INSTRUCTIONS Remember to latex the file twice to get correct cross references The text between the lines BODY and ENDBODY is made of 4538 lines and 138546 bytes (not counting or ) In the following table this count is broken down by ASCII code; immediately following the code is the corresponding character. 75523 lowercase letters 3327 uppercase letters 8828 digits 13291 ASCII characters 32 66 ASCII characters 33 ! 39 ASCII characters 34 " 29 ASCII characters 35 # 3696 ASCII characters 36 $ 652 ASCII characters 37 % 227 ASCII characters 38 & 275 ASCII characters 39 ' 2395 ASCII characters 40 ( 2392 ASCII characters 41 ) 695 ASCII characters 42 * 173 ASCII characters 43 + 2305 ASCII characters 44 , 758 ASCII characters 45 - 998 ASCII characters 46 . 57 ASCII characters 47 / 329 ASCII characters 58 : 74 ASCII characters 59 ; 66 ASCII characters 60 < 638 ASCII characters 61 = 86 ASCII characters 62 > 2 ASCII characters 64 @ 63 ASCII characters 91 [ 8432 ASCII characters 92 \ 63 ASCII characters 93 ] 562 ASCII characters 94 ^ 2487 ASCII characters 95 _ 73 ASCII characters 96 ` 4613 ASCII characters 123 { 605 ASCII characters 124 | 4613 ASCII characters 125 } 114 ASCII characters 126 ~ BODY \documentstyle[12pt,amssymbols]{article} \setlength{\hoffset}{-1cm} \setlength{\voffset}{-2cm} \setlength{\textwidth}{16cm} \setlength{\textheight}{23.5cm} \setlength{\parindent}{8mm} \frenchspacing %****************************** %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% %%% %%% FOR SHORT %%% %%% \newcommand{\beq}{ \begin{equation} } \newcommand{\eeq}{ \end{equation} } \newcommand{\beqno}{ \begin{equation*} } \newcommand{\eeqno}{ \end{equation*} } \newcommand{\beqa}{ \begin{eqnarray} } \newcommand{\eeqa}{ \end{eqnarray} } \newcommand{\beqano}{ \begin{eqnarray*} } \newcommand{\eeqano}{ \end{eqnarray*} } %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% %%% %%% THEOREMS AND ALIKE %%% %%% \renewcommand{\theequation}{\arabic{section}.\arabic{equation}} \newtheorem{theorem}{Theorem}[section] \newtheorem{definition}{Definition}[section] \newtheorem{proposition}{Proposition}[section] \newtheorem{lemma}{Lemma}[section] \newtheorem{remark}{Remark}[section] \newtheorem{sublemma}{Sublemma}[section] \newtheorem{corollary}{Corollary}[section] \newcommand\dfn[1]{ \begin{definition}\label{#1} \rm} \newcommand\dfntwo[2]{ \begin{definition}[#1]\label{#2} \rm} \newcommand\edfn{ \end{definition} } \newcommand\rem[1]{ \begin{remark}\label{#1} \rm} \newcommand\erem{ \end{remark} } \newcommand\thm[1]{ \begin{theorem}\label{#1}} \newcommand\thmtwo[2]{ \begin{theorem}[#1]\label{#2}} \newcommand\ethm{ \end{theorem} } \newcommand\pro[1]{ \begin{proposition}\label{#1}} \newcommand\protwo[2]{ \begin{proposition}[#1]\label{#2}} \newcommand\epro{ \end{proposition} } \newcommand\lem[1]{ \begin{lemma}\label{#1}} \newcommand\lemtwo[2]{ \begin{lemma}[#1]\label{#2}} \newcommand\elem{ \end{lemma} } \newcommand\sublem[1]{ \begin{sublemma}\label{#1}} \newcommand\sublemtwo[2]{ \begin{sublemma}[#1]\label{#2}} \newcommand\esublem{ \end{sublemma} } \newcommand\cor[1]{ \begin{corollary}\label{#1}} \newcommand\cortwo[2]{ \begin{corollary}[#1]\label{#2}} \newcommand\ecor{ \end{corollary} } %%% %%% %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% %%% %%% REFERRING TO %%% % \newcommand\equ[1]{{\rm (\ref{#1})}} % % % %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% %%% %%% VARIOUS %%% \newcommand{\ie}{{\it i.e. }} \newcommand{\eg}{{\it e.g. }} \newcommand{\etc}{{\it etc}} \newcommand{\HB}{\hfill\break} \newcommand{\acapo}{\linebreak} \newcommand{\HALF}{{\textstyle{1\over 2}}} %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% %%% A CAPO..... %%% %%% %%% \newcommand{\giu}{{\medskip\noindent}} \newcommand{\Giu}{{\bigskip\noindent}} \newcommand{\nl}{{\smallskip\noindent}} \newcommand{\nin}{{\noindent}} % %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% % \newcommand{\qed}{\hskip.5truecm \vrule width 1.7truemm height 3.5truemm depth 0.truemm} \newcommand{\qedd}{\vrule width 1.7truemm height 1.7truemm depth 0.truemm} \newcommand{\QED}{\hfill\smallskip \line{\hfill\vrule height 1.8ex width 2ex depth +.2ex \ \ \ \ \ \ } \bigskip} \newcommand{\PROOF}{\medskip\noindent{\bf Proof\ }} \newcommand{\REMARK}{\medskip\noindent{\bf Remark\ }} % %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% %%% %%% MATH MODE DEFINITIONS: %%% % % \newcommand{\ci}{ {C^\infty} } \newcommand{\ig}{ {\int} } \newcommand{\dpr}{ {\partial} } % \newcommand{\torus}{ {\Bbb T} } \renewcommand{\natural}{ {\Bbb N} } \newcommand{\real}{ {\Bbb R} } \newcommand{\integer}{ {\Bbb Z} } \newcommand{\complex}{ {\Bbb C} } \newcommand{\tN}{ {\torus^N } } \newcommand{\rN}{ {\real^N} } \newcommand{\cN}{ {\complex^N } } \newcommand{\zN}{ {\integer^N } } % \renewcommand{\a }{ {\alpha} } \renewcommand{\b}{ {\beta} } \newcommand{\g}{ {\gamma} } \newcommand{\G}{ {\Gamma} } \renewcommand{\d}{ {\delta} } \newcommand{\D}{ {\Delta} } \newcommand{\e }{ {\varepsilon} } \newcommand{\et}{ {\eta} } \newcommand{\th }{ {\theta} } \newcommand{\vth }{ {\vartheta} } \newcommand{\k}{ {\kappa} } \renewcommand{\l}{ {\lambda} } \renewcommand{\L}{ {\Lambda} } \newcommand{\m}{ {\mu} } \newcommand{\n}{ {\nu} } \newcommand{\x }{ {\csi} } \newcommand{\X }{ {\Csi} } \newcommand{\p}{ {\pi} } \renewcommand{\P}{ {\Pi} } \newcommand{\r}{ {\rho} } \newcommand{\s}{ {\sigma} } \newcommand{\Si}{ {\Sigma} } \renewcommand{\t}{ {\tau} } \newcommand{\f}{ {\varphi} } \newcommand{\ph}{ {\phi} } \newcommand{\F}{ {\Phi} } \renewcommand{\c}{ {\chi} } \newcommand{\ps}{ {\psi} } \renewcommand{\o}{ {\omega} } \renewcommand{\O}{ {\Omega} } % \renewcommand{\Im}{{\rm Im}} \renewcommand{\Re}{{\rm Re}} \renewcommand{\=}{ {\ \equiv\ } } % \newcommand{\implies}{\ \Longrightarrow\ } \newcommand{\sse}{\ \Longleftrightarrow\ } %%% \newcommand{\Xk}{ X^{(k)} } \newcommand{\Xj}{ X^{(j)} } \newcommand{\Xh}{ X^{(h)} } \newcommand{\Tr}{ {T_r} } % \newcommand{\cA}{ {\cal A} } \newcommand{\cT}{ {\cal T} } \newcommand{\cR}{ {\cal R} } \newcommand{\cH}{ {\cal H} } \newcommand{\cC}{ {\cal C} } \newcommand{\cF}{ {\cal F} } \newcommand{\calN}{ {\cal N} } % \newcommand{\noin}{ {\not\in} } \renewcommand{\lg}{ {\langle} } \newcommand{\rg}{ {\rangle} } \newcommand\agl[1]{ |\lg{#1}\rg| } \newcommand\gl[1]{ \lg{#1}\rg } \newcommand{\getr}{{{\ge}_{{\phantom{.}}_{\!\Tr}}}} \newcommand{\baj}{ {\bar\jmath}} \newcommand{\bai}{ {\bar\imath}} \newcommand\bks{\backslash} \newcommand{\zNn}{ {\zN\backslash \{0\} } } % \newcommand\bcR{ {\overline {\cal R}} } \newcommand\bR{ {\overline R} } \newcommand\bT{{ \overline T}} \newcommand\bu{{ \bar u}} \newcommand\bv{{ \bar v}} \newcommand\bw{{ \bar w}} \newcommand\bk{{ \bar \k}} \newcommand\bl{{ \bar \l}} \newcommand\bs{{ \bar \s}} \newcommand\tT{{ \widetilde T}} \newcommand\tcT{{ \widetilde \cT}} \newcommand\hT{{ \widehat T}} \newcommand\hcT{{ \widehat \cT}} %******************************************** \begin{document} %%%%%%%%%%%%%%%%%%%%%% sections: \title{A Direct Proof of a Theorem by Kolmogorov \\ in Hamiltonian Systems} \author{ L. Chierchia \\ {\footnotesize Dipartimento di Matematica} \\ {\footnotesize Universit\`a di Genova} \\ {\footnotesize via L.B.Alberti 4, 16132 Genova (Italy) }\\ {\footnotesize (Internet: chierchia@mat.utovrm.it)} \and C. Falcolini \\ {\footnotesize Dipartimento di Matematica} \\ {\footnotesize Universit\`a di Roma ``Tor Vergata"}\\ {\footnotesize via della Ricerca Scientifica, 00133 Roma (Italy)} \\ {\footnotesize (Internet: falcolini@mat.utovrm.it)}} \date{June 8, 1993} \maketitle \begin{abstract} \nin We present a direct proof of Kolmogorov's theorem on the persistence of quasi-periodic solutions for nearly integrable, real--analytic Hamiltonian systems with Hamiltonians of the form $\frac{1}{2} y\cdot y-\e f(x)$ where $(y,x)$ $\in$ $\rN \times\tN$ are standard symplectic coordinates. The method of proof consists in constructing, via graph theory, the formal solution as a formal power series in $\e$ and to show that the $k^{\rm th}$ coefficient of such formal series can be bounded by a constant to the $k^{th}$ power. All details are presented in a self contained way (included what is needed from the theory of graphs). \end{abstract} {\footnotesize \tableofcontents} %******************************************** \section{Introduction} % \label{sec:intro} \setcounter{equation}{0} % Each Section of this paper begins with a short summary of what is done in that Section: Here we present a bit of history of Kolmogorov's Theorem on the stability of quasi--periodic solutions in Hamiltonian systems followed by a rough outline of a novel proof, based on a tree representation of formal series. \giu A 1954 theorem by Kolmogorov \cite{Ko} guarantees the existence of infinitely many quasi--periodic solutions for the (standard) Hamilton equations associated to real--analytic, ``spatially periodic" Hamiltonians of the form $H(y,x;\e)$ $=$ $ h(y)+\e f(y,x)$, ($y\in U$ open subset of $\rN$, $x\in\tN\=\rN/(2\pi\zN)$, $\e\ge 0$), provided the Hessian matrix $(\frac{\dpr h}{\dpr y_i \dpr y_j})$ is invertible on $U$ and provided $|\e|$ is sufficiently small. We recall that a solution $(y(t),x(t))$ is called (maximal) quasi--periodic if there exists a rationally independent vector $\o\in\rN$ and smooth functions $Y,X:\tN\to\rN$ such that $(y(t),x(t))=(Y(\o t),\o t+ X(\o t))$. A detailed proof, different from that outlined by Kolmogorov in \cite{Ko}, was provided, in 1963, by V.~I.~Arnold in \cite{A1} and J.~Moser \cite{Mo1} proved, in the same period, an analogous theorem for symplectic diffeomorphisms removing the hypothesis of analyticity of the perturbation. The {\it corpus} of results and methods stemmed out from Kolmogorov's original ideas is now known as ``KAM (Kolmogorov--Arnold--Moser) theory". It is not difficult to write down {\em formal $\e$--expansion} of quasi--periodic solutions; the problem is then turned into whether such series are convergent or not. This question was extensively and thoroughly investigated by H.~Poincar\'e in his {\it m\'ethodes nouvelle de la m\'ecanique c\'eleste} \cite{Po}. Poincar\'e, following M.~Lindstedt ({\em M\'emoires de l'Acad\'emie de Saint--P\'etersbourg}, 1882), considered formal quasi--periodic solutions for which {\em $\o$ depends on $\e$} and showed that such series are {\em divergent} (\cite{Po}, vol. II, \S XIII). The main problem in this context is that the formal solution has Fourier coefficients which are divided by terms of the type $\o\cdot n$ $\equiv$ $\sum_{i=1}^N \o_i n_i$ with $n\in \zNn$, and such factors (``small divisors") become arbitrarily small as $|n|\to\infty$. In fact (see Section~\ref{sec:diverge} below), the {\em repeated occurrence of small divisors} (``resonances") in the $k^{\rm th}$ coefficient of the formal solution leads to contributions of the order of $(k!)^a$ with $a>0$. However, in 1967 Moser \cite{Mo2} showed {\em indirectly} (see below) that the formal series {\em converges} leading to analytic solutions, provided the {\em $\e$--independent} vector $\o$ satisfies certain number theoretic assumptions (verified by almost all (with respect to Lebesgue measure) vectors in $\rN$; see the ``Diophantine condition" \equ{dioph} below). This means that the $k^{\rm th}$ coefficient of the formal solution contains {\em many huge contributions which compensate among themselves} producing terms that behaves like a constant to the $k^{\rm th}$ power. Moser's proof, as well as {\em all} proofs in KAM theory (up to the 1988 Eliasson work \cite{E}), are based on a ``rapidly convergent" iteration technique. The strategy, similarly to Newton's method of tangents, consists in finding solutions of a nonlinear differential equation $\calN(u)=0$ by solving recursively a sequence of approximate equations of the form $\calN(u_j)=e_j$ where the size of the ``error function" $e_j$ becomes {\em quadratically smaller} at each step of the procedure. Such an approach has many advantages and can be used very effectively (see \cite{CC} for accurate estimates and Appendix~2 in \cite{CP} for a five--page proof) but is indirect and hides the mechanism beyond the above mentioned compensations. In the different context of linearization of germs of analytic diffeomorphisms, C.