446:   similar to the proof of Proposition~\ref{Pintegr}.  We have for some
1447:   $n$ $\anytr(x)=\sum_{\bk\ge\bk_0}\smallmo^{\bk}(L_n(x))$ where
1448:   $\smallmo_j\in\stackrel{\approx}{\mathcal{T}}$.
1449: \begin{multline}
1450:   \label{Eintg}
1451: \mathcal{P}e^{-x}(\anymo\circ L_n)=\mathcal{P}\left(e^{-E_{n+2}}\prod_{1\le
1452:     j\le
1453:     n+2}E_j\exp(\largetr\circ E_2)\right)\circ L_{n+2}\\=
1454: \left[\frac{e^{-E_{n+2}}\prod_{1\le
1455:     j\le
1456:     n+2}E_j\exp(\largetr\circ E_2)}{-E_{n+2}'+\sum_{0\le
1457:     j\le
1458:     n+1}E'_j+\largetr'\circ E_2'}\left(1+\smalltr\right)\right]\circ L_{n+2}\\ \ll
1459: \prod_{1\le
1460:     j\le
1461:     n+2}E_j\exp(\largetr\circ E_2)
1462: \end{multline}
1463:   
1464: \z The proof of (ii) is similar.
1465: 
1466: \end{proof}
1467: \section{Equations
1468:   in $\mathcal{T}$: examples}
1469: \begin{Remark}
1470:   The general contractivity principle stated in Theorem~\ref{PFp},
1471:   which we have used in proving closure of transseries with respect to
1472:   a number of operations can be used to show closure under more
1473:   general equations. Our main focus is on differential systems.
1474: \end{Remark}
1475: \subsubsection{Nonlinear ODEs in $\mathcal{T}$}
1476: \label{deg1}
1477: We start with an example, a first order equation:
1478: 
1479: \begin{equation}
1480:   \label{Eqord1}
1481: f'=J_1(f)=F_0(x^{-1})-f-\frac{\beta}{x}f-g(x^{-1},f)  
1482: \end{equation}
1483: 
1484: 
1485: \z where 
1486: 
1487: \begin{align}
1488:   \label{Edefterms}
1489: F_0(x^{-1})=\sum_{k\ge 2}\frac{F_{0k}}{x^k}\nonumber\\
1490: g(x^{-1},f)=\sum_{k\ge 0;\ l\ge 1}g_{kl}x^{-k}f^l
1491: \end{align}
1492: 
1493: \z where the sums are assumed to converge and $g_{01}=g_{11}=0$.
1494: 
1495: \z We see that $J_1$ is well defined if $f=\smalltr\in\mathcal{T}$
1496: (cf. Proposition~\ref{genclos}), and it is under this assumption that
1497: we study $J_1$.\footnote{If there are infinitely many nonzero terms in
1498:   the sum in (\ref{Edefterms}), $J_1$ is not in $\mathcal{T}$ if $f\gg
1499:   1$ (since, in this case, $\mag(f^n)$ is unbounded).}
1500: 
1501: 
1502: (1). Solutions of (\ref{Edefterms}) in $\tilde{\mathcal{T}}^{[0]}$.
1503: The equation 
1504: 
1505: \begin{eqnarray}
1506:   \label{Elev0}
1507: f=J_2(f)=-{\mathcal{D}} f+F_0(x^{-1})-\frac{\beta}{x}f-g(x^{-1},f)  
1508: \end{eqnarray}
1509: 
1510: 
1511: 
1512: \z is contractive\index{asymptotically contractive} in $\tilde{\mathcal{T}}^{[0]}$ (this follows
1513: immediately from \S\ref{SubContr}). Thus there exists in
1514: $\tilde{\mathcal{T}}^{[0]}$ a unique solution $\tilde{f}_0$. Since 
1515: (\ref{Elev0}) is also contractive\index{asymptotically contractive} in the subspace of
1516: $\tilde{\mathcal{T}}^{[0]}$
1517: of series of the form $\sum_{k=2}^{\infty}\frac{c_k}{x^k}$ we have
1518: 
1519: 
1520: \begin{eqnarray}
1521:   \label{Eformy0}
1522:   \tilde{f}_0=\sum_{k=2}^{\infty}\frac{c_k}{x^k}
1523: \end{eqnarray}
1524: 
1525: 
1526: {\bf Note.}  The iteration $f_{n+1}=J_1 f_n$, $f_1=x^{-1}$ is convergent
1527: in $\mathcal{T}$ and, if $f_i=\sum_{k=2}^{i}c_k^{[i]}x^{-k}$ then
1528: $c_k^{[i]}=c_k$ for $k\le i$, and this is a very
1529: convenient way to calculate the coefficients $c_i$.
1530: 
1531: 
1532: (2)  Let now $\delta=f-\tilde{f}_0$. Then
1533: 
1534: \begin{align}
1535:   \label{Eeqdelta}
1536: \delta'=-\delta-\frac{\beta}{x}\delta-g(x^{-1},\tilde{f}_0+\delta)
1537: +g(x^{-1},\tilde{f}_0+\delta)\nonumber\\
1538: =-\delta-\frac{\beta}{x}\delta+\sum_{k\ge 0;\ l\ge 1}c_{kl}x^{-k}\delta^l
1539: \end{align}
1540: \z with 
1541: \begin{align}
1542:   \label{Ecoeffdelta}
1543: c_{01}=c_{11}=0
1544: \end{align}
1545: 
1546: 
1547: \z or
1548: 
1549: 
1550: 
1551: \begin{align}
1552:   \label{Eeqdelta1}
1553: \frac{\delta'}{\delta}=-1-\frac{\beta}{x}-\sum_{k\ge
1554:   2}\frac{c_{k1}}{x^k}+
1555: \sum_{k\ge 0;\ l\ge 1}c_{k;l+1}x^{-k}\delta^l
1556: \end{align}
1557: 
1558: 
1559: \z Since by assumption $\delta\ll 1$ we have
1560: $$\ln\delta=C_0-x+\beta\ln x+\sum_{k\ge
1561:   1}\frac{c_{k+1;1}}{kx^k} + x\smalltr (x)$$
1562: 
1563: 
1564: 
1565: \z and thus $\delta\ll \exp(-cx)$ for any $c<1$ so that 
1566: 
1567: $$\ln\delta=C_0-x+\beta\ln x+\sum_{k\ge
1568:   1}\frac{c_{k+1;1}}{kx^k}+\exp(-cx)\smalltr (x)$$
1569: 
1570: \z whence, by composition with $\exp$ we get
1571: 
1572: $$\delta=C_1x^{\beta}e^{-x}\sum_{k\ge
1573:   1}\frac{d_{k+1;1}}{kx^k}+\exp(-cx)\smalltr (x)$$
1574: 
1575: \z Equation (\ref{Eeqdelta1}) implies 
1576: 
1577: \begin{align}\label{Eequint}
1578:   \delta=Cx^{\beta}e^{-x}\tilde{y}_0\exp\left(\int\sum_{k\ge 0;\ l\ge
1579:     1}c_{k;l+1}x^{-k}\delta^l\right);\ \ \left(\tilde{y}_0=\sum_{k\ge
1580:   0}\frac{d_{k+1;1}}{kx^k}\right)
1581: \end{align}
1582: 
1583: 
1584: 
1585: \z and (\ref{Eequint}) is contractive\index{asymptotically contractive} by Remark~\ref{Rlinearity} and
1586: Remark~\ref{Rexpcontrac}.  In particular, for every $C$ there is a
1587: unique $\delta(x;C)$ satisfying (\ref{Eequint}).
1588: 
1589: \begin{Proposition}
1590:   \label{Rspecialform}
1591: The general transseries solution of (\ref{Eqord1}) is $\tilde{f_0}+\delta$
1592: where  
1593: \begin{align}
1594:  \label{Especialform} 
1595: \delta=\sum_{k=1}^{\infty}C^kx^{\beta k}e^{-kx}\tilde{f}_k(x)
1596: \end{align}
1597: 
1598: 
1599: \z with $\tilde{f}_k\in\tilde{\mathcal{T}}^{[0]}$ and
1600: 
1601: $$\tilde{f}_k(x)=\sum_{j=0}^{\infty}\frac{f_{k;j}}{x^j}$$
1602: \end{Proposition}
1603: \begin{proof}
1604:   This is a straightforward consequence of discussion of this section
1605:   and of (\ref{Eequint}) .
1606: \end{proof}
1607: 
1608: 
1609: \subsubsection{Formal linearization}
1610: \label{sec:Lineariz}
1611: Let $z=Cx^{\beta}e^{-x}$. We have $$C(x,\delta)=x^{-\beta}e^{x}\sum_{k\ge
1612:   1}\delta^k\tilde{g}_k(x)$$ A direct calculation
1613: shows that $C'=C_x+C_\delta\delta'=0$. The transformation
1614: $(x\mapsto x;y\mapsto C(x,y-f_0)) $ formally linearizes (\ref{Eqord1}).
1615: 
1616: \subsection{Multidimensional systems: transseries solutions at irregular
1617:   singularities of rank one}\label{Sysnonlin} Consider the differential
1618: system
1619:             
1620: \begin{eqnarray}
1621:  \label{eqor1}
1622:   \mathbf{y}'=\mathbf{f}(x^{-1},\mathbf{y})  \qquad \mathbf{y}\in\CC^n               \end{eqnarray}
1623:    
1624:  
1625:    We look at solutions $\mathbf{y}$ such that $\mathbf{y}(x)\rightarrow
1626:    0$ as $x\rightarrow\infty$ along some direction
1627:    $d=\{x\in\CC:\arg(x)=\phi\}$. The
1628:    following conditions are assumed
1629:  
1630: 
1631: 
1632: 
1633: \z (a1) The function $\mathbf{f}$ is analytic
1634: at $(0,0)$.
1635: 
1636: \z (a2) Nonresonance: the eigenvalues $\lambda_i$ of the linearization
1637: 
1638: \begin{eqnarray}
1639:   \label{linearized}
1640:   \hat{\Lambda}:=-\left(\frac{\partial f_i}{\partial
1641:     y_j}(0,0)\right)_{i,j=1,2,\ldots n}
1642: \end{eqnarray}
1643: 
1644: \z are linearly independent over $\ZZ$ (in particular nonzero) and such
1645: that $\arg\lambda_i$ are different from each other (i.e., the Stokes
1646: lines are distinct; we will require somewhat less restrictive
1647: conditions, see \S~3.1). %\S\ref{nonres}).
1648: 
1649: By relatively straightforward algebra, (\cite{Wasow} and also \S~3.2
1650: where all this is exemplified in a two-dimensional case), the system
1651: (\ref{eqor1}) can then be brought to the form
1652: 
1653: 
1654: \begin{eqnarray}\label{eqor}
1655: {\bf y}'=-\hat\Lambda {\bf y}+ \frac{1}{x}\hat A {\bf y}+{\bf
1656: g}(x^{-1},{\bf y})
1657: \end{eqnarray}
1658: 
1659: \z where $\hat{\Lambda}=\mbox{diag}\{\lambda_i\},\ 
1660: \hat{A}=\mbox{diag}\{\alpha_i\}$ are constant matrices, $${\bf
1661:   g}(x^{-1},{\bf y})= O(x^{-2},\mathbf{y}^2),\ 
1662: (x\rightarrow\infty,\mathbf{y}\rightarrow 0)$$
1663: \begin{Remark}\label{invert}
1664:   (i) If ${\bf g}(x^{-1},{\bf y})\equiv 0$. In this case the system
1665:   (\ref{eqor}) is linear and has the general transseries solution
1666: 
1667: $${\bf y}=\erm^{-x\hat{\Lambda}}\mathbf{C}
1668:   x^{\hat{A}}$$ 
1669: 
1670: (ii) More generally, if ${\bf g}(x^{-1},{\bf y})={\bf G}(x^{-1})$ is a
1671: transseries,
1672: then the general solution of (\ref{eqor}) is
1673: 
1674: \begin{equation}
1675:   \label{eq:inhom}
1676:   {\bf y}=\erm^{-x\hat{\Lambda}}
1677:   x^{\hat{A}}\mathbf{C}+\erm^{-x\hat{\Lambda}}
1678:   x^{\hat{A}}\mathcal{P}\Big(\erm^{x\hat{\Lambda}} x^{-\hat{A}}{\bf
1679:     g}\Big)
1680: \end{equation}
1681: 
1682: \end{Remark}
1683: 
1684: 
1685: \begin{proof}
1686:   In both cases the system is diagonal and the result follows
1687:   immediately from the case when $n=1$, i.e. from
1688:   Proposition~\ref{Rspecialform}.
1689: \end{proof}
1690: 
1691: 
1692: The general solution of (\ref{eqor}) in $\mathcal{T}_{\CC}$  (cf.
1693: \S\ref{complextr}) is an
1694: $n_1\le n$ parameter transseries, as shown in the sequel.
