INSTRUCTIONS To produce the compuscript, Latex twice the file (so as to get correct cross references). The text between the lines BODY and ENDBODY is made of 1864 lines and 68955 bytes (not counting or ) In the following table this count is broken down by ASCII code; immediately following the code is the corresponding character. 41142 lowercase letters 1954 uppercase letters 2090 digits 8349 ASCII characters 32 17 ASCII characters 33 ! 43 ASCII characters 34 " 52 ASCII characters 35 # 1720 ASCII characters 36 $ 734 ASCII characters 37 % 46 ASCII characters 38 & 169 ASCII characters 39 ' 715 ASCII characters 40 ( 764 ASCII characters 41 ) 144 ASCII characters 42 * 92 ASCII characters 43 + 830 ASCII characters 44 , 293 ASCII characters 45 - 634 ASCII characters 46 . 3 ASCII characters 47 / 127 ASCII characters 58 : 75 ASCII characters 59 ; 5 ASCII characters 60 < 309 ASCII characters 61 = 19 ASCII characters 62 > 2 ASCII characters 64 @ 56 ASCII characters 91 [ 3574 ASCII characters 92 \ 56 ASCII characters 93 ] 395 ASCII characters 94 ^ 906 ASCII characters 95 _ 84 ASCII characters 96 ` 1734 ASCII characters 123 { 53 ASCII characters 124 | 1734 ASCII characters 125 } 35 ASCII characters 126 ~ BODY %%% This is a Latex file producing the compuscript entitled %%% "A Direct Proof of a Theorem by Kolmogorov %%% in Hamiltonian Systems" %%% by L.Chierchia and C. Falcolini %%% %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% FORMATO GRANDE % \documentstyle[amssymbols,12pt]{article} \setlength{\hoffset}{-1.cm} \setlength{\voffset}{-1.8cm} \setlength{\textwidth}{ 15.8cm} \setlength{\textheight}{22cm} \setlength{\parindent}{8mm} \setlength{\footskip}{2.truecm} \frenchspacing %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% %%% %%% FOR SHORT %%% %%% \newcommand\beq[1]{ \begin{equation}\label{#1} } \newcommand{\eeq}{ \end{equation} } \newcommand{\beqno}{ \[ } \newcommand{\eeqno}{ \] } \newcommand\beqa[1]{ \begin{eqnarray} \label{#1}} \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\ } \allowbreak} \newcommand{\REMARK}{\medskip\noindent{\bf Remark\ }} % %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% %%% %%% MATH MODE DEFINITIONS: %%% % % \newcommand{\io}{ {\infty}} \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 } } \newcommand{\tM}{ {\torus^M } } \newcommand{\rM}{ {\real^M} } \newcommand{\cM}{ {\complex^M } } \newcommand{\zM}{ {\integer^M } } % \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{\ppk}{ { (k)}} \newcommand{\ppo}{ { (0)}} \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{\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}} \newcommand\agl[1]{ |\lg{#1}\rg| } \newcommand\gl[1]{ \lg{#1}\rg } %******************************************** \newcommand\sumsud[2]{\sum_{ \scriptstyle {#1} \atop \scriptstyle {#2} }} \newcommand\sumsut[3]{\sum_{ {\scriptstyle {#1} \atop \scriptstyle {#2}} \atop \scriptstyle {#3} }} \newcommand\sumsuq[4]{\sum_{ { {\scriptstyle {#1} \atop \scriptstyle {#2}} \atop \scriptstyle {#3} } \atop \scriptstyle {#4} }} % \newcommand\littlesum{ {\scriptscriptstyle \sum }} \newcommand\wt{\widetilde} %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% \begin{document} \setcounter{page}{1} \renewcommand{\thepage}{\roman{page}} \title{Compensations in small divisor problems} \author{ L. Chierchia and C. Falcolini\\ {\footnotesize Dipartimento di Matematica} \\ {\footnotesize Universit\`a di Roma ``Tor Vergata"}\\ {\footnotesize via della Ricerca Scientifica, 00133 Roma (Italy)} \\ {\footnotesize (Internet: chierchia@mat.utovrm.it and falcolini@mat.utovrm.it)} \thanks{The authors gratefully acknowledge helpful discussions with C.~Liverani.} } \maketitle \date{August 1994} \begin{abstract} \nin {\footnotesize Several small divisor problems arising in the perturbative theory of Hamiltonian and Lagrangian systems are considered. A general method that allows to prove compensations among the elementary contributions of the formal power series expansions associated to invariant surfaces is presented. } \end{abstract} {\footnotesize \tableofcontents} %******************************************** \newpage \setcounter{page}{1} \renewcommand{\thepage}{\arabic{page}} \section{Introduction} \label{sec:par1} \setcounter{equation}{0} % C.L.~Siegel \cite{S} was the first to solve a small divisor problem (linearization of germs of analytic functions). His solution was based on a ``direct method" consisting in computing the $k^{\rm th}$ coefficient of the formal power series solution and estimating such coefficient by a constant to the $k^{\rm th}$ power. The solution of small divisor problems arising in perturbative series in the theory of Hamiltonian systems has been one of the major contributions to dynamical systems of this century: the techniques employed for such a task, started by A.N.~Kolmogorov in the early fifties and developed in the early sixties by V.I.~Arnold and J.K.~Moser, are known as ``Kolmogorov--Arnold--Moser (KAM) theory" (see \eg \cite{A1} and references therein). Such methods show {\sl indirectly} the convergence of the formal expansions of the perturbative series (longly before considered by astronomers and especially by H.~Poincar\`e \cite{Po}). In 1988 H.~Eliasson proposed a direct proof of the convergence of formal expansions in the Hamiltonian context (\cite{E1}, \cite{E2}, \cite{E3}). More recently direct proofs (in the Hamiltonian context) have been reconsidered in several papers: \cite{CF1}, \cite{Ga1}, \cite{Ga2}, \cite{Ga3} \cite{GG}, \cite{Ge1}, \cite{Ge2}; a small divisor problem arising in elliptic systems of PDE's has been solved by similar methods in \cite{CF2}. \giu In this paper we consider several small divisor problems and discuss a general method that allows to prove compensations among ``elementary" contributions to the $k^{\rm th}$ order of the associated formal expansion. Such compensations, modulo technical estimates (which have been thoroughly discussed in \cite{CF1}), yield new proofs of the {\sl convergence} of the formal power series. \giu The starting point in this subject is a formal power series whose coefficients are {\sl bona fide} functions determined by linear recursive equations. ``Recursive" means that the $k^{\rm th}$ coefficient can be expressed (typically by Taylor's expansions and by inversion of suitable differential operator $D$) in terms of the derivatives of a given function (``the Hamiltonian" or ``the Lagrangian") and in terms of the $h^{\rm th}$ coefficients with $h0$); the second sum runs over all possible functions assigning to each vertex $v\in V$ of a rooted tree $T$ an integer vector $\a_v\in\zN$ with the constrain $\sum_{v\in V} \a_v=n$; the third sum runs over a suitable set of (possibly) vector--valued indices taking a finite number of values ($\# B<\io$); $\L$ is a complex vector depending on $T$, $\{\a_v\}_{v\in V}$, $\{\b_v\}_{v\in V}$ and on the Hamiltonian (or Lagrangian); finally $\g_v\in\real$ are {\sl divisors} that are described as follows. Rooted trees can be naturally equipped with a partial order: we say that $v'\le v$ if the path joining the root $r$ of $T$ with $v'$ contains $v$; (obviously $v'< v$ means $v'\le v$ and $v'\neq v$ and $r\ge v$ $\forall$ $v\in V$ \ie the root $r$ is the first vertex of the rooted tree $T$). Given a function $\a$, we define \beq{1.dl} \d_v\=\d_v(T,\a)\=\sum_{v'\in V:\ v'\le v} \a_v\ . \eeq The divisors $\g_v$, which may assume arbitrarily small values, are defined in terms of a real--valued function $\lg \cdot \rg$ which satisfies the {\sl Diophantine condition} \beq{1.dp} |\lg n \rg|^{-1} \le \g |n|^\t\ ,\qquad \forall\ n\in \zN\backslash\{0\} \eeq with suitable positive constants $\g,\t$.