~L.~Si\-e\-gel \cite{S} succeeded in 1942 in proving directly the convergence of formal power series involving small divisors: the crucial difference being that the small divisors are of the form $\o\cdot n$ {\em with $n\in\natural^N$} so that a given divisor {\em occurs at most once} in each coefficient of the formal solution (\ie ``there are no resonances"). In 1988 Eliasson, in a Report of the University of Stockholm \cite{E} (which to the best of our knowledge has not been published), extended Siegel's method so as to cover the Hamiltonian case. We present a different version of Eliasson's proof based on a tree representation of the formal solutions and on the explicit exhibition of compensations of huge contributions. We follow closely \cite{E} in its convenient reformulation of Siegel's method (see Lemma~\ref{lem:se} below) which allows to bound product of (possibly) small divisors whenever (certain dangerous) repetition do not occur; however for the crucial part (grouping together the huge contributions) we adopt a different approach which now we outline. In order to simplify the presentation {\em we consider the particular model with Hamiltonian} $H= \frac{1}{2}y\cdot y-\e f(x)$: The extension of our method to the general situation is rather straightforward but would lead to heavier notations without providing any new insight. The {\em starting point} is to express the coefficients of the formal solution in terms of {\em labeled rooted trees}\footnote{The use of graph theory in connection with formal power series is natural and very old (see \eg \cite{GJ} and references therein); for connections with KAM theory see \cite{CG} and \cite{G}.}. This allows to express the $k^{\rm th}$ coefficient in terms of a sum over all possible labeled rooted trees of order $k$, and over all possible (and ``admissible") integers $\a_v\in\zNn$ ($v$ denoting a vertex of a tree) of \beq \prod_V f_{\a_v} \prod_E \a_v\cdot\a_{v'}\ \prod_{V}\d_v^{-2} \label{1.1} \eeq where $V$ and $E$ denote, as customary, the set of vertices $v$ and edges $vv'$ of a given tree (see Appendix~\ref{app:tree}), $f_n$ denote Fourier coefficients of $f$ and $\d_v$ represents the divisor associated to the vertex $v$ \ie $\d_v\=\o\cdot \sum_{v'\le v} \a_{v'}$ (rooted trees have a natural order according to which the root $r$ is $>v$ for all vertices $v\in V$, $v\neq r$; the square in \equ{1.1} comes from the fact that the Hamilton equations for our model can be immediately written as the second order system $\ddot x = f_x$; ``admissible" means that $\d_v\neq 0$ $\forall$ $v$ \ie $\sum_{v'\le v} \a_{v'}\neq 0$ $\forall$ $v\in V$). Now, the {\em main point} is to make a {\em partition} of the trees of order $k$ into families $\cF$ (called below ``complete families") so as to be able to bound sums over such families by a constant to the $k^{\rm th}$ power (recall that even single contributions given by \equ{1.1} may have size of order $(k!)^a$ as already mentioned). The main technical estimate is % \beq \big| \sum_{T\in\cF} \prod_{vv'\in E(T)} \a_v\cdot\a_{v'} \prod_V \d_v^{-2}\big| \le \prod_{v\in V} |\a_v|^{\deg_\cF v} \ c_4^k\ \prod_{v\in V} |\a_v|^{\b_4} \label{1.2} \eeq for suitable constants $c_4$ and $\b_4$, determined below, depending on $N$ and on the numerical properties of the vector $\o$; $\deg_\cF v$ is defined as the maximum degree of $v$ when $T$ varies in the family $\cF$. Obviously, it is crucial that complete families either do not intersect or coincide (in fact it is much easier to find families of trees for which \equ{1.1} holds but which do not form a partition: clearly such families are of no use for our purposes). The construction of the partition of complete families is the delicate part of our paper\footnote{In \cite{G} the problem analyzed here is also investigated with similar tools; however the families of trees considered there, as well as most of the technical aspects, are different from ours.}. Given Eliasson's version of Siegel's method (which we include for completeness) and the identification of complete families, the convergence of the formal solutions follows very easily (under the Diophantine condition on $\o$). All the constants are computed explicitly (see Remark~\ref{rem:constants} below). \giu We close this introduction by mentioning possible directions of future researches. \nin (i) The method of proof presented here, being based on a very direct approach, seems particularly suitable for computer--aided implementations and might shed some new light on the difficult problem of the break--down of stability of quasi--periodic solutions in connection with the $\e$--singularities of the function $X$ (see, \eg, \cite{CC}, \cite{BC}, \cite{BCCF}, \cite{FL} and references therein). \nin (ii) Let $(x^1,x^2)$ with $x^i\in \torus^{N_i}$ and $N_1+N_2=N$, let $\o\in \real^{N_1}$ and let $x^2_0$ be a nondegenerate critical point of the periodic function $x^2\in\torus^{N_2}\to\int_{\torus^{N_1}} f(x^1,x^2)dx^1$. Then, it is not difficult to see that (if $\o$ satisfies a Diophantine condition) there are formal (non maximal) quasi--periodic solutions whose first term ($\e=0$) is given by $(\o t,x_0^2)$. It would be nice to extend the proof in this paper so as to establish convergence for such formal series. \nin (iii) It is well known (\cite{Ru}) that maximal quasi--periodic solutions are stable under weaker assumptions on the vector $\o$ than the classical one made here. It might be interesting to see how far, in this direction, can lead the technique worked out in this paper. \giu The paper is organized as follows. In Section~\ref{sec:formal} existence and uniqueness of formal quasi--periodic solutions are established. In Section~\ref{sec:tree} the tree expansion of formal quasi--periodic solutions is given (the purely combinatorics aspects are proven in Appendix~\ref{app:combina}). The occurrence of huge terms is explicitly exhibited in Section~\ref{sec:diverge}. The construction of complete families is carried out in Section~\ref{sec:resona} allowing to formulate the results in Section~\ref{sec:estima} divided in four Lemmas and one Theorem. Detailed proofs are given in Section~\ref{sec:proofs}. Graph theory is used here mainly as a (very useful!) language and the basic definitions (enough to read our paper without any knowledge of graph theory) are presented in Appendix~\ref{app:tree}. %******************************************** \section{Formal Quasi--Periodic Solutions} % \label{sec:formal} \setcounter{equation}{0} % We recall the notions of quasi--periodic and formal quasi--periodic solution for a ``spatially periodic" Hamiltonian. \giu Let $H(y,x)$ be a $\ci(U\times \tN)$ Hamiltonian where $U$ is an open set of $\rN$ and $\tN$ $\=$ $\rN/2\pi\zN$ (\ie $H$ can be seen as a function of the $2N$ variables $y_1,...y_N$,$x_1,...,x_N$, $2\pi$--periodic in $x_i$). Consider the standard Hamilton equations: \begin{equation} \dot y= - H_x\ ,\qquad \dot x= H_y \label{2.1} \end{equation} where, as usual, $H_x\=\dpr_x H\=(\frac{\dpr H}{\dpr x_1},..., \frac{\dpr H}{\dpr x_N})$ and $H_y\=\dpr_y H\=(\frac{\dpr H}{\dpr y_1},..., \frac{\dpr H}{\dpr y_N})$. A solution $(y(t),x(t))$ of \equ{2.1} is called {\em (maximal) quasi--periodic} with frequency $\o\in\rN$ if $\o$ is rationally independent (\ie $\o\cdot n\=\sum_{i=1}^N\o_i n_i=0$ for some $n\in \zN$ implies $n=0$) and if there exist smooth functions $Y,X:\tN\to\rN$ such that \begin{equation} y(t)=Y(\o t)\ ,\qquad x(t)=\o t+ X(\o t) \label{2.2} \end{equation} The rational independence of $\o$ easily implies that the functions $\th\in\tN\to Y(\th),X(\th)$ satisfy the system of equations: \newpage \begin{eqnarray} D Y =& -H_x(Y,\th+X)\ , &\qquad( D\=\sum_{i=1}^N\o_i\dpr_{\th_i}) \nonumber\\ \o + D X =& H_y(Y,\th+X) \label{2.3} \end{eqnarray} Viceversa, given a solution of \equ{2.3}, \equ{2.2} (or more generally \equ{2.2} with $\o t$ replaced by $\th+\o t$ for any $\th\in\tN$) is a quasi--periodic solution of \equ{2.1}. ``Le probl\`eme g\'en\'eral de la dynamique" according to Henri Poincar\'e (\cite{Po}, vol. I, chapter I, \S 13) is the study of the equations governed by the ``nearly--integrable" Hamiltonian \begin{equation} H(y,x;\e)\=h(y)+\e f(y,x) \label{h} \end{equation} for small values of the parameter $\e$. As we shall see below, if the Hessian matrix $h''$ $\=$ $h_{yy}$ $\=$ $\big(\frac{\dpr^2 h}{\dpr y_i \dpr y_j}\big)$ is invertible, then there are ``a lot" of ``formal quasi--periodic solutions" of the equations \equ{2.1} with Hamiltonian \equ{h}. A {\em formal $\e$--power series $F$, over $\tN$}, is an infinite sequence of $\ci(\tN)$ functions $\{ F_k\}_{k\ge 0}$, $F_k=F_k(\th)$, and it is customary to write $F\sim \sum_{k\ge 0} F_k \e^k $. If $g$ is a $\ci$ function and $F\sim \sum_{k\ge 0} F_k \e^k$ a formal series, one naturally defines the formal series $g\circ F\sim \sum_{k\ge 0} G_k \e^k$ by setting \begin{equation} G_k= \left[ g\big( \sum_{h=0}^k F_h \e^h\big) \right]_k \= \frac{1}{k!}\ \frac{d^k}{d\e^k}\ g\big(\sum_{h=0}^k F_h \e^h\big) \Big|_{\e=0} \label{2.4} \end{equation} A formal solution of \equ{2.3} with $H$ as in \equ{h} is a couple of (vector--valued) formal $\e$--power series over $\tN$, $Y\sim \sum_{k\ge 0} Y^{(k)} \e^k$, $X\sim \sum_{k\ge 0} X^{(k)} \e^k$, ($Y^{(k)}, X^{(k)}:$ $\tN\to\rN$), verifying \equ{2.3} in the sense of formal series (\ie ``$=$" should be replaced by ``$\sim$") or equivalently: \begin{eqnarray} D Y^{(0)} =& 0 \ ,\phantom{Y^{(h)}} & D Y^{(k)} = \left[ -f_x(\sum_{h=0}^{k-1} Y^{(h)} \e^h , \th+\sum_{h=0}^{k-1} X^{(h)}\e^h )\right]_{k-1} \nonumber\\ \o + D X^{(0)} =& h_y(Y^{(0)}) \ ,\ & D X^{(k)} = \left[ H_y(\sum_{h=0}^{k} Y^{(h)}\e^h , \th+\sum_{h=0}^{k-1} X^{(h)}\e^h )\right]_{k} \label{2.5} \end{eqnarray} where $k\ge 1$; notice the different ranges of indices in the equation for $D \Xk$ due to the particular form of \equ{h}. \begin{proposition} \label{pro:ef} Let $H$ as in \equ{h} be $\ci$ in a neighborhood of $\{y_0\}\times \tN$ with $y_0$ such that $h''(y_0)$ is invertible and $\o\=h'(y_0)\=h_y(y_0)$ is ``Diophantine" \ie such that % \footnote{\rm It is well known that for any $\t>N$, almost all (with respect to Lebesgue measure) $\o\in \rN$ satisfy \equ{dioph} with some $\g>0$ (see, \eg, \cite{A2}, chapter I, \S 3) while if $\t=N-1$ then for any $\o\in\rN$ there exist a $\g>0$ and an infinite number of $n\in\zN$ such that \equ{dioph} is violated (this is a theorem by Dirichlet, see, \eg, \cite{Sc}).