1695: 
1696: 
1697: \begin{Proposition}
1698:   \label{GTS}
1699:   
1700:   Let $d$ be a ray in $\CC$.  The general solution of (\ref{eqor}) in
1701:   $\mathcal{T}_\CC$ with the restriction $\mathbf{y}\ll 1$ is of the
1702:   form
1703: 
1704: \begin{gather}
1705:   \label{transsf} \tilde{\mathbf{y}}(x)=\sum_{\mathbf{k}\ge 0}
1706:   \mathbf{C}^{\mathbf{k}}\erm^{-\boldsymbol{\lambda}\cdot\mathbf{k}x}
1707:   x^{\boldsymbol{\alpha}\cdot\mathbf{k}}\tilde{\mathbf{s}}_{\mathbf{k}}(x)=
1708:   \sum_{\mathbf{k}\ge 0}
1709:   \mathbf{C}^{\mathbf{k}}\erm^{-\boldsymbol{\lambda}\cdot\mathbf{k}x}
1710:   x^{\mathbf{m}_0\cdot\mathbf{k}}\tilde{\mathbf{y}}_{\mathbf{k}}(x)
1711: \end{gather}
1712: 
1713: \end{Proposition}
1714: \z where $C_i= 0$ for all $i$ so that $\erm^{-\lambda_i
1715:   x}\not\rightarrow 0$ as $x\rightarrow\infty$ in $d$.
1716: 
1717: 
1718: 
1719: \begin{proof}
1720: If $\mathbf{y}$ is a solution of (\ref{eqor}) then we have, by
1721: Remark~\ref{invert}
1722: 
1723: \begin{equation}
1724:   \label{eq:invert}
1725:   \mathbf{y}=\erm^{-x\hat{\Lambda}}
1726:   x^{\hat{A}}\mathbf{C}+\erm^{-x\hat{\Lambda}}
1727:   x^{\hat{A}}\mathcal{P}\Big(\erm^{x\hat{\Lambda}} x^{-\hat{A}}\mathbf
1728:     {g}(x^{-1},\mathbf{y})\Big)
1729: \end{equation}
1730: 
1731: \z for some $\mathbf{C}$. Since $\mathbf{y}\ll 1$ we have
1732: $\mathbf{g}(x^{-1},\mathbf{y})\ll 1$ and thus $$\mathcal{P}\Big(\erm^{x\hat{\Lambda}} x^{-\hat{A}}\mathbf
1733:     {g}(x^{-1},\mathbf{y})\Big)\ll \erm^{x\hat{\Lambda}} x^{-\hat{A}}$$ 
1734: Again since $\mathbf{y}\ll 1$, we then have $C_i=0$ for all 
1735: $i$ for which $e^{-\lambda_i x}\not\ll 1$.
1736: 
1737: \z {\bf Note}.  With the condition $\mathbf{y}\ll 1$,
1738: eq. (\ref{eq:invert}) has a unique solution. 
1739: 
1740: Indeed, the difference of
1741: two solutions $\mathbf{y}_1 -\mathbf{y}_2$ satisfies the equation
1742: \begin{equation}
1743:   \label{eq:invert2}
1744:  \mathbf{y}_1
1745: -\mathbf{y}_2=\erm^{-x\hat{\Lambda}}
1746:   x^{\hat{A}}\mathcal{P}\Big(\erm^{x\hat{\Lambda}} x^{-\hat{A}}\big[\mathbf
1747:     {g}(x^{-1},\mathbf{y}_1)-\mathbf
1748:     {g}(x^{-1},\mathbf{y}_2)\big]\Big)
1749: \end{equation}
1750: 
1751: \z Since $\mathbf {g}(x^{-1},\mathbf{y})=O(x^{-2},\mathbf{y}^2)$ we
1752:     have $$\mathbf {g}(x^{-1},\mathbf{y}_1)-\mathbf
1753:     {g}(x^{-1},\mathbf{y}_2)=O(x^{-2}\boldsymbol{\delta},|\mathbf{y}||\boldsymbol{
1754:     \delta}|)$$ which by Proposition~\ref{Ppropint} implies
1755:     $\boldsymbol{\delta}=o(\boldsymbol{\delta})$, i.e., $\delta=0$.
1756: 
1757: 
1758: Using Remark~\ref{Rexpcontrac} it is easy to check that
1759: (\ref{eq:invert2}) is an asymptotically contractive\index{asymptotically contractive} equation, in the
1760: space of $\mathbf{y}$ which are $\ll x^{-2}$ thus it has a solution
1761: $\mathbf{y}^{[0]}$ with this property. Since the previous note shows the
1762: solution of (\ref{eqor}) with $\mathbf{y}\ll 1$ is unique, we have
1763: $\mathbf{y}=\mathbf{y}^{[0]}$. Formula (\ref{transsf}) is obtained
1764: by straightforward iteration of (\ref{eq:invert2}).
1765: 
1766: 
1767: 
1768: \end{proof}
1769: %% sample LaTeX REFERENCES file for English Physics Books Dr. Kölsch
1770: %% to be compiled via the sample file input.tex
1771: %% put together by J. Lenz & V. Wicks, Team Physik 09:46 26.06.96
1772: 
1773: 
1774: 
1775: %%\subsubsubsection{When is a function determined by its local properties at
1776: %%  a point? Functions suitable for local analysis, or analyzability.} Let
1777: %%$f$ be defined in some open set $\mathcal{O}$ in $\RR$ or $\CC$ and let
1778: %%$x_0\in\partial \mathcal{O}$. We can broadly call local the quantities
1779: %%obtained from $f$ as (generalized) limits involving values of $f(x_n)$
1780: %%as $x_n\rightarrow x_0$. which can be defined solely in
1781: %%terms limits that Local properties There are infinitelIf $f$ is analytic
1782: %%at $0$ then $f$ is uniquely determined by $\{f^{(n)}(0):n\in\NN\}$. The
1783: %%absence of analytic
1784: 
1785: %%% Local Variables: 
1786: %%% mode: latex
1787: %%% TeX-master: t
1788: %%% End: 
1789: 
1790: \section{Borel summation techniques} 
1791: \subsubsection{Introduction}
1792: 
1793: 
1794: 
1795: 
1796:  
1797: In this section we discuss a Borel summation--induced isomorphism
1798: between transseries and functions in the setting \S\ref{Sysnonlin}.
1799: The goal is to show, on simple examples, how Borel summation is proved
1800: and used.  For more general results we refer to \cite{Duke}.
1801: 
1802: \begin{Definition}
1803: A Borel-summable series $\tilde y:=\sum_{k=K}^\infty y_kx^{-k}$,
1804: $K\in\ZZ$ is a formal
1805: power series with the following properties
1806: \begin{enumerate}
1807: \item the truncated Borel transform $Y=\mathcal{B}\tilde
1808: y:=\sum_{k>0}\frac{y_{k}}{(k-1)!}t^{k-1}$ of $\tilde y$ has a
1809: nonzero radius of convergence, \item $Y$ can be analytically
1810: continued along $[0,+\infty)$ and \item the analytic continuation
1811: $Y$ grows at most exponentially along $[0,+\infty)$ and is
1812: therefore Laplace transformable along $[0,+\infty)$.
1813: \end{enumerate}
1814: The Borel sum $y$ of $\tilde y$ is then given by
1815: \begin{equation}
1816:   \label{eq:defLB}
1817:   y = \lap\bor\tilde{y}:=\sum_{k=K}^{0}y_{k}x^{-k} +
1818:   \mathcal{L}Y,
1819: \end{equation}
1820: where the sum is understood to be zero if $K>0$ and $\mathcal{L}$
1821: denotes the usual Laplace transform.
1822: \end{Definition}
1823: {\em Example} The formal solution for large $x$ of the equation
1824: $f'-f=x^{-1}$ is Borel summable. Indeed
1825: $$\lap\bor\sum_{k=0}^{\infty}\frac{-k!}{(-x)^{k+1}}\\=\lap\frac{1}{1+p}=\int_0^{\infty}\frac{e^{-px}}{1+p}dp$$
1826: and it can be checked that the Borel sum is a solution of the given
1827: equation; see also (i) in the note below. 
1828: \begin{Note}{\rm 
1829:     (i) It can be shown that Borel summation is an extended
1830:     isomorphism: in particular it commutes with algebraic operations,
1831:     including multiplication and complex conjugation, and with
1832:     differentiation \cite{Balser}. This can be expected from the fact
1833:     that, formally, Borel summation is the composition of Laplace
1834:     transform with its inverse, and the identity obviously commutes
1835:     with the operations mentioned above.
1836:   
1837:   (ii) Borel summation has to be generalized in a number of ways in
1838:   order to be used for solving more general equations. Indeed, an
1839:   equation as simple as $f'+f=x^{-1}$ has a formal solution which is
1840:   not Borel summable: $\bor \sum_{k=0}^{\infty}k! x^{-k-1}=(1-p)^{-1}$
1841:   is not Laplace transformable. If the Laplace transform integral is
1842:   taken along a ray above or below $\RR^+$, then $\lap\bor (1-p)^{-1}$
1843:   is a solution of the given equation. None of these path choices
1844:   however yields a real valued function, whereas the formal series has
1845:   real coefficients. The resulting summation operator would not be an
1846:   isomorphism since commutation with complex conjugation fails. In
1847:   this simple example the half sum of the upper and lower integrals is
1848:   real valued and solves the equation, but this ad hoc procedure
1849: would not work for nonlinear equations. More complicated averages 
1850: need to be introduced, see \S\ref{Ave}. }
1851: \end{Note}
1852:  \subsection{Borel summation of transseries: a first order example}\label{Ord1}
1853:  
1854:  Consider the (first order) differential equation (\ref{Eqord1}) with,
1855:  say, $F_0=x^{-2}$ and $g=a f^2+b f^3$:
1856: 
1857:  \begin{equation}
1858:    \label{eq:eq*}
1859:    f'+(1-\beta x^{-1})f=x^{-2}+af^2+bf^3
1860:  \end{equation}
1861: 
1862: \z  and $a,b$ some constants. We first look at solutions, both
1863:  formal and actual, which go to zero as $x\rightarrow\infty$ in some
1864:  direction in $\CC$.
1865: 
1866: We have seen in \S~\ref{deg1} that the solution $\tilde{f}_0$ as a
1867:  formal power series (i.e., in $\mathcal{T}^{[0]}$), is unique and
1868:  $\tilde{f}_0=\sum_{k=2}^{\infty}c_kx^{-k}$. By Remark~\ref{Rspecialform}
1869: the general transseries solution is 
1870: 
1871: 
1872: \begin{equation}
1873:   \label{eq:gtrns}
1874:   \tilde{f}=\sum_{k=0}^{\infty}\xi^k\tilde{f}_k(x)\ \ \ \ (\xi:=Ce^{-x}x^\beta)
1875: \end{equation}
1876: 
1877: \z where $\tilde{f}_k$ are integer power series.  It is important to
1878: note that in (\ref{eq:gtrns}) only the constant $C$ depends on the
1879: solution. We will see that the $\tilde{f}_k$ are simultaneously Borel
1880: summable, to $f_k=\mathcal{LB}\tilde{f}_k$, and that the sum
1881: 
1882: \begin{equation}
1883:   \label{eq:gtrns,sum}
1884:  f=\sum_{k=0}^{\infty}\xi^k{f}_k(x)
1885: \end{equation}
1886: 
1887: 
1888: \z is {\em convergent} and provides the general solution of
1889: (\ref{eq:eq*}) with the property $f\rightarrow 0$ as
1890: $x\rightarrow\infty$ in some direction in which (\ref{eq:gtrns}) is a
1891: valid complex transseries.
1892: 
1893: 
1894: Assuming for the moment we proved that (\ref{eq:gtrns,sum}) indeed
1895: provides a solution of (\ref{eq:eq*}) for any $C$ it is not difficult
1896: to show that there are no further solutions:
1897: 
1898: \begin{Lemma}
1899:   \label{uniquen}
1900: If $f_1$ is any solution of (\ref{eq:eq*}) with the stated condition 
1901: in some direction at
1902: infinity, 
1903: then $f_1-f_0=Ce^{-x}x^\beta(1+o(1))$ as $x\rightarrow\infty$.
1904: If $C=0$ then $f_1=f_2$.
1905: 
1906: \end{Lemma}
1907: 
1908: 
1909: \begin{proof}
1910:  Writing the equation for $\delta=f_1-f_0$, multiplying
1911: with the integrating factor  of the ``dominant''
1912: part of the equation, $Ce^{t}t^{-\beta}$, and integrating
1913: we get
1914: 
1915:  \begin{equation}
1916:    \label{eq:intequ}
1917:    \delta = Ce^{-x}x^\beta+ e^{-x}x^\beta\int_{a}^x
1918: e^t t^{-\beta}\left[\left(2af_0 +3bf_0^2\right)\delta+(a+3b f_0)\delta^2+\delta^3\right]dt
1919:  \end{equation}
1920:  
1921:  \z which is contractive in the sup norm on $(x_0,\infty)$, for $|x_0|$
1922:  large enough, in a ball of radius $\epsilon>0$ small enough.
1923: The solution of (\ref{eq:intequ}) is thus unique and it is easy
1924: to see that $\delta =C_1e^{-x}x^\beta(1+o(1))$ for large $x$
1925: ($C_1$ is not in general equal to $C$).