\footnote{ Typically, in dynamical systems, $\lg n \rg = \o \cdot n$ (where the dot denotes the standard inner product in $\rN$) but in other situations (\eg \cite{CF2}) the function $\lg \cdot \rg$ might be more complicate (\eg non linear); in the case $\lg n \rg = \o \cdot n$ it is well known that, if $\t>N-1$ , up to a set of Lebesgue measure zero, all $\o\in\rN$ satisfy \equ{1.dp} for some $\g$.} Then \beq{1.ga} \g_v\=\g_v(T,\a,\b)\=\cases{ \lg \d_v\rg^{-\s_v} & if $\d_v\neq 0$\cr 1 & if $\d_v=0$.\cr} \eeq where $\s_v$ is, say, the first component of the index $\b_v$ and takes value $0$, $1$ or $2$. In \equ{1.tr} only the second sum runs over an infinite set of indices: therefore we assume that there exist positive numbers $\xi,\xi',a>0$ such that, if we set \beq{1.co} a^k_{n}\= \frac{1}{k!}\sum_{T\in \cT^k_*} \sum_{ {\a:V\to \zN}\atop{\Sigma \a_v=n}} \sum_{\b:V\to B} |\L(T,\a,\b)| \prod_{v\in V} e^{\xi'|\a_v|} \ , \eeq then, for all $k$ and $n$ one has \beq{1.co'} a^k_{n}\le a^k e^{-\xi|n|}\ . \eeq In the models considered here, such an assumption is an immediate consequence of the well--posedness of the (formal) problem and of the analyticity assumptions on the Hamiltonian (Lagrangian). >From \equ{1.co} it follows at once that if one could bound the product of the divisors as\footnote{From now on we will adhere to the the common abuse of notation $v\in T$ in place of the more proper $v\in V(T)$ and also if $vv'=v'v$ denotes an edge of $T$, we shall denote $vv'\in T$ rather than the more proper $vv'\in E(T)$.} \beq{1.si} \prod_{v\in T} |\g_v|\le c^k_3 \ \prod_{T} (1+|\a_v|^b) \eeq for some $c_3 > 1$ and $b > 0$, then from \equ{1.ft}, \equ{1.co} and \equ{1.si} it would follow \beq{1.5} |Z^k_{n}| \le \frac{1}{k!}\sum_{T\in \cT^k_*} \sum_{ {\a:V\to \zN}\atop{\Sigma \a_v=n}} \sum_{\b:V\to B} |\L(T,\a,\b)| \ c^k_4 \ \prod_{v\in V} (1+|\a_v|^b) \le c^k_5 e^{-\xi |n|} , \eeq (for suitable $c_1 > 0$) leading to ``absolute" convergence (\ie convergence without compensations) of the formal expansion $Z$. Indeed Siegel's original proof is based on a similar argument, even though the set up is slightly different (and simpler). Technically Siegel's problem corresponds to $\th$ varying in a small (complex) ball so that Fourier series is replaced by Taylor series and $\a_v$ ranges over $\integer^N_+$: in such a case $\d_v\neq \d_{v'}$ whenever $v>v'$ and Siegel's method \cite{S} yields the estimates \equ{1.si}\footnote{For a detailed discussion, in the present language, of Siegel's methods see Appendix C of \cite{CF1}; for a different approach see \cite{Br}.}. The problem with $\a_v\in \zN$ is that one can have ``resonances" \ie $\d_v=\d_{v'}$ for $v>v'$ which may lead to obstinate repetitions of particularly small divisors. It is well known (see \eg \cite{CF1}, Appendix B) that, in general, one has, for arbitrarily large $k$, subfamilies $\cF_{\rm div}\subset \cT^k_*$ and a choice of $\bar \a$ and $\bar \b$ (depending only on the subfamily) such that, for suitable $\bar a,\bar b>0$, \beq{1.di} \frac{1}{k!}\sum_{T\in \cF_{\rm div}} \L(T,\bar \a,\bar \b) \ \ \prod_{v\in V} \g_v \ge \bar a^k k!^{\bar b}\ . \eeq Such families are obtained by taking {\sl chains of resonances} which are defined as follows. Given $T\in \cT^k_*$ and $\a$ (\ie $\{\a_v\}_{v\in T}$), a {\sl resonance} is % a subtree\footnote{When referred to trees the notation $T'\subset T$ will always mean ``$T'$ (unrooted) subtree of $T$". Other special conventions we are adopting are the following: a) for rooted trees the root (usually denoted $r$) may be identified by adding an extra edge $\eta r$ where $\eta$ is a symbol (not a vertex of the tree) sometimes called the ``earth"; with these positions one has, for trees, $\# E= \#V-1$ and, for rooted trees, $\# E= \#V$; b) the degree of a vertex $v$ is the number of edges $vv'$ incident with $v$ and if $v=r$ is the root, the edge $\eta r$ is included in the count; c) the degree of a subtree $T'\subset T$ is the number of edges connecting $T'$ with $T\backslash T'$; if $T$ is rooted and the root $r$ belongs to $T'$ the edge $\eta r$ must be included in the count.} % $R\subset T$ such that: i) $R$ is of degree two (\ie $R$ is connected to $T\backslash R$ by two edges); ii) if $u$ is the first vertex\footnote{Recall that $T$ is a rooted tree and hence partially ordered (the order being such that the first vertex of $T$ is always its root).} in $R$ and $z$ is the first vertex following $R$, then $\d_u=\d_z\neq 0$; iii) $R$ cannot be disconnected, by removal of one edge, into two subtrees of degree two satisfying i) and ii). It will be important to consider different choices of the index $\b$ (\ie $\{\b_v\}_{v\in T}$): in particular given a resonance $R$ and given $\b$ we call {\sl order of the resonance} the number (see \equ{1.ga}) $\s\=\s_u$ ($u$ being the first vertex of $R$). A {\sl chain of resonances} is a maximal series of resonances $R_1,...,R_h$ with $R_i$ adjacent\footnote{\ie connected by one edge.} to $R_{i+1}$; given a choice of $\b$, the {\sl order of the chain} is defined to be $\bar \s\=\s_1+\cdots +\s_h$ where $\s_i\=\s_{u_i}$, $u_i$ being the first vertex of $R_i$. >From these positions it follows that if $C$ is a chain of order $\bar \s$, if $n=\d_z$ where $z$ is the first vertex following the chain (\ie following the last resonance, which by convention will be $R_h$), then \beq{1.ch} \prod_{v\in C} \g_v = \lg n \rg^{-\bar \s} \prod_{ {v\in C}\atop {v\neq u_i}} \g_v \eeq (where $v\in C$ means $v\in \bigcup_i V(R_i)$). The examples for which \equ{1.di} holds are based on chains with $\bar \s\sim k$ and $|\lg n \rg|\sim k^{-1}$. This phenomenon may be cured through compensations. To be more precise we introduce the notion of ``compensable chain". Consider a chain $C=(R_1,...,R_h)$, ($h\ge 1$) and fix the indices $\b_v$. Let, as above, $u_i$ be the first vertex in $R_i$, let $R_i$ be connected to $R_{i+1}$ by the edge $w_i u_{i+1}$ with $w_i\in R_i$ and $u_1\ge w_1>u_2\ge ...