} % \begin{equation} |\o\cdot n| \ge \frac{1}{\g |n|^\t}\ ,\qquad \forall\ n\in\zN\backslash\{0\}\ \quad {\rm for\quad some}\quad \g,\t>0\ . \label{dioph} \end{equation} Then there exists a unique formal quasi--periodic solution $Y,X$ of \equ{2.5} with \begin{equation} \ig_{\tN} X^{(k)} d\th=0\ ,\qquad \forall\ k\ . \label{av} \end{equation} \end{proposition} % \PROOF We construct the formal solution by induction over $k$. For $k=0$ we get immediately from \equ{2.5} that $Y^{(0)}$ and $X^{(0)}$ are constant vectors and that $h'(Y^{(0)})=\o$ (as if $g$ is a smooth function over $\tN$ and $Dg=a$ with some constant $a$, then $g\=0=a$) so that, from our hypotheses it follows that $Y^{(0)}=y_0$; requirement \equ{av} fixes the constant vector $X^{(0)}$ to be zero. Let, now, $k\ge 1$ and assume that $y_0, Y^{(1)},...,Y^{(k-1)}$, $X^{(1)},...,X^{(k-1)}$, are smooth functions over $\tN$ solving \equ{2.5} with $\int_\tN X^{(h)}=0$. We claim that \begin{equation} \int_\tN \phi^{(k)}(\th)\= \int_\tN\Big[ f_x \big( \sum_{h=0}^{k-1} \e^h Y^{(h)}, \th +\sum_{h=0}^{k-1} \e^h X^{(h)} \big) \Big]_{k-1} = 0 \label{clm} \end{equation} (note that the claim is obvious for $k=1$ as in such a case \equ{clm} is just the average of the gradient of a periodic function). If the claim is true, then the proposition follows easily: $\phi^{(k)}(\th)$ would be a $\ci$ (vector--\-valued) function over $\tN$ with zero average and therefore the solutions of the equation $DY^{(k)}= \phi^{(k)}$ are given by $Y^{(k)}=D^{-1}\phi^{(k)} + c_k$ with $c_k$ constant and $D^{-1}\phi^{(k)}$ the smooth function with zero average given in Fourier expansion by \begin{equation} D^{-1}\phi^{(k)}\= \sum_{{n\in\zN}\atop{n\neq 0}} \frac{\phi^{(k)}_n} {i\o\cdot n} e^{i n\cdot \th} \label{phi} \end{equation} where $\phi^{(k)}_n$ are the Fourier coefficients of $\phi^{(k)}$ (notice that the fast decay of the Fourier coefficients of the smooth function $\phi^{(k)}$ yields the fast decay, in view of \equ{dioph}, of the coefficients in \equ{phi}, ensuring that $D^{-1}\phi^{(k)}$ is $\ci(\tN)$). The constant $c_k$ is not arbitrary, as the right hand side of the second of \equ{2.5}, in order to make sense, must have vanishing mean value over $\tN$; rewriting such equation we get \begin{eqnarray} D X^{(k)} & = & h_{yy}(y_0) Y^{(k)} + \big[ h_y (\sum_{h=0}^{k-1} \e^h Y^{(h)} ) \big]_k + \left[ f_y (\sum_{h=0}^{k-1} \e^h Y^{(h)}, \th+\sum_{h=0}^{k-1} \e^h X^{(h)} )\right]_{k-1} \nonumber \\ & \equiv & h_{yy} (y_0) c_k + \psi^{(k)}\nonumber \\ \label{Xk} \end{eqnarray} where $\psi^{(k)}$ is a smooth function (depending on $H$ and $X^{(h)},Y^{(h)}$ for $h\le k-1$) so that \begin{equation} c_k= h_{yy}(y_0)^{-1} \int_\tN\psi^{(k)} \label{ck} \end{equation} in which case \equ{Xk} has a unique solution with $\int_\tN X^{(k)}=0$. \nin It remains to prove \equ{clm}. Observe that for any (smooth) functions $\bar X, \bar Y$ over $\tN$ one has \begin{equation} \int \Big\{ \big( D \bar Y + H_x(\bar Y, \th+\bar X) \big) \ \big( I + \dpr_\th \bar X \big) \ - \ \big( \o + D \bar X - H_y(\bar Y, \th + \bar X)\big)\ \dpr_\th \bar Y \Big\} d\th = 0 \label{av2} \end{equation} where $\dpr_\th \bar X$ is the matrix $(\dpr_\th \bar X)_{ij}\=\dpr_{\th_j} \bar X_i$ and we adopted the standard convention about row--by--column multiplication of matrices (interpreting vectors as (respectively) $1\times N$ ($N\times 1$) matrices if they are to the left (right) of an $N\times N$--matrix); identity \equ{av2} follows immediately if one notices that $(H_x)(I+\dpr_\th \bar X)+(H_y)(\dpr_\th Y)$ is just the $\th$--gradient of $\th\to H(\bar Y,\th+\bar X)$ (so that its average vanishes) and that the remaining terms in \equ{av2} disappear by integration by parts. Now, we use \equ{av2} with $\bar Y=\sum_{h=0}^{k-1} \e^h Y^{(h)}$, $\bar X= \sum_{h=1}^{k-1} \e^h X^{(h)}$ to get an identity in $\e$; differentiating such identity with respect to $\e$ $k$ times and setting $\e=0$ (\ie evaluating $[\cdot]_k$ of the identity) and using the fact that $Y^{(h)},X^{(h)}$ solve \equ{2.5}, for $h\le k-1$ we easily obtain \equ{clm}. \qed %******************************************** % \section{Tree expansion of formal series} % \label{sec:tree} \setcounter{equation}{0} % A very explicit representation of the formal solutions described in Proposition~\ref{pro:ef} can be obtained by mean of {\em trees}, as explained below. For a different representation see \cite{V}. \giu {\em From now on we restrict our attention on Hamiltonian \equ{h} of the simple form} \begin{equation} H(y,x;\e)= \frac{1}{2}\ y\cdot y\ - \ \e f(x)\ . \label{h1} \end{equation} Such a choice (besides the harmless change of sign in front of the perturbation $f$) has the advantage of simplifying sensibly the notations without introducing any essential modification: in other words, all the arguments and proofs presented below could be extended, at the expense of a heavier formalism, to cover the general case \equ{h} (with $\det h''\neq 0$). Thus, in the present case \equ{h1}, Hamilton equations and the equations characterizing quasi--periodic solutions with frequencies $\o$ are given respectively (see \equ{2.1}, \equ{2.3}) by \beq \ddot x=\e f_x(x)\ ,\qquad D^2 X=\e f_x(\th+X) \label{he} \eeq The formal power series $X\sim \sum_{k\ge 1} X^{(k)} \e^k$, with $\int_\tN X^{(k)}=0$, whose existence and uniqueness (for $f\in\ci(\tN)$) is guaranteed by Proposition \ref{pro:ef}, satisfies \beq D^2 X \sim \e f_x(\th+X)\qquad {\rm or}\qquad D^2 \Xk = \left[ f_x (\th+\sum_{h=1}^{k-1}\Xh \e^h)\right]_{k-1} \label{qpe} \eeq The rest of this section is devoted to a {\em tree representation} of the formal solution $X$ of \equ{qpe} (with $\int_\tN \Xk=0$). We recall that a tree $T$ is a connected acyclic graph (see Appendix~\ref{app:tree} for the fundamentals used here or see any introductory book on graph theory such as \cite{Ha1} or \cite{Bo}). We denote respectively by $V=V(T)$ and $E=E(T)$ the set of {\em vertices} and {\em edges} (or {\em points} and {\em lines}) of the tree $T$. A {\em rooted tree} is a tree with one distinguished vertex called {\em root}; we shall usually denote $r$ such a point and $T_r$ the rooted tree obtained by selecting, as root, the vertex $r$ of the tree $T$. It will also be useful to regard a rooted tree $\Tr$ as a tree with one extra point $\eta\noin V(\Tr)$, called the {\em earth}, and one more edge $\eta r\in E(\Tr)$ connecting the earth $\eta$ to the root $r$. It is natural to define on rooted trees a {\em partial ordering}: Given $T_r$ and $u,v\in V$ we say that \beq u\ge v \qquad {\rm or}\qquad u \ \getr v \label{ord} \eeq if the path with endpoints $r$ and $v$ passes through $u$; $u>v$ means obviously $u\ge v$ and $u\neq v$. In particular $r\ge v$ for any $v\in V$. {\em We fix once and for all a Diophantine vector $\o\in\rN$ satisfying} \equ{dioph} and denote the inner product between $n\in\zN$ and $\o$ by \beq \lg n \rg \= \o\cdot n \=\sum_{i=1}^N \o_i n_i \label{n} \eeq Given a rooted tree $\Tr$ and a function $\a:v\in V\to \a_v\in\zN$ we denote by $\d_v(T_r;\a)$ (or $\d_v(T_r)$ or $\d_v$ when there is no ambiguity) the number \beq \d_v(T_r;\a)\=\lg \sum_{{v'\in V}\atop{v'\le v}} \a_{v'}\rg \label{del} \eeq Given a rooted tree $\Tr$ and an integer--valued function $\a:V\to\zN$ we say that $\a$ is {\em $\Tr$--admissible} if, for all vertices of $\Tr$ one has \beq \a_v\neq 0\ ;\qquad \sum_{v'\le v} \a_{v'} \neq 0 \label{adm} \eeq We shall denote by $\cA(\Tr)$ the set of all $\Tr$--admissible functions over $\Tr$. Finally we let $\cT^k$denote the set of {\em rooted, labeled trees} with $k$ vertices (see Appendix~\ref{app:tree} for more informations); and, for any integer--valued function $\a$ and any subset $S\subset V$ of vertices of a tree $T$, we denote \beq \a(S) \= \sum_{v\in S} \a_v \label{S} \eeq % \begin{proposition} \label{pro:exptree} The $j^{\rm th}$ component of the $n$--Fourier coefficients of $\Xk$ is given by \beqa \Xk_{jn} & = &\frac{1}{k!} \sum_{\Tr\in\cT^k} F_{jn}(\Tr)\ ,\qquad{\rm with:}\nonumber\\ F_{jn}(\Tr)& \equiv & (-i) \sum_{{\a\in\cA(\Tr)}\atop {\a(\Tr)=n,\a_\eta\equiv e_j}} \prod_{v\in V} f_{\a_v}\ \prod_{vv'\in E} \a_v\cdot\a_{v'} \ \prod_{v\in V} \d_v^{-2} \label{exptree} \eeqa where $e_j$ is the unit vector with all zeroes except in the $j^{\rm th}$ entry. \end{proposition} % Note that in the above formula $E=E(T_r)$ {\em includes} the edge $\eta r$ and for this reason the function $\a$ has been extended on $\eta$ (which is outside $V=V(T_r)$). % \rem{rem:exptree} The above function $F^{(k)}_{jn}$ is obviously a function of rooted trees rather than labeled rooted trees and the introduction of the labels has the only task of simplifying the combinatorial factors entering in the formula. In terms of rooted trees, \equ{exptree} can be rewritten as follows. Let $\tT_r\=[\Tr]$ denote the rooted tree obtained by removing the labels from $\Tr$ (clearly $\tT_r$ can be seen as the equivalence class generated by $\Tr$), and let $\ell(\tT_r)$ denote the number of ways of putting $k$ labels on $\tT_r$ (\ie $\ell(\tT_r)=\#\{T'_r\in\cT^k:$ $T'_r\in\tT_r\}$). For example, if $\tT_r$ is the path of order $k$ rooted in one of its endpoints, then $\ell(\tT_k)=k!$. The ``$\tT_r$--admissible" class of $\zN$--valued functions is defined in exactly the same way (see \equ{adm} replacing $T_r$ by $\tT_r$ (as the labeling did not enter in the definition of $\cA(\Tr)$). Finally denote by $\widetilde\cT^k$ the class of all rooted trees of order $k$. With this notations we see that \beqa \Xk_{jn} & = & \sum_{\tT_r\in\widetilde\cT^k} \frac{\ell(\tT_r)}{k!}\ \f_{jn}(\tT_r)\ ,\qquad {\rm with:}\nonumber\\ \f_{jn}(\tT_r)& \equiv & (-i) \sum_{{\a\in\cA(\tT_r)}\atop {\a(\tT_r)=n,\a_\eta\equiv e_j}} \prod_{v\in V} f_{\a_v}\ \prod_{vv'\in E} \a_v\cdot\a_{v'} \ \prod_{v\in V} \d_v^{-2} \label{exptreebis} \eeqa % \erem % Below it will be useful to exchange the sums (over trees and over integers $\a$'s) in \equ{exptree}, in which case we obtain the following formula: \beq \Xk_{jn}=\frac{1}{k!} \sum_{ {n_1,...,n_k}\atop{ n_i\in\zN\backslash \{0\}, \sum n_i=n} } \sum_{\Tr\in\cT^k}\bar F_{j\{n_i\}} (\Tr) \label{exptr1} \eeq where if $v_1,...,v_k$ are the labels of $\cT^k$ and if we set $\a_{v_i}\=n_i$ for any choice of $\{n_i\}$, the function $\bar F$ is defined by \beq \bar F_{j\{n_i\}} (\Tr)\= 0\ ,\qquad {\rm if}\quad \exists\ v:\ \sum_{v'\le v} \a_{v'} =0 \label{exptr2} \eeq and otherwise by \beq \bar F_{j\{n_i\}} (\Tr)\= \prod_{v\in V} f_{\a_v}\ \prod_{vv'\in E} \a_v\cdot\a_{v'} \ \prod_{v\in V} \d_v^{-2} \label{exptr3} \eeq \giu Expansions like \equ{exptree} are quite general; another tree expansion, which we shall need later, for a series satisfying an equation simpler than (but related to) \equ{qpe} is described in the following % \begin{proposition} \label{pro:exptree2} Let $n\in\zN\to \phi_n\in\complex$ decay faster than any power of $|n|$ (~\ie $\forall$ $s>0$, $\sup_n |n|^s |\phi_n|<\infty$) and let $g=g(\e)\sim\sum_{k\ge 1}g_k\e^k$ be the unique formal solution of the equation: \beq g\sim \e \sum_{n\in\zN} \phi_n |n| \exp(|n| g)\ . \label{gphi} \eeq Then \beqa g_k & = &\frac{1}{k!} \sum_{T_r\in\cT^k} \sum_{{\a:T_r\to\zN}\atop {\a_\eta\equiv 1}} \prod_{v\in V} \phi_{\a_v}\ \prod_{vv'\in E} |\a_v|\ |\a_{v'}| \nonumber\\ & = & \frac{1}{k!} \sum_{T_r\in\cT^k} \sum_{\a:T_r\to\zN} \prod_{v\in V} \phi_{\a_v}\ \prod_{v\in V} |\a_v|^{\deg v} \label{exptree2} \eeqa where $\deg v$ denotes the degree of $v$ (\ie the number of edges incident with $v$). \end{proposition} % Actually, if the $\phi_n$'s decay exponentially (\ie $|\phi_n|\le M \exp(-\xi |n|)$ for some $M,\xi>0$ and all $n\in\zN$), it follows immediately from the (analytic) Implicit Function Theorem (applied to the analytic function of two complex variables $(x,\e)\in\complex^2\to G(x,\e)$ $\=$ $x-\e\sum \phi_n |n| \exp(|n|x)$) that $g$ converges absolutely near $\e=0$ (see also Subsection~\ref{subsec:1} below). \giu The proofs (by induction on $k$) of Proposition~\ref{pro:exptree} and Proposition~\ref{pro:exptree2} are of pure combinatorial character. Here we outline the main steps and refer to Appendix~\ref{app:combina} for complete details. The proofs of the two Propositions are very similar but, obviously, the proof of Proposition~\ref{pro:exptree2} is simpler and we discuss it first. The starting point is so to get a recursive formula for the coefficients $g_k$ defined implicitly by \equ{gphi}: Expanding the right hand side of \equ{gphi} in Taylor series and comparing equal powers of $\e$ one immediately obtain \beqa g_1 & = & \sum_{n\in\zN} \phi_n |n| \nonumber\\ g_k & = & \sum_{j=2}^k \sum_{n\in\zN} \phi_n\ \frac{|n|^j}{(j-1)!