1926: \end{proof}
1927: 
1928: \z 
1929: \begin{Corollary}
1930:   Formula (\ref{eq:gtrns,sum}) provides the most general solution
1931: of (\ref{eq:eq*}) with the property $f\rightarrow 0$ in
1932: some direction in $\CC$.
1933: 
1934: \end{Corollary}
1935: 
1936: 
1937:  A convenient way to
1938:  generate $\tilde{f}_0$ is the iteration(\ref{Elev0}), which we start
1939:  with $\tilde{f}_0^{[0]}=0$.  Denoting
1940:  $\delta^{[k]}=\tilde{f}_0^{[k+1]}-\tilde{f}_0^{[k]}$ we have
1941: 
1942: 
1943: 
1944: $$\delta^{[k]}=(-\mathcal{D}-\beta x^{-1}+O(x^{-2}))\delta^{[k-1]}$$
1945: 
1946: \z whence $\delta^{[k]}\sim const. \Gamma(k-\beta)x^{-k}$ and thus
1947: Borel summation appears natural.  Since $\tilde{f}_0$ is defined
1948: through a differential equation it is natural to Borel transform the
1949: equation itself. The result is
1950: 
1951: \begin{equation}
1952:   \label{eq:{Eqord1,B}}
1953:   -pF+F=p-\beta F*1+a F^{*2}+b F^{*3}
1954: \end{equation}
1955: 
1956: \z where convolution is defined by
1957: \begin{equation}
1958:   \label{eq:defConvo}
1959: (  f*g)(p)=\int_0^pf(s)g(p-s)ds
1960: \end{equation}
1961: and we write
1962: $$F^{*k}=\underbrace{F*F*\cdots*F}_{\mbox{k times}}$$
1963: 
1964: 
1965: \z  A solution 
1966: $F$ that is Laplace transformable along any ray {\em other than
1967: $\RR^+$} is obtained by noting that the equation  
1968: \begin{equation}
1969:   \label{eq:{Eqord1,B1}}
1970:  F=(1-p)^{-1}\left(p-\beta F*1+a F^{*2}+b F^{*3})\right)=\mathcal{N}(F)
1971: \end{equation}
1972: 
1973: \z is contractive in an appropriate norm.
1974: 
1975: 
1976: \begin{Proposition}
1977:   \label{C*a}
1978:   The space $L^1_\nu$ of  functions along a ray 
1979: $d=\{p:\arg(p)=a$, such that $\|f\|_\nu<\infty$ with
1980:   the norm $\|f\|_\nu=\int_{t\in d} |f(t)|e^{-\nu |t|}d|t|$
1981:   is a Banach algebra with respect to convolution.
1982: \end{Proposition}
1983: 
1984: \begin{proof}
1985:   All the properties are verified in a straightforward manner. 
1986: In particular, we have $\|f*g\|_\nu=\|f\|_\nu \|g\|_\nu$. 
1987: \end{proof}
1988: 
1989: \z \z Therefore, with $\rho=\|F\|_\nu$ and $d_1=\mbox{dist}(1,d)\ne 0$
1990: we have
1991: 
1992: 
1993: 
1994: \begin{multline}
1995:  \left\| \mathcal{N}(F)\right\|_\nu =\left\|(1-p)^{-1}\left(p-\beta
1996: F*1+a F^{*2}+b F^{*3}\right)\right\|_\nu
1997: \\\le \frac{1}{d_1}\left(\|p\|_\nu+|\beta| \rho
1998:   \|1\|_\nu+ |a|\rho^2+ |b|\rho^3\right)\\=
1999: \frac{1}{d_1}\left(\frac{1}{\nu^2}+\frac{|\beta|\rho}{\nu}
2000: + |a|\rho^2+ |b|\rho^3\right)\rightarrow 0 \ \ \mbox{as}\ \ \nu\rightarrow\infty
2001: \end{multline}
2002: 
2003: \z and clearly, if $\nu$ and $1/\epsilon$ are large enough, the image
2004: $\mathcal{N}B_\epsilon$
2005: of the ball $B_\epsilon=\{F:\|F\|_\nu<\epsilon\}$ is contained 
2006: in $B_\epsilon$. Similarly, it can be seen that $\mathcal{N}$ is contractive
2007: in $B_\epsilon$. Thus the following conclusion.
2008: 
2009:  
2010: \begin{Proposition}
2011:   \label{L1loc}
2012: There is a unique solution of
2013: (\ref{eq:{Eqord1,B}}) in $\cup_{\nu\ge\nu_0}L^1_{\nu}$.
2014: 
2015: \end{Proposition}
2016: 
2017: 
2018: This however does not yet imply Borel summability of the series
2019: $\tilde{f}_0$; to show this we must prove appropriate analyticity 
2020: properties for $F$ and this can be done in essentially the same manner. 
2021: 
2022: \begin{Proposition}
2023:  \label{C*a1}
2024:  The space of analytic functions in a region of the form
2025:   
2026: 
2027: $$S_M=\{p:\arg(p)\in (a_1,a_2))\not\ni 0\ \mbox{and }|p|<M\}\cup\{p:|p|<1-\epsilon\}$$
2028: 
2029: \z vanishing at $p=0$, continuous in $\overline{S}$, and such that
2030: $\|f\|_\nu<\infty$ with the norm
2031: $\|f\|_{\nu;\infty}=M^{-1}\sup_{\overline{S}}|f(p)|e^{-\nu|p|}$ is a
2032: Banach algebra with respect to convolution. In addition, if $g(\cdot
2033: e^{i\phi})\in L^1_{\nu}(\RR^+)$ for any $\phi\in (a_1,a_2)$ and $g$ is
2034: analytic in $S$ then we have
2035: 
2036:  \begin{align}
2037:   \|f*g\|_{\nu;\infty} \le \|f\|_{\nu;\infty}\|g\|_{\nu;1}\\
2038:  \end{align}
2039: 
2040: \z The function $F_0$ is analytic in $\mathcal{S}_M$,
2041:  Laplace transformable in any direction other
2042: than $\RR^+$ and $y_0=\lap \{F_0\}$ is a solution of (\ref{eq:eq*}).
2043: \end{Proposition}
2044: 
2045: 
2046: \begin{proof}
2047:   For large enough $\nu$ and for any $M$ there is a unique analytic
2048:   solution in $S_M$, $F_0$, and $F_0$ is thus independent of $M$. Since
2049:   we have $|F_0(p)|\le |p|e^{\nu |p|}$ for $p\in S=\cup_{M>0} S_M$ it
2050:   follows that $F_0=F$.  Using Proposition~\ref{L1loc} the proof is
2051:   complete.
2052: \end{proof}
2053: 
2054: \subsubsection{Summability of the transseries $\tilde{f}$}
2055: 
2056: A straightforward calculation shows that the series
2057: $\tilde{f}_k$ in (\ref{eq:gtrns}) satisfy the system of equations
2058: 
2059: 
2060: \begin{eqnarray}
2061:   \label{eq:syste}
2062:   \tilde{f}_0'+(1-\beta x^{-1})\tilde{f}_0 & = & x^{-2}+a\tilde{f}_0^2+b\tilde{f}_0^3\nonumber\\
2063:   \tilde{f}_1'+(2a \tilde{f}_0 + 3 b \tilde{f}_0^2)\tilde{f}_1&=&0\\
2064:  \tilde{f}_k'+\Big((1-k)(1-\beta x^{-1})+2a \tilde{f}_0 + 3 b
2065:   \tilde{f}_0^2\Big)\tilde{f}_k
2066:   &=&a{\sum}^\dagger\tilde{f}_{k_1}\tilde{f}_{k_2}
2067:   +b{\sum}^{\ddagger}\tilde{f}_{k_1}\tilde{f}_{k_2}\tilde{f}_{k_3}\nonumber
2068: \end{eqnarray}
2069:  where in the sum $\sum^\dagger$ the indices satisfy $k_1+k_2=k;\ k_1>0;
2070: k_2>0$ while in $\sum^{\ddagger}$ the condition is $ k_1+k_2+k_3=k ;\ 
2071: k_i>0$. We have already seen that $\tilde{f}_0$ is summable. The
2072: equation for $\tilde{f}_1$ is special, and we treat it separately. To
2073: ensure Borel transformability, since $\tilde{f}_1=O(1)$, it is
2074: convenient to take $\tilde{f}_1= x^2\tilde{y}_1$, which gives
2075: 
2076: $$\tilde{y}_1'+2x^{-1}\tilde{y}_1+(2a \tilde{f}_0 + 3 b
2077: \tilde{f}_0^2)\tilde{y}_1=0$$
2078: 
2079: \z which, in Borel transform, with $Y_1= \bor\tilde{y}_1$,
2080: 
2081: \begin{equation}
2082:   \label{eq:eqY10}
2083:   -pY_1 +2\int_0^p Y_1(s)ds+(2a _0 F_0 + 3 b
2084: F_0^{*2})*Y_1=0
2085: \end{equation}
2086: 
2087: \z which implies, denoting $2a _0 F_0 + 3 b
2088: F_0^{*2}=G$
2089: 
2090: \begin{equation}
2091:   \label{eq:eqY1}
2092:   -pY_1'+Y_1 =-G'*Y_1
2093: \end{equation}
2094: 
2095: \z in which the leading behavior of $Y_1$ is expected to be
2096: $Y_1=p+\ldots$ and thus the dominanat balance is between the terms on
2097: the l.h.s. of (\ref{eq:eqY1}). In integral form we have, with $Y_1=pQ$,
2098: 
2099: 
2100: \begin{equation}
2101:   \label{eq:eqY11}
2102:  Q=1+\int_0^p ds s^{-2}\int_0^suG'(u)Q(s-u)du=
2103: 1+\int_0^p\int_0^1   v G'(sv)Q(sv-v) dv ds
2104: \end{equation}
2105: 
2106: 
2107: \z It is easy to see that (\ref{eq:eqY1}) is contractive 
2108: in a space of analytic functions for $|p|<\epsilon$
2109: if $\epsilon$ is small. To find exponential bounds for large $p$
2110: we first restrict to a ray $p=t e^{i\phi}$ with $\phi\ne 0$.
2111: 
2112: For $x=p e^{-i\phi}\in [0,\infty)$ it is useful to write $Q=Q_0+Q_1$
2113: where $Q_0=0$ for $|p|>\epsilon$ and $Q_1=0$ for $|p|<\epsilon$. Noting
2114: that $\int_0^s Q_1(t)G'(s-t)dt=0$ if $|s|<\epsilon$, the equation for
2115: $Q_1$ takes the form, for $|p|>\epsilon$, 
2116: 
2117: $$Q_1=F(p)+\int_{\epsilon e^{i\phi}}^p 
2118: \int_{\epsilon e^{i\phi}}^s s^{-2} G'(u)Q_1(s-u)\,\,\drm u\,\drm s $$
2119: 
2120: 
2121: \z where 
2122: 
2123: $$F(p)=1+\int_0^p\int_0^{\epsilon e^{i\phi}}
2124: s^{-2}G'(s-u)Q_0(u)\,\drm u\,\drm s
2125: $$
2126: 
2127: \z Taking $Y(p)=Q_1(\epsilon e^{i\phi} +p)$ we get a convolution
2128: equation
2129: of the form
2130: 
2131: $$Y=F_1+Y*\{(s+\epsilon e^{i\phi})^{-2}\}*F_2$$
2132: 
2133: \z which is manifestly contractive in the norm $\|\cdot\|_\nu$
2134: for large $\nu$. Since $Q_0$ is manifestly in $\mathcal{A}_\nu$
2135: for any $\nu$, it follows that $Q\in\mathcal{A}_\nu$ as well.
2136: 
2137: 
2138: For $k>1$ it is convenient to take $\tilde{f}_k=x^{2k}\tilde{y}_k$ 
2139: and we have 
2140: 
2141: \begin{multline}\label{eqeqk}
2142:   \tilde{y}_k'+\Big(1-k+(2k+(k-1)\beta) x^{-1}+2a \tilde{y}_0 + 3 b \tilde{y}_0^2\Big)\tilde{y}_k
2143: =a{\sum}^\dagger
2144: \tilde{y}_{k_1}\tilde{y}_{k_2}
2145: \\+b{\sum}^{\ddagger}\tilde{y}_{k_1}\tilde{y}_{k_2}\tilde{y}_{k_3}
2146: \end{multline}
2147: (cf. (\ref{eq:syste})) which, after Borel transform becomes
2148: 
2149: \begin{multline}
2150:   \label{eqeqYk}
2151:   Y_k=(p+k-1)^{-1}\Bigg((2k+k\beta -\beta+G)*Y_k+
2152:  a{\sum}^\dagger
2153:  Y_{k_1}* Y_{k_2}
2154: +b{\sum}^\ddagger
2155:  Y_{k_1}* Y_{k_2}* Y_{k_3}\Bigg)\\=
2156: \mathcal{J}Y_k+(p+k-1)^{-1}a{\sum}^\dagger
2157:  Y_{k_1}* Y_{k_2}+b{\sum}^{\ddagger}
2158:  Y_{k_1}* Y_{k_2}* Y_{k_3}
2159: \end{multline}
2160: Note now that the operator on the r.h.s. of 
2161: (\ref{eqeqYk}) is contractive in the norms introduced, for large $\nu$. 