\ge w_h$; let $z$ be the first vertex following the chain (\ie following $w_h$) and $n\=\d_z$; and, finally, let $P_i$ be the path joining $u_i$ with $w_i$. Consider the following function of $x\in \complex$ \beq{1.pi} \pi_C(x;T,\a,\b)\=\prod_{i=1}^h \pi_{R_i}(x;T,\a,\b)\ , \qquad \pi_{R_i}(x) \=\prod_{v\in (R_i\backslash P_i)} \g_v\ \prod_{{v\in P_i}\atop{v\neq u_i}} \bar \g_v(x)\ , \eeq where, if $u_i=w_i$ (\ie $P_i=\{u_i\}$), the product over $P_i$ is absent, while if $v\in P_i\neq \{u_i\}$ we set \beq{1.pi'} \bar \g_v(x)\= \lg \sum_{ {v'\in R_i}\atop {v'\le v}} \a_{v'} + x n \rg^{-\s_v} \ . \eeq Thus, $\g_v=\bar \g_v (1)$ and \beq{1.10} \prod_{v\in C} \g_v= \lg n \rg^{-\bar \s} \pi_C(1)\ . \eeq We say that {\sl a chain $C\subset T\in \cT^k_*$ {\rm (\ie $C=(R_1,...,R_h)$ with $R_i\subset T$)} is compensable} if there exists a family of trees $\cF_C\subset \cT^k_*$ whose elements $T'$ have $C$ as common chain of resonance, and, for each $T'$, there exists a choice of indices $\b'\=\b'(T')$, such that the function\footnote{ In other words, the elements $T'$ of $\cF_C$ are obtained from $T$ and $C$ by (possibly) changing the edges connecting $T\backslash C$ with $R_1$, $R_1$ with $R_2$,...,$R_h$ with $T\backslash C$, and by (possibly) changing the values of the indices $\b$ on $C$.} \beq{1.11} \bar \pi_C(x)\=\sum_{T'\in \cF_C} \L(T',\a,\b')\ \pi_C(x;T',\a,\b') \eeq has a zero in $x$ of order at least $\bar \s-1$. >From the detailed estimates in \cite{CF1}, it follows easily that {\sl if in a small divisor problem all chains of resonances are compensable then the associated formal series \equ{1.1} is in fact convergent}. In the rest of the paper we shall prove the following statements. \giu {\bf P1)} Let $H=H_0(y)+\e H_1(x,y)$ be a real--analytic Hamiltonian with $y\in B({y_0})\subset \rN$, $x\in \tN$ and $(x,y)$ standard symplectic coordinates\footnote{This simply means that the Hamiltonian (evolution) equations are given by $\dot x= \dpr_y H$, $\dot y=- \dpr_x H$.}; here $B({y_0})$ denotes some $N$--sphere centered at $y_0$. Assume the following standard ``non degeneracy conditions" on the ``integrable part" $H_0$: i) the Hessian matrix $\dpr^2_y H_0(y_0)$ is invertible; ii) $\o\=\dpr_y H_0\in\rN$ is a Diophantine vector \ie there exists $\g,\t>0$ such that\footnote{Compare with \equ{1.dp}.} \beq{1.om} |\o \cdot n|^{-1} \le \g |n|^\t\ , \qquad \forall\ n\in \zN\backslash \{0\}\ . \eeq Then, quasi--periodic solutions with frequencies $\o$, \ie solutions of the form $(x(t),y(t))=Z(\o t)$ with $Z^0(\th)=(\th,y_0)$ and $Z^k:\th\in \tN \to Z^k(\th)\in \real^{2N}$, satisfy the equations \beq{1.h1} D Z = J \dpr H\Big( Z(\th) \Big)\ , \eeq where $D\=\o\cdot \dpr_\th$, $J$ is the standard symplectic matrix $\pmatrix{0 & I\cr -I & 0\cr}$ and $\dpr$ is the gradient $\dpr_{(x,y)}$ with respect to the variables $(x,y)$. Then it is well known\footnote{See \cite{Po} or Appendix B of \cite{CF1}.} that there exists a unique formal solutions $Z\sim \sum_{k\ge 0} \e^k Z^k$ of \equ{1.h1} with the normalization condition \beq{1.av} \ig_{\tN} \p_1\circ Z^k d\th= 0\ \quad (k\ge 1) , \eeq where $\p_1$ is the projection onto the first coordinates: $\p_1(x,y)=x$. We can then prove \thm{thm1} There exists a tree expansion \equ{1.tr} for $Z$ such that all chains of resonances are compensable. \ethm % {\bf P2)} The formulation for the Lagrangian case is very similar. We let $L\=L_0(y)+\e L_1(x,y)$ be a real--analytic Lagrangian\footnote{The Lagrangian (evolution equations) are $\frac{d}{dt} \dpr_y L(x,\dot x)= \dpr_x L(x,\dot x)$.} with $L_i$ satisfying the same hypotheses assumed above for $H_i$ and $\o$, except that now $\o \= y_0$. Quasi--periodic solutions $x(t)=Z(\o t)=\sum_{k\ge 0} \e^k Z^k(\o t)$, where now $Z^0(\th)=\th$ and $Z^k:\th\in \tN\to Z^k(\th)\in \rN$, satisfy the equations \beq{1.h2} D \ \dpr_y L\big(Z(\th), DZ(\th)\big)= \dpr_x L\big(Z(\th), DZ(\th)\big) \ ,\qquad (D\=\o\cdot\dpr_\th)\ . \eeq Mimicking the proof of Appendix B in \cite{CF1} one can show that there exists a unique formal solution $Z\sim \sum_{k\ge 0} \e^k Z^k$ of \equ{1.h2} with the normalization condition as in \equ{1.av} but without the projection $\p_1$. Then, {\sl Theorem~\ref{thm1} holds also in this case}. \giu {\bf P3)} Maximal invariant tori correspond to analytic continuation (in $\e$) of unperturbed tori having ``all frequencies excited" \ie tori run by linear flow $t\to \o t$ with $\o\=\dpr_y H_0(y_0)$ rationally independent over $\zN$ and $N=\#$ of degrees of freedom. More delicate is what happens to unperturbed ``resonant" tori\footnote{Here, the word ``resonant" which is related to the ``resonances" of celestial mechanics, is used in a technically different meaning from the one used throughout this paper.} \ie invariant unperturbed tori for which there exist $n\in \zNn$ such that $\o\cdot n=0$. We shall argue (see next item P4)) that {\sl in general, such tori are not analytically continuable for $\e\neq 0$}. Nevertheless we will show, under suitable conditions, how to construct, with the (intrinsically analytic) methods outlined above, lower dimensional invariant (unstable) tori when $\e\neq 0$. For simplicity, we shall discuss only a special case of ``second order" nearly integrable Hamiltonian systems with Hamiltonian function \beq{1.12} H\= \frac{1}{2} y^2 + \frac{1}{2} p^2 + \e f(x,q) \eeq where $(x,y)$ are symplectic variables as above (\ie with $x\in \tN$) and so are $(q,p)$ with $q\in \tM$; hence the number of degrees of freedom is $N+M$ and we shall study $N$--dimensional invariant tori. The specification ``second order" is due to the particular form in which can be cast the Hamilton equations for $H$: \beq{1.13} \ddot x= -\e \dpr_x f\ ,\qquad \ddot q = -\e \dpr_q f\ . \eeq For $\e=0$, $N$--dimensional invariant tori (up to a trivial linear and symplectic change of coordinates\footnote{A transformation of (standard) symplectic coordinates $(q,p)$ (of a $2d$--dimensional phase space) is called {\sl symplectic} if it preserves the (standard) two form $\sum_{i=1}^d dq_i\wedge dp_i$.}) are spanned by solutions of the form $(y,p)=(\o,0)$, $(x,q)=(x_0+\o t, q_0)$. % >From classical transformation theory it follows that (if $\o$ satisfies \equ{1.om}) the evolution equations of $H$ in \equ{1.12} are equivalent to the evolution equations for a Hamiltonian of the form $\frac{y^2}{2} +\frac{p^2}{2} +\e f_0(q) +o(\e)$ where $f_0$ is the average (w.r.t. $x$) over $\tN$ of $f$. Motivated by this observation, we consider the Hamiltonian \beq{1.12'} \frac{1}{2} y^2 + \frac{1}{2} p^2 + \e f_0(q) + \e^{2} \tilde f(x,q) \= H_0(q,y,p;\e) + H_1(x,q;\e) \ ,\qquad H_1\=\e^2 \tilde f\ . \eeq Even though $H_0$ is not, in general, integrable, {\sl if $q_0$ is a critical point for $f_0$}, $H_0$ still admits the invariant $N$--torus $\cT_0$ spanned by $(y,p)=(\o,0)$, $q=q_0$ and $x= x_0+ \o t$ and we want to study the persistence of such torus for the full Hamiltonian. To attack the problem perturbatively, we introduce a new analyticity parameter $\m$ with respect to which we shall make formal (and eventually convergent with radius of convergence greater than one) power series expansion and consider the Hamiltonian $H_0+\m H_1$ (so that for $\m=1$ we recover the Hamiltonian \equ{1.12'}). We also make the following {\sl hyperbolicity assumption}: we assume that $q_0$ is such that the matrix $\e \dpr^2_q f_0(q_0)$ is {\sl positive definite}. Under these hypotheses it is easy to prove that there exists a (unique) formal expansion \beq{1.15} Z\=(Z_1,...,Z_d)\sim \sum_{k\ge 0} Z^k(\th;\e) \m^k\ ,\qquad \th\in \torus^N\ ,\qquad d=2(N+M) \eeq such that $t\to Z(\o t)$ is a formal quasi--periodic