} \sum_{ {h_1+\cdots h_{j-1} =k-1} \atop{h_i\ge 1}} \prod_{i=1}^{j-1} g_{h_i}\ \quad (k\ge 2) \label{3.1} \eeqa Now, there is a natural way of linking (labeled rooted) trees of order $k$ with all possible trees of order $h_1,...,h_{j-1}$ with $h_1+\cdots+ h_{j-1}=k-1$: This link is based on the following construction. Let $\tcT^k$ denote the rooted trees of order $k$ (non labeled) and let $\tcT^k_0$ denote the trees of order $k$ (no labels, no root). \dfn{def:3.1} Let $s\ge 1$, let $\tT_i\in\tcT^{h_i}$ where $i=1,...,s$ and $h_i\ge 1$ with $h\=\sum_{i=1}^s h_i$. Let $\tT_i^0$ $\in$ $\tcT^{h_i}_0$ be the (unrooted) tree obtained from $\tT_i$ by not distinguishing the root. We define the rooted tree $\tT_1 * \cdots * \tT_s$ $\in \tcT^{h+1}$ by setting\footnote{ For the (standard) notation see Appendix~\ref{app:tree}} \beq \tT_1 * \cdots *\tT_s\= \Big( \tT_1^0\cup \cdots \cup \tT_s^0 \cup \{r\} + \sum_{i=1}^s rr_i \Big)_r \label{3.2} \eeq where $r$ is the root of $\tT_1*\cdots *\tT_s$ ($r$ is an extra vertex \ie $r\noin \tT_i$) and the $r_i$'s are the roots of $\tT_i$ \ie $(\tT_i^0)_{r_i}=\tT_i$. \edfn % \begin{figure}[hbt] \begin{center} \begin{picture}(385,120) \thicklines \put(10,20){\begin{picture}(70,100) \put(30,25){\circle*{3}} \put(30,25){\circle{8}} \put(30,25){\line(-1,1){18}} \put(12,43){\circle*{3}} \put(30,25){\line(1,1){36}} \multiput(48,43)(18,18){2}{\circle*{3}} \put(35,20){$r_1$} \put(25,0){$\tT_1$} \end{picture}} \put(80,20){\begin{picture}(70,100) \put(30,25){\circle*{3}} \put(30,25){\circle{8}} \put(30,25){\line(-1,1){18}} \put(12,43){\circle*{3}} \put(30,25){\line(1,1){18}} \put(48,43){\circle*{3}} \put(35,20){$r_2$} \put(25,0){$\tT_2$} \end{picture}} \put(150,20){\begin{picture}(70,100) \put(30,25){\circle*{3}} \put(30,25){\circle{8}} \put(30,25){\line(0,1){15}} \put(30,40){\line(-1,2){20}} \multiput(30,40)(-10,20){3}{\circle*{3}} \put(30,40){\line(1,2){20}} \multiput(40,60)(10,20){2}{\circle*{3}} \put(35,20){$r_3$} \put(25,0){$\tT_3$} \end{picture}} \put(240,20){\begin{picture}(145,100) \put(60,25){\circle*{3}} \put(60,25){\circle{8}} \put(60,25){\line(-4,3){20}} \put(40,40){\line(0,1){36}} \multiput(40,40)(0,18){3}{\circle*{3}} \put(40,40){\line(-1,0){18}} \put(22,40){\circle*{3}} \put(60,25){\line(0,1){23}} \put(60,48){\circle*{3}} \put(60,48){\line(1,2){10}} \put(70,68){\circle*{3}} \put(60,48){\line(-1,2){10}} \put(50,68){\circle*{3}} \put(60,25){\line(2,1){40}} \multiput(80,35)(20,10){2}{\circle*{3}} \put(100,45){\line(1,0){36}} \multiput(118,45)(18,0){2}{\circle*{3}} \put(100,45){\line(0,1){36}} \multiput(100,63)(0,18){2}{\circle*{3}} \thinlines \put(-15,50){\vector(1,0){20}} \put(55,14){$r$} \put(62,40){$r_2$} \put(82,25){$r_3$} \put(35,30){$r_1$} \put(35,0){$\tT_1 * \tT_2 * \tT_3$} \end{picture}} \parbox{12cm}{\caption{\label{fig:3.1} The $*$ operation}} \end{picture} \end{center} \end{figure} % \giu And here it is the combinatorics (recall that the number of labeled rooted trees of order $h$ is $h^{h-1}$ (see Appendix~\ref{app:tree}): \lem{lem:3.1} For $k\ge j\ge 2$, denote by $t_{kj}$ the number of labeled rooted trees of order $k$ with root of degree $j$. Then \beqano & {\rm (i)} &\quad t_{kj}= k\ {k-2\choose j-2} \ (k-1)^{k-j} \\ & {\rm (ii)} &\quad t_{kj}= \frac{k!}{(j-1)!} \sum_{ {h_1+\cdots h_{j-1} =k-1} \atop{h_i\ge 1}} \prod_{i=1}^{j-1} \frac{h_i^{h_i-1}}{h_i!}\\ & {\rm (iii)} &\quad \frac{k^{k-1}}{k!} = \sum_{j=2}^k \frac{1}{(j-1)!} \sum_{ {h_1+\cdots h_{j-1} =k-1} \atop{h_i\ge 1}} \prod_{i=1}^{j-1} \frac{h_i^{h_i-1}}{h_i!} \eeqano \elem The proof is given in Appendix~\ref{app:combina}. The main point of the above Lemma is (ii): (i) is just a curiosity and will not be used and (iii) is simply obtained by summing (ii) over $j$. An immediate corollary of this Lemma (better: ``of the proof of this Lemma") is the following Corollary. % \cor{cor:3.1} Let $G:\tcT^k\to\complex$. Then \beq \frac{1}{k!} \sum_{T\in\cT^k} G(\tT)= \sum_{j=2}^k \frac{1}{(j-1)!} \sum_{ {h_1+\cdots h_{j-1} =k-1} \atop{h_i\ge 1}} \prod_{i=1}^{j-1} \frac{1}{h_i!}\ \ \sum_{ T_1\in \cT^{h_1},...,T_{j-1}\in \cT^{h_{j-1}} }\ G(\tT_1 *\cdots *\tT_{j-1}) \label{3.3} \eeq \ecor >From \equ{3.3}, Proposition~\ref{pro:exptree2} will follow at once. To get also Proposition~\ref{pro:exptree} from Corollary~\ref{cor:3.1}, one needs only to rewrite \equ{exptree} properly. To do this, let for any $m,n\in\zNn$ and $k\ge 1$ \beq \xi^m\=i \ m\cdot X\ ,\quad \xi^{(m,k)}\=i \ m\cdot \Xk\ ,\quad \xi_n^{(m,k)}\=i \ m\cdot \Xk_n \label{3.4} \eeq Then from \equ{qpe} and Taylor expansion (in $\e$) we obtain easily \beqa \xi_n^{(m,1)} & = & - \gl{n}^{-2} \ m\cdot n\ f_n \nonumber\\ \xi_n^{(m,k)} & = & - \gl{n}^{-2} \sum_{j=2}^k \sum_{n\in\zNn} \sum_{ {n_1+\cdots +n_j=n}\atop{n_i\in\zNn}} \ m\cdot n_j \ f_{n_j}\times \nonumber\\ &&\qquad \times \frac{1}{(j-1)!} \sum_{ {h_1+\cdots h_{j-1} =k-1} \atop{h_i\ge 1}} \prod_{i=1}^{j-1} \xi_{n_i}^{(n_j,h_i)} \quad (k\ge 2) \label{3.5} \eeqa The similarity of \equ{3.5} and \equ{3.1} is transparent\footnote{$|n|\phi_n$ corresponds to $\gl{n}^{-2} m\cdot n_j f_{n_j}$, while $|n|g_{h_i}$ corresponds to $\xi_{n_i}^{(n_j,h_i)}$.} and it is enough to change the extensions of the $\a$ functions on the earth $\eta$ to generalizes \equ{exptree} to \beqa \xi^{(m,k)}_{n} & = &\frac{1}{k!} \sum_{\Tr\in\cT^k} F_{mn}\ ,\qquad{\rm with:}\nonumber\\ F_{mn}(\Tr)& \equiv & \sum_{{\a\in\cA(\Tr)}\atop {\a(\Tr)=n,\a_\eta\equiv m}} \prod_{v\in V} f_{\a_v}\ \prod_{vv'\in E} \a_v\cdot\a_{v'} \ \prod_{v\in V} \d_v^{-2} \label{exptreeg} \eeqa For full details see Appendix~\ref{app:combina}. %******************************************** % \section{Divergences} % \label{sec:diverge} \setcounter{equation}{0} % {\em From now on $f$ in \equ{qpe} will be assumed real--analytic}. In this section we illustrate the well known mechanism of divergences of series with coefficients of type \equ{exptree}, (for an $f$ in \equ{qpe} real analytic), {\em if signs are disregarded} \ie when absolute values are introduced within the sums. This fact, based on the repetition of particularly small divisors $\d_v$ (``resonances"), shows the need for detecting the compensations among all the terms whose size in absolute value grows faster than exponentially in $k$ and whose actual occurrence will readily be seen. \giu More explicitely, let, for $1\le \baj\le N$: \beq A^{(k)}_{\baj n} = \frac{1}{k!} \sum_{\Tr\in\cT^k} \sum_{{\a\in\cA(\Tr)}\atop {\a(\Tr)=n,\a_\eta\equiv e_\baj}} \mid\prod_{v\in V} f_{\a_v}\ \prod_{vv'\in E} \a_v\cdot\a_{v'}\mid \ \prod_{v\in V} \d_v^{-2} \label{abstree} \eeq we want to show that \beq \sup_{\baj,k,n} A^{(k)}_{\baj n}\ e^{-\s|n|}\ \e_0^k=+\infty \ ,\qquad \forall\ \ \e_0,\ \s>0 \label{diverge} \eeq where in the supremum $1\le \baj\le N$, $k\ge 1$ and $n\in\zN$. We shall prove \equ{diverge} in the case \beqa & & N =2\ ,\qquad f(x_1,x_2)\=\cos x_1 \ + \cos(x_2-x_1)\ , \nonumber\\ & & \o_2 >\o_1>0\ ,\qquad {\o_1}/{\o_2}\ \quad{\rm irrational} \label{model} \eeqa since it will be clear that the same proof can easily be adapted to the general analytic case {\em as long as $f$ has two Fourier coefficients with linearly independent (in $\zN$) indeces.} \giu {\bf Proof of \equ{diverge} in case \equ{model}} % >From an elementary number theoretical theorem by Dirichlet (see \cite{Sc}) it follows that there exist $p_j,q_j\ge 1$, with $q_j\nearrow \infty$ such that \beq \mid \frac{\o_1}{\o_2} \ q_j \ - \ p_j \mid < \frac{1}{q_j} \label{Dirichlet} \eeq and we can assume without loss of generality that for all $j\ge 1$ \beq q_j > \frac{\o_2}{\o_2-\o_1} \label{condqj} \eeq Observing the the left hand side of \equ{Dirichlet} is bigger or equal than $p_j-q_j(\o_1/\o_2)$ we see immediately that \equ{condqj} implies that \beq p_ju_{k-1}>...>u_1$. Let $s_j\=q_j-p_j$ and, for any $h\ge 0$ let $k\=q_j+2h$: \beq q_j=p_j+s_j\ ,\qquad 1\le s_j,p_j< q_j\ ,\qquad k=q_j+2h \label{defsh} \eeq Let $\a_{u_i}\=n_i\in\integer^2$ be defined as follows: \beq n_i\= \cases{ (1,0) & if \ $1\le i\le s_j$\cr (1,-1) & if \ $s_jFrom \equ{Dirichlet}, \equ{condqj} and \equ{di} it follows that \beqa & & |\o_1 q_j- \o_2 p_j + \o_1|<\o_2\nonumber \\ & & |\d_{u_i}| < 2 \o_2\ i\ ,\qquad {\rm for} \quad 1\le i\le q_j \label{4.1} \eeqa which, together with \equ{Dirichlet} yields \beq \prod_{i=1}^k \d_{u_i}^{-1} > \o_2^{-k}\ 2^{-q_j}\ \frac{q_j^h}{q_j!} \label{4.2} \eeq Furthermore, since for \equ{model} it is \beq |f_{\pm(1,0)}|=|f_{\pm(1,-1)}| =\frac{1}{2} \label{4.3} \eeq we have also \beq \prod_{i=1}^k |f_{n_i}| = 2^{-k}\ ,\qquad \prod_{E(P)} |\a_v\cdot\a_{v'}|\ge 1 \label{4.4} \eeq having set $\baj=1$ \ie $\a_\eta\=(1,0)$. We now choose \beq h=3q_j\ ,\qquad {\rm so\quad that}\qquad k\=k_j=7q_j \label{4.5} \eeq Then, if in the sum \equ{abstree} (which is finite in the case of \equ{model}), we keep only the terms corresponding to the unlabeled rooted tree $P$ with the above choice of Fourier indices, recalling from Remark~\ref{rem:exptree} that the number of labeled rooted trees corresponding to $P$ are exactly $k!$, we obtain (recall that $\baj=1$ and that $n=(q_j,-p_j)$) \beq A^{(k_j)}_{1 n}\ > 2^{-k_j} \ (2\o_2)^{-2k_j}\ (k_j!)^{4/7} \label{4.6} \eeq from which \equ{diverge} follows at once. Notice that by increasing $h$, the exponent of $k_j!$ in \equ{4.6} becomes arbitrarly close to 1. \qed %******************************************** \section{Resonances} \label{sec:resona} \setcounter{equation}{0} % Repetitions of small divisors correspond to {\em resonant subtrees} \ie to subtrees $S$ of a given (labeled rooted) tree $T$ for which $\a(S)\=\sum_{v\in S}\a_v=0$: In the example of Section~\ref{sec:diverge} resonant subtrees are the subtrees with endpoints $u_{q_j+s}$ and $u_{q_j+s+1}$ for $1\le s\le k-1$, or the subtrees with endpoints $u_{q_j+s}$ and $u_{q_j+s+3}$ \etc. Throughout the rest of the paper (unless otherwise stated), given $T\in\cT^k$ (we shall often omit the explicit indication of the root $r$ appearing elsewhere as index of $T$ when this does not lead to confusion), the function $\a:V\=\{v_1,...,v_k\}\to\zN$, $\{\a_{v_i}\}$ $\=$ $\{n_i\}$, is {\em admissible} in the usual sense that $n_i\neq 0$ and $\sum_{v'\le v} \a_v\neq 0$ for all $v\in V$ (in which case $\d_v\neq 0$ by the rational independence of $\o$). In this section we develop the tools needed to {\em make a partition} of $\cT^k$, for a given choice of $\{n_i\}$ (see \equ{exptr1}$\div$ \equ{exptr3}), into {\em ``complete families"}. In the next section we shall see that the main property of complete families is that (possibly) huge terms (coming from repetition of small divisors) {\em compensate among themselves} within the same class of the partition. This fact will allow us to bound the contribution to \equ{exptr1}, coming from the sum over trees belonging to the same class of the partition, by a constant to the $k^{\rm th}$ power. \giu The rest of this section consists basically in a sequel of definitions and checks of elementary properties of trees with a given admissible $\a_{v_i}\=n_i$, $v_1,...,v_k$ being the labels of $\cT^k$ (see Appendix~\ref{app:tree} for the fundamentals from graph theory used here). Let us begin (with a bit of patience). \dfntwo{Degree of a subtree}{dfn:1} Given a (possibly rooted) tree $T$ and $S\subset T$ (\ie $S$ is a subtree of $T$ \ie $S$ is a connected subgraph of $T$) we call {\em degree of $S$}, $\deg S$, the number of edges connecting $S$ with $T\backslash S$ (if $T$ is rooted and $r\in S$ the edge $\eta r$ has to be counted). \edfn \dfntwo{Resonances}{dfn:2} Given a rooted tree $T$ and an admissible function $\a$ on it, we say that the subtree $R\subset T$ is {\em resonant} if: (i) $\deg R=2$; (ii) $R$ is null \ie $\a(R)=0$; (iii) $R$ cannot be disconnected by removal of one edge into two null subtrees. A resonant subtree will also be called a {\em resonance}. \edfn \dfntwo{Resonant couples}{dfn:4} Given a rooted tree $T$, $\l>0$ we say that a couple of points $(u,w)\in V\times V$ is {\em $\l$--resonant} if: (i) $u>w$; (ii) $\d_u=\d_w$; (iii) $u$ and $w$ are not adjacent and $|\d_w|\le \l |\d_v|$ for all $v$ between $u$ and $w$. \edfn % Such a notion is given also in \cite{E} where the couple $(u,w)$ is called a critical resonance; we reserve such a name for a different object (see Definition \ref{dfn:6} below). \dfntwo{$\l$--Resonances}{dfn:5} Given a