2162: It is then clear that $Y_k$ are analytic and Laplace transformable
2163: for large enough $\nu$. It remains to show they are
2164: simultaneously Laplace transformable.
2165: 
2166: Note that for large $\nu$, denoting $\|Y_k\|=x_k$ we have, with
2167: $\lambda_\nu=\max\{\|\mathcal{J}\|,\|Y_1\|_\nu\}$ arbitrarily small if
2168: $\nu$ is large,
2169: \begin{equation}
2170:   \label{eqeqYk1}
2171:   x_k\le \lambda_\nu x_k+a{\sum}^\dagger
2172: x_{k_1}x_{k_2}+b{\sum}^\ddagger
2173:  x_{k_1}x_{k_2}x_{k_3}
2174: \end{equation}and the coefficients $x_k$ are majorized by the Taylor coefficients
2175: of the analytic solution of the algebraic equation
2176: $$\psi(z)=\lambda_\nu z+\lambda_\nu\psi(z)+a\psi(z)^2+b\psi(z)^3$$
2177: Thus, for large enough $\nu$ we have that $x_k\le \rho_\nu^k$
2178: ($\rho_\nu=o(1)$ for large $\nu$) and thus, for $x$ large enough
2179: $$\|\lap\{Y_k\}\|_{\infty}\le \rho_\nu^k$$
2180:   (An inductive calculation
2181: shows that $Y_k=O(p^{2k-1})$ for small $p$.) The sum
2182: $$\sum_{k=0}^{\infty}(\xi x^2)^k \lap\{Y_k\}$$
2183: is then uniformly convergent. It is easy to see that it gives
2184: therefore a solution of (\ref{eq:eq*}). 
2185: \subsection{Generalized Borel summation for rank one ODEs}
2186: 
2187:  We look at the differential system (\ref{eqor1}) under the same
2188: assumptions and normalization as in \S~\ref{Sysnonlin}. 
2189: 
2190: 
2191: \z {\em Further normalizing transformations}.  For convenience, we
2192: rescale $x$ and reorder the components of $\mathbf{y}$ so that
2193: 
2194: 
2195: (n3) $\lambda_1=1$, and, with $\phi_i=\arg(\lambda_i)$, we have
2196: $\phi_i<\phi_j$ if $i<j$. To simplify notations, we formulate some of
2197: our results relative to $\lambda_1$; they can be easily adapted to any
2198: other eigenvalue.
2199: 
2200:   To unify the treatment we
2201: make, by taking $\mathbf{y}=\mathbf{y}_1 x^{-N}$ for some $N>0$,
2202: 
2203: (n4) $\Re(\beta_j)<0,\ j=1,2,\ldots,n$.
2204: 
2205: 
2206: \z {\bf Note}: there is an asymmetry at this point: the opposite
2207: inequality cannot be achieved, in general, as simply and without
2208: violating analyticity at infinity. In some instances this
2209: transformation is not convenient since it makes more difficult the
2210: study of certain properties, see \cite{Invent}.
2211: 
2212: 
2213: 
2214: 
2215:  Finally, through a transformation of the form
2216: $\mathbf{y}\leftrightarrow\mathbf{y}-\sum_{k=1}^M\mathbf{a}_k x^{-k}$ we arrange that
2217: 
2218: (n5) $ \mathbf{f}_0=O(x^{-M-1})\mbox{ and }\mathbf{g}(x,\mathbf{y})=
2219: O(\mathbf{y}^2,x^{-M-1}\mathbf{y}) $. We choose $M>1+\max_i\Re(-\beta_i)$.
2220: 
2221: 
2222: {\em Formal solutions.} The transseries solutions of (\ref{eqor1}) 
2223: were studied in \S~\ref{Sysnonlin}. More generally, there 
2224: is an n--paramter family of formal exponential series
2225: solutions of (\ref{eqor1}):
2226: \begin{eqnarray}
2227:   \label{eqformgen,n}
2228:    \tilde{\mathbf{y}}_0+\sum_{\mathbf{k}\ge 0; |\mathbf{k}|>0}C_1^{k_1}\cdots C_n^{k_n}
2229: \mathrm{e}^{-(\bfk\cdot\bflam) x}x^{\bfk\cdot\bfm}\tilde{\mathbf{y}}_{\bfk}
2230: \end{eqnarray}
2231: 
2232: \z (see \cite{Wasow} below) where $m_i=1-\lfloor\beta_i\rfloor$,
2233: ($\lfloor\cdot\rfloor=$ integer part), $\bfC\in\CC^n$ is an arbitrary
2234: vector of constants, and
2235: $\tilde{\mathbf{y}}_\bfk=x^{-\bfk(\bfbet+\bfm)}
2236: \sum_{l=0}^{\infty}\mathbf{a}_{\bfk;l} x^{-l}$ are formal power
2237: series.  
2238: 
2239: \subsubsection{Summability} We give a brief  overview
2240: of the results in \cite{Duke}.  The details of the proof follow the general
2241: strategy presented in \S\ref{Ord1}. Let
2242: 
2243: \begin{eqnarray}
2244:   \label{defW}
2245:   \mathcal{W}=\left\{p\in\CC:p\ne k\lambda_i\,,\forall
2246: k\in\NN,i=1,2,\ldots,n\right\}
2247: \end{eqnarray}
2248: (see Fig. 1) The directions $d_j=\{p:\arg(p)=\phi_j\}, j=1,2,\ldots,n$
2249: are the {\em Stokes lines}.
2250: 
2251: 
2252: 
2253: We construct over $\mathcal{W}$ a surface $\mathcal{R}$,
2254: consisting of homotopy classes of smooth curves in $\mathcal{W}$
2255: starting at the origin, moving away from it, and crossing at most one
2256: Stokes line, at most once (Fig. 1):
2257: 
2258: \begin{eqnarray}\label{defpaths}
2259: {\mathcal{R}}:=\Big\{\gamma:(0,1)\mapsto \mathcal{W}:\ 
2260: \gamma(0_+)=0;\ \frac{\mathrm{d}}{\mathrm{d}t}|\gamma(t)|>0;\ \arg(\gamma(t))\ \mbox{monotonic}\Big\}\cr
2261: \end{eqnarray} 
2262: Define $\mathcal{R}_1$ as the restriction of
2263: $\mathcal{R}$ to $\arg(\gamma)\in(\psi_n-2\pi,\psi_2)$ where
2264: $$\psi_n=\max\{-\pi/2,\phi_n-2\pi\}\ \text{and} \ \psi_2=\min\{\pi/2,\phi_2\}$$
2265: 
2266: \scalebox{0.5}[0.5]{\includegraphics{median4}}
2267: 
2268: \centerline{{{Fig 1.} \emph{The paths
2269: near $\lambda_2$ belong to $\mathcal{R}$.   }}}\nobreak
2270: \centerline{\em The paths
2271: near $\lambda_1$ relate to the balanced average}
2272: \subsubsection{Singularities of $\bf Y_k$} The Borel transforms of $\bf \tilde{y}_k$ are analytic at zero and the only possible singularities 
2273: are at multiples of the eigenvalues $\lambda_j$ of the linearized
2274: system.  These singularities are ``regular'' in the sense that there
2275: exist {\em convergent} local expansions at these singularities in
2276: terms of powers and possibly logarithms. The following theorem makes
2277: this statement precise.
2278:  \begin{Theorem}[\cite{Duke}]\label{AS} (i) $\mathbf{Y}_0=\bor\tilde{\mathbf{y}}_0$
2279: is analytic in $\mathcal{R}\cup \{0\}$. The singularities of $\mathbf{Y}_0$
2280: (which are contained in the set
2281: $\{l\lambda_j:l\in\NN^+,j=1,2,\ldots,n\}$) are described as follows.
2282: For $l\in\NN^+$ and small $z$
2283: 
2284:  \begin{multline}
2285:    \label{SY0}
2286:    \mathbf{Y}_0^{\pm}(z+l\lambda_j)=\pm\Big[(\pm S_j)^l\ln(z)^{0,1}
2287:    \mathbf{Y}_{l\mathbf{e}_j}(z)\Big]^{(lm_j)}+\mathbf{B}_{lj}(z)=\cr
2288:    \Big[z^{l\beta_j'-1}\ln
2289:    z^{0,1}\,\mathbf{A}_{lj}(z)\Big]^{(lm_j)}+\mathbf{B}_{lj}(z) \ (l=1,2,\ldots)
2290:  \end{multline}
2291: 
2292:  \z where the power of $\ln(z)$ is one iff
2293:  $l\beta_j\in\ZZ$, and $\mathbf{A}_{lj},\mathbf{B}_{lj}$ are analytic for small
2294:  $z$. The functions $\mathbf{Y}_\bfk$ are, in addition, analytic at
2295:  $p=l\lambda_j$, $l\in\NN^+$, iff, exceptionally,
2296: 
2297: \begin{eqnarray}\label{defSj}
2298: S_j=r_j\Gamma(\beta'_j)\left(\mathbf{A}_{1,j}\right)_j(0)=0
2299: \end{eqnarray}
2300: 
2301: \z where $r_j=1-\mathrm{e}^{2\pi \mathrm{i}(\beta'_j-1)}$
2302: if $l\beta_j\notin\ZZ$ and $r_j=-2\pi \mathrm{i}$
2303: otherwise. The $S_j$ are Stokes constants,
2304: see theorem~\ref{Stokestr}.
2305: 
2306: (ii) $\mathbf{Y}_\bfk=\bor{\tilde{\mathbf{y}}}_\bfk$, $|\bfk|>1$, are analytic in
2307: $\mathcal{R}\backslash
2308: \{-\bfk'\cdot{\bflam}+\lambda_i:\bfk'\le\bfk,1\le i\le n\}$.  For
2309: $l\in\NN$ and $p$ 
2310:  near $l\lambda_j$, $j=1,2,\ldots,n$ there exist $\mathbf{A}=\mathbf{A}_{\bfk jl}$ and
2311: $\mathbf{B}=\mathbf{B}_{\bfk jl}$ analytic at zero so that ($z$ is as above)
2312: 
2313:  \begin{multline}
2314:    \label{SYK}
2315:   \mathbf{Y}_\bfk^{\pm}(z+l\lambda_j)=
2316: \pm\Big[(\pm S_j)^l{k_j+l\choose
2317:      l}\ln(z)^{0,1}
2318:    \mathbf{Y}_{\bfk+l\mathbf{e}_j}(z)\Big]^{(lm_j)}+l\mathbf{B}_{\bfk lj}(z)=\cr
2319:   \Big[z^{\bfk\cdot{\bfbet}'+l\beta_j'-1}(\ln z)^{0,1}\,\mathbf{A}_{\bfk l
2320:     j}(z)\Big]^{(lm_j)}+l\mathbf{B}_{\bfk l j}(z)\ (l=0,1,2,\ldots)
2321:        \end{multline}
2322: 
2323: 
2324:        \z where the power of $\ln z$ is $0$ iff $l=0$ or
2325:        $\bfk\cdot{\bfbet}+l\beta_j-1\notin\ZZ$ and $\mathbf{A}_{\bfk 0
2326:          j}=\mathbf{e}_j/\Gamma(\beta'_j)$.  Near
2327:        $p\in\{-\bfk'\cdot{\bflam}:0\prec\bfk'\le\bfk\}$, (where
2328:        $\mathbf{Y}_0$ is analytic) $\mathbf{Y}_\bfk,\,\bfk\ne 0$ have convergent
2329:        Puiseux series.
2330: 
2331: 
2332: \end{Theorem} 
2333: \subsubsection{Averaging}\label{Ave} It is possible to take Laplace transforms
2334: of $\bf Y_0$ along a ray avoiding $\RR^+$ from above or below. However
2335: it can be checked that, assuming $\bf Y_0$ is singular at $p=1$ and
2336: that the series $\tilde{y}_0$ has real coefficients, neither of them
2337: would yield a real valued function.