solution of the Hamiltonian equations governed by $H_0+\m H_1$ and such that the set $\{ Z^{0}(\th):\th\in \tN\}$ coincides with the unperturbed torus $\cT_0$ introduced above. Uniqueness is achieved by requiring \equ{1.av} where $\p_1$ denotes the projection onto the $x$ variable. Then, {\sl Theorem~\ref{thm1} holds also in this case}. >From this result, as already remarked, it follows that the formal power series is actually convergent but, what is more interesting in this case, one can show that for $\e\neq 0$ small enough, {\sl the radius of convergence (in $\m$) is greater than one} so that the set $\{ Z(\th):\th\in \tN, \m=1\}$ is an invariant $N$--torus for the Hamiltonian \equ{1.12'}. We finally mention that from \cite{Gr} it follows easily that these tori are {\sl whiskered} in the sense of \cite{A2}. \giu {\bf P4)} Consider again the Hamiltonian \equ{1.12}. Indeed, one can consider formal power series in $\e$ and one can show that {\sl if $q_0$ is a non degenerate critical point of the $x$--average of $f$ then there exists a (unique) formal power series \equ{1.1} (with $d=2(N+M)$) such that $t\to Z(\o t)$ ($\o$ satisfying \equ{1.om}) is a formal quasi--periodic solution for \equ{1.12} and the set $\{Z^{0}(\th):\th\in\zN\}$ coincide with the torus spanned by $y=\o$, $p=0$, $q=q_0$, $x= x_0 +\o t$.} Uniqueness, again, is enforced by the requirement \equ{1.av}. Also for such a formal series one can write down a tree expansion completely analogous to those referred to in P1)$\div$P3) above. However, we shall prove that there exist chains $C=(R_1,...,R_h)$ such that if $\cF_0$ denotes the family of {\sl all} trees with chain $C$ then \beq{1.16} \sum_{T'\in\cF_0} \L (T')\ \p_C(0)\neq 0\ . \eeq In view of this fact it seems natural to {\sl conjecture that in the present case the formal power series $Z$ is divergent}. \section{Weighted trees} \label{sec:par3} \setcounter{equation}{0} % Here we describe the tree family $\cT^k_*$, which appears in the basic formula \equ{1.tr}. \giu For the models P3) and P4) introduced in \S~\ref{sec:par2}, $\cT^k_*$ is simply the family of all labeled, rooted trees with $k$ vertices\footnote{The basic terminology can be found in any introductory book on graphs or in Appendix~A of \cite{CF1}.}, which we will denote by $\cT^k$. In this case, as is well known (see \eg \cite{Bo}), $\# \cT^k = k^{k-1}$ and \equ{1.ft} is clearly satisfied. \giu To treat the cases P1) and P2) one has to distinguish, in the Taylor's expansion, the contributions coming from $H_0$ and $L_0$ from those coming from $H_1$ and $L_1$. To do this we introduce the following family of ``weighted trees". Given a rooted (unlabeled) tree $T$ we call a function of the vertices of $T$ \beq{3.1} \chi : v\in V(T) \to \chi_v\in\{0,1\} \eeq a {\sl weight function}. A {\sl weighted rooted tree} is a couple $(T,\chi)$ with $T$ a rooted tree and $\chi$ a weight function. We now denote $\widetilde \cT_k$ the set of weighted rooted trees satisfying \beq{3.2} (i) \quad \deg v\le 2 \ \implies \ \chi_v=1\ ,\qquad (ii) \quad \sum_{v\in T} \chi_v = k\ . \eeq Notice that, in particular, all final vertices (\ie vertices of degree 1) have weight 1 and that, for any $T\in \widetilde \cT_k$, $\# V(T) \le 2k-1$ as is easy to verify\footnote{ Let $V_i=\{v\in V:\chi_v=i\}$, and let $k_i=\#V_i$, so that $k=k_1$. It is well known (see, \eg, \cite{Bo} and recall our convention on degree of the root) that $\sum\limits_{V_0} \deg v + \sum\limits_{V_1} \deg v= 2(k_0+k_1)-1$. If $v\in V_0$ then $\deg v\ge 3$, thus the sum over $V_0$ can be bounded from below by $3 k_0$ while the sum over $V_1$ can be bounded from below by $k_1$. This leads to $k_0\le k_1-1$ which is the claim. Such inequality is optimal.}. We now define the class $\cT_k$ of {\sl labeled, weighted rooted trees} obtained from $\widetilde \cT_k$ by labelling with $k$ different labels the $k$ vertices with weight 1. \giu In cases P1) and P2) we let $\cT^k_*=\cT_k$; it is easy to check that \equ{1.ft} holds also in this case\footnote{ Since (\cite{Bo}) $\# \widetilde \cT^h\le 4^h$ and $\# V(T)\le 2k-1$ for any $T\in \widetilde \cT_k$ one sees that $\# \widetilde \cT_k\le 4^{2k}$. Since the labels are attached to $k$ vertices, it is $\# \cT_k\le k! 4^{2k}$.}. \giu The relation between Taylor's formula and trees may be based on the following operation $*$ (that we now briefly discuss for the case of $\cT_k$; for the case $\cT^k$ see Appendix~B of \cite{CF1}). If $T\in\cT_k$ we denote by $\widetilde T$ the tree in $\widetilde \cT_k$ obtained by removing the labels from $T$; we also denote $T^0$ (or $\widetilde T^0$) the unrooted tree obtained from $T$ (or $\widetilde T$) by removing the edge $\eta r$ \ie by not distinguishing any more the root from the other vertices; finally if $T$ (or $\widetilde T$) is an unrooted labeled (or unlabeled) tree and $r$ is one of its vertices, we denote by $T_r$ (or $\widetilde T_r$) the rooted tree obtained by adding the edge $\eta r$ (\ie by decreeing that $r$ is the root). Let $\chi\in \{0,1\}$ and $\ell\ge 1$, let $h_i$ be $\ell$ positive integers such that $h_1+\cdots+h_\ell=k-\chi$ and pick trees $T_i\in\widetilde \cT_{h_i}$. We can form a tree $T\in \widetilde \cT_k$ with root $r$ ($r$ being a vertex different from the vertices of $T_i$, $\forall$ $i$) of weight $\chi_r\=\chi$ by setting \beq{3.4} T\=T_{h_1}*\cdots *T_{h_\ell}\= \Big(T_{h_1}^0\cup\cdots\cup T^0_{h_\ell} \cup \{r\}+ \sum_{i=1}^\ell rr_i\Big)_r \eeq where $r_i$ is the root of $T_i$ (and summing an edge $e$ to a tree $S$ means, obviously, to add $e$ to $E(S)$). Then, one has the following \pro{pro:1} Let $F$ be a complex valued function defined on trees in $\widetilde \cT_k$ (for any $k$). Then \beq{3.5} \frac{1}{k!} \sum_{T\in \cT_k} F(\widetilde T) = \sum_{\chi=0,1} \sum_{\ell=2-\chi}^{k-\chi} \frac{1}{\ell!} \ \sum_{h_1+\cdots+h_\ell=k-\chi}\prod_{i=1}^\ell \frac{1}{h_i!} \ \sum_{ T_i\in \cT_{h_i}} F(\widetilde T_{h_1}*\cdots * \widetilde T_{h_\ell})\ . \eeq \epro For the proof we refer to \cite{CF1} (Corollary~B.1 of Appendix~B)\footnote{In \cite{CF1} it is treated the case of $\cT^k$ (the $*$ operation in $\widetilde \cT^k$ is defined as above, replacing $\chi$ systematically by 1); adapting the proof to $\cT_k$ is a trivial exercise.