rooted tree $T$, $\l>0$ and a resonance $R\subset T$ let \beq n\= \sum_{vw$ such that $\d_u=\d_w$ but $\d_v\neq \d_u$ for any $v$ between $u$ and $w$ (if there are any) it is always possible to associate a resonant subtree $R\=R(u,w)$: \beq R(u,w)\= \{ v\in V:\ v\ge u\}\backslash \{v\in V:\ v\ge w\} \label{5.3} \eeq Notice however that $(u,w)$ may be $\l$--resonant while the associated resonance $R(u,w)$ is not $\l$--resonant as shown in the example in Figure~\ref{fig:1}: it is easy to see that one can choose $n_1,n_2,n_3$ so that $n_1+n_2+n_3=0$ (implying $\d_u=\d_w$) and $|\lg n\rg|\le \l|\lg n_2+n_3+n\rg |$ (implying $(u,w)$ $\l$--resonant) but $|\lg n\rg |>\l |\lg n_3\rg|$ (implying $R(u,w)$ not $\l$--resonant). \erem % \begin{figure}[hbt] \begin{center} \begin{picture}(350,120) \put(100,20){\begin{picture}(150,100) \thicklines \put(11,40){\circle*{3}} \put(11,40){\circle{8}} \multiput(11,40)(3,0){9}{\circle*{1}} \put(35,40){\line(1,0){85}} \multiput(50,40)(35,0){3}{\circle*{3}} \put(85,40){\line(1,3){10}} \put(95,70){\circle*{3}} \thinlines \put(72.5,55){\oval(70,62)} \put(44,44){$n_1$} \put(73,44){$n_2$} \put(90,75){$n_3$} \put(116,44){$n$} \put(46,32){$u$} \put(116,32){$w$} \put(70,13){$R$} \end{picture}} \parbox{12cm}{\caption{\label{fig:1} $R=R(u,w)$}} \end{picture} \end{center} \end{figure} % \protwo{Non overlapping of resonant couples}{pro:nonover1} Let $(u_i,w_i)$ \allowbreak for $i=1,2$ be couples of $\l_i$--resonant points. If $u_1>u_2>w_1>w_2$ then either $\l_1\ge 1$ or $\l_2\ge 1$. \epro \PROOF By contradiction: Assume both $\l_1$ and $\l_2$ are less than one.Then: \beq |\d_{u_2}|=|\d_{w_2}| \le \l_2 |\d_{w_1}|<|\d_{w_1}|\ ,\qquad |\d_{w_1}| \le \l_1 |\d_{u_2}| <|\d_{u_2}| \eeq which is absurd. \qed \protwo{Non overlapping of resonances}{pro:nonover2} Let $R_i$ for $i=1,2$ be $\l_i$--re\-so\-na\-n\-ces which are not one subtree of the other and with nonempty intersection (\ie with at least one common vertex). Then either $\l_1\ge 1/2$ or $\l_2\ge 1/2$. \epro \PROOF Two subtrees of degree 2 which are not one a subtree of the other can intersect in three ways, see Figure~\ref{fig:2}. % \begin{figure}[hbt] \begin{center} \begin{picture}(385,145) \put(0,45){\begin{picture}(180,100) \thicklines \put(10,50){\line(1,0){120}} \multiput(10,50)(30,0){5}{\circle*{6}} \put(130,50){\line(3,1){15}} \multiput(145,55)(3,1){6}{\circle*{1}} \put(160,60){$\odot$} \thinlines \put(55,50){\oval(50,40)} \put(85,50){\oval(50,40)} \put(5,57){$n_1$} \put(35,57){$n_2$} \put(65,57){$n_3$} \put(95,57){$n_4$} \put(55,15){$R_1$} \put(85,15){$R_2$} \put(80,-15){(a)} \end{picture}} \put(190,45){\begin{picture}(95,100) \thicklines \put(20,65){\line(2,1){40}} \put(20,65){\circle*{6}} \put(15,72){$n_1$} \put(40,75){\circle*{6}} \put(35,82){$n_2$} \put(60,85){\circle*{6}} \put(55,92){$n_3$} \put(40,75){\line(2,-1){20}} \put(60,65){\circle*{6}} \put(55,72){$n_4$} \put(60,65){\line(1,-1){15}} \put(75,50){\circle*{6}} \put(75,50){\line(0,-1){10}} \multiput(75,40)(0,-3){5}{\circle*{1}} \put(71,20){$\odot$} \thinlines \put(30,72.5){\oval(45,45)} \put(53,78){\oval(50,50)} \put(25,35){$R_1$} \put(50,40){$R_2$} \put(40,-15){(b)} \end{picture}} \put(295,45){\begin{picture}(90,100) \thicklines \put(15,75){\line(1,0){60}} \put(15,75){\circle*{6}} \put(10,82){$n_1$} \put(45,75){\circle*{6}} \put(40,82){$n_2$} \put(75,75){\circle*{6}} \put(70,82){$n_3$} \put(45,75){\line(0,-1){40}} \put(45,45){\circle*{6}} \multiput(45,35)(0,-3){6}{\circle*{1}} \put(41,12){$\odot$} \thinlines \put(30,75){\oval(52,40)} \put(60,75){\oval(52,40)} \put(20,40){$R_1$} \put(55,40){$R_2$} \put(40,-15){(c)} \end{picture}} \parbox{12cm}{\caption{\label{fig:2} Intersections of resonances}} \end{picture} \end{center} \end{figure} % \giu In drawing pictures we are using the following \giu {\bf Notational Remark} {\em Thick points correspond in general to subtrees (a subtree of degree $d$ collapse to a thick point of the same degree) and the $n$ written above a thick point correspond to the sum of $\a_v$ when $v$ varies in the corresponding (collapsed) subtree. In the figures, null subtrees (indicated usually by $R$) are encircled (while the root, as custumary, is distinguished by a small circle around it.)} \giu Since $R_i$ are resonant subtrees $n_2+n_3=0$, $n_3+n_4=0$ in case (a) while $n_1=-n_2$ in case (b) and (c). We proceed now by contradiction: Assume $\l_i<1/2$. \nin In case (a) \beq |\lg n_1\rg|\le \l_1 |\lg n_2 \rg|\ ,\qquad |\lg n_1+n_2\rg| \le\l_2|\lg n_3\rg|=\l_2|\lg n_2\rg| \eeq and \beq |\lg n_1+n_2\rg| \ge (1-\l_1)|\lg n_2\rg|\quad \implies \quad |\lg n_2 \rg|\le \frac{\l_2}{1-\l_1} \ |\lg n_2\rg| <|\lg n_2 \rg| \eeq which is absurd. \giu In case (b) and (c): $|\lg n_1\rg|\le \l_1 |\lg n_2\rg|=\l_1|\lg n_1 \rg|$ which is absurd because $\l_1<1$. \qed \dfntwo{Critical Resonances}{dfn:6} A resonance $R$ is called {\em critical} (or $\l$--critical) if it is $\l$--resonant with $\l<1/2$ and if it is maximal \ie it is not properly contained into another $\l$--resonant subtree. \nin Obviously: (i) Critical resonances cannot intersect (by definition and by Proposition \ref{pro:nonover2}). (ii) Critical resonances may contain $\l'$--resonances with any $\l'$. \edfn % Critical resonances may appear in sequels where each resonance is a subtree of another resonance (creating ``hierarchies of subresonances"). In order to classify them, we introduce the following concept. \dfntwo{$\l$--Subresonances}{dfn:7} Let $R'$ be a null subtree of degree two of a $\l$--critical resonance $R$ and let $u_1v_1$, $u_2v_2$, with $v_i\in R'$, be the two edges connecting $R'$ with $R\bks R'$ (obviously $u_i\in R$ and $u_1\neq u_2$). Define \beq m\=\sum_{ {v\in R_{v_1}} \atop {v\le u_1} } \a_v= - \sum_{ {v\in R_{v_2}} \atop {v\le u_2} } \a_v \label{defm} \eeq where $R_{v_i}$ is the rooted tree $R$ with root in $v_i$ and the order $\le$ in each sum is relative to $R_{v_i}$ (we are using the standard convention that $v\in T$ means $v\in V(T)$). The integer $m\in \zN\bks \{0\}$ is defined up to sign (as the roles of the $v_i$ can be exchanged). We then say that $R'$ is a {\em $\l$--subresonance} if \beq \agl{m} \le \l \ \D(R') \label{subres} \eeq \edfn % As in case of resonances we have a simple non overlapping criterion for subresonances: \protwo{Non overlapping of subresonances}{pro:nonover3} Let $R_i'$ for $i=1,2$ be $\l$--\-sub\-re\-so\-nan\-ces of a $\l$--critical resonance $R$. Assume that $R_i$ are not one subtree of the other. Then $V(R_1)\cap V(R_2)=\emptyset$. \epro % The proof is very similar to the proof of Proposition \ref{pro:nonover2} and is left to the reader. \dfntwo{Hierarchies of critical subresonances}{dfn:8} Let $R$ be a $\l$--cri\-ti\-cal resonance (\ie $R$ is a maximal $\l$--resonance with $\l<1/2$). Let $R_1$ be a maximal $\l$--subresonance (if it exists) of $R$ (maximal means that $R_1$ is not properly contained in another $\l$--subresonance of $R$). Note that, by Proposition \ref{pro:nonover3}, $R_1$ cannot overlap with another $\l$--subresonance of $R$. We can now {\em define subresonances of $R_1$ replacing, in Definition~\ref{dfn:7}, $R$ with $R_1$}. We then let $R_2$ be a maximal $\l$--subresonance (if it exists) of $R_1$. And so on, till $R_h$, $h\ge 1$, does not contain any $\l$--subresonance. We consider also the case in which $R$ does not contain any $\l$--subresonance setting in such a case $h=0$ and $R_0\=R$. The sequence \beq \cH\=\cH_\l(R)\=\{R_1,...,R_h\}\ ,\qquad R\supset R_1\supset \cdots\supset R_h \label{hierarchy} \eeq is called a {\em critical hierarchy of $\l$--subresonances} (or simply a hierarchy of subresonances) and the elements of the hierarchy, $R_i$ with $1\le i\le h$, are called {\em critical subresonances}. Thus by definition a critical subresonance of the rooted tree $T$ is an element of a hierarchy associated to some critical resonance. Obviously a critical resonance may contain more than one hierarchy of subresonances (see Figure~\ref{fig:5} below). \edfn % \rem{rem:5.2} In general a critical $\l$--subresonance need not be a $\l$--resonance as shown by the following example. % \begin{figure}[hbt] \begin{center} \begin{picture}(395,100) \thicklines \put(110,60){\circle{8}} \put(110,60){\line(1,0){175}} \multiput(110,60)(35,0){6}{\circle*{3}} \thinlines \put(197.5,60){\oval(70,35)} \put(197.5,60){\oval(128,60)} \put(136,67){$-3n$} \put(170,67){$-9n$} \put(210,67){$9n$} \put(235,67){$3n$} \put(280,67){$n$} \put(230,35){$R_1$} \put(260,25){$R$} \parbox{14cm}{\caption{\label{fig:3} $R_1$ is not a $\frac{1}{3}$--resonance}} \end{picture} \end{center} \end{figure} \giu % Fix $\l=1/3$ and let $n\in\zN\bks\{0\}$. Then one checks immediatly that $R$, in Figure~\ref{fig:3}, is a $\l$--critical resonance and that $\cH_\l(R)=\{R_1\}$. However the critical subresonance {\em $R_1$ is not a $\l$--resonance} since $\agl{4n}>\l\D(R_1)=3\agl{n}$. \erem \rem{rem:5.3} If $R'$ is a $\l$--critical subresonance then $R'=R_i$ for some $1\le i\le h$ where $\{R_1,...,R_h\}$ is a $\l$--hierarchy associated to some $\l$--critical resonance $R$. To each $R_j$ we can associated (up to sign) an integer $m_j\in\zNn$ as in \equ{defm} (see Figure~\ref{fig:4}). % \begin{figure}[hbt] \begin{center} \begin{picture}(385,140) \thicklines \multiput(45,80)(-3,0){9}{\circle*{1}} \put(21,80){\circle*{3}} \put(21,80){\circle{8}} \put(45,80){\line(1,0){270}} \multiput(70,80)(35,0){8}{\circle*{6}} \thinlines \put(175,80){\oval(20,20)} \put(175,80){\oval(100,40)} \put(175,80){\oval(163,60)} \put(175,80){\oval(245,100)} \put(60,87){$-m_1$} \put(95,87){$-m_2$} \put(130,87){$-m_3$} \put(205,87){$m_3$} \put(235,87){$m_2$} \put(275,87){$m_1$} \put(310,87){$m_0$} \put(183,65){$R_3$} \put(220,55){$R_2$} \put(255,45){$R_1$} \put(295,25){$R$} \parbox{12cm}{\caption{\label{fig:4} A hierarchy of resonances with $h=3$}} \end{picture} \end{center} \end{figure} % \nin Then $\agl{m_j}\le \l \D(R_j)$ for any $j=1,...,h$ and in particular $\agl{m_j}\le \l \agl{m_{j+1}}$ so that \beq \agl{m_i}\le \l^{j-i}\agl{m_j}\ ,\qquad \forall\ \ 1\le i\le j \le h \label{5.4} \eeq and \beq \sum_{i=1}^j \agl{m_i} < \frac{1}{1-\l}\ \agl{m_j} \le \frac{\l}{1-\l} \D(R_j) \label{5.5} \eeq Also, if $\s_h$ is plus or minus one and $\s_i=0$ or $\pm 1$ ($i=1,...h-1$) then \beq \agl{\sum_{i=1}^h \s_i m_i} \ge \frac{1-2\l}{1-\l} \agl{m_h} \label{5.6} \eeq Furthermore, if we let $m_0\=\sum_{v'< R} \a_{v'}$ (since $\agl{m_0}\le \l\agl{m_1}$ we get as above: \beq \sum_{i=0}^h \agl{m_i}< \frac{1}{1-\l} \agl{m_h} \le \frac{\l}{1-\l} \D(R_h) \label{5.7} \eeq \erem \dfntwo{The set of critical (sub)resonance of $T$}{dfn:9} We shall often use the \allowbreak word\- ``(sub)resonance" to indicate either a resonance or a subresonance. Given a rooted tree $T$, an admissible $\a$ and a $\l<1/2$ we let \beq \cR\=\cR(T)\=\{\rm \ all\ \l\!-\!critical\ resonances\ and\ \l\!-\!critical\ subresonances\ of \ T\} \label{defR} \eeq and \beq \cR_0\=\cR_0(T)\=\{ { \rm \ all\ \l\!-\!critical\ resonances\ of \ T} \} \label{defR0} \eeq If $2\cdot\cdot\cdot > \bR_s$ (as in the chain $C_1$ of the above example) {\em or} $\bR_1<\cdot\cdot\cdot < \bR_s$ (as in $C_2$). \erem % Finally, we are ready to define {\em complete families of trees generated by a tree $T$ and a given $T$--admissible function $\a$.