2338: 
2339:  Let $\bor\tilde{\mathbf{y}}_\bfk$ be extended along $d_j$ by
2340: the ``balanced average'' of analytic continuations
2341: 
2342: 
2343: 
2344: \begin{eqnarray}
2345: \label{defmed}
2346: \bor\tilde{\mathbf{y}}_\bfk=\mathbf{Y}_\bfk^{ba}=
2347: \mathbf{Y}_\bfk^++\sum_{j=1}^{\infty}\frac{1}{2^j}\left(\mathbf{Y}_\bfk^{-}
2348: -\mathbf{Y}_\bfk^{-({j-1})+}\right)
2349: \end{eqnarray}
2350: 
2351: \z The sum above coincides with the one in which $+$ is exchanged with
2352: $-$, accounting for the
2353: reality-preserving property. Clearly, if $\mathbf{Y}_\bfk$ is analytic along
2354: $d_j$, then the terms in the infinite sum vanish and
2355: $\mathbf{Y}_\bfk^{ba}=\mathbf{Y}_\bfk$; we also let $\mathbf{Y}_\bfk^{ba}=\mathbf{Y}_\bfk$ if
2356: $d\ne d_j$, where again $\mathbf{Y}_k$ is analytic. It follows from
2357: (\ref{defmed}) and theorem~\ref{CEQ} below that the Laplace integral
2358: of $\mathbf{Y}^{ba}_\bfk$ along $\RR^+$ can deformed into contours
2359: as those depicted in Fig. 1, with weight $2^{-k}$ for a contour turning around
2360: $(k+1)\lambda_1$. More generally, we consider the averages
2361: 
2362: \begin{eqnarray}
2363: \label{defmedC0}
2364: \bor_\alpha\tilde{\mathbf{y}}_\bfk=\mathbf{Y}_\bfk^{\alpha}=
2365: \mathbf{Y}_\bfk^++\sum_{j=1}^{\infty}\alpha^j\left(\mathbf{Y}_\bfk^{-}
2366: -\mathbf{Y}_\bfk^{-({j-1})+}\right)
2367: \end{eqnarray}
2368: 
2369: \z and correspondingly
2370: 
2371: \begin{eqnarray}
2372:   \label{defmedC}
2373: (\mathcal{LB})_\alpha\tilde{\mathbf{y}}_\bfk:=\mathcal{L}\mathbf{Y}_\bfk^{\alpha}
2374: \end{eqnarray}
2375: 
2376: \z With $\alpha\in\RR$, this represents the most general family of
2377: averages of Borel summation formulas which commute with complex
2378: conjugation, with the algebraic and analytic operations and have good
2379: continuity properties \cite{Duke}.  The value $\alpha=1/2$ is special
2380: in that it is the only one compatible with optimal truncation.
2381: 
2382: \begin{Note}
2383:   For rank one ODEs the balanced average mentioned above can be shown
2384:   to coincide with the ``median'' summation of Ecalle, who has
2385:   introduced and studied a wide variety of averages \cite{EcalleMenous}, suitable
2386: in more complicated settings.
2387: \end{Note}
2388: 
2389: 
2390: 
2391: \begin{Theorem}\label{CEQ} (i) The branches of $(\mathbf{Y}_\bfk)_\gamma$ in $\mathcal{R}_1$
2392:   have limits in a $C^*$-algebra of distributions,
2393:   $\mathcal{D}'_{m,\nu}(\RR^+)\subset\mathcal{D}'$. Their Laplace transforms in
2394:   $\mathcal{D}'_{m,\nu}(\RR^+)$ $\lap(\mathbf{Y}_\bfk)_\gamma$ exist
2395:   simultaneously and with $x\in\mathcal{S}_x$ and for any $\delta>0$
2396:   there is a constant $K$ and an $x_1$ large enough, so that for
2397:   $\Re(x)>x_1$ we have $\left|\lap(\mathbf{Y}_\bfk\right)_\gamma(x)|\le
2398:   K\delta^{|\bfk|}$.
2399: 
2400: In addition, $\mathbf{Y}_\bfk(p\mathrm{e}^{\mathrm{i}\phi})$ are continuous
2401: in $\phi$ with respect to the $\mathcal{D}'_{m,\nu}$ topology,
2402: (separately) on $[\psi_n-2\pi,0]$ and $[0,\psi_2]$.
2403: 
2404: If
2405: $m>\max_i(m_i)$ and $l<\min_i |\lambda_i|$ then
2406: $\mathbf{Y}_0(p\mathrm{e}^{\mathrm{i}\phi})$ is continuous in
2407: $\phi\in[0,2\pi]\backslash\{\phi_i:i\le n\}$ in the
2408: $\mathcal{D}'_{m,\nu}(\RR^+,l)$ topology and has (at most) jump
2409: discontinuities for $\phi=\phi_i$. For each $\bfk$, $|\bfk|\ge 1$ and
2410: any $K$ there is an $l>0$ and an $m$ such that
2411: $\mathbf{Y}_k(p\mathrm{e}^{\mathrm{i}\phi})$ are continuous in
2412: $\phi\in[0,2\pi]\backslash\{\phi_i; -\bfk'\cdot\bflam+\lambda_i:i\le n
2413: ,\bfk'\le\bfk\}$ in the $\mathcal{D}'_{m,\nu}((0,K),l)$ topology and
2414: have (at most) jump discontinuities on the boundary.
2415: 
2416: (ii) The sum (\ref{defmed}) converges in $\mathcal{D}'_{m,\nu}$ (and
2417: coincides with the analytic continuation of $\mathbf{Y}_\bfk$ when
2418: $\mathbf{Y}_\bfk$ is analytic along $\RR^+$). For any
2419: $\delta$ there is a large enough $x_1$ {\em
2420:   independent of $\bfk$} so that  $\mathbf{Y}^{ba}_\bfk(p)$ with
2421: $p\in\mathcal{R}_1$ are
2422: Laplace transformable
2423: for
2424: $\Re(xp)>x_1$ and furthermore $|(\lap\mathbf{Y}^{ba}_\bfk)(x)|\le
2425: \delta^{|\bfk|}$. In addition, if $d\ne\RR^+$, then for large $\nu$,
2426: $\mathbf{Y}_\bfk\in L^1_\nu(d)$.
2427: 
2428: The functions
2429: $\lap\mathbf{Y}_\bfk^{ba}$ are analytic for $\Re(xp)>x_1$. For any
2430: $\bfC\in\CC^{n_1}$ there is an $x_1(\bfC)$ large enough so that the
2431: sum
2432: 
2433: \begin{eqnarray}
2434:   \label{soleqn}
2435:   \mathbf{y}=\lap\mathbf{Y}_0^{ba}+\sum_{|\bfk|> 0}\bfC^{\bfk}\mathrm{e}^{-\bfk\cdot\bflam
2436:     x}x^{-\bfk\cdot\bfbet}\lap\mathbf{Y}_\bfk^{ba}
2437: \end{eqnarray}
2438: 
2439: 
2440: \z converges uniformly for $\Re(xp)>x_1(\bfC)$, and $\mathbf{y}$ is a solution
2441: of (\ref{eqor}).  When the direction
2442: of $p$ is not the real axis then, by definition, $\mathbf{Y}^{ba}_\bfk=\mathbf{Y}_\bfk$,
2443: $\mathcal{L}$ is the usual Laplace transform 
2444: and (\ref{soleqn}) becomes
2445: 
2446: 
2447: \begin{eqnarray}
2448:   \label{soleqnpm}
2449:   \mathbf{y}=\lap\mathbf{Y}_0+\sum_{|\bfk|> 0}\bfC^{\bfk}\mathrm{e}^{-\bfk\cdot\bflam
2450:     x}x^{-\bfk\cdot\bfbet}\lap\mathbf{Y}_\bfk
2451: \end{eqnarray}
2452: 
2453: 
2454: 
2455: In addition, $\lap\mathbf{Y}_\bfk^{ba}\sim \tilde{\mathbf{y}}_\bfk$ for large $x$
2456: in the half plane $\Re(xp)>x_1$, for all $\bfk$, uniformly.
2457: 
2458: iii) The general
2459: solution of (\ref{eqor}) that is asymptotic to $\tilde{\mathbf{y}}_0$ for
2460: large $x$ along a ray
2461: in $S_x$ can be equivalently written in  the form
2462: (\ref{soleqn}) or as
2463: \begin{eqnarray}
2464:   \label{soleqnp}
2465:   \mathbf{y}=\lap\mathbf{Y}_0^{\pm}+\sum_{|\bfk|> 0}\bfC^{\bfk}\mathrm{e}^{-\bfk\cdot\bflam
2466:     x}x^{-\bfk\cdot\bfbet}\lap\mathbf{Y}_\bfk^{\pm}
2467: \end{eqnarray}
2468: 
2469: \z for some $\bfC$ (depending on the solution and chosen form). With
2470: the convention binding the directions of $x$ and $p$ and the
2471: representation form being fixed the representation of a solution is
2472: unique.
2473: 
2474: \end{Theorem}
2475: 
2476: 
2477: 
2478: \begin{Theorem}\label{RE}
2479: i) For all $\bfk$ and $\Re(p)>j,\Im(p)>0$ as well
2480: as in $\mathcal{D}'_{m,\nu}$ we have
2481: 
2482: \begin{eqnarray}
2483:   \label{mainresur}
2484:   \mathbf{Y}_{\bfk}^{\pm j\mp}(p)-\mathbf{Y}_{\bfk}^{\pm (j-1) \mp}(p) = (\pm S_1)^j\binom{k_1+j}{j}
2485:   \left(\mathbf{Y}^\pm_{\bfk+j\mathbf{e}_1}(p-j)\right)^{(mj)}
2486: \end{eqnarray}
2487: 
2488: \z and also, 
2489: 
2490: \begin{eqnarray}
2491:   \label{thirdresu}
2492:   \mathbf{Y}_{\bfk}^\pm=\mathbf{Y}_\bfk^{\mp}+\sum_{j\ge 1} {j+k\choose k}(\pm S_1)^{j}(\mathbf{Y}^\mp_{\bfk+j\mathbf{e}_1}(p-j))^{(mj)}
2493: \end{eqnarray}
2494: 
2495: ii) {\em Local Stokes transition.}
2496: 
2497: \z Consider the expression of a fixed solution $\mathbf{y}$ of (\ref{eqor})
2498: as a Borel summed transseries (\ref{soleqn}). As $\arg(x)$ varies,
2499: (\ref{soleqn}) changes only through $\bfC$, and that change occurs
2500: when the Stokes lines are crossed. We have, in the neighborhood
2501: of $\RR^+$, with $S_1$
2502: defined in (\ref{defSj}):
2503: 
2504:  
2505: \begin{eqnarray}
2506:   \label{microsto}
2507:   \bfC(\xi)=\left\{\begin{array}{lll} \bfC^-=\bfC(-0)\qquad&\mbox{for
2508:     $\xi<0$}\\ \bfC^0=\bfC(-0)+\frac{1}{2}S_1\mathbf{e}_1\qquad&\mbox{for
2509:     $\xi=0$}\\ \bfC^+=\bfC(-0)+S_1\mathbf{e}_1\qquad&\mbox{for
2510:     $\xi>0$}\end{array}\right.
2511: \end{eqnarray}
2512: \end{Theorem}
2513: 
2514: \begin{Remark}\label{R1} In view of (\ref{mainresur}) the different analytic
2515: continuations of $\mathbf{Y}_0$ along paths crossing
2516: $\RR^+$ at most once can be expressed in terms of
2517: $\mathbf{Y}_{j\mathbf{e}_1}$. The most general formal solution of (\ref{eqor})
2518: that can be formed in terms of $\mathbf{Y}_{j\mathbf{e}_j}$ with $j\ge 0$ is
2519: (\ref{eqformgen,n}) with $C_1=\alpha$ arbitrary and $C_j=0$ for $j\ne
2520: 1$. Any true solution of (\ref{eqor}) based on such a transseries
2521: is given in (\ref{soleqnp}) with 
2522: $\bfC$ as above. Any average $\mathcal{A}\mathbf{Y}_0$  along paths
2523: going forward in $\RR^+$ such that 
2524: $\lap \mathcal{A}\mathbf{Y}_0$ is thus of the form (\ref{defmedC}).
2525: \end{Remark}
2526: 
2527: 
2528: \begin{Theorem}\label{Stokestr} Assume only $\lambda_1$ lies
2529:   in the right half plane. Let $\gamma^{\pm}$ be two paths in the
2530:   right half plane, near $\pm i\RR^+$ such
2531:   that $|x^{-\beta_1+1}\mathrm{e}^{-x\lambda_1}|\rightarrow 1$ as
2532:   $x\rightarrow\infty$ along $\gamma^{\pm}$. Consider the solution
2533:   $\mathbf{y}$ of (\ref{eqor}) given in (\ref{soleqn}) with
2534:   $\mathbf{C}=C\mathbf{e}_1$ and where the path of integration is
2535:   $p\in\RR^+$. Then
2536: 
2537: \begin{eqnarray}\label{classicS}
2538: \mathbf{y}=
2539: (C\pm\frac{1}{2}S_1)\mathbf{e}_1 x^{-\beta_1+1}\mathrm{e}^{-x\lambda_1}(1+o(1))
2540: \end{eqnarray}
2541: 
2542: 
2543: \z for large $x$ along $\gamma^{\pm}$, where $S_1$ is the same as in
2544: (\ref{defSj}), (\ref{microsto}).
2545: 
2546: 
2547:  
2548: \end{Theorem}
2549: 
2550: \begin{Proposition}\label{asymptrick} i)  Let $\mathbf{y}_1$ and $\mathbf{y}_2$ be solutions of (\ref{eqor})
2551: so that $\mathbf{y}_{1,2}\sim\tilde{\mathbf{y}}_0$ for large $x$ in an
2552: open sector $S$ (or in some direction
2553: $d$);
2554: then $\mathbf{y}_1-\mathbf{y}_2=\sum_{j}C_j\mathrm{e}^{-\lambda_{i_j}
2555:   x}x^{-\beta_{i_j}}(\mathbf{e}_{i_j}+o(1))$ for some constants $C_j$, where the indices run over
2556:   the eigenvalues $\lambda_{i_j}$ with the property $\Re(\lambda_{i_j}
2557:   x)>0$ in $S$ (or $d$). 