} . % \begin{figure}[hbt] \begin{center} \begin{picture}(350,80) \thicklines \put(5,20){\begin{picture}(30,60) \put(0,40){\circle{8}} \multiput(0,40)(15,0){3}{\circle*{3}} \multiput(-2,48)(15,0){3}{{\scriptsize $1$}} \put(0,40){\line(1,0){30}} \end{picture}} \put(60,20){\begin{picture}(30,60) \put(15,40){\circle{8}} \multiput(0,40)(15,0){3}{\circle*{3}} \multiput(-2,48)(15,0){3}{{\scriptsize $1$}} \put(0,40){\line(1,0){30}} \end{picture}} \put(115,20){\begin{picture}(30,60) \put(0,40){\circle{8}} \multiput(0,40)(15,0){2}{\circle*{3}} \put(0,40){\line(1,0){15}} \put(-2,48){{\scriptsize $1$}} \put(13,48){{\scriptsize $0$}} \put(35,50){{\scriptsize $1$}} \put(35,28){{\scriptsize $1$}} \put(30,50){\circle*{3}} \put(15,40){\line(3,2){15}} \put(30,30){\circle*{3}} \put(15,40){\line(3,-2){15}} \end{picture}} \put(170,20){\begin{picture}(30,60) \put(15,40){\circle{8}} \multiput(0,40)(15,0){2}{\circle*{3}} \put(0,40){\line(1,0){15}} \put(-2,48){{\scriptsize $1$}} \put(13,48){{\scriptsize $0$}} \put(35,50){{\scriptsize $1$}} \put(35,28){{\scriptsize $1$}} \put(30,50){\circle*{3}} \put(15,40){\line(3,2){15}} \put(30,30){\circle*{3}} \put(15,40){\line(3,-2){15}} \end{picture}} \put(225,20){\begin{picture}(45,60) \put(15,40){\circle{8}} \multiput(0,40)(15,0){4}{\circle*{3}} \put(-2,48){{\scriptsize $1$}} \put(13,48){{\scriptsize $0$}} \put(28,48){{\scriptsize $1$}} \put(43,48){{\scriptsize $1$}} \put(0,40){\line(1,0){45}} \end{picture}} \put(295,20){\begin{picture}(45,60) \put(15,40){\circle{8}} \multiput(0,40)(15,0){3}{\circle*{3}} \put(0,40){\line(1,0){30}} \put(-2,48){{\scriptsize $1$}} \put(13,48){{\scriptsize $0$}} \put(29,48){{\scriptsize $0$}} \put(45,50){\circle*{3}} \put(30,40){\line(3,2){15}} \put(45,30){\circle*{3}} \put(30,40){\line(3,-2){15}} \put(50,50){{\scriptsize $1$}} \put(50,28){{\scriptsize $1$}} \end{picture}} \parbox{12cm}{\caption{\label{fig:wt1}The elements of $\widetilde \cT_3$. (The numbers $0$, $1$ are the values of the index $\chi$ for the corresponding vertices and the encircled vertex is the root of the tree).}} \end{picture} \end{center} \end{figure} % % \begin{figure}[hbt] \begin{center} \begin{picture}(350,80) \thicklines \put(115,20){\begin{picture}(30,60) \put(0,40){\circle{8}} \multiput(0,40)(20,0){2}{\circle*{3}} \multiput(-2,48)(20,0){2}{{\scriptsize $1$}} \put(-4,25){$u_2$} \put(16,25){$u_1$} \put(0,40){\line(1,0){20}} \end{picture}} \put(170,20){\begin{picture}(30,60) \put(0,40){\circle{8}} \multiput(0,40)(20,0){2}{\circle*{3}} \multiput(-2,48)(20,0){2}{{\scriptsize $1$}} \put(-4,25){$u_1$} \put(16,25){$u_2$} \put(0,40){\line(1,0){20}} \end{picture}} \put(225,20){\begin{picture}(30,60) \put(0,40){\circle*{3}} \put(0,40){\circle{8}} \put(-2,48){{\scriptsize $0$}} \put(13,55){{\scriptsize $1$}} \put(13,35){{\scriptsize $1$}} \put(20,50){$u_1$} \put(20,30){$u_2$} \put(15,50){\circle*{3}} \put(0,40){\line(3,2){15}} \put(15,30){\circle*{3}} \put(0,40){\line(3,-2){15}} \end{picture}} \parbox{12cm}{\caption{\label{fig:wt2}The elements of $\cT_2$ ($u_1$ and $u_2$ are the labels of $\cT_2$).}} \end{picture} \end{center} \end{figure} % \section{Compensations I (Maximal Hamiltonian tori)} \label{sec:par4} \setcounter{equation}{0} \subsection{Tree expansion} \label{subsec:4.1} % Consider the model introduced in P1) of \S~\ref{sec:par2} and recall that there exists a unique formal solutions $Z\sim\sum_{k\ge 0} \e^k Z^k$ satisfying \equ{1.h1} and \equ{1.av}. Denote by $Z^{(1)k}$ the $x$--component (\ie the first $N$ components) of the vector $Z^k$ and by $Z^{(2)k}$ the $y$--component; consistently, let $\dpr^{(1)}\=\dpr_x$ and $\dpr^{(2)}\=\dpr_y$. We also let \beqno [\ \cdot\ ]_k\= \frac{1}{k!} \frac{d^k}{d\e^k}\ (\cdot)|_{\e=0} \eeqno denote the $k^{\rm th}$ order operator which to a (possibly formal) power series $a\sim \sum a_k \e^k$ associates its $k^{\rm th}$ order coefficient: $[ a ]_k\= a_k$. Finally let \beq{4.0} A \= \dpr^2_y H_0(y_0)\ . \eeq With these definitions, we can rewrite \equ{1.h1} as \beq{4.1} D Z^{(1)k} = A Z^{{(2)} k} + [\dpr^{(2)} H_0]_k^{(k-1)} + [\dpr^{(2)} H_1]_{k-1}^{(k-1)}\ , \qquad D Z^{{(2)} k} = - [\dpr^{(1)} H_1]_{k-1}^{(k-1)}\ , \eeq where the suffix $\phantom{[\cdot]}^{(k-1)}$ means that the argument of the function within square brackets is, for $k\ge 2$, the polynomial in $\e$ of degree $(k-1)$ given by \beq{4.pol} x=\th+\sum_{h=1}^{k-1} \e^h Z^{{(1)} k}\ , \qquad y= y_0+ \sum_{h=1}^{k-1} \e^h Z^{{(2)} k}\ , \eeq and, for $k=1$, is $(x,y)=(\th,y_0)$. Since the average of $[\dpr^{(1)} H_1]_{k-1}^{(k-1)}$ vanishes (as is clear from the second of \equ{4.1}) we can apply to it the operator $D^{-1}$ obtained by inverting the constant coefficient operator\footnote{If $f$ is a (smooth) function on $\tN$ with vanishing mean value we denote by $D^{-1} f$ the unique solution with vanishing mean value of the equation for $g$: $D g = f$. Expanding in Fourier series one has $g(\th)=D^{-1}f(\th)= -i \sum\limits_{n\in\zNn} \frac{f_n}{\o\cdot n} \exp({i n\cdot \th})$, where $i=\sqrt{-1}$: the inversion of $D$ introduces the small divisors.} $D$ and rewrite \equ{4.1} in a more compact way as \beq{4.2} D Z^{(\r)k}= (2-\r) A Z^{(2)k}+ \sum_{\chi=0,1} (-1)^{3-\r} [\dpr^{(3-\r)} H_\chi]^{(k-1)}_{k-\chi}\ , \qquad (\r=1,2)\ . \eeq Notice that while, by \equ{1.av}, the average of $Z^{(1)k}$ vanishes, the average of $Z^{(2)k}$ can be read (for $\r=1$) from \equ{4.2} by integrating over $\tN$: \beq{4.3} \ig_{\tN} Z^{(2)k}=- A^{-1} \sum_{\chi=0,1} \ig_{\tN} [\dpr^{(2)} H_\chi]^{(k-1)}_{k-\chi}\ . \eeq Taking the $n$--Fourier coefficient of \equ{4.2}, \equ{4.3} one gets \beq{4.4} Z^{(\r)k}_n= \sumsud{\s\in \{0,1,2\}^*}{\chi\in\{0,1\} } \lg n\rg^{-\s}\ \Big\{ [D^{(\s,\r)} H_\chi]_{k-\chi}^{(k-1)}\Big\}_n\ ,\qquad \big(\lg n\rg\= \o\cdot n\big)\ , \eeq where $D^{(\s,\r)}$ is the vector--valued operator\footnote{ For example the $j^{\rm th}$ component of $D^{(2,1)}$ is given by $D_{j}^{(2,1)}=\sum\limits_{s=1}^N \frac{\dpr^2 H_0}{\dpr y_j\dpr y_s}(y_0) \frac{\dpr}{\dpr x_s}$.} \beq{4.5} D^{(\s,\r)}\=(-1)^{1-\s\r} \ i^{-\s}\ A^{\s-1}\ \dpr^{(4-\s-\r)}\ , \eeq and the $*$ attached to the range of $\s$ means that the following {\sl constraints} have to be satisfied: $\s+\r\in \{2,3\}$ and $\s=0$ $\iff$ $n=0$ in which case we adopt the convention that $\lg n\rg^\s=0^0\=1$. Notice that $D^{(\s,\r)} H_\chi=0$ if $\chi=0$ and $\s+\r=3$ (as in such a case $\dpr^{(4-\s-\r)}=\dpr_x$ and $H_0$ is independent of $x$). \giu We shall now use Taylor's formula in the following form. If $f:x\in \real^m \to f(x)\in \real$ is a $C^\io$ function and if $a(\e)\sim\sum_{s\ge 1} \e^s a^{(s)}$ is a $\real^m$--valued (possibly formal) power series, then \beq{4.6} f\big(a(\e)\big) \sim f(0) + \sum_{h\ge 1} \e^h \sum_{\ell=1}^h \frac{1}{\ell!} \sum_{{h_1+\cdots+h_\ell=h}\atop{1\le h_i\le \ell}} \sum_{{j_1,...,j_\ell}\atop{1\le j_i\le m}} \frac{\dpr^\ell f}{\dpr x_{j_1}\cdots\dpr x_{j_\ell}}(0)\ a^{(h_1)}_{j_1}\cdots a^{(h_\ell)}_{j_\ell}\ . \eeq Thus, if $z\=(x,y)$ and $z^{(1)}\=x$, $z^{(2)}\=y$, by \equ{4.4} and \equ{4.6} we get for the $j^{th}$ component of the vector $Z^{(\r)k}_n$, and for $k\ge 2$, \beqa{4.7} && Z^{(\r)k}_{nj}= \\ && \sum_{ {\s\in \{0,1,2\}^*}\atop{\chi\in \{0,1\}}} \sum_{\ell=2-\chi}^{k-\chi} \frac{1}{\ell!} \sumsut{1\le h_i\le k-1}{\Sigma h_i=k-\chi}{(1\le i\le \ell)} \sumsut{n_i\in\zN}{\Sigma n_i=n}{(0\le i\le \ell)} \sum_{{\r_1,...,\r_\ell}\atop{\r_i\in \{1,2\}}} \sum_{ {j_1,...,j_\ell}\atop{1\le j_i\le N}} \lg n\rg^{-\s} \Big\{ \frac{ \dpr^\ell\ D_{j}^{(\s,\r)} H_\chi}{\dpr z^{(\r_1)}_{j_1} \cdots \dpr z^{(\r_\ell)}_{j_\ell}} \Big\}_{n_0} \prod_{i=1}^\ell Z^{(\r_i)h_i}_{n_i j_i}\nonumber\ , \eeqa where the derivatives of $H_\chi$ are evaluated at $(\th,y_0)$ and then one takes the $n_0$--Fourier coefficient (with respect to $\th$); for $k=1$, since $[D^{(\s,\r)} H_0]_1^{(0)}=0$, one has the simple formula \beq{4.7'} Z^{(\r)1}_{nj}= \sum_{\s\in \{0,1,2\}^*} \lg n\rg^{-\s} \{D^{(\s,\r)}_j H_1\}_n\ . \eeq We are ready to prove the following {\sl tree expansion formula} (recall \equ{1.ga}, \equ{1.dl}) \beq{4.8} Z^{(\r_0)k}_{nj_0} = \frac{1}{k!} \sum_{T_r\in\cT_k} \quad \sum_{\a:V\to \a_v\in \zN \atop \Sigma_v \a_v = n}\quad \sumsud{\b:V\to\b_v\in B}{\r_r=\r_0} \quad \sumsud{j:V\to j_v\in\{1,...,N\}}{j_r=j_0}\ \prod_{v\in T_r} \big\{ \L_v(T_r,\b) H_{\chi_v} \big\}_{\a_v}\ \prod_{v\in T_r} \g_v \ , \eeq where: the index set $B$, which depends on the function $\d_v$, is defined as \beqa{4.B} & B & \=\Big\{\b=(\s,\r):\ \s\in\{0,1,2\};\ \r\in\{1,2\};\ {\rm s.t.}\nonumber\\ && \qquad \s+\r\in\{2,3\}\ ,\s=0\ \iff \ \d_v= 0\Big\}\ ; \eeqa the scalar operator $\L_v(T_r,\b)$ is given by \beq{4.9} \L_v(T_r,\b)\= D^{(\s_v,\r_v)}_{j_v} \prod_{v'\in \calN_v} \dpr_{j_{v'}}^{(\r_{v'})}\ , \qquad \calN_v\=\{v'v>w'$. % We shall classify resonances by assigning to them an integer $s\=s_R\in\{0,1,2\}$, which we shall call {\sl index of the resonance $R$}. Then, to each resonance $R\subset T$ we shall associate a family $\cF_R$ of trees $T'$ obtained by (possibly) changing the edges connecting $R$ with $T\bks R$ and choosing a suitable set of indices . The family $\cF_R$ and the index $\b'\=\b'(T')$ will be chosen so that (recall \equ{1.pi}) \beq{4.12} \sum_{T'\in\cF_R} \L(T',\a,\b')\ \pi_R(x;T',\a,\b')=O(x^s) \eeq and the family $\cF_C$ will simply be given by \beq{4.13} \cF_C=\bigcup_{i=1}^h \cF_{R_i} \ . \eeq Let $R$ be a resonance and let $u$ be its first vertex and $w'$ the first vertex following $R$. We define the {\sl index of $R$} as \beq{4.in} s_R\=\s_u+\r_u-\r_{w'}\ . \eeq Thus if $C=(R_1,...,R_h)$ is a chain of resonances and if $\bar \s$ denotes its order (see \S~\ref{sec:par2}), then \beq{4.14} \sum_{i=1}^h s_{R_i}= \bar \s +\r_{u_1} - \r_{z} \eeq where $z$ is the first vertex following $R_h$ (which, by convention, is the last resonance in the chain $C$). Hence, if \equ{4.12} holds, from \equ{4.13}, \equ{4.14}, \equ{1.pi} and the definition of $\L$ \equ{4.11'} it follows easily that \beq{4.15} \sum_{T'\in \cF_C} \L(T',\a,\b')\ \pi_{C}(x;T',\a,\b')= O(x^{\bar \s +\r_{u_1} - \r_{z}}) \eeq which implies that the chain $C$ is compensable. \giu Let $R\subset T$ be a resonance and let $\hat u u$ and $w {w'}$ be the edges connecting $R$ with $T\bks R$, with $\hat u>u\ge w>{w'}$ (hence $u,w\in R$). We proceed by constructing the family $\cF_R$. If $\r_{w'}\neq 1$ we set $\cF_R'=\{T\}$; if $\r_{w'}=1$ we let $\cF_R'$ be the family of all trees $T'$ obtained from $T$ by replacing the edge $w {w'}$ with the edge $\bar w {w'}$, as $\bar w$ varies in $R$. Hence, $T\in\cF_R'$ and if $\r_{w'}=1$, $\# \cF_R'=\# R$. If $\s_u+\r_u\neq 3$ we set $\cF_R''=\{T\}$; if $\s_u+\r_u=3$ (\ie $(\s_u,\r_u)=(1,2)$ or $(\s_u,\r_u)=(2,1)$) we let $\cF_R''$ be the family of all trees $T'$ obtained from $T$ by replacing the edge $\hat u u$ with the edge $\hat u\bar u$, as $\bar u$ varies in $R$. As above, $T\in\cF_R''$ and if $\s_u+\r_u=3$, $\# \cF_R'=\# R$. In the first case (\ie for $T'\in\cF_R'$) we do not modify the values $\b_v$ (\ie $\b'(T')\=\b$). In the second case (\ie for $T'\in \cF_R''$) we define $\b'\=\b'(T')$ as follows. If $v\notin R$ then $\b'_v=\b_v$. Recall that, by definition, the first vertex of $R$ considered as subtree of $T'$ is $\bar u$ while the first vertex of $R$ considered as subtree of $T$ is $u$. Let $\bar u\neq u$ (otherwise, obviously, we set $\b'\=\b$) and consider the path $P(u,\bar u)$ connecting $u$ with $\bar u$. The path $P$ will be formed by $p\ge 2$ ordered vertices that we denote $v_i$: the order is such that $v_1=u$, $v_p=\bar u$ and the edges of the path are $v_1v_2$,...,$v_{p-1}v_p$. We then set \beqa{4.be} & \b_{v_p}'\= \b_{v_1}\ , &\nonumber\\ & \b_{v_i}'\=\big( \s_{v_{i+1}}, 4-\s_{v_{i+1}}-\r_{v_{i+1}}\big)\ , &{\rm for}\ 1\le i\le p-1\nonumber\\ & \b_v'\=\b_v\ , &\forall\ v\notin P(u,v)\ . \eeqa It is easy to see that this definition is well posed (see also (ii) of the following Remark). Finally, we define \beq{4.fa} \cF_R\=\cF_R' \cup \cF_R''\ . \eeq \rem{rem:4.1} (i) If $s_R=0$, it is $\cF_R=\{T\}$ and $\b'\=\b$. \noindent (ii) The map $\b\to \b'$ is involutive: more precisely, if we denote by $\b'(\b;u,\bar u)$ the map defined in \equ{4.be} (definition which depends on the ordered path $P(u,\bar u)$) then $\b'(\b'(\b;u,\bar u);\bar u,u)=\b$. \noindent (iii) One might say that the families $\cF_R'$ and $\cF_R''$ are constructed ``going around" the resonance $R$ with a ``discrete curve" obtained by moving, respectively, the edge connecting $R$ with $w'$ and the edge connecting $\hat u$ with $R$. This interpretation might explain the name ``index" given to the quantity $s_R$: $\cF_R$ is obtained by ``going around" $R$ exactly $s_R$ times. \noindent (iv) (On the definition of $\cF_R'$) In practice, moving around the edge $\bar w z$ produces a factor proportional to $\a_{\bar w}$ in \equ{4.11'} coming from the $\dpr^{(\r_z)}=\dpr^{(1)}$ appearing in \equ{4.9} (as $z\in \calN_{\bar w}$). It is easy to see that, for $x=0$, summing $\L$ $\pi_R(x)$ over the family $\cF_R'$ produces a {\sl common} factor $\sum_{\bar w\in R}\a_{\bar w}$ which vanishes by definition of resonance. \noindent (v) (On the definition of $\cF_R''$) The idea is similar: one wants to produce a factor proportional to $\a_{\bar u}$ when moving around the ``first" edge $\hat u \bar u$ connecting $R$ with $\hat u$. But now the situation is more delicate as changing the connection $\hat u\bar u$ {\sl changes the order} in $R$ and, consequently, change the small divisors and also the structure of the derivatives (since both $\g_v$ and $\L_v$ depend on the order). Since one wants eventually to set $x=0$ and collect the factor $\sum_{\bar u\in R}\a_{\bar u}$ one sees the necessity of changing the values of the indices $\b$. In fact $\b'$ is defined in such a way that, when $x=0$, one can factor out the {\sl product} of divisors and the {\sl products} of the operators (derivatives) $\L_v$. \erem % With the above remarks in mind, it is not difficult to verify that \equ{4.12} holds, proving Theorem~\ref{thm1} in case P1). \section{Compensations II (Maximal Lagrangian tori)} \label{sec:par5} \setcounter{equation}{0} % Consider the model introduced in P2) of \S~\ref{sec:par2} for a real--analytic Lagrangian $L\= L_0(y)+\e L_1(x,y)$. Formal quasi--periodic solutions of the Lagrangian equations have the form $x(t)=Z(\o t)$ for an $\o$ satisfying \equ{1.om} and a function $Z(\th)$ which is the (unique) formal solution $Z\sim\sum_{k\ge 0} \e^k Z^k$, with $Z\in\real^N$, of \equ{1.h2}. The vector valued function $Z^k$ can be put in a form which is amazingly similar to the Hamiltonian case P1). \nin Let us denote by $Z^{(1)k}$ the full vector $Z^k \in \real^N$ and by $Z^{(2)k}$ the vector $DZ^k=DZ^{(1)k}$ (where $D\= \o\cdot\partial_\th$) and, as before, let $\dpr^{(1)}\=\dpr_x$, $\dpr^{(2)}\=\dpr_y$ and $A_L \= \dpr^2_y L_0(y_0)$. Equation \equ{1.h2} can then be written as \beq{5.1} A_L D Z^{{(2)} k} = - D[\dpr^{(2)} L_0]_k^{(k-1)} - D[\dpr^{(2)} L_1]_{k-1}^{(k-1)} + [\dpr^{(1)} L_1]_{k-1}^{(k-1)}\ . \eeq adopting the same convention used in \equ{4.1}: in particular the arguments of the functions within square brackets are as in \equ{4.pol}. As for the previous case (see \equ{4.4}), the $n$--Fourier coefficient of \equ{5.1} is \beq{5.2} Z^{(\r)k}_n= \sumsud{\s\in \{0,1,2\}^*}{\chi\in\{0,1\} } \lg n\rg^{-\s}\ \Big\{ [D_L^{(\s,\r)} L_\chi]_{k-\chi}^{(k-1)}\Big\}_n \eeq where $D_L^{(\s,\r)}$ is now the vector--valued operator \beq{5.3} D_L^{(\s,\r)}\=(-1)^{1-(\s+\r)} \ i^{-\s}\ A_L^{-1}\ \dpr^{(4-\s-\r)}\ , \eeq and $\{0,1,2\}^*$ is the same set of equation \equ{4.4}. \nin The Lagrangian problem is now in a form which is identical to the Hamiltonian case, except for the definition of $D_L^{(\s,\r)}$. Therefore the tree expansion formula for the component $j_0$ of the $n$--Fourier coefficient of $Z^k$, with fixed values of $\r_0$ and $j_0$ at the root $r$, is given again by \equ{4.8} with the only proviso of replacing $D^{(\s_v,\r_v)}_{j_v}$ in \equ{4.9} with $D^{(\s_v,\r_v)}_{L,j_v}$. \noindent Also, the families $\cF$ of trees for which there are compensations are found exactly as in the Hamiltonian case and we refer to \S~\ref{subsec:4.2} for details. \section{Compensations III (Lower dimensional tori)} \label{sec:par6} \setcounter{equation}{0} % Recall the notations of \S~\ref{sec:par2}, P3). Because of the particular form of the Hamiltonian, the formal solution\footnote{Recall that here $\e$ is a fixed real number different from zero, while the (complex) perturbation parameter appearing in the formal power series is $\m$.} $Z(\th,\m)$ is of the form $Z\=(X,Q,DX,DQ)$ where, as usual, $D\=\o\cdot \dpr_\th$, $\th\in\tN$ and $X\in \rN$, $Q\in \rM$. Denote \beq{6.2} Z^{(1)}\=X\ ,\quad Z^{(2)}\=Q\ ,\quad \dpr^{(1)}\=\dpr_x\ ,\quad \dpr^{(2)}\=\dpr_q\ ,\quad A\=\e \dpr^2_q f(q_0)\ (>0)\ . \eeq One checks immediately that the recursive equations for $Z^{(\r)k}$ ($\r=1,2$) are \beq{6.eq} \Big( -D^2+(\r-1) A\Big) Z^{(\r)k}= \sum_{\chi=0,1}\big[\dpr^{(\r)} H_\chi \big]^{(k-1)}_{k-\chi}\ ,\qquad \r=1,2 \eeq where the argument of the derivatives of $H_\chi$ is $x=\th$, $q=q_0$. From \equ{6.eq} one can see that the average (over $\tN$) of $Z^{(1)k}$ vanishes, while the average of $Z^{(2)k}$ is as in \equ{4.3} but with the minus sign replaced by a plus. Taking Fourier coefficients of \equ{6.eq} we get the analogous of \equ{4.4}, namely \beq{6.fo} Z^{(\r)k}_n= \sumsud{\s\in \{0,2\}^*}{\chi\in\{0,1\} } \lg n\rg^{-\s}\ \Big\{ [D^{(\r)}_n H_\chi]_{k-\chi}^{(k-1)}\Big\}_n\ ,\qquad \big(\lg n\rg\= \o\cdot n\big)\ , \eeq which differs from \equ{4.4} for the range of $\s$ and relative constraints: \beq{6.co} \s \in \{0,2\}^*\ \iff \ \s+\r\in\{2,3\}\ ,\qquad n=0 \ \implies \ \s=0\ , \eeq and for the definition of the vector valued operator $D^{(\r)}_n$: \beq{6.dr} D^{(\r)}_n\= \Big( \lg n\rg^2+ A\Big)^{1-\r} \ \dpr^{(\r)}\ . \eeq Notice that the components $D^{(\r)}_{nj}$ are $N$ if $\r=1$ and $M$ if $\r=2$; we therefore let \beq{6.N} N_1\=N\ ,\qquad N_2\= M \ \implies j\in\{1,...,N_\r\}\ . \eeq Recall that $A=\e \dpr^2 f_0(q_0)$ is positive definite, and so is $a+A$ for any $a\ge 0$; thus \beq{6.aa} \|(a+A)^{-1}\| \le \frac{b}{|\e|} \eeq for a suitable constant $b$ depending only on $\dpr^2 f_0(q_0)$. We also remark that, by \equ{6.co}, $\r=2$ implies $\s=0$ (hence no divisors) while if $\r=1$ then $\s=2$ and $n\neq 0$, in which case the divisors are $\lg n\rg^2$. In view of these remarks, we see that \equ{4.8} holds also in the present case provided we change the following items: $N$ in the fourth sum is replaced (see \equ{6.N}) by $N_{\r_v}$; the index set $B$ is defined as \beqa{6.B} & B & \=\Big\{\b=(\s,\r):\ \s\in\{0,2\};\ \r\in\{1,2\};\ {\rm s.t.}\nonumber\\ && \qquad \s+\r\in\{2,3\}\ ,\d_v=0\ \implies \ \s=0 \Big\}\ ; \eeqa finally the operator $\L_v$ now depends also on $\a$: $\L_v(T_r,\b)$ in \equ{4.8} is now replaced by \beq{6.L} \L_v(T_r,\a,\b)\= D^{(\r_v)}_{\d_v j_v} \prod_{v'\in \calN_v} \dpr_{j_{v'}}^{(\r_{v'})}\ . \eeq Also \equ{4.11'} is readily adapted replacing $N$ by $N_{\r_v}$ and $\L_v$ with \equ{6.L}. We can proceed to define the families of trees $\cF$ and relative indices $\b'$ which exhibit compensations. Given a resonance $R$ (and a choice of $\a$ and $\b$), we define the index of $R$ and the family $\cF_R''$ and relative indices $\b'$ exactly in the same way we did in \S~\ref{subsec:4.2}. Also the family $\cF_R'$ is defined in the same way but the relative indices $\b'$ are now defined in a slightly different way (due to the different definition of $D^{(\r)}_n$): we let (same notations as in \S~\ref{subsec:4.2}) \beqa{6.be} & \b_{v_p}'\= \b_{v_1}\ , &\nonumber\\ & \b_{v_i}'\=\b_{v_{i+1}}\ , &{\rm for}\ 1\le i\le p-1\nonumber\\ & \b_v'\=\b_v\ , &\forall\ v\notin P(u,v)\ . \eeqa With these definitions it is easy to check that \equ{4.12} holds and hence that Theorem~\ref{thm1} is valid in case P3) too. \giu We close by a remark on the $\m$--radius of convergence of $\sum_{k\ge 1} \m^k Z^k$. It is an easy exercise to adapt the estimates in \cite{CF1} to the present case and to check how the radius of convergence depends on $\e$. In fact, observing that, by \equ{6.aa}, one has \beq{6.es} |\lg n\rg^{-\s}|\ |\big(\lg n\rg^2+A\big)^{1-\r}| \le \max\{\g^2, \frac{b}{|\e|}\}\ \ \max\{|n|^{2\t}\ ,\ 1\}\ , \eeq leading to an estimate on the radius of convergence $\m_0$ of the type \beq{6.mu} \m_0\ge {\rm const}\ \min\{\e\ ,\ \g^{-2}\}\ . \eeq Thus $\m=\e^2$ (or $\m=\e^c$ with any $c>1$) is within the domain of analyticity provided $\e$ is small enough. \section{An example with no compensations} \label{sec:par7} \setcounter{equation}{0} % Consider the Hamiltonian \equ{1.12} with \beq{p.1} f(x,q) =\sum_{s\geq 1} f_s e^{i (n^{(s)}\cdot x + m^{(s)}\cdot q)} \eeq where $n^{(s)} \in \integer^N$, $m^{(s)} \in \integer^M$ are given integer vectors (with $|n^{(s)}|+|m^{(s)}|>0$) and the Fourier coefficients $f_s$ decay exponentially fast with $|n^{(s)}|+|m^{(s)}|$. Let $q_0$ be a non degenerate critical point of the $x$--average of $f$ (i.e. of $f_0=\sum_{s\geq 1} f_s e^{i m^{(s)}\cdot q}$); then, as for the previous cases, there exists a (unique) formal power series \beq{p.2} Z \= (Z_1,\dots , Z_d)\ \sim\ \sum_{k\geq 0} Z^k(\th)\e^k\ , \ \th\in\torus^N \eeq with $d=2(N+M)$, such that $t\to Z(\o t)$ is a formal quasi--periodic solution for \equ{1.12} and the set $\{Z^0(\th):\th\in\torus^N\}$ coincide with the torus spanned by $y=\o$, $p=0$, $q=q_0$, $x=x_0+\o t$, where $\o\in\real^N$ satisfies condition \equ{1.om}. \nin Expanding $Z^k$ in Fourier series also in the variable $q$, besides the variable $x$ (see \equ{1.2}), it is still possible to write $Z_n^k$ as in \equ{1.tr} provided one makes the following changes. $\cT^k_* = \cT^k$; $B$ is the trivial set $B=\{\b =\s = 2\}$; $\L(T,\a,\b)$ is replaced by \beq{p.3} \L(T,\hat \a,\b) = \sum_{\a':V\to \a_v'\in \integer^M}\ \prod_{v\in V} f_{\hat\a_v} \prod_{vv'\in E} \hat\a_v \cdot \hat\a_{v'} \eeq where $\hat\a_v\= (\a_v,\a_v')\in \integer^{N+M}$; finally $\g_v = \lg \d_v\rg^{-2}\=\bigl(\o\cdot\sum_{v'\leq v} \a_{v'}\bigr)^{-2}$. \nin It is then clear that in order to have compensations it would be sufficient to have resonances in the variable $\hat\a$ whenever there are resonances in the variable $\a$. In other words compensations take place if the Fourier modes in \equ{p.1} are such that \beq{p.4} \sum_{s\in I} n^{(s)} = 0 \quad \implies \sum_{s\in I} m^{(s)} = 0 \eeq $\forall I\subset \natural$; in fact in such a case one could repeat word--by--word the arguments in \cite{CF1}. \nin In general, if \equ{p.4} does not hold, compensations (of the type described in this paper) do not occur as it is shown by the following example. \giu Take $N=2$, $M=1$, fix a scalar integer $n\neq 0$ and let \beq{p.5} f(x_1,x_2,q) = 2\{\cos x_1 + \cos x_2 + \cos (x_1 + x_2 - q) + \cos (n x_1 + x_2 + q)\}\ . \eeq Hence, the range of $\hat\a$ is the set $\{\pm (1,0,0),\ \pm (0,1,0),\ \pm (1,1,-1),\ \pm (n,1,1)\}$; for definiteness we also fix $\o = (\sqrt{2},-1)$. % \begin{figure}[hbt] \begin{center} \begin{picture}(400,100) \thicklines \put(5,60){\circle{8}} \put(5,60){\circle*{3}} %\put(10,50){$u_1$} %\put(50,50){$w_1$} %\put(100,50){$u_2$} %\put(140,50){$w_2$} %\put(190,50){$u_3$} %\put(230,50){$w_3$} %\put(280,50){$u_4$} %\put(320,50){$w_4$} \multiput(65,60)(90,0){4}{\line(1,0){30}} \multiput(65,60)(90,0){4}{\circle*{3}} \multiput(95,60)(90,0){4}{\circle*{3}} \put(35,60){\oval (75,30)} \put(125,60){\oval (75,30)} \put(215,60){\oval (75,30)} \put(305,60){\oval (75,30)} \put(350,65){{\scriptsize ($n$,1,1)}} \put(365,45){$z$} \put(65,35){$R_1$} \put(155,35){$R_2$} \put(245,35){$R_3$} \put(335,35){$R_4$} \parbox{14cm}{\caption{\label{fig:d1} Divergent contribution: the chain $C=( R_1,R_2,R_3,R_4 )$ is made of 4 adjacent resonances with $|R_i|=3$ (in this symbolic picture only two vertex are drawn).}} \end{picture} \end{center} \end{figure} \nin We fix a tree $T\in \widetilde \cT^k$, with $k = 3h+1$, and a function $\hat \a_v:V(T)\to \integer^3$ so that $T$ contains a chain $C=( R_1,...,R_h )$ made of $h$ identical resonances $R_i\= \{v_1,v_2, v_3\}$, with $\hat\a_{v_1}=(-1,-1,1)$, $\hat\a_{v_2}=(0,1,0)$, $\hat\a_{v_3}=(1,0,0)$, and the last vertex $z$, following the chain, with $\hat\a_z = (n,1,1)$ (see Figure~\ref{fig:d1}). \noindent Let $\cF_C$ be the family of all trees $T\in\cT^k$ which contains the chain $C$. A lengthy but straightforward computation shows that the $i^{th}$ component of the vector valued function $\bar \pi_C(x)$ (see \equ{1.11}) for $h = 1$ is: \beqa{p.6} &(\bar\pi_{C}(x))_i \= (\bar \pi_{R_1}(x))_i & \= \prod_{v\in V} f_{\hat\a_v} \sum_{u,w\in R_1} \hat e_i \cdot \hat \a_u \ \Big( \prod_{vv'\in E(R_1)} \hat\a_v \cdot \hat\a_{v'}\Big) \ \hat \a_w \cdot \hat \a_z \ \pi_{R_1}(x)\nonumber \\ &&\ =\ 2\sum_{j=1}^3 (\d_{i3} \d_{j3} + x^2 B_{ij}(x))\hat \a_{zj} \eeqa where $\d_{ij}$ is the Kronecker's symbol, $\hat e_i$ is the vector of $j^{th}$ component $\d_{ij}$ and \beq{p.7} B(x)=\pmatrix{{{-2(6-x^2)}\over {(2-x^2)^2}}& {{8\sqrt{2}+18 x^2 - 9x^4 + x^6}\over {(2-x^2)^2 (1-x^2)^2}}& {{6-x^2}\over {(2-x^2)^2}}\cr &\phantom{a} &\cr {{8\sqrt{2}+18 x^2 - 9x^4 + x^6}\over {(2-x^2)^2 (1-x^2)^2}}& {{-2(3-x^2)}\over {(1-x^2)^2}}& {{3-x^2}\over {(1-x^2)^2}}\cr &\phantom{a} &\cr {{6-x^2}\over {(2-x^2)^2}}& {{3-x^2}\over {(1-x^2)^2}}&0} . \eeq Hence $\bar \pi_{R_1}(0)\neq 0$. For $h>1$, one simply has: \beq{p.8} \bar \pi_{C}(x) = 2^h (A+x^2 B)^h \hat\a_z\quad,\quad A_{ij}=\d_{i3} \d_{j3}\ . \eeq Thus, if $x = \o\cdot \a_z = \sqrt{2}n-1$, taking the third component of \equ{p.8} (the component for which condition \equ{p.4} is violated) and assuming that $n\geq h$, one easily checks that \beq{p.9} \bigl( (A+x^2 B)^h \hat\a_z\bigr)_3 \geq 1 + O({1\over n}) \eeq which implies that $C$ is a non compensable chain. %%%%%%%%%%%%%%%%%%% bibliography: \begin{thebibliography}{99} % {\small \bibitem{A1} Arnold, V.I. (Ed.): {\em Encyclopaedia of Mathematical Sciences, Dynamical Systems}, Vol. 3, Springer--Verlag, 1988 \bibitem{A2} Arnold, V.I.: {\em Instability of dynamical systems with several degrees of freedom} Sov. Math. Dokl. {\bf 5} 581--585 (1964) \bibitem{Bo} Bollobas, B.: {\em Graph Theory}, Springer (Graduate text in mathematics: 63), 1979 \bibitem{Br} Brjuno, A. D.: {\em Convergence of transformations of differential equations to normal form}, Dokl. Akad. Nauk SSSR {\bf 165}, 987--989 (1965); {\em Analytic form of differential equations}, Trans. Moscow Math. Soc. {\bf 25}, 131--288 (1971) and {\bf 26}, 199--239 (1972) \bibitem{CC} Celletti, A.; Chierchia, L.: {\em Construction of analytic KAM surfaces and effective stability bounds}, Commun. Math. Phys. {\bf 118}, 119--161 (1988) \bibitem{CF1} Chierchia, L.; Falcolini, C.: {\em A direct proof of a theorem of Kolmogorov in Hamiltonian systems}, to appear in Annali Sc. Norm. Super. Pisa, Cl. Sci. \bibitem{CF2} Chierchia, L.; Falcolini, C.: {\em Quasi--periodic solutions of some elliptic systems}, Preprint (1993) \bibitem{CG} Chierchia, L.; Gallavotti, G.: {\em Drift and Diffusion in phase space}, Ann. Inst. Henri Poincar\'e (Physique Th\'eorique) {\bf 60}, n$^{\rm o}$ 1, 1--144 (1994) \bibitem{E1} Eliasson, L. H.: {\em Absolutely convergent series expansions for quasi periodic motions}, Reports Department of Math., Univ. of Stockholm, Sweden, No. 2, 1--31 (1988) \bibitem{E2} Eliasson, L. H.: {\em Hamiltonian systems with linear normal form near an invariant torus}, in ``Nonlinear Dynamics", G. Turchetti (Ed.) World Scientific, Singapore (1989) \bibitem{E3} Eliasson, L. H.: {\em Generalization of an estimate of small divisors by Siegel}, in ``Analysis, et cetera", P.H.~Rabinowitz and E.~Zehnder (Eds.), Academic Press (1990) \bibitem{Ga1} Gallavotti, G.: {\em Twistless KAM tori, quasi flat homoclinic intersections, and other cancellations in the perturbation series of certain completely integrable hamiltonian systems. A review.} Preprint (1993) \bibitem{Ga2} Gallavotti, G.: {\em Twistless KAM tori} to appear in Commun. Math. Physics \bibitem{Ga3} Gallavotti, G.: {\em Invariant tori: a field theoretic point of view on Eliasson's work} Preprint (1993) \bibitem{GG} Gallavotti, G.; Gentile, G.: {\em Non recursive proof of the KAM theorem}, to appear in Ergodic Theory and Dynamical Systems \bibitem{Ge1} Gentile, G.: {\em A proof of existence of whiskered tori with quasi flat homoclinic intersections in a class of almost integrable hamiltonian systems} Preprint (1994) \bibitem{Ge2} Gentile, G.: {\em Whiskered tori with prefixed frequencies and Lyapunov spectrum} Preprint (1994) \bibitem{Gr} Graff, S.M.: {\em On the conservation of hyperbolic invariant tori for Hamiltonian systems}, J. Diff. Equations {\bf 15}, 1--69 (1974) \bibitem{Po} Poincar\'e, H.: {\em Les m\'ethodes nouvelles de la m\'ecanique c\'eleste}. Vols. 1--3. Paris: Gauthier--Villars. (1892/1893/1899) \bibitem{S} Siegel, C. L.: {\em Iterations of analytic functions}, Annals of Math. {\bf 43}, No. 4, 607--612 (1942) \bibitem{SZ} Salamon, D.; Zehnder, E.: {\em KAM theory in configuration space}, Comm. Math. Helv., {\bf 64}, 84 (1988) \bibitem{X} Xia, Z.: {\em Arnold diffusion and oscillatory solutions in the three--body problem}, Preprint (1993) } \end{thebibliography} \end{document}