} \dfntwo{Complete families}{dfn:13} Given $T\in\cT^k$, $\a$ $T$--admissible and $\l<1/2$ we define \beq \cF(T)\=\{ T\} \label{defcF1} \eeq if $\cR(T)=\emptyset$ (\ie if there are no $\l$--critical resonances in $T$), otherwise, letting $\cR\=\cR(T)$, $\cC\=\cC(T)$, we define \beqa \cF(T)& \= \{ T'\in\cT^k: & \ T'= \bT\cup \bigcup_{\bR\in\bcR} \bR + \nonumber\\ & & \sum_{C=(\bR_1,...,\bR_s)\in\cC} u(C)\bu_1+ w(C)\bw_s + \sum_{i=1}^{s-1} \bw_i\bu_{i+1} \ ,\nonumber\\ & &{\rm where}\ \bu_i,\bw_i\ {\rm vary\ in\ } \bR_i \} \label{defcF} \eeqa if $s=1$ for some $C$ the sum over $i$ has to be omitted. $\cF$ {\em is called the complete family associated to $(T,\a)$}. \edfn % Obviously, the integers $n_i\=\a_{v_i}$, which together with the ``topological" structure of the tree $T$ generates $\cF$, may be thought of as fixed ``attributes" of the labels $v_i$ and, {\em by construction, $\{\a_{v_i}\}$ is admissible for any tree $T'\in\cF(T)$}. The following Proposition collects a few elementary properties of complete families. \begin{figure}[hbt] \begin{center} \begin{picture}(385,340) \multiput(5,20)(0,80){4}{\begin{picture}(380,80) \multiput(0,0)(95,0){4}{\begin{picture}(90,80) \thicklines \put(20,50){\circle*{3}} \put(40,40){\line(0,1){20}} \multiput(40,40)(0,20){2}{\circle*{3}} \put(60,50){\circle*{3}} \thinlines \put(40,45){\oval(26,40)} \put(40,42.5){\oval(60,60)} \put(46,18){$R_2$} \put(65,10){$R_1$} \put(15,40){$v_1$} \put(35,30){$v_2$} \put(42,48){$v_3$} \put(55,40){$v_4$} \end{picture}} \put(0,0){\begin{picture}(380,80) \thicklines \put(20,50){\line(2,-1){20}} \put(40,60){\line(2,-1){20}} \put(115,50){\line(2,1){20}} \put(135,60){\line(2,-1){20}} \put(210,50){\line(2,-1){20}} \put(230,40){\line(2,1){20}} \put(305,50){\line(2,1){20}} \put(325,40){\line(2,1){20}} \end{picture}} \end{picture}} \multiput(5,260)(95,0){4}{\begin{picture}(90,80) \put(20,50){\circle{6}} \put(60,50){\line(1,0){20}} \put(80,50){\circle*{3}} \put(75,40){$v_5$} \end{picture}} \multiput(5,180)(95,0){4}{\begin{picture}(90,80) \put(20,50){\circle{6}} \put(0,50){\line(1,0){20}} \put(0,50){\circle*{3}} \put(0,40){$v_5$} \end{picture}} \multiput(5,100)(95,0){4}{\begin{picture}(90,80) \put(60,50){\circle{6}} \put(60,50){\line(1,0){20}} \put(80,50){\circle*{3}} \put(75,40){$v_5$} \end{picture}} \multiput(5,20)(95,0){4}{\begin{picture}(90,80) \put(60,50){\circle{6}} \put(0,50){\line(1,0){20}} \put(0,50){\circle*{3}} \put(0,40){$v_5$} \end{picture}} \parbox{15cm}{\caption{\label{tab:1} A complete family $\cF$ (Example 1)}} \end{picture} \end{center} \end{figure} % \begin{figure}[hbt] \begin{center} \begin{picture}(385,340) \multiput(5,20)(0,80){4}{\begin{picture}(380,80) \multiput(0,0)(95,0){4}{\begin{picture}(90,80) \thicklines \put(20,40){\line(0,1){20}} \multiput(20,40)(0,20){2}{\circle*{3}} \put(50,40){\line(0,1){20}} \multiput(50,40)(0,20){2}{\circle*{3}} \put(70,50){\circle*{3}} \thinlines \put(20,45){\oval(26,40)} \put(50,45){\oval(26,40)} \put(10,10){$R_1$} \put(40,10){$R_2$} \put(15,30){$v_1$} \put(8,48){$v_2$} \put(45,30){$v_3$} \put(52,49){$v_4$} \put(65,40){$v_5$} \end{picture}} \put(0,0){\begin{picture}(380,80) \thicklines \put(20,60){\line(1,0){30}} \put(115,60){\line(3,-2){30}} \put(210,40){\line(3,2){30}} \put(305,40){\line(1,0){30}} \end{picture}} \end{picture}} \multiput(5,260)(95,0){4}{\begin{picture}(90,80) \put(20,40){\circle{6}} \put(50,60){\line(2,-1){20}} \end{picture}} \multiput(5,180)(95,0){4}{\begin{picture}(90,80) \put(20,40){\circle{6}} \put(50,40){\line(2,1){20}} \end{picture}} \multiput(5,100)(95,0){4}{\begin{picture}(90,80) \put(20,60){\circle{6}} \put(50,60){\line(2,-1){20}} \end{picture}} \multiput(5,20)(95,0){4}{\begin{picture}(90,80) \put(20,60){\circle{6}} \put(50,40){\line(2,1){20}} \end{picture}} \parbox{15cm}{\caption{\label{tab:2} A complete family $\cF$ (Example 2)}} \end{picture} \end{center} \end{figure} % \protwo{Properties of $\cF$}{pro:procF} Let $T\in\cT^k$, $\a$ a $T$--admissible function and $\l<1/2$. Then: \beqano & {\rm (i)} & \cF(T)=\{ T\} \sse \ \cR(T)=\emptyset\ ;\\ % & {\rm (ii)} & \cF(T') = \cF(T)\qquad \forall\ T'\in\cF(T)\ ;\\ % & {\rm (iiii)} & \cF(T)\cap\cF(T')\neq \emptyset \implies\ T'\in \cF(T)\ ;\\ % & {\rm (iv)} & \bT,\cR(T),U(T)\quad{\rm is\ a\ complete\ and \ minimal}\\ & &{\rm set\ of\ invariants\ for\ \cF(T)}\ ;\\ % & {\rm (v)} & |\cF|=\prod_{\bR\in\bcR} |V(\bR)|^2\ ;\\ % & {\rm (vi)}& \deg_{T'} v\le \deg_{T''} v+2 \qquad \forall\ T',T''\in\cF(T)\\ \eeqano \epro The proof of the Proposition is a simple consequence of the various definitions. \rem{rem:5.6} >From (i), (ii), (iii) above it follows immediately that the property $T'\in\cF(T)$ is an {\em equivalence relation}. Thus the set of all labeled rooted trees, given $\{n_i\}\=\{\a_{v_i}\}$, can be {\em partitioned into disjoint complete families}: here we have preassigned the function $\a_{v_i}\=n_i$ and we use the convention that if $\a$ is not admissible for some $T$ we are omitting the (meaningless) contribution coming from that tree. \erem % A final comment: the set of labels $\{v_1,...,v_k\}$ of $\cT^k$ can be thought of as the set of vertices $V$ of $\cF$ and for any $v\in V$ we set \beq \deg_\cF v= \max_{T'\in\cF} \deg_{T'} v \label{defdegF} \eeq and from (vi) of Proposition \ref{pro:procF} it follows immediately that: \beq \deg_\cF v - 2 \le \deg_{T'} v\le \deg_\cF v\ ,\qquad \forall\ T'\in\cF \label{degF} \eeq In Figure~\ref{tab:1} and Figure~\ref{tab:2} we show the complete families associated to order 5 trees with two resonances. % \nin In Figure~\ref{tab:1} we have $\cR_0=R_1$, $\cR=\{R_1,R_2\}$, $C_1 \equiv \overline{R}_1$, $C_2 \equiv R_2$, $\overline{T} = \{\eta\} \cup \{v_5\}$, $\overline{R}_1 \equiv \{v_1\} \cup \{v_4\}$, $u(C_1)=\eta$, $w(C_1)=v_5$, $u(C_2)=v_1$, $w(C_2)=v_4$, $|\cF|=|V(\bR_1)|^2 |V(R_2)|^2=16$. \giu In Figure~\ref{tab:2} we have $\cR_0=\cR=\{ R_1, R_2\}$, $C_1=(R_1,R_2)$, $u(C_1)=u(R_1)=\eta$, $w(C_1)=w(R_2)=v_5$, $|\cF|=|V(R_1)|^2 |V(R_2)|^2=16$. %******************************************** \newpage \section{Estimates} \label{sec:estima} \setcounter{equation}{0} % In this section we formulate the estimates on (sums) of products of small divisors needed to prove Kolmogorov's (KAM) theorem following Siegel's strategy. The first two lemmas are a simple generalization of Siegel's argument and deal with situations without resonances (or better with non critical resonances). The first lemma is taken from \cite{E}. The third and fourth lemmas are the crucial point of our paper: it is shown that sums over complete families obey the same bounds that hold for products of small divisors without resonances (\ie without repetitions). In other words, the ``divergent terms" whose occurrence has been discussed in Section~\ref{sec:diverge} compensate (and in fact {\em they do not cancel exactly}) within complete families. The four lemmas will easily yield Kolmogorov's theorem for the case under consideration \equ{he}; see Theorem~\ref{thm:kam} below. The proofs are presented in the next section. \giu Recall \equ{dioph} and the definition of admissible functions \equ{adm}. % \lemtwo{Siegel, Eliasson}{lem:se} Let $T\in\cT^k$ be a rooted tree of order $k$ (\ie $k$ $\=$ cardinality of $V(T)$), $\a$ a $T$--admissible function and $0<\l<1$. Assume that \beq {\rm if}\ \d_u=\d_w\ {\rm for\ some\ }u>w \ \ {\rm then\ } \exists\ v\ {\rm s.t.} \ u>v>w\ {\rm and}\ |\d_v|>\l |\d_u| % \label{hy1} \eeq Then there exist constants $c_1>\g$, $\b_1>\t$ such that \beq \prod_{v\in T} |\d_v|^{-1}\le c_1^k \ \prod_{v\in T} |\a_v|^{\b_1} \label{e.1} \eeq The constant $c_1,\b_1$ can be taken to be \beq c_1=\g\ 2^{6\t}\ \frac{1+\l}{\l}\ ,\qquad \b_1=3\t \label{const1} \eeq \elem % Next Lemma extend the above type of estimate to products of small divisors arising in trees with (possible) resonant subtrees which are non $\l$--resonant (for some prefixed $\l<1/2$); recall that such trees may contain couples of points which violate \equ{hy1} (see Remark~\ref{rem:5.1}) so that the result described in the next Lemma is a genuine generalization of the estimates described in Lemma~\ref{lem:se}. % \lem{lem:2} Let $T$ and $\a$ be as in Lemma~\ref{lem:2} and let $0<\l<1/2$. Assume that there are no $\l$--resonance (Definition~{\rm\ref{dfn:4}}). Then there exist constants $c_2>\g$ and $\b_2>\t$ such that \beq \prod_{v\in T} |\d_v|^{-1}\le c_2^k \ \prod_{v\in T} |\a_v|^{\b_2} \label{e.2} \eeq The constant $c_2,\b_2$ can be taken to be \beq c_2=c_1\ 2^{\t}\ \frac{1-\l}{1-2\l}\ ,\qquad \b_2=\b_1+\t \label{const2} \eeq \elem % \lemtwo{Small divisor compensations}{lem:3} \HB\nin Let $T$, $\a$ and $\l$ be as in Lemma~{\rm\ref{lem:2}} and let $0<\bl<1$. Let $R$ be a minimal $\l$--(sub)resonance of $T$ (\ie $R$ does not contain any $\l$--subresonance, see Definition~{\rm\ref{dfn:7}}) and let $n\in\zNn$ be such that $\agl{n}\le \bl \D(R)$. Let $z$ be an extra vertex (not in $T$) and set $\a_z=n$. Finally, for any $u,w\in R$, denote by $R_u^w$ the tree\footnote{\rm If $R$ is rooted, $R=R_r$, first remove the edge $\eta r$ so as to obtain an unrooted subtree.} $R\cup \{z\}+wz$ rooted at $u$. Then there exist constants $c_3>\g^2$, $\b_3>2\t$ such that for all $1\le i,j\le N$ \beq |\sum_{u,w\in R} \a_{ui} \a_{wj} \prod_{v\in R} \d_v(R_u^w)^{-2}| \le c_3^{|V(R)|} \ \prod_{v\in R} |\a_v|^{\b_3} \label{e.3} \eeq where $\a_{ui}$ (respectively $\a_{uj}$) denote the $i^{\rm th}$ (respectively the $j^{\rm th}$) component of the integer vector $\a_u$. The constant $c_3,\b_3$ can be taken to be \beq c_3=\big( c_2\ 2^{\t+1} (1-\bl)^{-1} \big)^2\ ,\qquad \b_3=2(\b_2+\t+1) \label{const3} \eeq \elem % The estimates \equ{e.3} lead easily to control the contribution to \equ{exptree} coming from complete families. % \lem{lem:4} Let $T$, $\a$ and $\l<1/2$ be as above. Let $\cF\=\cF(T)$ be the complete family defined in Definition~{\rm\ref{dfn:13}}, $V\=V(T)=V(\cF)$ and recall the definition of $\deg_\cF$ given in \equ{defdegF}. Then there exist constants $c_4>\g^2$, $\b_4>2\t$ such that \beq \big| \sum_{T'\in\cF} \prod_{vv'\in E(T')} \a_v\cdot\a_{v'} \prod_V \d_v^{-2}\big| \le \prod_{v\in V} |\a_v|^{\deg_\cF v} \ c_4^{|V|}\ \prod_{v\in V} |\a_v|^{\b_4} \label{e.4} \eeq The constants $c_4,\b_4$ can be taken to be \beq c_4\=(c_2\ 2^{\t+1}\ \frac{1-\l}{1-2\l} )^2\ , \qquad \b_4= \b_3 \label{const4} \eeq \elem % Kolmogorov's theorem follows now easily. We introduce the following norms on analytic functions $g:\tN\to\complex$: \beq \| g\|_\s \= \sum_{n\in\zN} |g_n| e^{|n|\s} \label{norm} \eeq Clearly, for any analytic function on $\tN$ there exist a $\s>0$ such that the above norm is finite; viceversa if $g$ is such that $\|g\|_\s<\infty$ for some $\s>0$ then $|g_n|\le \|g\|_\s \exp(-|n|\s)$ so that $g$ is analytic. If $g=(g_1,...,g_M):\tN\to\complex^M$ we set \beq \|g\|_\s\=\sup_{1\le j\le M} \|g_j\|_\s \label{norm2} \eeq % \thm{thm:kam} Let $f$ in \equ{he} be real--analytic with $\| f\|_\bs<\infty$ for some $\bs>0$. Then the formal solution $X(\th,\e)$ described in Proposition~{\rm\ref{pro:ef}} is real--analytic. Furthermore, fix $0<\s<\bs$, $0<\l<1/2$ and set \beq \e_0\=\frac{(\bs-\s)^{\b_5}}{c_5 \|f\|_\bs}\ \quad {\rm with}\quad\cases{ \b_5\=\b_4+4\cr c_5\= c_4 2^{\b_5+1}\ \b_5!\cr} \label{const5} \eeq where the constants $\b_4,c_4$ are defined above. Then $X(\th,\e)$ is jointly analytic in the domain $\{|\Im \th_i|\le \s\}$ $\times$ $\{|\e|\le \e_0\}$ and for any (complex) $\e$ with $|\e|<\e_0$ the following bound holds \beq \|X(\cdot,\e)\|_\s \le \frac{\bs-\s}{2}\ \frac{|\e|}{\e_0-|\e|} \label{est} \eeq \ethm % \rem{rem:constants} Choosing $\l=1/5$ (so as to minimize the constant $c_5$) one obtains \beq \b_5= 10\t+12\ ,\qquad c_5< \g^2 \ 2^{26 \t+23}\ (10\t+12)! \eeq Such constants are certainly not optimal. \erem % %******************************************** \section{Proofs} \label{sec:proofs} \setcounter{equation}{0} % In the following five Subsections we give the proofs of the four Lemmas and of the Theorem of Section~\ref{sec:estima}. In the first Subsection we show how Kolmogorov's Theorem can be easily derived from Lemma~\ref{lem:4} and in the remaining Subsections we give the proofs of the four Lemmas. % \subsection{Proof of Theorem {\bf 6.1}} %\ref{thm:kam} \label{subsec:1} % Recall the form of the coefficients $\Xk_{jn}$ given in \equ{exptr1} $\div$ \equ{exptr3}. By Remark~\ref{rem:5.6} we know that, for a given choice of $\{n_1,...,n_k\}$ with $n_i\in\zNn$,the labeled rooted trees in $\cT^k$ can be partitioned into complete families having as common admissible $\a$ function $\a_{v_i}=n_i$. Denote by $\calN\=\calN(n_1,...,n_k)$ the set of all such complete families. Then, by \equ{exptr1} $\div$ \equ{exptr3}, by Lemma~\ref{lem:4} and by \equ{defdegF}, \equ{degF} we see that, for any $0<\s<\bs$, \beqa &&\sum_{n\in\zNn} e^{\s|n|} |\Xk_{jn}| \nonumber\\ &&= \sum_n e^{\s|n|} \frac{1}{k!}\ \Big| \sum_{ {n_1,...,n_k}\atop{n_i\neq 0,\sum n_i=n}} \sum_{\cF\in\calN} \prod_{i=1}^k f_{n_i} \sum_{T\in \cF} \prod_{vv'\in E(T)} \a_v\cdot\a_{v'} \prod_{V} \d_v^{-2}\Big| \nonumber\\ && \le \sum_{n} \frac{1}{k!} \sum_{ {n_1,...,n_k}\atop{n_i\neq 0,\sum n_i=n}} \sum_{\cF\in\calN} \prod_{i=1}^k \big(f_{n_i}e^{\s|n_i|}\big) \prod_{v\in V} |\a_v|^{\deg_\cF v} c_4^k \prod_{V} |\a_v^{\b_4}|\nonumber\\ &&\le \frac{1}{k!} \sum_{T\in\cT^k} \sum_{\a:T\to\zN} \prod_{V} \big( c_4 |f_{\a_v} |\a_v|^{\b_4+2} e^{\s |\a_v|}\big) \prod_V |\a|^{\deg v} \label{6.1} \eeqa In view of Proposition~\ref{pro:exptree2} we see that, if we define \beq \phi_n\=c_4|f_n| e^{\s|n|} |n|^{\b_4+2} \label{6.2} \eeq then, last line of \equ{6.1} is exactly the formal solution $g$ of the equation \equ{gphi} with coefficient $g_k$ given in \equ{exptree2}. But since (by the analyticity of $f$) the nonnegative numbers $\phi_n$ decay exponentially fast as $|n|\to\infty$, the function of two complex variables \beq G(x,\e)\= x-\e\sum_{n\in\zN} \phi_n |n| e^{|n|x} \label{6.3} \eeq is holomorphic and bounded in the complex $2$--disk \beq D\=\{ x\in\complex:\ |x|\le s\} \times \{\e\in\complex: \ |\e|\le\e_0\} \label{6.4} \eeq for any $\e_0>0$ and any $0w\ \ {\rm then}\ \exists \ v \ {\rm between} \ u\ {\rm and}\ w \ {\rm s.t.}\ |\d_w|>\l|\d_v| \label{6.8} \eeq Then there exist positive constants $D_0\ge B_0\ge \g$ such that \beq \prod_P |\d_v|^{-1}\le B_0 D_0^{k-1} \prod_V |\a_v|^\t \label{6.9} \eeq The constants $B_0,D_0$ can be taken to be \beq B_0=\g\ ,\qquad D_0=\g\ 2^\t \ \frac{1+\l}{\l} \label{6.10} \eeq \esublem % \PROOF By induction on $k$. If $k=1$, \equ{6.9} follows at once from the Diophantine inequality \equ{dioph}. Let $k\ge 2$ and assume the statement true for all rooted paths with up to $k-1$ vertices satisfying \equ{6.8}. Let $(P,\a)$ as above. Define $s$ to be the greatest integer between $1$ and $k$ such that \beq |\d_{v_s}|\ge \max_{1\le i\le k} |\d_{v_i}| \label{6.11} \eeq There are four cases: \beqa & {\rm (i)} \ s=1 \ , &{\rm(ii)}\ \a_{v_s}+\a_{v_{s-1}}\neq 0\ , \label{6.12} \\ &{\rm (iii)} \ \a_{v_s}+\a_{v_{s-1}}= 0\ , s+11$, by the maximality of $|\d_{v_s}|$ we have \beq \agl{\a_{v_i}} = |\d_{v_i}-\d_{v_{i+1}}|\le|\d_{v_i}|+|\d_{v_{i+1}}| \le 2|\d_{v_s}| \label{6.13} \eeq hence, in case (i), by \equ{dioph}, \beq |\d_{v_1}|^{-1}\le 2 \agl{\a_{v_1}}^{-1} \le 2 \g|\a_{v_1}|^\t \label{6.14} \eeq while in case (ii) \beq |\d_{v_s}|\ge \frac{1}{2} \max\{\agl{\a_{v_s}}, \agl{\a_{v_{s-1}}}\} \label{6.15} \eeq In case (iii) and (iv) we have that $\d_{v_{s+1}}=\d_{v_{s-1}}$, and by \equ{6.8} \beq |\d_{v_{s+1}}|>\l|\d_{v_s}|>\l|\d_{v_{s+2}}| \label{6.16} \eeq where last inequality holds only in case (iii). In all cases mimicking \equ{6.13} replacing $2$ by $(1+\l)^{-1}$ we obtain \beq |\d_{v_{s+1}}|\ge \frac{\l}{1+\l} \max \{ \agl{\a_{v_{s+1}}}, \agl{\a_{v_s}}\}\ . \label{6.17} \eeq Let now $\bv\=v_s$ in case (i), (ii) and $\bv\=v_{s+1}$ in case (iii), (iv) and let, in the cases (ii), (iii), (iv), $v'>\bv$ be adjacent to $\bv$. Then in case (ii), (iii), (iv) we obtain from \equ{6.15} and \equ{6.17}, using the Diophantine inequality \equ{dioph} (and the fact that $\l<1$), that \beq |\d_\bv|^{-1}\ \frac{|\a_\bv+\a_{v'}|^\t}{|\a_\bv|^\t |\a_{v'}|^\t} \le \frac{1+\l}{\l} 2^\t \g \label{6.18} \eeq Now, if we let $\bar P\= P/\{\bv\}$ (recall Definition~\ref{dfn:11}) and $\bar \a:\bar P\to \zNn$ defined by \beq \bar a_v\=\a_v \ {\rm in\ case\ (i)}\ ; {\rm and \ in\ case\ (ii),(iii),(iv):}\quad \bar \a_v\equiv \cases{ \a_v & if $v\neq v'$\cr \a_v+\a_\bv & if $v=v'$\cr} \label{6.19} \eeq Then \beq \prod_P |\d_v|^{-1}= |\d_\bv|^{-1} \prod_P |\d_v(\bar P;\bar \a)|^{-1} \label{6.20} \eeq If $(\bar P,\bar \a)$ verifies \equ{6.8} then we can apply the inductive hypothesis and, using \equ{6.15} or \equ{6.16}, the Sublemma would follow immediately. It remains to check that $(\bar P,\bar \a)$ satisfy \equ{6.8}. \nin Case (i) is obvious as $\bar \a$ coincide with $\a$ on $\bar P=v_2...v_k$. Also case (iv) is easily checked as any couple of resonant points (\ie points where the divisors $\d$ coincide) in $\bar P$ would also be resonant for $P$. Consider case (ii) and assume that $u>w$ form a couple of resonant points for $\bar P$. If either $w>v_{s-1}$ or $v_{s-1}>u$, then \equ{6.8} follows immediately from the validity of \equ{6.8} for $P$. If $u>v_{s-1}>w$, let $v_i\in P$ be such that $u>v_i>w$ and such that $|\d_w|>\l |\d_{v_i}|$. Then, if $v_i\neq v_s$, \equ{6.8} holds for $\bar P$ as $v_i\in \bar P$; on the other hand if $v_i=v_s$ then \beq |\d_{v_s}|\ge |\d_{v_{s-1}}|\ \implies |\d_w|>\l |v_s|\ge \l |\d_{v_{s-1}}| \label{6.21} \eeq showing that \equ{6.8} holds for $\bar P$ also in this case. Consider case (iii). Arguing as above we see that we only have to check the case when, in $P$, we have $\d_u=\d_w$ with $u>v_s>w$. Let as above $v_i\in P$ be such that $|\d_w|>\l |\d_{v_i}|$. If $v_i=v_{s+1}$ then either $u>v_{s-1}$ or $u=v_{s-1}$. If $u>v_{s-1}$ then $|\d_w|>\l|\d_{v_{s+1}}|=\l|\d_{v_{s-1}}|$. If $u=v_{s-1}$, then \beq \sum_{ {v\in P}\atop{v_s>v>w}} \a_v= \sum_{ {v\in P}\atop{v_{s+1}>v>w}} \a_v=-\a_{v_{s+1}} \eeq Hence $v_{s+1}>w$ for a resonant couple in $P$ and there exists by \equ{6.8} a point $v_j$ between $v_{s+1}$ and $w$ ($\implies v_j\in\bar P$) such that $|\d_w|>\l |\d_{v_j}|$ and we see that \equ{6.8} holds for $(\bar P,\bar \a)$ also in this final case. \qed % \sublem{sublem:2} Let $r\ge 0$ and $s\ge 2$ be integers and let $x_1,...,x_r$, $y_1,...,y_s$ be real numbers greater or equal than 1 and let $z$ be their sum $z\=\sum x_i +\sum y_j$ ($r=0$ means that there are no $x$'s). If \beq \sum_{j=1}^s y_j\ge \frac{z}{2}\qquad {\rm and} \qquad y_j\le \frac{z}{2}\quad\forall\ j \label{6.22} \eeq then \beq \prod_{i=0}^r x_i^{-1} \ \prod_{j=1}^s y_j^{-2} \le 2^{r+2s} z^{-3} \label{6.23} \eeq \esublem % \PROOF The proof is taken from \cite{E}: we include it for completeness. We first observe that it is enough to prove the statement for $r\le 1$ (as $\sum x_i\le r \prod x_i \le 2^{r-1} \prod x_i$). We can also assume that $s\le 3$ (in fact, since $(y_i+y_j)^2\le 4(y_i y_j)^2$, we can replace $y_i$ and $y_j$ by their sum if such sum is $\le z/2$; hence we can assume $y_i+y_j>z/2$ for all $i\neq j$ and repeating the argument we end up with at most three $y_j$). Now, let $s=3$. For $z=3$ the result holds therefore we assume that $z>3$ and $y_1\le y_2\le y_3$. Clearly $y_2 \frac{1}{2} |\a|(T)\} \label{defP} \eeq It is easy to check that {\em $P$ is a path}: $P=u_0...u_p$ with $p\ge 0$, $u_0$ being the root $r$ of the tree $T$. To each $u_i$ we associate $m_i\ge 0$ subtrees $S_{ij}$ as follows. Let $m_i\= \deg u_i-2$ when $i\le p-1$, $m_p\=\deg u_p-1$ and let, for $m_i>0$, $v_{ij}$ be the $m_i$ vertices adjacent to $u_i$ with $v_{ij}\notin P$. We then set $S_{ij}\=S_{v_{ij}}$ when $m_i\neq 0$ and $S_{ij}=\emptyset$ otherwise. We also define \beqa & & y_{ij}\=|\a|(S_{ij})\ ,\quad y_i\=\sum_{j=1}^{m_i} y_{ij}\ ,\quad x_i\=|\a_{u_i}|\nonumber\\ & & I\=[0,...,p]\ ,\quad I_0\=\{i\in I:\ m_i=0\}\ , \quad I_1\=\{i\in I:\ m_i>0\}\nonumber\\ &&m\=\sum_I m_i\ ,\quad k_{ij}\=|S_{ij}|\ , \quad k_i\=\sum_{j=1}^{m_i} k_{ij}\ ,\quad p_0\=|I_0|\ ,\ p_1\=|I_1|\nonumber\\ && \phantom{aaaaaaaaaa} \label{defvarie} \eeqa It then follows \beq |\a|(T)=\sum_I x_i + \sum_{I_1} y_i\ ,\quad p_0+p_1=p+1\ ,\quad k=p+1+\sum_{i\in I_1} k_i\ ,\quad m\ge p_1 \label{6.30} \eeq and, by the definition of $P$ \beqa & &\sum_{j\ge i} x_j + \sum_{ {j\ge i}\atop{j\in I_1}} > \frac{1}{2} |\a|(T)\ ,\qquad \forall\ i\in I\nonumber\\ & &x_i\le \frac{1}{2} |\a|(T)\ \ \forall\ iFrom now on we assume that $I_1\neq \emptyset$ \ie $m\ge p_1>0$. We shall adopt the convention that {\em sums or products over empty set of indices have to be disregarded} (\ie replaced, respectively, by $0$ or $1$). Next, let \beq \bar \a_{u_i}\=\a_{u_i}+\a(S_{ij})\ \qquad \implies \quad |\bar \a_{u_i}|\le x_i+y_i \label{6.33} \eeq then $\d_{u_i}\=\d_{u_i}(T;\a)=\d_{u_i}(P;\bar \a)$, whence \beq \prod_T |\d_v|^{-1}= \prod_P |\d_{u_i}(P;\bar \a)|^{-1} \prod_{i\in I_1} \prod_{j=1}^{m_i} \prod_{v\in S_{ij}} |\d_v|^{-1} \label{6.34} \eeq and we can use Sublemma~\ref{sublem:1} to bound the product over $P$ (as $(P,\bar \a)$ obviously satisfy \equ{6.8}) and the inductive hypotheses to bound the product over $S_{ij}$ (note that $\d_v(T)=\d_v(S_{ij}$)). Recalling the definitions in \equ{defvarie}, we obtain from \equ{6.34} and \equ{6.33} \beq \prod_{v\in T} |\d_v|^{-1} \le M B D^{k-1} \big(\prod_T |\a_v|^{3\t}\big) |\a|(T)^{-2\t} \label{6.35} \eeq with \beq M\= \frac{B_0}{B} \big(\frac{B}{D}\big)^m \big(\frac{D_0}{D}\big)^p \big[ \prod_{I_0} x_i^{-2}\cdot\prod_{I_1}\big(\frac{x_i+y_i}{x_i^3} \prod_jy_{ij}^{-1}\big)\cdot|\a|(T)^2\big]^\t \label{6.34bis} \eeq and we have to show that $M\le 1$ if we choose suitably $D,B$. Suppose first that $p\in I_0$. Then by the first inequality in \equ{6.31} with $i=p$ we have \beq x_p>\frac{1}{2} |\a|(T) \label{6.35bis} \eeq Then using repeatedly the bound $\sum_{i=1}^m b_i\le m\prod_{i=1}^m b_i$ $\le$ $2^{m-1} \prod_{i=1}^m b_i$ (valid for real numbers $b_i\ge 1$) and estimating $|\a|(T)^2$ in \equ{6.34bis} by $4x_p^2$ we get \beq M\le \frac{B_0}{B} \big(\frac{D_0}{D}\big)^p \big[ \prod_{{i\in I}\atop{i\neq p}} x_i^{-2} 2^{p_1+2}\ 2^{m-p_1}\ \prod_{i\in I_1,j} y_{ij}^{-1} \big]^\t\le 1 \label{6.36} \eeq provided \beq D\ge 2^\t \max\{B,2^\t D_0\}\ ,\qquad B\ge B_0 2^{2\t} \label{6.37} \eeq (which is stronger than \equ{6.32}). Suppose now that $p\in I_1$. We can assume that $x_p\le |\a|(T)/2$ (otherwise \equ{6.35bis} holds also in this case and we can proceed as above). Then \beq M\le \frac{B_0}{B} \big(\frac{B}{D}\big)^m \big(\frac{D_0}{D}\big)^p \big[ 2^{2m}\ \prod_{{i\in I}\atop{i\neq p}}x_i^{-1} \cdot\prod_{{i\in I_1}\atop{i=neq p}} y_i^{-1} \cdot x_p^{-2} \prod_{j=1}^{m_p} y_{pj}^{-2}\cdot y_p |\a|(T)^2\big]^\t \label{6.38} \eeq and we see that we are in position to apply Sublemma~\ref{sublem:2} (with $r=p+p_1-1$, $s=m_p+1$, $z=|\a|(T)=\sum_I x_i+\sum_{I_1} y_i$) obtaining \beq M\le \frac{B_0}{B} \big(\frac{B}{D}\big)^p \big(\frac{D_0}{D}\big)^p \big[ 2^{2m}\ 2^{p+p_1-1}\ 2^{2(m_p+1)}\ \frac{y_p}{|\a|(T)}\big]^\t\le 1 \label{6.39} \eeq provided we take \beq B\= B_0 \ 4^\t\ ,\qquad D\= 4^\t \max\{ D_0,B_0 2^{4\t} \} \label{6.40} \eeq (which satisfy also \equ{6.37}). Thus, by \equ{6.10}, we can take \beq \b_1\=3\t\ ,\qquad c_1=\g\ 2^{6\t}\ \frac{1+\l}{\l}\ \label{6.41} \eeq This concludes the proof of Lemma~\ref{lem:se}. \qed % \subsection{Proof of Lemma {\bf 6.2} } %\ref{lem:2} \label{subsec:3} % We prove the Lemma by induction on $k\=|V|$. For $k=1$, \equ{e.2} follows from \equ{dioph} (as $c_2>\g$ and $\b_2>\t$). Let $k\ge 2$, assume that the Lemma holds for all trees with up to $k-1$ vertices and let $T\in\cT^k$ have no $\l$--resonances. Recall that even though $T$ does not contain $\l$--resonances, it may contain either $\l$--resonant couples of points (compare Remark~\ref{5.1}), or {\em adjacent points} $u,w$ such that $\d_u=\d_w$ (recall that the main hypothesis of Lemma~\ref{lem:se}, \equ{hy1}, excludes the presence of adjacent points of constancy for $\d$). For any couple $(u,w)$ of $\l$--resonant points let $R(u,w)$ be the associated resonant subtree (see \equ{5.3}). Let $R$ be a minimal resonant tree associated to a couple of $\l$--resonant points (minimal meaning that $R$ does not contain $\l$--resonant points); if there are no $\l$--resonant points we set $R=\emptyset$. There are two cases: \nin (i) either $R=\emptyset$ or $R$ contains two adjacent points $u>w$ such that $\d_u=\d_w$, or \nin (ii) $R\neq \emptyset$ and $R$ does not contain any couple of adjacent points on which $\d$ is constant. \nin Consider case (i): if there are no couples of adjacent points $u,w$ such that $\d_u=\d_v$, then Lemma~\ref{lem:2} follows from Lemma~\ref{lem:se}, otherwise let $S=S(v_1,v_2)$ be a minimal resonant subtree such that $v_1$ and $v_2$ are adjacent and of constancy for $\d$ (minimal means that $S$ does not contain any couple of adjacent points on which $\d$ is constant). Since $\a_v\neq 0$ for all vertices $v$, $S$ contains at least two points and by hypothesis $S$ is not $\l$--resonant. Thus observing that \beq \prod_T |\d_v|^{-1} = \prod_{T/S} |\d_v|^{-1} \prod_S |\d_v|^{-1} \= \prod_{T/S} |\d_v|^{-1} \prod_S |\d_v(S_{v_1})|^{-1} \label{6.42} \eeq we see that we can apply Lemma~\ref{lem:se} to the tree $S_{v_1}$ while to the tree $T/S$ we can use the inductive %FOOTNOTE hypothesis\footnote{Note that contracting resonant trees associated to adjacent points where $\d$ is constant cannot introduce new resonances, thus all resonances in $T/S$ are not $\l$--resonant.} % so that \equ{e.2} holds in case (i) provided \beq c_2\ge c_1\ ,\qquad \b_2\ge \b_1 \label{6.43} \eeq % Consider case (ii). It is easy to see that the tree $T/R$ does not contain any $\l$--resonance and since \beq \prod_T |\d_v|^{-1} = \prod_{T/R} |\d_v|^{-1} \prod_R |\d_v|^{-1} \label{6.44} \eeq we see that we can estimate, by the inductive hypothesis, the product over $T/R$. To bound the product over $R=R(u,w)$, let $n\=\sum_{v'\le w} \a_{v'}$. Then, since $(u,w)$ form a couple of $\l$--resonant points, for any $u\l^{-1} \g$). \qed % \subsection{Proof of Lemma {\bf 6.3} } %\ref{lem:3} \label{subsec:4} % The heart of the matter is the following remarkable {\em algebraic fact.} \protwo{Compensations}{pro:comp} Let $T,R,n$ be as in Lemma~\ref{lem:3} and set (for prefixed $1\le i,j\le N$) \beq \s(x)\= \sum_{u,v\in R} \a_{ui} \a_{uj} \prod_{ {v\in P(u,w)}\atop{v\neq u}} \big( \d_v+x\big)^{-2} \prod_{ v\in R\bks P(u,w)} \d_v^{-2} \label{6.51} \eeq where $P(u,w)$ is the path joining $u$ and $w$ (if $u=w$ the first product is missing, while if $R$ is a path the second product is missing). Then \beq \s(0)=\s'(0)=0 \qquad ( {}'\ \=\frac{d}{dx})\ . \label{6.52} \eeq \epro % Note that \beq \s(\gl{n})= \sum_{u,w\in R} \a_{ui} \a_{uj} \prod_R \d_v(R_u^w)^{-2} \label{6.53} \eeq (the definition of $R_u^w$ is given in Lemma~\ref{lem:3}). % \giu\PROOF Recall that $\a(R)=0$ and define, for any $u$, $w$ in $R$ and any $S\subset V(R)$ such that $\a(S)=0$ \beqa && \m(u) \= \prod_{ {v\in R}\atop{ v\neq u}} |\d_v(R_u)| \nonumber\\ && \n(w)\= \sum_{u\in S,u\neq w} \a_{wi} \sum_{{v\in P(u,w)}\atop{ v\neq u}} |\d_v(R_u)|^{-1} \label{6.54} \eeqa We claim that \beqa && {\rm the\ functions\ }\ \m\ {\rm and}\ \n \ {\em are\ \ independent \ of\ } u \ {\em and}\ w\ : \nonumber\\ && \m(u)\=\m\=\m_R\ ,\qquad \n(u)\=\n\=\n_{R,S} \label{6.55} \eeqa The independence of $\m(u)$ from $u$ comes immediately from the identities \beq \d_v(R_u)=\d_v(R_w)\ ,\quad\forall\ v\notin \ P(u,w) \label{6.56} \eeq (where as usual $P(u,w)$ denotes the path joining $u$ and $w$) and \beq \prod_{ {v\in P(u,w)}\atop{v\neq u}} \d_v(R_u) = \prod_{ {v\in P(u,w)}\atop{v\neq w}} \big(-\d_v(T_w)\big) \ ,\quad \forall \ u\neq w \label{6.57} \eeq Note that \equ{6.56} holds for any rooted tree (\ie not necessarily null) while in \equ{6.57} it is important that $R$ is null. Now, for any $u$ and $w$ with $u\neq w$ \equ{6.56} and \equ{6.57} imply \beqa \m(u) & =& \prod_{ {v\in P(u,w)}\atop{ v\neq u}} |\d_v(R_u)| \prod_{v\notin P(u,w)} |\d_v(R_u)|\nonumber\\ &=& \prod_{ {v\in P(u,w)}\atop{ v\neq w}} |\d_v(R_w)| \prod_{v\notin P(u,w)} |\d_v(R_w)| = \m(w) \label{6.58} \eeqa which proves the independency of $\m$ from points in $R$. Next, observe that to prove the independence of $\n$ from points in $R$ it is enough to check that $\n(w)=\n(w')$ for {\em adjacent points} $w$ and $w'$. Thus, let $w$, $w'$ be adjacent points in $R$. Then \beqa \n(w)-\n(w') &=& \sum_{ {u\in S}\atop{u\neq w,w'}} \a_{ui} \Big[ \sum_{ {v\in P(u,w)}\atop{v\neq u}} \d_v(R_u)^{-1}- \sum_{ {v\in P(u,w')}\atop{v\neq u}} \d_v(R_u)^{-1}\Big]\ +\nonumber\\ &&\qquad + \chi_{{}_S}(w') \a_{w' i} \d_w(R_{w'})^{-1} - \chi_{{}_S}(w) \a_{wi} \d_{w'}(R_{w})^{-1}\nonumber\\ &&\phantom{a}\nonumber\\ &=& \sum_{ {u\in S}\atop{u\neq w,w'}} \a_{ui} \d_w(R_{w'})^{-1}+ \chi_{{}_S}(w') \a_{w' i} \d_w(R_{w'})^{-1} +\nonumber\\ &&\qquad + \chi_{{}_S}(w) \a_{wi} \d_{w}(R_{w'})^{-1}\nonumber\\ &&\phantom{a}\nonumber\\ &=& \Big( \sum_{u\in S} \a_{ui}\Big)\ \d_w(R_{w'})^{-1}=0 \label{6.59} \eeqa where in the second equality we used $\d_{w'}(R_w)=-\d_w(R_{w'})$ which is a particular case of \equ{6.57} and $\chi_{{}_S}$ is the characteristic function of the set $S$. This finishes the proof of the claim \equ{6.55}. \nin The check of \equ{6.52} is now trivial: \beqano \s(0) &=& \sum_{u,w\in R} \a_{ui} \a_{wj} \prod_{ {v\in R}\atop{v\neq u}} \d_v(R_u)^{-2} \\ &=& \m_R^{-2} \sum_{u,w\in R} \a_{ui} \a_{wj}=0 \eeqano and \beqano \frac{d\s}{dx}(0) &=& -2 \sum_{{u,w\in R}\atop{u\neq w}} \a_{ui} \a_{wj} \prod_{ {v\in R}\atop{v\neq u}} \d_v(R_u)^{-2} \sum_{ {v\in P(u,w)}\atop{v\neq u}} \d_v(R_u)^{-1}\\ &=& -2 \ \m_R^{-2} \sum_{w\in R} \a_{wj} \Big( \sum_{u\neq w} \a_{ui} \sum_{ {v\in P(u,w)}\atop{v\neq u}}\d_v(R_u)^{-1} \Big) \\ &=& -2 \m_R^{-2}\ \n_{R,R}\ \sum_{w\in R} \a_{wj}=0 \eeqano finishing the proof of Proposition~\ref{pro:comp}. \qed \rem{rem:comp} In fact, from the above proof it follows that \equ{6.52} holds also if the sum over $R$ is replaced by the sum over any null subset of $R$. \erem % We can now proceed with the Proof of Lemma~\ref{lem:3}. \giu Since $\a(R)=0$ we can rewrite the left hand side of \equ{e.3} as follows \beqa \sum_{u,w\in R} \a_{ui} \a_{wj} \prod_R \d_v(R_u^w)^{-2} & =& \gl{n}^{-2} \sum_{u,w\in R} \a_{ui} \a_{wj} \prod_{ {v\in R}\atop {v\neq u}} \d_v(R_u^w)^{-2} \nonumber\\ &\equiv & \gl{n}^{-2} \s(\gl{n}) \label{6.60} \eeqa see \equ{6.53}. By Proposition~\ref{pro:comp} we see that \beqa \s(x) & =& x^2\ \int_0^1 \s''(tx) (1-t)\ dt \nonumber\\ &\equiv & x^2 \bs(x) \label{6.61} \eeqa where $\bs$ is defined here. By assumption we have \beq \agl{n}\le \bl \ \D \= \bl \D(R) \label{6.62} \eeq and we see that $\s$ {\em is holomorphic and bounded} in the complex disk $\{x\in\complex: |x|\le \frac{1+\bl}{2}\ \D\}$. Recall the {\em ``Cauchy estimate"}: \beq \sup_{D'} |g^{(k)}|\le \ k!\ r^{-k} \sup_D |g|\ , \qquad r\={\rm dist} \{D',\dpr D\} \label{cauchy} \eeq valid for any holomorphic function $g$ bounded in a complex domain $D$, for any $k\ge 0$ and for any subdomain $D'$ whose closure is contained in $D$ (above $\dpr$ denotes ``boundary of")\footnote{The proof of \equ{cauchy} is standard: it follows immediately by expressing the $k^{\rm th}$ derivative of $g$ at a point $x\in D'$ in terms of Cauchy's integral formula (taking as path of integration a disk centered at $x$ and of radius $r$).}. Then, letting \beq D\=\{x\in \complex:\ |x|\le \frac{1+\bl}{2}\ \D\}\ ,\quad D'\=\{x\in \complex:\ |x|\le \bl\ \D\}\ \label{6.disk} \eeq (recall that $\bl<1$ so that $D'$ is smaller that $D$) we see that \beq \sup_{D'} |\s''| \le 8 \ (1-\bl)^{-2}\ \D^{-2}\ \sup_D |\s| \label{6.63} \eeq Furthermore, for any $x\in D$ and any $u,v\in R$ we have (recall that from the definition of $\D(R)$ it follows that $\D\le \d_v(R_u)$). \beqa \Big( \d_v(R_u)+x\Big)^{-2} &\le & \Big( \d_v(R_u) - \frac{1+\bl}{2}\ \D \Big)^{-2} \nonumber\\ &\le & 4 (1-\bl)^{-2} \ \d_v(R_u)^{-2} \label{6.64} \eeqa Thus, by \equ{6.61}, \equ{6.62}, \equ{6.63}, \equ{6.64}, \equ{6.51} and Lemma~\ref{lem:2} \beqa \gl{n}^{-2} |\s(\gl{n})| &=& |\bs(\gl{n})|\le \frac{1}{2} \sup_{D'} |\s''|\nonumber\\ &\le & 4 (1-\bl)^{-2}\ \D^{-2}\ \sup_D |\s|\nonumber\\ &\le & \Big(4 (1-\bl)^{-2}\Big)^{|V(R)|} \D^{-2}\ \sum_{u,w\in R} |\a_u|\ |\a_w|\prod_{v\neq u} |\d_v(R_u)|^{-2}\nonumber\\ &\le & \Big(4 (1-\bl)^{-2}\Big)^{|V(R)|} \D^{-2}\ \Big( |V(R)|\ \prod_{v\in R} |\a_v|\Big)^2 \ c_2^{2(|V(R)|-1)}\ \prod_{ {v\in R}\atop{v\neq u}} |\a_v|^{2\b_2} \nonumber\\ &\le & \Big( (1-\bl)^{-2}\ 2^{\t+1}\ c_2\Big)^{2|V(R)|} \ \prod_{v\in R} |\a_v|^{2 (\b_2+\t+1)} \label{6.65} \eeqa where in the last estimate we have applied Lemma~\ref{lem:2} to the tree(s) $R\bks \{u\}$ (if $\deg_{R_u} u>2$ then $R\bks \{u\}$ is the disjoint union of $\deg_{R_u}-1$ trees) and the estimate \beq \D^{-1}\le \g\ \big( \sum_R |\a_v|\big)^\t \le \g\ 2^{(|V(R)|-1)\t} \prod_R |\a_v|^\t \eeq The proof of Lemma~\ref{lem:3} is complete. \qed \subsection{Proof of Lemma {\bf 6.4}} %lem:4 \label{subsec:5} % If $T'\in \cF(T)$, denote by \beq \G(T')\=\{vv'\in E(T'):\ \exists\ R\in \cR(T)\ \ {\rm with} \ \ v'\in\bR \ {\rm and}\ v\notin R\} \label{6.66} \eeq and by $E(\bcR)\=\cup_{\bR\in\bcR} E(\bR)$. Then \beq \prod_{vv'\in E(T')} \a_v\cdot\a_{v'} = \prod_{vv'\in \G(T')} \a_v\cdot\a_{v'} \ \prod_{vv'\in E(\bcR)\cup \bT} \a_v\cdot\a_{v'} \label{6.67} \eeq and the second product does not depend on $T'$ (\ie it is common to all elements of $\cF$). Furthermore, recalling \equ{defcF}, we see that \beq \prod_{vv'\in \G(T')} \a_v\cdot\a_{v'}= \prod_{C=(\bR_1,...,\bR_s)\in\cC} \a_{u(C)} \cdot \a_{\bu_1}\ \a_{w(C)}\cdot \a_{\bw_s} \ \prod_{i=1}^{s-1} \a_{\bw_i}\cdot\a_{\bu_{i+1}} \label{6.68} \eeq last product being absent if $s=1$. Fixing a set of indices (depending upon $C$ and taking the values $1,...,N$), $j_i=j_i(C)$, for $i=0,...,s$, we can make more explicit \equ{6.67} by writing out the scalar products: \beq \prod_{vv'\in \G(T')} \a_v\cdot\a_{v'} = \sum_{ { \{j_0,...,j_s\}_{C\in\cC}}\atop{1\le j_i\le N}} \prod_{C\in\cC} \a_{u(C)j_0} \a_{w(C)j_s}\ \prod_{C\in\cC} \ \prod_{i=1}^{s} \a_{\bu_i j_{i-1} }\a_{\bw_i j_i} \label{6.69} \eeq Now observe that $|\bcR|\le (|V|-1)/2$, which implies \beq \sum_{ \{j_o,...,j_s\}_{C\in\cC}} 1 \le N^{|V|}\ , \label{6.70} \eeq and notice that, for all $T'\in\cF$ (recall \equ{defdegF}) \beq \Big| \prod_{vv'\in E(T')} \a_v\cdot\a_{v'}\Big| \le \prod_{v\in V} |\a_v|^{\deg_{T'} v} \le \prod_{v\in V} |\a_v|^{\deg_{\cF} v} \label{6.71} \eeq Then \beqa && \Big| \sum_{T'\in\cF} \prod_{vv'\in E(T')} \a_v\cdot\a_{v'} \prod_V \d_v^{-2} \Big|\nonumber\\ &&\quad \le \prod_{vv'\in E(\bcR)\cup E(\bT)} |\a_v| \ |\a_v'|\ \sum_{ \{j_0,...,j_s\} }\prod_{C\in \cC} |\a_{u(C) j_0}|\ |\a_{w(C) j_s} | \ \times\nonumber\\ &&\quad\qquad \times \prod_{\bR\in\bcR} \Big| \sum_{\bu,\bw\in\bR} \a_{\bu i_0} \a_{\bw i_0'} \prod_{v\in T'} \d_v(T')^{-2}\Big| \nonumber\\ &&\quad \le \prod_V |\a_v|^{\deg_\cF v} \ N^{|V|}\ \sup_{ {T'\in \cF}\atop{1\le i(\bR),i'(\bR)\le N}} \Big| \sum_{\bu,\bw\in\bR} \a_{\bu i} \a_{\bw i'} \prod_V \d_v(T')^{-2}\Big| \label{6.72} \eeqa where the supremum is taken over all choices of the indices $\{i(\bR)$, $i'(\bR)\}_{\bR\in\bcR}$ and $i_0,i_0'$ coincide with suitable indices $j$'s. Fix $T'\in\cF$, fix indices $i\=i(\bR)$ and $i'\=i'(\bR)$ ($1\le i,j\le N$) and let $R_*$ be a {\em minimal critical (sub)resonance}: \ie either $R_*\=R_0\in\cR_0$ is a critical resonance which does not contain any critical subresonance or $R_*$ is the smallest element of a hierarchy $\cH_\l(R_0)$ for some $R_0\in\cR_0$: $\cH_\l(R_0)=\{R_1,...,R_h\}$ and $R_*=R_h$. Now, observing that \beq \bT=\overline{T'/R_*}\ ,\qquad \bcR(T'/R_*)\=\bcR\bks\{R_*\} \label{6.73} \eeq we have \beqa && \prod_{R\in\bcR(T')} \sum_{\bu,\bw\in \bcR} \a_{\bu i} \a_{\bw i'} \prod_{v\in T'} \d_v(T')^{-2}\nonumber\\ && \qquad = \Big( \prod_{R\in\bcR(T'/R_*)} \sum_{\bu,\bw\in \bcR} \a_{\bu i} \a_{\bw i'} \prod_{v\in T'} \d_v(T'/R_*)^{-2}\Big) \Big(\sum_{\bu,\bw\in R_*} \a_{\bu i} \a_{\bw i'} \prod_{v\in R_*} \d_v(T')^{-2}\Big) \nonumber\\ &&\phantom{aa} \label{6.74} \eeqa If $h>0$ (\ie $R_*$ is a critical subresonance) and if $m_i$, for $i=0,1,...,h$, are the integer vectors associated to the hierarchy $\{R_1,...,R_h\}$ (see Remark~\ref{5.3} of Section~\ref{sec:resona}) we have for some $\s_h=\pm 1$ and some $\s_i=0,\pm 1$, (for $i=0,...,h-1$) \beq m_0\= \sum_{v