2558: If $\mathbf{y}_1-\mathbf{y}_2=o(\mathrm{e}^{-\lambda_{i_j}
2559:     x}x^{-\beta_{i_j}})$ for all $j$, then $\mathbf{y}_1=\mathbf{y}_2$.
2560:  
2561: ii) Let $\mathbf{y}_1$ and $\mathbf{y}_2$ be solutions of
2562: (\ref{eqor1}) and assume that $\mathbf{y}_1-\mathbf{y}_2$ has differentiable
2563: asymptotics of the form 
2564: $\mathbf{K}a\exp(-ax)x^b(1+o(1))$ with  $\Re(ax)>0$ and $\mathbf{K}\ne 0$, for large $x$.
2565: Then $a=\lambda_i$ for some $i$.
2566: 
2567: iii) Let $\mathbf{U}_\bfk\in\mathcal{T}_{\{\cdot\}}$ for all $\bfk$,
2568: $|\bfk|>1$.  Assume in addition that for
2569: large $\nu$ there is a function $\delta(\nu)$ vanishing as
2570: $\nu\rightarrow\infty$ such that
2571: 
2572: 
2573: \begin{gather}
2574: \label{cd1}
2575: \sup_{\bfk}\delta^{-|\bfk|}\int_{d}\left|\mathbf{U}_{\bfk}(p)\mathrm{e}^{-\nu p}\right|\mathrm{d}|p|<K<\infty
2576: \end{gather}
2577: 
2578: \z Then, if $\mathbf{y}_1,\mathbf{y}_2$ are solutions of (\ref{eqor}) in $S$ where in addition
2579: 
2580: \begin{eqnarray}
2581:   \label{lapcond}
2582:   \mathbf{y}_1-\mathbf{y}_2=\sum_{|\bfk|>1}
2583: \mathrm{e}^{-\bflam\cdot\bfk x}x^{\bfm\cdot\bfk}\int_d\mathbf{U}_\bfk(p)
2584: \exp(-xp)\mathrm{d}p
2585: \end{eqnarray}
2586: 
2587: \z where $\bflam,x$ are as in (n6), then $\mathbf{y}_1=\mathbf{y}_2$, and
2588: $\mathbf{U}_\bfk=0$ for all $\bfk$, $|\bfk|>1$.
2589: 
2590: \end{Proposition}
2591: 
2592: 
2593: 
2594: Given $\mathbf{y}$, the value of $C_i$ can change only when
2595: $\xi+\arg(\lambda_i-\bfk\cdot\bflam)=0$, $k_i\in\NN\cup\{0\}$, i.e.
2596: when crossing one of the (finitely many by (c1)) Stokes lines.  The
2597: procedure is similar to the medianization \index{medianization}  proposed by \'Ecalle, but (due
2598: to the structure of (\ref{eqor})) requires substantially fewer analytic
2599: continuation paths. Resurgence \index{resurgence} relations are found and in addition we
2600: provide a complete description, needed in applications, of the
2601: singularity structure of the Borel transforms of $\tilde{\mathbf{y}}_\bfk$.
2602: 
2603: 
2604: \subsubsection{More general equations amenable to first rank. Examples}\label{sec:ExaNorm} 
2605: 
2606: Many classical equations which are not presented in the form studied
2607: in \S~\ref{Sysnonlin} can be brought to that form by elementary
2608: transformations. We look at a few examples, to illustrate this and the
2609: normalization process.
2610: 
2611: \z {\bf (1)}. The equation
2612: 
2613: \begin{equation}
2614:   \label{erf}
2615:   f'+2xf=1
2616: \end{equation}
2617: 
2618: \z for large $x$ is not of the form (\ref{eqor1}). It can nevertheless
2619: be brought to that form by simple changes of variables. The
2620: normalizing change of variables of a given is most conveniently
2621: obtained in the following way.  
2622: 
2623: {\em A transformation which brings the
2624:   equation to its normal form also brings its transseries solutions to
2625:   the form (\ref{transsf})}. 
2626: 
2627:  It is simpler to look for substitutions
2628: with this latter property, and then the first step is to find the
2629: transseries solutions of the equation.
2630: 
2631: At level zero, differentiation is contractive and thus, within power
2632: series there is a unique solution of (\ref{erf}), obtained as a fizxed
2633: point of
2634: \begin{equation}
2635:   \label{fixp}
2636:   f=\frac{1}{x}-\frac{1}{x}f'
2637: \end{equation}
2638: Denoting this series by $f_1$ we look for further transseries solutions in the form $f=f_1+\delta$. We get
2639: 
2640: \begin{equation}
2641:   \label{delta1}
2642:     \delta'+2x\delta=0
2643: \end{equation}
2644: where we look for higher level transseries solutions (we choose to
2645: ignore the fact that (\ref{delta1}) is solvable explicitly).  We
2646: therefore take $\delta=e^w$ and look for level zero solutions.  We
2647: have $w'+2x=0$ with the one parameter family of solutions
2648: $w=-x^2/2+K$.  Therefore
2649: \begin{equation}
2650:   \label{tr2}
2651:   f=f_1+Ce^{-x^2/2}
2652: \end{equation}
2653: The exponent is supposed to be linear in $x$ which suggests the change of variable $x^2=z$. In terms of $z$ we get
2654: \begin{equation}
2655:   \label{zz}
2656:   g'+g=\frac{1}{2\sqrt{z}}
2657: \end{equation}
2658: which is not of the required form because of the noninteger power of $z$.
2659: This can be easily take care of by the substitution $g=\sqrt{z} h$
2660: which leads to
2661: \begin{equation}
2662:   \label{fin3}
2663:   h'+\Big(1+\frac{1}{2z}\Big)h=\frac{1}{2z}
2664: \end{equation}
2665: 
2666: {\bf (2)} A simple Schrodinger equation in one dimension (the harmonic
2667: oscillator).
2668: 
2669: 
2670: \begin{equation}
2671:   \label{eq:Schr}
2672: y''-x^2y=\lambda y  
2673: \end{equation}
2674: 
2675: A convenient way to find transseries solutions is to use a WKB-like
2676: procedure. Substituting $y(x)=e^{g(x)}$ in (\ref{eq:Schr}) one obtains
2677: 
2678: 
2679: $$g'=\pm \sqrt{\lambda+x^2-g''}$$
2680: 
2681: \z which is a contractive equation in the space
2682: $\tilde{\mathcal{T}}^{[0]}$, cf. \S~\ref{deg1}. It follows 
2683: in particular that $\tilde{g}'=\pm x+O(1/x)$ and thus $\ln
2684: y(x)=\pm\frac{1}{2}x^2+O(\ln x)$. Since for (\ref{eqor1})
2685: the exponentials have linear exponents, the natural variable
2686: of (\ref{eq:Schr}), for our purpose, is $t=\frac{1}{2} x^2$. In this
2687: variable, (\ref{eq:Schr}) becomes 
2688: 
2689: 
2690: \begin{equation}
2691:   \label{eq:Schr,norm}
2692: h''+\frac{1}{2t}h'-\left(1+\frac{\lambda}{2t}\right)h=0
2693: \end{equation}
2694: 
2695: \z which in vector form, with $y_1=h+h', y_2=h-h'$, reads
2696: 
2697: \begin{equation}
2698:   \label{eq:SchrSyst}
2699:   \mathbf{y}'=
2700: \left\{\begin{pmatrix}
2701:   1&0\cr 0&-1
2702: \end{pmatrix}+\frac{1}{4t}
2703: \begin{pmatrix}
2704:   \lambda-1&\lambda+1\cr
2705: 1-\lambda&-1-\lambda 
2706: \end{pmatrix}\right\}
2707: \mathbf{y}=\left(\hat{\Lambda}+\frac{1}{t}\hat{C}\right)
2708: \mathbf{y}
2709: \end{equation}
2710: 
2711: The matrix $\hat{C}$ can (always) be diagonalized, by taking
2712: $\mathbf{y}=(I+t^{-1}\hat{S})\mathbf{Y}$ for a suitable matrix
2713: $\hat{S}$, and the diagonalized matrix is
2714: $\hat{B}=\mbox{diag}(C_{11},C_{22})$.  Indeed, the equation for
2715: $\hat{S}$ is
2716: $\hat{\Lambda}\hat{S}-\hat{S}\hat{\Lambda}=\hat{B}-\hat{C}$, which can
2717: be solved whenever the eigenvalues of $\hat{\Lambda}$ are distinct.  The
2718: normal form of (\ref {eq:SchrSyst}) is thus
2719: 
2720: \begin{equation}
2721:   \label{eq:SchrSystN}
2722:   \mathbf{y}'=\left(\hat{\Lambda}+\frac{1}{t}\hat{C}\right)
2723: \mathbf{y}+t^{-2}\hat{g}(t^{-1})\mathbf{y}
2724: \end{equation}
2725: 
2726: \z with $\hat{g}$ analytic. 
2727: 
2728: {\bf (3)} A nonintegrable case of Abel's equation
2729: 
2730: 
2731: \begin{gather}
2732:   \label{(*)}
2733:  w'=w^3-z 
2734: \end{gather}
2735: {\em Power series solutions}. As before, (\ref{(*)}) written in the form
2736: 
2737: $$w=(w'+z)^{1/3}$$
2738: 
2739: \z is contractive in $\tilde{\mathcal{T}}^{[0]}$. By iteration,  one
2740: obtains the power series formal solutions
2741: $\tilde{w}_0=Az^{1/3}(1+\sum_{k=1}^{\infty}{\tilde{w}}_{0,k}z^{-5k/3})$
2742: ($A^3=1$).
2743: 
2744: \z {\em General transseries solutions of (\ref{(*)})}. Once we determined
2745: $\tilde{w}_0$ we look for possible small corrections to $\tilde{w}_0$;
2746: since these are usually exponentially small, formal WKB 
2747: is again useful. We substitute $w=\tilde{w}_0+e^g$. The 
2748: equation for $g$
2749: 
2750: $$\tilde{g}=C+\mathcal{P}\left(3\tilde{w}_0^2+3\tilde{w}_0 e^{\tilde{g}}
2751:   +e^{2\tilde{g}}\right)$$
2752: 
2753: \z is contractive in any sector where $\Re(\tilde{w}_0)<0$ and in this
2754: case $e^{\tilde{g}}\propto z^{2/3}\exp\left(\frac{9}{5}A^2z^{5/3}\right)$
2755: 
2756: 
2757: Since the exponentials in a transseries solution of a normalized system
2758: have {\em linear} exponent, with negative real part, the independent
2759: variable should be $x=-(9/5)A^2z^{5/3}$ and $\Re(x)>0$. Then
2760: $\tilde{w}_0=x^{1/5}$ $\sum_{k=0}^{\infty} w_{0;k} x^{-k}$, which
2761: compared to (\ref{transsf}) suggests the change of dependent variable
2762: $w(z)=Kx^{1/5}h(x)$. Choosing for convenience $K=A^{3/5}(-135)^{1/5}$
2763: yields
2764: 
2765: \begin{gather}
2766:   \label{tr12}
2767: h'+\frac{1}{5x}h+3h^3-\frac{1}{9}=0
2768: \end{gather}
2769: 
2770: \z The next step is to achieve leading behavior $O(x^{-2})$. This is
2771: easily done by subtracting out the leading behavior of $h$ (which can be
2772: found by maximal balance, as above). With $h=y+1/3-x^{-1}/15$ we
2773: get the normal form
2774: 
2775: \begin{align}
2776:   \label{eqfu}
2777: y'=-\left(1-\frac{1}{5x}\right)y+g(y,x^{-1})
2778: \end{align}
2779: 
2780: \z where
2781: \begin{equation}
2782:   \label{eq:defg}
2783:   g(y,x^{-1})=-3(y^2+y^3)+\frac{3y^2}{5x}-\frac{1}{15x^2}-\frac{y}{25x^2}+\frac{1}{3^2
2784:   5^3 x^3}
2785: \end{equation}
2786: 
2787: 
2788: {\bf (3)} The Painlev\'e equation P1. 
2789: 
2790: 
2791: \begin{gather}
2792:   \label{eP1}
2793: \frac{d^2y}{dz^2}=6y^2+z
2794: \end{gather}
2795: 
2796: The transformations needed to normalize
2797: (\ref{eP1}) are derived in the same way as in Example 1.  After the
2798: change of variables
2799: \begin{gather*}
2800: x=\frac{(-24z)^{5/4}}{30};\ y(z)=\sqrt{\frac{-z}{6}}\left(1-\frac{4}{25x^2}+h(x)\right)
2801: \end{gather*}
2802: P1 becomes 
2803: 
2804: \begin{gather}\label{eqp1n}
2805: h'' +\frac{1}{x}h'-h-\frac{1}{2}h^2-\frac{392}{625x^4}=0
2806: \end{gather} 
2807: 
2808: \z Written as a system, with $\mathbf{y}=(h,h')$ this equation satisfies
2809: the assumptions in \S~\ref{Sysnonlin} with $\lambda_{1,2}=\pm 1$,
2810: $\alpha_{1,2} =-1/2$, and then $\xi(x)=C\erm^{-x}x^{-1/2}$.
2811: 
2812: {\bf (4)} The Painlev\'e equation P2. 
2813: 
2814: This equation reads:
2815: 
2816: \begin{eqnarray}
2817:   \label{eqP2}
2818:   y''=2y^3+xy+\alpha
2819: \end{eqnarray}
2820: 
2821: \z This example also shows that for a given equation distinct solution
2822: manifolds associated to distinct asymptotic behaviors may lead to
2823: different normalizations. After the change of variables
2824: 
2825: $$x=(3t/2)^{2/3};\ \ \
2826: y(x)=x^{-1}(t\,h(t)-\alpha)$$
2827: 
2828: \z one obtains the normal form equation
2829: \begin{align}
2830:   \label{eq:p2n0}
2831:   h''+\frac{h'}{t}-\left(1+\frac{24\alpha^2+1}{9t^2}\right)h-\frac{8}{9}h^3+
2832: \frac{8\alpha}{3t}h^2+\frac{8(\alpha^3-\alpha)}{9t^3}=0
2833: \end{align}
2834: 
2835: Distinct  normalizations (and sets of solutions)
2836: are provided by 
2837: 
2838: $$x=(At)^{2/3};\ \
2839: y(x)=(At)^{1/3}\left(w(t)-B+\frac{\alpha}{2At}\right)$$
2840: 
2841: \z if $A^2=-9/8,B^2=-1/2$. In this case, 
2842: 
2843: \begin{multline}
2844:   w''+\frac{w'}{t}+w\left(1+\frac{3B\alpha}{tA}-
2845: \frac{1-6\alpha^2}{9t^2}\right)w\\
2846: -\left(3B-\frac{3\alpha}{2tA}\right)w^2+w^3+
2847: \frac{1}{9t^2}\left(B(1+6\alpha^2)-t^{-1}\alpha(\alpha^2-4)
2848: \right)
2849: \end{multline}
2850: 
2851: \z so that
2852: 
2853: $$\lambda_1=1, \alpha_1=-\frac{1}{2}-\frac{3}{2}\frac{B\alpha}{A}$$
2854: 
2855: 
2856: The first normalization applies for the manifold of solutions such
2857: that $y\sim-\frac{\alpha}{x}$ (for $\alpha=0$ $y$ is exponentially
2858: small and behaves like an Airy function) while the second one
2859: corresponds to $y\sim -B-\frac{\alpha}{2}x^{-3/2}$.
2860: \subsection{Difference equations and PDEs}
2861: \subsubsection{Difference equations}  Transseries and
2862: Borel summation techniques can be successfully used for difference equations.
2863: Consider difference systems of equations which can be brought to
2864: the form
2865: \begin{equation}
2866:   \label{eq:Bra1}
2867:   \mathbf{x}(n+1)=\hat\Lambda\left(I+ \frac{1}{n}\hat{A}\right)\mathbf{x}(n) +\mathbf{g}(n,\mathbf{x}(n))
2868: \end{equation}
2869: where $\hat\Lambda $ and $\hat{A} $ are constant coefficient matrices, $\mathbf{g}$ is
2870: convergently given for small $\mathbf{x}$ by
2871: \begin{equation}
2872:   \label{eq:defg1}
2873: \mathbf{g}(n,\mathbf{x})=\sum_{\mathbf{k}\in\NN^m}
2874: \mathbf{g}_\mathbf{k}(n)\mathbf{x}^\mathbf{k} 
2875: \end{equation}
2876: with $\mathbf{g}_\mathbf{k}(n)$ analytic in $n$ at infinity and
2877: \begin{equation}
2878:   \label{eq:condg}
2879:   \mathbf{g}_\mathbf{k}(n)=O(n^{-2}) \ \mbox{as }n\rightarrow\infty,\
2880:   \mbox{if }\sum_{j=1}^m k_j\le 1
2881: \end{equation}
2882:  under nonresonance conditions: Let
2883: $\boldsymbol{\mu}=(\mu_1,...,\mu_n)$ and $\mathbf a=(a_1,...,a_n)$ where
2884: $e^{-\mu_k}$ are the eigenvalues of $\hat{\Lambda}$ and the $a_k$ are
2885: the eigenvalues of $\hat{A}$.  Then the nonresonance condition is
2886: \begin{equation}
2887:   \label{eq:nonres}
2888:   \left(\mathbf{k}\cdot \boldsymbol\mu=0 \mod 2\pi i \ \ \mbox{with}\ \ \mathbf{k} \in\ZZ^{m_1}
2889:   \right)
2890: \Leftrightarrow \mathbf{k}=0.
2891: \end{equation}
2892: We consider the solutions of (\ref{eq:Bra1}) which are small as $n$
2893: becomes large.  \ Braaksma and Kuik \cite{Braaksma} \cite{Kuik}
2894: showed that the recurrences (\ref{eq:Bra1}) posess $l$-parameter
2895: transseries solutions of the form
2896: \begin{equation}
2897:   \label{eq:transszf}
2898:   \tilde{{\bf x}}(t)
2899: :=\sum_{{\bf k}\in\NN^{m}}{\bf C}^{\bf k} e^{-{\bf k}\cdot
2900:   \boldsymbol{\mu} t}
2901: t^{{\bf k}\cdot {\bf a}} \tilde{{\bf x}}_\bfk(t)
2902: \end{equation}
2903: with $t=n$ where $ \tilde{{\bf x}}_\bfk(n)$ are formal power series in
2904: powers of $n^{-1}$ and $l\le m$ is chosen such that, after reordering
2905: the indices, we have $\Re(\mu_j)>0$ for $1\le j\le l$.
2906:  
2907:  It is shown in \cite{Braaksma}, \cite{Kuik} that these transseries are generalized
2908:  Borel summable in any direction and Borel summable in all except $m$
2909:  of them and that
2910: \begin{equation}
2911:   \label{eq:transsSn}
2912:   {\bf x}(n)
2913: =\sum_{{\bf k}\in\NN^{l}}{\bf C}^{\bf k} e^{-{\bf k}\cdot
2914:   \boldsymbol{\mu} n}
2915: n^{{\bf k}\cdot {\bf a}}{\bf x}_\bfk(n)
2916: \end{equation} 
2917: is a solution of (\ref{eq:Bra1}), if $n>y_0$, $t_0$ large enough.
2918: \subsubsection{PDEs} 
2919: 
2920: 
2921: \subsubsection{Existence, uniqueness, regularity, asymptotic behavior}
2922: For partial differential equations with analytic coefficients which
2923: can be transformed to equations in which the differentiation order in
2924: a distinguished variable, say time, is no less than the one with
2925: respect to the other variable(s), under some other natural
2926: assumptions, Cauchy-Kowalevski theory (C-K) applies and gives
2927: existence and uniqueness of the initial value problem. A number of
2928: evolution equations do not satisfy these assumptions and even if
2929: formal power series solutions exist their radius of convergence is
2930: zero.  
2931: The paper \cite{Invent2} provides a C-K type theory in such
2932: cases, providing existence, uniqueness and regularity of the
2933: solutions.  Roughly, convergence is replaced by Borel summability,
2934: although the theory is more general.
2935: 
2936: Unlike in C-K, solutions of nonlinear evolution equations develop
2937: singularities which can be more readily studied from the local
2938: behavior near $t=0$, and this is useful in determining and proving
2939: spontaneous blow-up. 
2940: 
2941:   
2942: 
2943: In the following, $\partial_{\bf x}^{\bf j} \equiv
2944: \partial_{x_1}^{j_1} \partial_{x_2}^{j_2} ...\partial_{x_d}^{j_d}$,
2945: $|{\bf j}|= j_1 + j_2 + ... + j_d$, ${\bf x}$ is in a poly-sector
2946: $\mathcal S=\{\mathbf x:|\arg x_i|<\frac{\pi}{2}+\phi;|\mathbf x|>a\}$
2947: in $\CC^d$ where $\phi<\frac{\pi}{2n}$, ${\bf g} \left ( {\bf x}, t,
2948:   \{\mathbf y_{\bf j}\}_{|{\bf j}|=0}^{n-1} \right )$ is a function
2949: analytic in $\{\mathbf y_{\bf j}\}_{|{\bf j}|=0}^{n-1}$ near $\bf 0$
2950: vanishing as $|\bf x|\to \infty$.  The results in \cite{Invent2} hold
2951: for $n$-th order nonlinear {\em quasilinear} partial differential
2952: equations of the form
2953: \begin{equation}
2954:   \label{geneq}
2955: {\bf u}_t + \mathcal{P}(\partial_{\bf x}^{\bf j}){\bf u}+{\bf g}
2956:   \left ( {\bf x}, t, \{\partial_{\bf x}^{{\bf j}} {\bf u}\} \right )
2957:   =0
2958: \end{equation}
2959: \z where $\mathbf u\in\CC^m$, for large $|\bf x|$ in $S$.
2960: Generically, the constant coefficient operator $\mathcal{P}(\partial_{\bf
2961:   x}) $ in the linearization of $\mathbf g(\infty,t,\cdot)$ is
2962: diagonalizable. It is then taken to be diagonal, with eigenvalues
2963: $\mathcal{P}_j$.  $\mathcal{P}$ is subject to the requirement that for
2964: all $j\le m$ and ${\bf p}\ne 0$ in $\mathbb{C}^d$ with $|\arg p_i|\le
2965: \phi $ we have
2966: \begin{equation}
2967:   \label{conecond}
2968:   \Re\mathcal{P}^{[n]}_j (-{\bf p} )> 0
2969: \end{equation}
2970: where $\mathcal{P}^{[n]}(\partial_{\bf x})$ is the principal symbol of
2971: $\mathcal{P}(\partial_{\bf x})$. Then the following holds. (The
2972: precise conditions and results are given in \cite{Invent2}.)
2973: \begin{Theorem}[large $|\bf x|$ existence] 
2974:  \label{T1} Under the assumptions above, for any $T>0$   (\ref{geneq}) 
2975:  has a unique solution ${\bf u}$ that for $t\in [0,T]$ is $O(|\mathbf x|^{-1})$  and analytic in $\mathcal S$.
2976: \end{Theorem}
2977: 
2978: Determining asymptotic properties of solutions of PDEs is
2979: substantially more difficult than the corresponding question for ODEs.
2980: Borel-Laplace techniques however provide a very efficient way to
2981: overcome this difficulty.  The paper shows that formal series
2982: solutions are actually Borel summable, a fortiori asymptotic, to
2983: actual solutions.
2984: \begin{Condition}\label{Cond 2}
2985:   The functions ${\bf b}_{{\bf q}, {\bf k}} ({\bf x}, t)$ and ${\bf r}
2986:   ({\bf x}, t)$ are analytic in
2987:   $(x_1^{-\frac{1}{N_1}},...,x_d^{-\frac{1}{N_d}})$ for large $\bf|x|$
2988:   and some $N\in\NN$.
2989: \end{Condition}
2990: 
2991: \begin{Theorem}\label{TrB}
2992:   If Condition~\ref{Cond 2} and the assumptions of Theorem~\ref{T1}
2993:   are satisfied, then the unique solution $\mb f$ found there
2994:   can be written as
2995:   \begin{equation}
2996:     \label{acc3}
2997:     \mathbf{f}(\mathbf{x},t)=
2998: \int_{{\RR^+}^d}e^{-\mathbf{p}\cdot\mathbf{x}^{\frac{n}{n-1}} }{\mathbf{F}_1}(\mathbf{p},t)d\mathbf{p}
2999:   \end{equation}
3000:   where ${\mathbf{F}_1}$ is (a)  analytic at zero in
3001:   $(p_1^{\frac{1}{nN_1}},...,p_d^{\frac{1}{nN_d}})$; (b) analytic in $\bf p\ne 0$ in the
3002:   poly-sector $|\arg p_i|<\frac{n}{n-1}\phi+\frac{\pi}{2(n-1)}$, $i\le
3003:   d$; and (c) exponentially bounded in the latter poly-sector.
3004: \end{Theorem}
3005: Existence and asymptoticity of the formal power series follow as a
3006: corollary, using Watson's Lemma. 
3007: 
3008: Eariler, Borel summability has been shown for heat equation by Lutz,
3009: Miyake and Sch\"afke \cite{Lutz} and generalized to linear PDEs with
3010: constant coefficients by Balser \cite{Balser} and in special classes
3011: of higher order nonlinear PDEs in \cite{CPAM}.
3012: \subsection{More general irregular singularities and multisummability} The
3013: simple normalizing procedure described above does not always work;
3014: although this situation is in some sense nongeneric it plays an
3015: important role in some applications. In some instances mixed type
3016: transseries; for instance, by taking $e^{-x}Ei(x)$ which is a solution
3017: of a rank one differential equation as the rhs of (\ref{erf}) we get
3018: 
3019: \begin{equation}
3020:   \label{erfmix}
3021:   f'+2xf=e^{-x}Ei(x)
3022: \end{equation}
3023: which can be brought to a second order meromorphic equation by using
3024: the equation of $e^{-x}Ei(x)$, i.e. by adding (\ref{erfmix}) to its
3025: derivative,
3026: \begin{equation}
3027:   \label{erfmix2}
3028:   Df:=f''+(2x+1)f'+2(1+x)f=\frac{1}{x}
3029: \end{equation}
3030: The power series solution can be obtained contractively as in the
3031: previous examples. For the complete transseries solution, the
3032: substitution $f=e^w$ now yields
3033: $$\frac{1}{2}{w'}^2+xw'+x=0$$
3034: which has two solutions of level zero,
3035: \begin{eqnarray}
3036:   \label{degsolns}
3037:   & \displaystyle w_1=-x^2+x+\frac{1}{2}\ln x+K_1-\frac{1}{2x}-...\nonumber \\
3038:   & \displaystyle w_2=-x-\frac{1}{2}\ln x+K_2+\frac{1}{2x}+...
3039: \end{eqnarray}
3040: and thus 
3041: \begin{equation}
3042:   \label{gentr}
3043:   f=f_1+C_2 e^{-x^2+x} x^{\frac{1}{2}}f_2+ C_3 e^{-x}x^{-\frac{1}{2}}f_3
3044: \end{equation}
3045: with $f_1, f_2, f_3$ power series. 
3046: 
3047: Clearly, no change of independent variable brings (\ref{gentr}) to the
3048: form (\ref{transsf}). On the other hand, neither of variable $x$ nor
3049: $x^2-x$ (equivalently, $x^2$) can be used for Borel summation:
3050: 
3051: (i) Borel transform in the variable $x^2$ (or $x^2-x$) yields a
3052: function that has still a {\em divergent} power series at the origin.
3053: Indeed the the coefficients $c_n$  of the power series
3054: $f_1$, which satisfy the recurrence relation
3055: \begin{equation}
3056:   \label{recrel}
3057:   c_{n+1}=(n-1)\Big[c_{n}+\frac{1}{2}c_{n-1}-\frac{1}{2}(n-2)c_{n-2}\Big]
3058: \end{equation}
3059: (with $c_0=0=c_1=0, c_2=\frac{1}{2}$) grow like $n!$, while a Borel
3060: transform in $x^2$ only decreases the growth of the coefficients by a
3061: factor of, roughly, $(n!)^{-1/2}$. At a closer look it can be shown
3062: seen that the factorial growth of the coefficients and the presence of
3063: the term $e^{-x}$ are interrelated.
3064: 
3065: (ii) It can be checked that in the variable $x$, the Borel transform
3066: $F$ of the power series solution $f$ of (\ref{erfmix2}) is not Laplace
3067: transformable because of faster than exponential growth.  Indeed, $F$
3068: satisfies the equation
3069: \begin{equation}
3070:   \label{Borelt1}
3071:  2F'-pF =\frac{1}{1-p};\ \ \ F(0)=0
3072: \end{equation}
3073: The superexponential growth, in turn, can be related to the presence
3074: of $e^{-x^2+x}$ in the transseries. 
3075: 
3076: \subsubsection{}\label{751} This suggests decomposing the formal power series solution into two
3077: parts, one summable in the variable $x$, the other one in the variable
3078: $x^2$. We can adjust the parameters $a$ and $b$ so that
3079: 
3080: (i) The initial condition $c_0=0,c_1=-a/2,
3081: c_3=-\frac{a}{4}+\frac{b}{2}$ in the recurrence (\ref{recrel}) of the
3082: equation $Df=-a-bx^{-2}$ implies at most  $(n!)^{-1/2}$ growth of $c_n$
3083: and at the same time,
3084: 
3085: (ii) The Borel transform
3086: of equation $Df=a+x^{-1}+bx^{-2}$ has a solution with subexponential
3087: growth. 
3088: 
3089: The formal series solution of $Df=x^{-1}$ is thus decomposed in a sum
3090: of separately Borel summable series in the variables $x^2$ and $x$
3091: respectively.  That this (rather ad-hoc) procedure works is a
3092: reflection of general theorems in multisummability.
3093: 
3094: \subsubsection{Multisummability} A powerful and general technique, that of {\em acceleration and
3095:   multisummability} introduced by \'Ecalle, see \cite{Eca84} and
3096: \cite{EcalleNATO}, adequately deals with mixed divergences in very
3097: wide generality.  The procedure is universal and works in the same way
3098: for very general systems.
3099: 
3100: 
3101: In the case of solutions of meromorphic differential equations, mixed
3102: types of diveregence relate to the presence of exponential terms with
3103: exponents of different powers. Then multisummation consists in Borel
3104: transform with respect to the lowest power of $x$ in the exponents of
3105: the transseries, usual summation $\mathcal{S}$, a sequence of
3106: transformations called accelerations (which mirror in Borel space the
3107: passage from one power in the exponent to the immediately larger one)
3108: followed by a final Laplace transform in the largest power of $x$.
3109: More precisely (\cite{EcalleNATO}):
3110: \begin{equation}
3111:   \label{multisum}
3112:   \mathcal{L}_{k_1}\circ \mathcal{A}_{k_2/k_1}\circ \cdots\circ \mathcal{A}_{k_{q}/k_{q-1}} \mathcal{S}\mathcal{B}_{k_q}
3113: \end{equation}
3114: where $ (\mathcal{L}_{k}f)(x)=(\mathcal{L}f)(x^k)$, $\mathcal{B}_{k}$ is the
3115: formal inverse of $\mathcal{L}_{k}$, $\alpha_i\in (0,1)$ and the
3116: acceleration operator $\mathcal{A}_{\alpha}$ is formal the image, in
3117: Borel space, of the change of variable from $x^{\alpha}$ to $x$ and is
3118: defined as
3119: \begin{equation}
3120:   \label{accel}
3121:   \mathcal{A}_{\alpha}\phi=\int_0^{\infty}C_{\alpha}(\cdot,s)\phi(s)ds
3122: \end{equation}
3123: and where, for $\alpha\in(0,1)$,  the kernel $C_{\alpha}$ is defined as
3124: \begin{equation}
3125:   \label{accel2}
3126:   C_{\alpha}(\zeta_1,\zeta_2):=\frac{1}{2\pi i}\int_{c-i\infty}^{c+i\infty}e^{\zeta_2 z-\zeta_1 z^{\alpha}}dz
3127: \end{equation}
3128: where we adapted the notations in \cite{Balser} to the fact that the
3129: formal variable is large.  In our example, $q=2, k_2=1, k_1=2$.
3130: 
3131: In \cite{[Br]} Braaksma proved of multisummability of series solutions
3132: of general nonlinear meromorphic ODEs using Borel transforms in the
3133: spirit of Ecalle'e theory.
3134: 
3135: 
3136: \begin{Note}
3137:   (i) Multisummability of type (\ref{multisum}) can be equivalently
3138:   characterized by decomposition of the series into terms which are
3139:   ordinarily summable after changes of independent variable of the
3140:   form $x\to x^\alpha$.  This is shown in \cite{Balser} where it is
3141:   used to give an alternative proof of multisummability of series
3142:   solutions of meromorphic ODEs, closer to the cohomological point of
3143:   view of Ramis and Sibuya.
3144:   
3145:   (ii) More general multisummability is described by Ecalle
3146:   \cite{EcalleNATO}, allowing, among others, for stronger than
3147:   power-like acceleration. This is relevant to more general
3148:   transseries equations.
3149:   
3150:   (iii) We expect that the finite generation property of transseries
3151:   allows for the cohomological approach (i) combined with (ii) to
3152:   prove multisummability of transseries solutions of more general
3153:   contractive equations, in the sense of Theorem~\ref{PFp}. This would
3154:   give a rigorous proof of closure of analyzable functions under 
3155:   all natural operations in the sense of Ecalle.
3156: \end{Note}
3157: \begin{thebibliography}{99}
3158: 
3159:   
3160: 
3161: 
3162: \bibitem{Balser} W Balser {\em From divergent power series to
3163:     analytic functions, Springer-Verlag, 1994.}
3164: 
3165: \bibitem{Berry-hypB} Berry, M. V.; Howls, C. J. 1990 Hyperasymptotics.
3166:   {\em Proc. R. Soc. Lond. A} {\bf 430}, 653--668.
3167: \bibitem{BerryKeating} Berry, M. V.; Keating, J. P. 1992
3168:  A new asymptotic representation for $\zeta(\frac12+it)$ and quantum
3169: spectral determinants. {\em Proc. Roy. Soc. London Ser. A \textbf{437} (1992),
3170: no. 1899, 151--173.}
3171: 
3172: \bibitem{[Br]} B L J     Braaksma, \emph{Multisummability of formal power series solutions of nonlinear meromorphic differential equations}, Ann.  Inst. Fourier, Grenoble,{\bf 42}, 3, 517-540 (1992).
3173: 
3174: \bibitem{Orszag} C M Bender and S A Orszag {\em Advanced mathematical
3175:     methods for scientists and engineers I; Asymptotic methods and
3176:     perturbation theory}, Mc Graw Hill (1978) and Springer (1999).
3177: 
3178: 
3179: \bibitem{BorelM} E Borel {\em Le\c cons sur les fonctions monog\`enes,
3180: Gauthier-Villars}, Paris (1917).
3181: 
3182: \bibitem{Braaksma} B L J Braaksma {\em Transseries for a class of
3183:     nonlinear difference equations}  J. Differ. Equations Appl. 7,
3184: no. 5, 717--750 (2001). 
3185: 
3186:  \bibitem{Braaksma-Kuik}  B. L. J. Braaksma,  R Kuik (submitted).
3187:  
3188: 
3189: 
3190: \bibitem{Conway} J H  Conway, {\em On numbers and games}, Academic Press (1976). \bibitem{CPAM} O. Costin, S. Tanveer, Existence and uniqueness for a
3191:   class of nonlinear higher-order partial differential equations in
3192:   the complex plane, Comm. Pure Appl. Math, Vol. LIII, 1092---1117
3193:   (2000).
3194: \bibitem{Invent2} O. Costin, S. Tanveer {\em On the existence and uniqueness of solutions of nonlinear
3195:   evolution systems of PDEs in $\RR^+\times\CC^d$, their asymptotic
3196:   and Borel summability properties} preprint (2003).
3197: 
3198: 
3199:   
3200: \bibitem{Duke} O Costin {\em On Borel summation and
3201:     Stokes phenomena for rank one nonlinear systems of ODE's} {\em Duke
3202:     Math. J. Vol. 93, No.2 pp. 289--344 (1998).}
3203:   
3204: \cite{Kuik} R Kuik {\em Transseries in Difference and Differential Equation} 
3205: Thesis, University of Groningen (2003).
3206: 
3207: 
3208: 
3209: 
3210: \bibitem{Invent} O Costin and R D Costin {\em On the formation of
3211: singularities of solutions of nonlinear differential systems in
3212: antistokes directions} To appear in Inventiones Mathematicae.
3213: 
3214: 
3215: 
3216: \bibitem{Eca84} J Ecalle {\em Les fonctions resurgentes, vol. I, II and III},
3217:   Publ. Math. Orsay, 1985.
3218:  \bibitem{EcalleNATO} 
3219:  J \'Ecalle, {\em in Bifurcations and periodic orbits of 
3220: vector fields, NATO ASI Series, Vol. 408,  1993}
3221: \bibitem{FlN} H Flashka and A C Newell {\em Monodromy and spectrum
3222:     preserving transformations}
3223: Commun. Math. Phys. {\bf 76} pp. 65--116 (1980).
3224: \bibitem{EcalleMenous} J Ecalle, F Menous, {\em Well-behaved
3225:     convolution averages and the non-accumulation theorem for
3226:     limit-cycles.}  The Stokes phenomenon and Hilbert's 16th problem,
3227:   71--101, World Sci. Publishing (1996)
3228:   \bibitem{Fuchs} L Fuchs {\em Sur quelques \'equations diff\'erentielles 
3229: lin\'eaires du second ordre} C. R. Acad. Sci., Paris {\bf 141}
3230: pp. 555-558 (1905).
3231: 
3232: 
3233: \bibitem{Hardy} G H Hardy. Orders of Infinity, Cambridge Univ. Press,
3234:   (1954).
3235: 
3236: \bibitem{Kuik} R Kuik, Transseries in differential and difference equations,
3237:   PhD Thesis, University of Groningen, ISBN 90-367-1771-x (2003).
3238:   \bibitem{Lutz} D. A. Lutz, M. Miyake and R.
3239:                  Sch\"afke On the Borel summability of divergent
3240:                  solutions of the heat equation, Nagoya Math. J. {\bf
3241:                  154}, 1, (1999).
3242: 
3243: \bibitem{VDD} L van den Dries, A Macintyre and D Marker {\em The
3244:     elementary theory of restricted analytic elds with exponentiation}
3245:   Ann. of Math. 140 (1994)
3246: 
3247: \bibitem{Wasow} W Wasow, {\em {Asymptotic expansions
3248: for ordinary differential equations}}, Interscience Publishers (1968).
3249: 
3250: \end{thebibliography}
3251: 
3252: \end{document}
3253: 
3254: 
3255: 
3256: 
3257: 
3258: 
3259: 
3260: