Content-Type: multipart/mixed; boundary="-------------0801230946404" This is a multi-part message in MIME format. ---------------0801230946404 Content-Type: text/plain; name="08-18.keywords" Content-Transfer-Encoding: 7bit Content-Disposition: attachment; filename="08-18.keywords" Aubry-Mather theory, KAM tori, Lagrangian graphs, Schwartzman unique ergodicity ---------------0801230946404 Content-Type: application/x-tex; name="FaGiSo22-01-2008_Submitted_Version.tex" Content-Transfer-Encoding: 7bit Content-Disposition: inline; filename="FaGiSo22-01-2008_Submitted_Version.tex" %\documentclass[english,a4paper,10pt]{article} \documentclass[10pt]{article} \setlength{\oddsidemargin}{0cm} \setlength{\evensidemargin}{0cm} \setlength{\textwidth}{14.5cm} \usepackage{amsmath,amssymb,amsbsy,amsfonts,latexsym,amsthm,amscd,amsxtra, amssymb,appendix} \usepackage[all]{xy} %\parindent 0pt %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% Definitions, Theorems etc...%%%%%%%%%%%%%%%%%% \newtheorem{sez}{}[section] \newtheorem{Teo}{Theorem}[section] \newtheorem{Def}[Teo]{Definition} \newtheorem{Cor}[Teo]{Corollary} \newtheorem{Lem}[Teo]{Lemma} \newtheorem*{acknow}{Acknowledgements} \newenvironment{Acknow}{\begin{acknow} \begin{upshape}} {\end{upshape} \end{acknow}} \newtheorem{Prop}[Teo]{Proposition} \newtheorem{Cong}[Teo]{Conjecture} \newtheorem{Claim}[Teo]{Claim} \newtheorem{Rema}[Teo]{Remark} \newenvironment{Rem}{\begin{Rema} \begin{upshape}} {\end{upshape}\end{Rema}} \newtheorem*{Pf}{Proof} \newenvironment{Proof}{\begin{Pf} \begin{upshape}} {\end{upshape} \qed\end{Pf}} \newenvironment{Proofnoqed}{\begin{Pf} \begin{upshape}} {\end{upshape} \end{Pf}} \newtheorem*{Teono}{Theorem} \newtheorem*{Lemno}{Lemma} \newtheorem*{Corno}{Corollary} \newtheorem*{Propno}{Proposition} \newtheorem*{mainTeo}{Main theorem} \newtheorem*{mainLem}{Main lemma} %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% Equazioni %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% \newcommand\beq[1]{ \begin{equation} \label{#1}} \newcommand{\eeq}{ \end{equation} } \newcommand{\beqno}{ \begin{equation*}} \newcommand{\eeqno}{ \end{equation*}} \newcommand\beqa[1]{ \begin{eqnarray} \label{#1}} \newcommand{\eeqa}{ \end{eqnarray} } \newcommand{\beqano}{ \begin{eqnarray*} } \newcommand{\eeqano}{ \end{eqnarray*} } \def\be{\begin{equation}} \def\ee{\end{equation}} \def\bea{\begin{eqnarray}}\def\eea{\end{eqnarray}} \def\nn{\nonumber} %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% Some Sets %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% \newcommand{\T}{ {\mathbb T} } \newcommand{\N}{ {\mathbb N} } \newcommand{\R}{ {\mathbb R} } \newcommand{\Z}{ {\mathbb Z} } \newcommand{\C}{ {\mathbb C} } \newcommand{\Q}{ {\mathbb Q} } %\renewcommand{\L}{ {\mathbb L} } \font\msytw=msbm9 scaled\magstep1 \font\cs=cmcsc10 \def\TTT{\T}%{\hbox{\msytw T}} \def\RRR{\R}%{\hbox{\msytw R}} \def\ZZZ{\Z}%{\hbox{\msytw Z}} %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% Greek letters %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% \renewcommand \a {\alpha} \newcommand \e {\varepsilon } \newcommand \x {\xi} \renewcommand \b {\beta} \renewcommand \r {\rho} \renewcommand \d {\delta} \newcommand \D{\Delta} \newcommand \m {\mu} \newcommand \n {\nu} \newcommand \om {\omega} \newcommand \Om {\Omega} \newcommand \ff {\phi} \newcommand \f {\varphi} \newcommand \F {\Phi} \newcommand \g {\gamma} \newcommand \tg {\tilde{\gamma}} \newcommand \G {\Gamma} \newcommand \s {\sigma} \renewcommand \S {\Sigma} \renewcommand \k {\chi} \renewcommand \t {\tau} \renewcommand \th {\theta} \renewcommand \l {\lambda} \renewcommand \L {\Lambda} \newcommand \ph {\phi} \newcommand \Ph {\phi} \newcommand \et {\eta} \newcommand \z {\zeta} \newcommand \p {\pi} \renewcommand \P {\Pi} %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% Various %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% \renewcommand \c[1]{\mathcal{#1}} \newcommand \cA {{\mathcal A}} \newcommand \cN {{\mathcal N}} \newcommand \cS {{\mathcal S}} \newcommand \cD {{\mathcal D}} \newcommand \cbA {\bar{{\mathcal A}}} \newcommand \bd {\bar{\d}} \newcommand \cL {{\mathcal L}} \newcommand \cE {{\mathcal E}} \newcommand \cR {{\mathcal R}} \newcommand \cLip {{\mathcal Lip}} \newcommand \dpr {\partial} %\newcommand{\ie}{{\it i.e.}} %\newcommand{\eg}{{\it e.g.\ }} \newcommand{\etc}{{\it etc.\ }} \newcommand{\tu}{\tilde{u}} \newcommand{\tx}{\tilde{x}} \newcommand \per {\times} \newcommand \se {\Longrightarrow} \newcommand \su[1]{\frac{1}{#1}} \renewcommand \iff{\Longleftrightarrow} \newcommand \der[2]{\frac{\partial #1}{\partial #2}} \newcommand \id {{\mathbb I} } \newcommand \idd {{\rm id}\, } \newcommand \ov {\overline} \newcommand \rH {{\rm H}} %\newcommand \tAeqc {{\sim_{\cA_{L,c}}}} %\newcommand \cLAo {\cL_{\cbA_{L,c},1}} \newcommand{\tw}{\tilde{w}} \newcommand \ctS {\tilde{\mathcal S}_{\cA}} \newcommand \ctD {\tilde{\mathcal D}_{\cA_{L,c}}} \newcommand \bx {\bar{x}} \newcommand{\tv}{\tilde{v}} \newcommand{\tf}{\tilde{f}} %\newcommand \cLA {\cL_{\cbA,\frac{1}{2}}} \newcommand \calM {\mathfrak{M}} \newcommand \cM {{\mathcal M}} \newcommand \cI {{\mathcal I}} \def\PPP{{\mathcal P}}\def\EE{{\mathcal E}}\def\MM{{\mathcal M}} \def\VV{{\mathcal V}}\def\CC{{\mathcal C}}\def\FF{{\mathcal F}} \def\HHH{{\mathcal H}}\def\WW{{\mathcal W}}\def\TT{{\mathcal T}} \def\NN{{\mathcal N}} \def\BBB{{\mathcal B}}\def\III{{\mathcal I}} \def\RR{{\mathcal R}}\def\LL{{\mathcal L}} \def\JJ{{\mathcal J}} \def\OO{{\mathcal O}}\def\DD{{\mathcal D}}\def\AAA{{\mathcal A}} \def\GG{{\mathcal G}} \def\SS{{\mathcal S}}\def\KK{{\mathcal K}} \def\UU{{\mathcal U}} \def\QQ{{\mathcal Q}} \def\XXX{{\mathcal X}} \let\io=\infty \def\Dpr{\V\dpr\,} \def\aps{{\it a posteriori\ }}\def\ap{{\it a priori\ }} \let\0=\noindent\def\pagina{{\vfill\eject}} \def\bra#1{{\langle#1|}}\def\ket#1{{|#1\rangle}} \def\media#1{{\langle#1\rangle}} \def\ie{\hbox{\it i.e.,\ }} \def\eg{\hbox{\it e.g.,\ }} \let\dpr=\partial \def\der{{\rm d}} \let\circa=\cong \def\arccot{{\rm arccot}} \let\o=\omega \let\==\equiv %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% \begin{document} \title{Uniqueness of Invariant Lagrangian Graphs in a Homology or a Cohomology Class.} \author{\sc{Albert Fathi, Alessandro Giuliani and Alfonso Sorrentino}} \date{22 January 2008} \maketitle %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% %%%%%%%%%%%%%%%%%% ABSTRACT %%%%%%%%%%%%%%%%%%%% \begin{center} \parbox[c]{4.5 in} {\small Given a smooth compact Riemannian manifold $M$ and a Hamiltonian $H$ on the cotangent space ${\rm T}^*M$, strictly convex and superlinear in the momentum variables, we prove uniqueness of certain ``ergodic'' invariant Lagrangian graphs within a given homology or cohomology class. In particular, in the context of quasi-integrable Hamiltonian systems, our result implies global uniqueness of Lagrangian KAM tori with rotation vector $\r$. This result extends generically to the $C^0$-closure of KAM tori.} \end{center} %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% %%%%%%%%%%%%%%%%%%%%%%%% SECTION 1%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% \section{Introduction}\label{intro} %%%%%%%%%%%%%%%%%%%%s+\t}%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% A particularly interesting and fruitful approach to the study of local and global properties of dynamical systems is concerned with the study of invariant submanifolds, rather than single orbits, paying particular attention to their existence, their ``fate'' and their geometric properties. In the context of quasi-integrable Hamiltonian systems, one of the most celebrated breakthroughs in this kind of approach was KAM theory, which provided a method to construct invariant submanifolds diffeomorphic to tori, on which the dynamics is conjugated to a quasi-periodic motion with rotation vector $\r$, sometimes referred to as {\it KAM tori}. KAM theory finally settled the old question about {\it existence} of such invariant submanifolds in ``generic'' quasi-integrable Hamiltonian systems, dating back at least to Poincar\'e, and opened the way to a new understanding of the nature of Hamiltonian systems, of their stability and of the onset of chaos in classical mechanics. However, the natural question about the {\it uniqueness} of these invariant submanifolds for a fixed rotation vector $\r$ remained open for many more years and, quite surprisingly, even nowadays, for many respects it is still unanswered. A possible reason for this is that the analytic methods, which the KAM algorithm is based on, are not well suited for studying global questions, while, on the other hand, the natural variational methods to approach this problem have been developed only much more recently and they are still not so widely well-known. In this paper we address the above problem and prove global uniqueness of certain ``ergodic'' invariant Lagrangian graphs within a given homology or cohomology class, for a large class of convex Hamiltonians, known as Tonelli Hamiltonians. Our work will be based on the variational approach provided by the so-called Aubry-Mather theory, as well as its functional side called weak KAM theory. The paper is organized as follows. In Section \ref{sec1} we define the geometric objects we shall look at (the invariant Lagrangian graphs), we introduce some concepts (homology class of an invariant measure and Schwartzman ergodicity), which will turn out to be useful for illustrating their dynamical properties, and we state our main uniqueness results. In Section \ref{sec2} we review the main definitions and results of Aubry-Mather and weak KAM theories, to the extent of what we need for our proofs. In Section \ref{sec3} we prove our main results. In Section \ref{sec4} we discuss in detail the implications of our results for KAM theory and compare them with some previous local uniqueness theorems for KAM tori. In Appendix \ref{app1} we discuss some details concerning the definition of Schwartzman ergodicity, give some examples and describe some properties of Schwartzman ergodic flows. In Appendices \ref{app2} and \ref{app3} we prove some properties of Lagrangian graphs and the Ma\~n\'e set. %In Appendix \ref{app2} we prove some properties of %Lagrangian graphs and their relation with KAM tori. %In Appendix \ref{app3} we prove a connection between invariant measures supported on minimizing orbits and Aubry sets. \vspace{.3truecm} {\bf Acknowledgements.} We would like to thank Giovanni Gallavotti for having drawn our attention to this problem and for useful discussions. AF and AS are grateful to John Mather for many fruitful conversations.\\ %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% %%%%%%%%%%%%%%%%%%%%%%%% SECTION 1%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% \section{Setting and Main results}\label{sec1} %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% Let $M$ be a compact and connected smooth manifold without boundary of dimension $n$. Denote by ${\rm T}M$ its tangent bundle and ${\rm T}^*M$ the cotangent one. A point of ${\rm T}M$ will be denoted by $(x,v)$, where $x\in M$ and $v\in {\rm T}_xM$, and a point of ${\rm T}^*M$ by $(x,p)$, where $p\in {\rm T}_x^*M$ is a linear form on the vector space ${\rm T}_xM$. Let us fix a Riemannian metric $g$ on $M$ and let $\|\cdot\|_x$ be the norm induced by $g$ on ${\rm T}_x M$; we shall use the same notation for the norm induced on the cotangent space ${\rm T}_x^*M$. This cotangent space ${\rm T}^*M$ can be canonically endowed with a symplectic structure, given by the exact $2$-form $\o=dx\wedge dp = -d(pdx)$, where $({\mathcal U},\,x)$ is a local coordinate chart for $M$ and $({\rm T}^*{\mathcal U},\,x, p)$ the associated cotangent coordinates. The $1$-form $\l = pdx$ is also called {\it tautological form} (or {\it Liouville form}) and is intrisically defined, \ie independently of the choice of local coordinates (see Appendix \ref{app2}). A distinguished role in the study of the geometry of a symplectic space is played by the so-called {\it Lagrangian submanifolds}. \begin{Def}\rm Let $\Lambda$ be an $n$-dimensional $C^1$ submanifold of $({\rm T}^*M,\omega)$. We say that $\Lambda$ is {\it Lagrangian} if for any $(x,p)\in \Lambda$, ${\rm T}_{(x,p)}\Lambda$ is a Lagrangian subspace, \ie $\omega \big|_{{\rm T}_{(x,p)}\Lambda} = 0$. \end{Def} We shall mainly be concerned with Lagrangian graphs, that is Lagrangian manifolds $\Lambda\subset T^*M$ such that $\Lambda=\{(x,\eta(x))\,,\ x\in M\}$. It is straightforward to check that the graph $\Lambda$ is Lagrangian if and only if $\eta$ is a closed $1$-form (see Appendix \ref{app2}). The element $c=[\eta]\in \rH^1(M;\R)$ is called the {\it cohomology class}, or {\it Liouville class}, of $\Lambda$. This motivates the following extension of the notion of Lagrangian graph to the continuous case. \begin{Def}\rm A continuous section $\L$ of ${\rm T}^*M$ is a $C^0$-Lagrangian graph if it locally coincides with the graph of an exact differential. \end{Def} We will consider the dynamics on ${\rm T}^*M$ generated by a Tonelli Hamiltonian. \begin{Def}\rm A function $H:\,{\rm T}^*M\longrightarrow \R$ is called a {\it Tonelli (or optical) Hamiltonian} if: \begin{itemize} \item[i)] the Hamiltonian $H$ is of class $C^k$, with $k\geq 2$; \item[ii)] the Hamiltonian $H$ is strictly convex in the fiber in the $C^2$ sense, \ie the second partial vertical derivative ${\dpr^2 H}/{\dpr p^2}(x,p)$ is positive definite, as a quadratic form, for any $(x,p)\in {\rm T}^*M$; \item[iii)] the Hamiltonian $H$ is superlinear in each fiber, \ie $$\lim_{\|p\|_x\rightarrow +\infty} \frac{H(x,p)}{\|p\|_x} = + \infty$$ (by the compactness of $M$, this condition is independent of the choice of the Riemannian metric). \end{itemize} \end{Def} Given $H$, we can define the associated {\it Lagrangian}, as a function on the tangent bundle: % \bea L:\; {\rm T}M &\longrightarrow & \R \nn\\ (x,v) &\longmapsto & \sup_{p\in {\rm T}^*_xM} \{\langle p,\,v \rangle_x -H(x,p)\}\,\nn \eea % where $\langle \,\cdot,\,\cdot\, \rangle_x$ represents the canonical pairing between the tangent and cotangent space. If $H$ is a Tonelli Hamiltonian, one can easily prove that $L$ is finite everywhere, of class $C^k$, superlinear and strictly convex in the fiber in the $C^2$ sense (\ie $L$ is a {\it Tonelli Lagrangian}) and the associated Euler-Lagrange flow $\Phi^L_t$ of $L$ is conjugated to the Hamiltonian flow $\Phi^H_t$ of $H$ via the {\it Legendre transform}: % \bea \cL:\; {\rm T}M &\longrightarrow & {\rm T}^*M \\ (x,v) &\longmapsto & \left(x,\,\frac{\dpr L }{\dpr v}(x,v) \right).\label{2.1} \eea % From now on we shall fix $H$ and denote by $L$ its conjugated Lagrangian and, when referring to an ``invariant'' measure or set, we will understand ``invariant with respect to the Hamilton flow generated by $H$'' or ``with respect to the Euler-Lagrange flow generated by $L$''. Given an invariant probability measure $\m$ on ${\rm T}M$ one can associate to it an element $\r(\m)$ of the homology group $\rH_1(M;\RRR)$, known as {\it rotation vector} or {\it Schwartzman asymptotic cycle}, which generalizes the notion of rotation vector given by Poincar\'e and describes how, asymptotically, a $\m$-average orbit winds around ${\rm T}M$. See Section \ref{sec2} and Appendix \ref{app1} for a precise definition. This allows us to define the homology class of certain invariant Lagrangian graphs. \begin{Def}\rm A Lagrangian graph $\L$ is called {\it Schwartzman uniquely ergodic} if all invariant measures supported on $\L$ have the same rotation vector $\r$, which will be called {\it homology class of} $\L$. Moreover, if there exists an invariant measure with full support, $\L$ will be called {\it Schwarztman strictly ergodic}. \end{Def} We are now ready to state our main results.\\ {\bf{Main Result.}} {\it For any given $\r\in \rH_1(M;\RRR)$, there exists at most one Schwarzman strictly ergodic invariant Lagrangian graph with homology class $\r$} [Theorem \ref{uniqschw}, Section \ref{sec3}]\\ For sake of completeness, we also recall the following well-known result, which is a corollary of the results in \cite{Mather91} (see also Section \ref{sec3} for a proof).\\ \\ {\bf{Well-known Result.}} {\it For any given $c\in \rH^1(M;\RRR)$, there exists at most one invariant Lagrangian graph $\Lambda$ with cohomology class $c$, carrying an invariant measure whose support is the whole of $\Lambda$.} [Theorem \ref{uniqcom}, Section \ref{sec3}]\\ If $M=\TTT^n$, it is natural to ask for the implications of our result for KAM theory. In this case, the homology group $\rH_1(\TTT^n;\RRR)$ is canonically identified with $\R^n$, and the invariant manifolds of interest are the so-called KAM tori, defined as follows. \begin{Def}\label{kamtorus}\rm $\TT\subset\TTT^n\times\RRR^n$ is a (maximal) KAM torus with rotation vector $\r$ if:\\ \0i) $\TT\subset\TTT^n\times\RRR^n$ is a continuous graph over $\TTT^n$;\\ \0ii) $\TT$ is invariant under the Hamiltonian flow $\Phi_t^H$ generated by $H$;\\ \0iii) the Hamiltonian flow on $\TT$ is conjugated to a uniform rotation on $\TTT^n$; \ie there exists a diffeomorphism $\f: \TTT^n\to \TT$ such that $\f^{-1}\circ\Phi_t^H\circ\f= R_\r^t$, $\forall t\in\RRR$, where $R_\r^t: x \to x+\r t$. \end{Def} The celebrated KAM Theorem, whose statement will be recalled in Section \ref{sec4}, gives sufficient conditions on $H$ and on the rotation vector $\r$, allowing one to construct a KAM torus with rotation vector $\r$ and prescribed regularity (depending on the regularity class of $H$). Its proof is constructive and the invariant torus one manages to construct is locally unique (in a sense that will be clarified in Section \ref{sec4}). In spite of the long history and the huge literature dedicated to the KAM theorem, the issue of global uniqueness of such tori is still object of some debate and study, see for instance \cite{BroerTakens}. Our main result settles such question in the case of Tonelli Hamiltonians. \begin{Cor}[Global uniqueness of KAM tori]\label{corollary1} Every Tonelli Hamiltonian $H$ on ${\rm T}^*\TTT^n$ possesses at most one Lagrangian KAM torus for any given rotation vector $\r$. In particular, if $H$ and $\r$ satisfy the assumptions of the KAM Theorem, then there exists one and only one KAM torus with rotation vector $\r$. \end{Cor} The property of being Lagrangian plays a crucial role. As was observed by Herman \cite{Herman} (see also Appendix \ref{app2}), when $\r$ is {\it rationally independent}, \ie $\media{\r,\n}\neq 0$, $\forall \n\in\ZZZ^n\setminus\{0\}$, as assumed in the KAM theorem, every KAM torus with frequency $\r$ is automatically Lagrangian. On the other hand, the existence of Lagrangian KAM tori with rationally dependent frequency is not typical. In some cases, a variant of the classical KAM algorithm allows one to construct {\it resonant} invariant tori with a given rational rotation vector $\r$, also known as low dimensional tori \cite{GG1,GG2}. However, typically they do not foliate any Lagrangian submanifold. Therefore, the question of uniqueness of resonant tori is more subtle and, to our knowledge, apart from a few partial results \cite{CGGG}, it is still open. Note that the fact that the orbits on a KAM torus are action-minimizing is a key remark used in the proof of the Corollary. We deduce this property from the fact that such a Lagrangian torus gives a solution to the Hamilton-Jacobi equation. It can be also deduced from a classical result by Weierstrass, as already pointed out by J\"urgen Moser. We thank John Mather for having drawn our attention to the remark in \cite{Mather91}. [In fact, Weierstrass method or the use of the Hamilton-Jacobi Equation are essentially two sides of the same coin]. In Section \ref{sec4} we will extend Corollary \ref{corollary1} to generic invariant tori contained in the $C^0$-closure of the set of KAM tori. As remarked by Herman \cite{HermanTori}, generically this set is much larger than the set of KAM tori, and typically the flow on such invariant manifolds is not conjugated to a rotation. See Section \ref{sec4} for a more detailed discussion of these issues. %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% %%%%%%%%%%%%%%%%%%%%%%%% SECTION 2%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% \section{An introduction to Aubry-Mather theory}\label{sec2} %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% This section is meant to provide a brief (but comprehensive) introduction to Mather's theory for Lagrangian systems and weak KAM theory. We shall recall most of the results that we are going to use, trying to give general ideas rather than rigourous proofs (for which we refer the reader to \cite{Fathibook, Mather91}). The setting is the same introduced in Section \ref{sec1}. Let $M$ be a compact and connected smooth manifold without boundary of dimension $n$. We denote by $H$ a fixed Tonelli Hamiltonian and $L$ the associated Tonelli Lagrangian. Before entering into the details of Mather's theory, let us make a crucial remark. Observe that if $\eta$ is a $1$-form on $M$, we can easily define a function on the tangent space (linear on each fiber) \beqano \hat{\eta}: {\rm T}M &\longrightarrow& \R \\ (x,v) &\longmapsto& \langle \eta(x),\, v\rangle_x \eeqano and consider a new Tonelli Lagrangian $L_{\eta}:= L - \hat{\eta}$. The associated Hamiltonian will be $H_{\eta}(x,p) = H(x,\eta(x) + p)$. Moreover, if $\eta$ is closed, then $\int L\, dt$ and $\int L_{\eta} dt$ will have the same extremals and therefore the Euler-Lagrange flows on ${\rm T} M$ associated to $L$ and $L_{\eta}$ will be the same. This last point may be seen by observing that the variational equations $\d \int L\, dt =0$ and $\d \int (L-\hat{\eta})\, dt =0$ for the fixed end-point problem clearly have the same solutions, since $\eta$ is closed. Although the extremals are the same, this is not generally true for the orbits ``minimizing the action'' (we shall give a precise definition of ``minimizers'' later in this section). What one can say is that they stay the same when we change the Lagrangian by an exact $1$-form. Thus, for a fixed $L$, the minimizers will depend only on the de Rham cohomology class $c=[\eta] \in \rH^1(M;\R)$. Considering modified Lagrangians corresponding to different cohomology classes represents a keystone of Mather's theory of Lagrangians systems.\\ %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% In order to generalize to more degrees of freedom Aubry and Mather's variational approach to twist maps, a first important notion is that of {\it minimal measures}, which replaces that of action minimizing orbits: Aubry-Mather theory in higher dimension cannot deal with such orbits, due to a lack of them; there is, in fact, a classical example due to Hedlund (see \cite{Hedlund}) of a Riemannian metric on a three-dimensional torus, for which minimal geodesics exist only in three directions. Instead, Mather proposed to look at the closely related notion of action minimizing invariant probability measures. Let $\calM(L)$ be the space of probability measures on ${\rm T}M$ invariant under the Euler-Lagrange flow of $L$. To every $\m \in \calM(L)$, we may associate its {\it average action} $$ A_L(\m) = \int_{{\rm T}M} L\,d\m\,. $$ Since $L$ is bounded below (because of the superlinear growth condition), this integral exists although it might be $+\infty$. In \cite{Mather91}, Mather showed the existence of $\m \in \calM(L)$ such that $A_L(\m)<+\infty$. The argument is mainly the same as Krylov-Bogoliubov's theorem concerning existence of invariant measures for flows on compact spaces. This argument is applied to a one-point compactification of ${\rm T}M$, and the main step consists in showing that the measure provided by this construction has no atomic part supported at $\infty$ (which is a fixed point for the extended system). Note that Mather's approach works for periodic time-dependent Lagrangians. For time independent Lagrangians, finding such a $\mu$ is much easier. By conservation of energy, the levels of the energy function are compact and invariant under the Euler-Lagrange flow $\Phi^L_t$, and therefore carry such measures. In case $A_L(\m)<\infty$, thanks to the superlinearity of $L$, for any closed 1-form $\eta$ on $M$, the integral $ \int_{{\rm T}M} \hat{\eta} d\m$ is well defined and finite, see \cite{Mather91}. Moreover, it is quite easy to show (again see \cite{Mather91}) that since $\m$ is invariant by the Euler-Lagrangian flow $\Phi^L_t$, if $\eta=df$ is an exact form, then $\int{\widehat{df}} d\m =0$. Therefore, we can define a linear functional \beqano {\rm H}^1(M;\R) &\longrightarrow& \R \\ c &\longmapsto& \int_{{\rm T}M} \hat{\eta} d\m\,, \eeqano where $\eta$ is any closed $1$-form on $M$, whose cohomology class is $c$. By duality, there exists $\rho (\m) \in {\rm H}_1(M;\R)$ such that $$ \int_{{\rm T}M} \hat{\eta} d\m = \langle c,\rho(\m) \rangle \qquad \forall\,c\in {\rm H}^1(M;\R)$$ (the bracket on the right--hand side denotes the canonical pairing between cohomology and homology). We call $\rho(\m)$ the {\it rotation vector} of $\m$. It is the same as the Schwartzman's asymptotic cycle of $\mu$ (see Appendix \ref{app1} and \cite{Schwartzman}). One can show that the action functional $A_L: \calM(L) \longrightarrow \R\cup \{+\infty\}$ is lower semicontinuous and that for every $h\in {\rm H}_1(M;\R)$ there exists $\m\in \calM(L)$ with $\rho(\m)=h$. In other words the map $\rho: \calM(L) \longrightarrow {\rm H}_1(M;\R)$ is surjective.\\ Following Mather, let us consider the minimal value of the average action $A_L$ over the probability measures with rotation vector $h$: \bea \b: {\rm H}_1(M;\R) &\longrightarrow& \R \nn\\ h &\longmapsto& \min_{\m\in\calM(L):\,\rho(\m)=h} A_L(\m)\,.\label{2.2} \eea In fact, the minimum above is actually achieved (see \cite{Mather91}). This function $\beta$ is what is generally known as {\it Mather's $\beta$-function}. A measure $\m \in \calM(L)$ realizing such a minimum, \ie $A_L(\m) = \b(\rho(\m))$, is called a {\it minimal} (or a {\it Mather}) {\it measure}. The $\beta$-function is convex, and therefore one can consider its {\it conjugate} function (given by Fenchel's duality) $ \a: {\rm H}^1(M;\R) \longrightarrow \R $ defined by \beqano \a(c) &:=& \max_{h\in {\rm H}_1(M;\R)} \left(\langle c,h \rangle - \b(h) \right) = \\ &=& - \min_{h\in {\rm H}_1(M;\R)} \left(\b(h) - \langle c,h \rangle\right) =\\ &=& - \min_{\m \in \calM(L)} \left( A_L(\m) - \langle c, \rho(\m)\rangle\right) =\\ &=& - \min_{\m \in \calM(L)} A_{L-c}(\m).\\ \eeqano Note that for a given $\mu$ invariant under the Euler-Lagrange flow $\Phi^L_t$, we have $\langle c, \rho(\m)\rangle= \int_{{\rm T}M} \hat{\eta}\, d\m$, for any closed 1-form $\eta$ on $M$ in the cohomology class $c$. Therefore, we have $A_{L-c}(\m)= \int_{{\rm T}M} (L-\hat{\eta})\, d\m$, for $\eta$ a closed $1$-form on $M$ such that $[\eta]=c$. It is important to notice that the value of this $\a$-function coincides with what is called {\it Ma\~n\'e's critical value}, which will be introduced later in this section. An important fact is the next Lemma. To state it, recall that, like any convex function on a finite-dimensional space, the Mather function $\b$ admits a subderivative at any point $h\in \rH_1(M;\R)$, \ie we can find $c\in \rH^1(M;\R)$ such that $$\forall h'\in \rH_1(M;\R), \quad \b(h')-\b(h)\geq \langle c,h'-h\rangle.$$ As it is usually done, we will denote by $\partial \b(h)$ the set of $c\in \rH^1(M;\R)$ which are subderivatives of $\b$ at $h$, \ie the set of $c$ which satisfy the inequality above. By Fenchel's duality, we have $$c\in \partial \b(h) \Longleftrightarrow \langle c,h\rangle=\a(c)+\b(h).$$ \begin{Lem}\label{CaracMinim} If $\m \in \calM(L)$, then $A_L(\m)=\b(\rho(\m))$ if and only if there exists $c\in {\rm H}^1(M;\R)$ such that $\m$ minimizes $A_{L-c}$ (\ie $A_{L-c}(\mu)=-\a(c)$). Moreover, if $\m \in \calM(L)$ satisfies $A_L(\m)=\b(\rho(\m))$, and $c \in \rH^1(M;\R)$, then $\m$ minimizes $A_{L-c}$ if and only if $c\in \partial \b(\rho(\mu))$ {\rm(}or $\langle c,h\rangle=\a(c)+\b(\rho(\m)\rm{)}$. \end{Lem} \begin{Proof} We will prove both statement at the same time. Assume $A_L(\m_0)=\b(\rho(\m_0))$. Let $c\in \partial \b(\rho(\mu_0))$, by Fenchel's duality this is equivalent to \begin{align*} \a(c)&=\langle c,\rho(\m_0)\rangle-\b(\rho(\m_0))\\ &=\langle c,\rho(\m_0)\rangle-A_L(\m_0)\\ &=-A_{L-c}(\m_0). \end{align*} Therefore by what was obtained above $A_{L-c}(\m_0)=\min_{\m \in \calM} A_{L-c}(\m)$. Assume conversely that $A_{L-c}(\m_0)=\min_{\m \in \calM} A_{L-c}(\m)$, for some given cohomology class $c$. Therefore by what was obtained above $$\a(c)=-A_{L-c}(\m_0),$$ which can be written as $$ \langle c,\rho(\m_0)\rangle=\a(c)+A_L(\m_0).$$ It now suffices to use the Fenchel inequality $ \langle c,\rho(\m_0)\rangle\leq \a(c)+\b(\rho(\m_0)$, and the inequality $\b(\rho(\m_0)\leq A_L(\m_0)$, given by the definition of $\b$, to obtain the equalities $$ \langle c,\rho(\m_0)\rangle=\a(c)+A_L(\m_0).$$ In particular, we have $A_L(\m_0)=\b(\rho(\m_0))$. \end{Proof} If $\m\in\calM(L)$ and $\m$ minimizes $A_{L-c}$, we shall say that $\m$ is a {\it $c$-action minimizing measure}. For such $\m$, $c$ is a subderivative of $\b$ at $\rho(\m)$, \ie the slope of a supporting hyperplane of the epigraph of $\b$ at $\rho(\m)$.\\ The above discussion leads to two equivalent formulations for the minimality of a measure $\m$: \begin{itemize} \item there exists a homology class $h \in {\rm H}_1(M;\R)$, namely its rotation vector $\rho(\m)$, such that $\m$ minimizes $A_L$ amongst all measures in $\calM(L)$, with rotation vector $h$; \ie $A_L(\m)=\b (\rho(\m))$; \item there exists a cohomology class $c \in {\rm H}^1(M;\R)$, namely any subderivative of $\b$ at $\rho(\m)$, such that $\m$ minimizes $A_{L-c}$ amongst all measures in $\calM(L)$; \ie $A_{L-c}(\m)=-\a (c)$. \end{itemize} For $h \in {\rm H}_1(M;\R)$ and $c \in {\rm H}^1(M;\R)$, let us define \beqano \calM^h &:=& \{\m \in \calM(L): \; A_L(\m)<+\infty, \;\rho(\m)= h \; {\rm and}\; A_L(\m)=\beta(h)\}\\ \calM_c &:=& \{\m \in \calM(L): \; A_{L}(\m)<+\infty \;{\rm and}\; A_{L-c}(\m) =-\a (c)\}. \eeqano Observe that because of the superlinear growth condition in the fiber, $A_L(\m)< +\infty$ implies $A_{L-c}(\m)< +\infty$. We have $$ \bigcup_{h \in {\rm H}_1(M;\R)} \calM^h = \bigcup_{c \in {\rm H}^1(M;\R)} \calM_c\,. $$ This leads to the definition of a first important family of invariant sets: {\it Mather sets}. For a cohomology class $c \in {\rm H}^1(M;\R)$, we call {\it Mather set of cohomology class} $c$ the set: % \be \widetilde{\cM}_c := \overline{\bigcup_{\m \in \calM_c} {\rm supp}\,\m} \subset {\rm T}M\,;\label{2.3}\ee % the projection on the base manifold $\cM_c = \pi \left(\widetilde{\cM}_c\right) \subseteq M$ is called {\it projected Mather set} (with cohomology class $c$). In \cite{Mather91}, Mather proved the celebrated {\it graph theorem}: \begin{Teo}\label{Theograph}Let $\widetilde{\cM}_c$ be defined as in (\ref{2.3}). The set $\widetilde{\cM}_c$ is compact, invariant under the Euler-Lagrange flow and $\pi|{\widetilde{\cM}_c}$ is an injective mapping of $\widetilde{\cM}_c$ into $M$, and its inverse $\pi^{-1}: \cM_c \longrightarrow \widetilde{\cM}_c$ is Lipschitz. Moreover this set is contained in the energy level corrisponding to the value $\a(c)$, \ie % \be H \circ \cL(x,v) = \a(c) \qquad \forall\,(x,v)\in \widetilde{\cM}_c\,. \label{2.4}\ee \end{Teo} % \begin{Rem}{\rm The last statement, corresponding to (\ref{2.4}), is due to Dias-Carneiro \cite{Carneiro}.} \end{Rem} Analogously, one can consider the {\it Mather set corresponding to a rotation vector} $h\in{\rm H}_1(M;\R)$ as % \be \widetilde{\cM}^h := \overline{\bigcup_{\m \in \calM^h} {\rm supp}\,\m} \subset {\rm T}M\,,\label{2.5}\ee % and the projected one $\cM^h = \pi \left(\widetilde{\cM}^h\right) \subseteq M$. Notice that by Lemma \ref{CaracMinim}, if $c\in\partial\b(h)$, we have $$\widetilde{\cM}^h\subseteq \widetilde{\cM}_c.$$ Therefore, although this was not shown in \cite{Mather91}, the set $\widetilde{\cM}^h$ also has a Lipschitz graph over the basis. \begin{Teo}\label{matherrotation} Let $\widetilde{\cM}^h$ be defined as in (\ref{2.5}). $\widetilde{\cM}^h$ is compact, invariant under the Euler-Lagrange flow and $\pi|{\widetilde{\cM}^h}$ is an injective mapping of $\widetilde{\cM}^h$ into $M$ and its inverse $\pi^{-1}: \cM^h \longrightarrow \widetilde{\cM}^h$ is Lipschitz. \end{Teo} \begin{Rem} Though the graph property for $\widetilde{\cM}^h$ is not proved in \cite{Mather91}, it was shown there that the support of an $h$-minimizing measure has the graph property. The graph property for $\widetilde{\cM}^h$ can be easily deduced from this last property. In fact, since the space of probability measures on ${\rm T}M$ is a separable metric space, one can take a countable dense set $\{\m_n\}_{n=1}^{\infty}$ of Mather's measures with rotation vector $h$ and consider the new measure $\tilde{\m} = \sum_{n=1}^\infty \frac{1}{2^n}\m_n$. This is still an invariant probability measure with rotation vector $h$ and ${\rm supp}\,\tilde{\m} = \widetilde{\cM}_h$. Therefore, as the support of a single $h$-minimizing measure, $\widetilde{\cM}^h $ has the graph property. Moreover, this remark points out that there always exist Mather's measures $\m^h$ and $\m_c$ of full support, \ie ${\rm supp}\,{\m^h} = \widetilde{\cM}^h$ and ${\rm supp}\,{\m_c} = \widetilde{\cM}_c$. We will say that an $h$-minimizing (resp.~$c$-minimizing) measure $\mu$ has maximal support if ${\rm supp}\,{\m} = \widetilde{\cM}^h$ (resp.~${\rm supp}\,{\m} = \widetilde{\cM}_c$). \end{Rem} %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% In addition to the Mather sets, one can construct other compact invariant sets, that play an interesting role both from a dynamical system and a geometric point of view: the {\it Aubry sets} and the {\it Ma\~n\'e sets}. Instead of considering action minimizing invariant probability measures, we now shift our attention to $c$-{\it minimizing curves}. Let us fix $\eta$ a closed $1$-form on $M$, with cohomology class $c$. As done by Mather in \cite{Mather93}, it is convenient to introduce, for $t>0$ and $x,\,y \in M$, the following quantity: $$ h_{\eta,t}(x,y) = \inf \int_0^t L_{\eta}(\g(s),\dot{\g}(s))\,ds\,,$$ where the infimum is taken over all piecewise $C^1$ paths $\g: [0,t]\longrightarrow M$, such that $\g(0)=x$ and $\g(t)=y$. We define the {\it Peierls' barrier} as: $$ h_{\eta}(x,y) = \liminf _{t \rightarrow +\infty} (h_{\eta,t}(x,y) + \a(c)t) \,.$$ %where $\a:\,H^1(M;\R) \longrightarrow \R$ is Mather's $\a$ function %(see \cite{Mather91}). It can be shown that this function is Lipschitz and that, only in the autonomous case, this $\liminf$ can be replaced with a $\lim$ (see \cite{Fathibook}). Observe that $h_{\eta}$ does not depend only on the cohomology class $c$, but also on the choice of the representant; namely, if $\eta' = \eta + df$, then $h_{\eta'}(x,y) = h_{\eta}(x,y) + f(y) - f(x)$. \begin{Prop} The values of the map $h_{\eta}$ are finite. Moreover, the following properties hold: \begin{itemize} \item[{\rm i)}] $h_{\eta}$ is Lipschitz; \item[{\rm ii)}] for each $x\in M$, $h_{\eta}(x,x)\geq 0$; \item[{\rm iii)}] for each $x,\,y,\,z \in M$, $h_{\eta}(x,y) \leq h_{\eta} (x,z) + h_{\eta}(z,y)$; \item[{\rm iv)}] for each $x,\,y \in M$, $h_{\eta}(x,y) + h_{\eta}(y,x) \geq 0$. \end{itemize} \end{Prop} For a proof of the above claims and more, see \cite{Mather93, Fathibook, ContrerasIturriaga}. Inspired by these properties, one can consider its symmetrization: \beqano \d_c:\; M\times M &\longrightarrow & \R \\ (x,y) &\longmapsto & h_{\eta}(x,y) + h_{\eta}(y,x) \eeqano % (observe that this function does depend only on the cohomology class). This function is positive, symmetric and satisfies the triangle inequality; therefore, it is a pseudometric on $$\cA_{c} =\{x\in M : \; \d_c(x,x)=0 \}\,. $$ $\cA_{c}$ is called the {\it projected Aubry set} associated to $L$ and $c$, and $\d_c$ is {\it Mather's pseudometric}. In \cite{Mather93}, Mather has showed that this is a non-empty compact subset of $M$, that can be Lipschitzly lifted to a compact invariant subset of ${\rm T}M$. \begin{Rem} One can easily construct a metric space out of it. We call {\it quotient Aubry set}, or {\it Mather quotient}, the metric space $(\cbA_{c},\, \bd_c)$ obtained by identifying two points in $\cA_{c}$, if their $\d_c$-pseudodistance is zero. We shall denote an element of this quotient by $\bar{x} = \{y\in \cA_{c}:\; \d_c(x,y)=0 \}$. These elements (that are also called {\it $c$-static classes}, see \cite{ContrerasIturriaga}) provide a partition of $\cA_{c}$ into compact subsets, that can be lifted to invariant subsets of ${\rm T}M$. \end{Rem} We say that an absolutely continuous curve $\g:\R \longrightarrow M$ is a $c$-{\it minimizer}, if for any interval $[a,b]$ and any other absolutely continuous curve $\g_1: [a,b] \longrightarrow M$ such that $\g(a)=\g_1(a)$ and $\g(b)= \g_1(b)$, we have $$ \int_a^b L_{\eta}(\g(t),\dot{\g}(t))dt \leq \int_a^b L_{\eta}(\g_1(t), \dot{\g}_1(t))dt\,.$$ In other words, $\g:\R \longrightarrow M$ is a $c$-minimizer, if for any $a}[d] \ar@{}[r]|\subseteq & \widetilde{\cA}_c \ar@{->}[d] \ar@{}[r]|\subseteq & \widetilde{\cN}_c \ar@{}[r]|\subseteq & \widetilde{\cE}_c \ar@{}[r]| \subseteq & {\rm T}M \ar@{->}[d]^{\pi} \\ \cM_c \ar@{}[r]|\subseteq & \cA_c \ar@{}[r]|\subseteq &&& \ar@{}[l]| \subseteq M\\} $$ where $\widetilde{\cE}_c$ is the energy level corresponding to $\a(c)$.\\ %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% Another interesting approach to the study of these invariant sets is provided by {\it weak KAM theory}. This is mainly based on the concept of ``critical'' subsolutions and ``weak'' solutions of Hamilton-Jacobi equation and can be interpreted, from a symplectic geometric point of view, as the study of particular Lagrangian graphs and their non-removable intersections (see \cite{Paternain-Siburg}). This latter approach is particularly interesting, since it relates the dynamics of the system to the geometry of the space and might potentially open the way to a ``symplectic'' definition of Aubry-Mather theory. \begin{Def}\rm A locally Lipschitz function $u: M \longrightarrow \R$ is a {\it subsolution} of $H_{\eta}(x,d_xu)=k$, with $k\in \R$, if $H_{\eta}(x,d_xu)\leq k$ for almost every $x\in M$. \end{Def} This definition makes sense because, by Rademacher's theorem, $d_xu$ exists almost everywhere. It is possible to show that there exists $c[\eta]\in \R$, such that $H_{\eta}(x,d_xu)=k$ admits no subsolutions for $k0$, there exists a $C^1$ function $\tu: M \longrightarrow \R$ such that: \begin{itemize} \item[{\rm i)}] $\tu(x)=u(x)$ and $H_{\eta}(x,d_x\tu)=\a(c)$ on $\cA_{c}$; \item[{\rm ii)}] $|\tu(x)-u(x)|<\e$ and $H_{\eta}(x,d_x\tu)<\a(c)$ on $M \setminus \cA_{c}$. \end{itemize} } This implies that $C^1$ $\eta$-critical subsolutions are dense in $\cS_{\eta}$ with the $C^0$-topology. This result has been recently extended by Patrick Bernard (see \cite{Bernard}), showing that every $\eta$-critical subsolution coincides, on the Aubry set, with a $C^{1,1}$ $\eta$-critical subsolution (that is the best regularity that one can generally expect). Denoting the set of $C^1$ $\eta$-critical subsolutions by $\cS^1_{\eta}$, one can rewrite: % \be{{\cA}}^*_c = \bigcap_{u\in \cS^1_{\eta}} {\rm Graph}(\eta + du); \label{2.8}\ee % in particular, there exists a $C^{1}$ $\eta$-critical subsolution $\tilde{u}$ such that: % \be{{\cA}}^*_c = {\rm Graph}(\eta + d\tilde{u}) \cap \cE^*_c. \label{2.9}\ee % In the same way, one obtains: \beqano h_{\eta}(x,y) &=& \max_{u \in \cS^{1}_{\eta}} \left\{u(y)-u(x)\right\}\\ \d_c(x,y) &=& \max_{u,v\in \cS^{1}_{\eta}} \{(u-v)(y)-(u-v)(x)\}\,, \eeqano for any $x,\,y \in \cA_{c}$. %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%SECTION 3%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% \section{Minimizing properties of Lagrangian graphs and uniqueness results.} \label{sec3} In this section we shall prove the main results announced in Section \ref{sec1}. Let us start by proving some action minimizing properties of probability measures supported on Lagrangian graphs. Given a Lagrangian graph $\Lambda$ with Liouville class $c$, we shall say that $\Lambda$ is $c$-{\it subcritical}, or simply {\it subcritical}, if $\Lambda \subset \{(x,p) \in {\rm T}^*M:\; H(x,p)\leq \a(c)\}$, where $\a :{\rm H}^1(M;\R) \longrightarrow \R $ is Mather's $\a$-function. Given a subcritical Lagrangian graph $\Lambda$ with Liouville class $c$, we shall call $\Lambda_{crit}=\{(x,p)\in \Lambda:\; H(x,p) = \a(c)\}$ its {\it critical} part. The key result we need to prove is the following characterization of minimizing measures. \begin{Lem} \label{lemmadellacontesa} Let $\m$ be an invariant probability measure on ${\rm T}M$ and $\m^*=\cL_*\m$ its push-forward to ${\rm T}^*M$, via the Legendre transform $\cL$. Then, $\m$ is a Mather's measure if and only if ${\rm supp}\,\m^*$ is contained in the critical part of a subcritical Lagrangian graph. In particular, any invariant probability measure $\m^*$ on ${\rm T}^*M$, whose support is contained in an invariant Lagrangian graph with Liouville class $c$, is the image, via the Legendre transform, of a $c$-action minimizing measure. \end{Lem} \begin{Proof} ({\it i}) If $\m$ is a Mather's measure with cohomology class $c$, then the support of $\m^*$ is contained in $\cL(\widetilde\cM_c)\subseteq \cA^*_c$. By (\ref{2.8}), $\cA^*_c$ is an intersection of subcritical Lagrangian graphs, so ${\rm supp}\,\m^*$ is contained in at least one subcritical Lagrangian graph $\Lambda$. In particular ${\rm supp}\,\m^*$ is contained in the critical part of $\Lambda$, simply because $\cL(\widetilde\cM_c)$ is contained in the energy level ${\cal E}^*_c = \{(x,p)\in {\rm T}^*M:\; H(x,p) = \a(c)\}$, see Theorem \ref{Theograph}. ({\it ii}) Let us fix $\eta$ to be a closed $1$-form with $[\eta]=c$. Since we are assuming that $\Lambda$ is a $c$-subcritical Lagrangian graph, we can write $\Lambda=\{(x,\eta(x)+du(x))\,,\, x\in M\}$, where $u:M\to\RRR$ is $C^1$. By Theorem \ref{emmeinenne}, in order to show that $\m$ is a $c$-action minimizing measure, it is enough to show that ${\rm supp}\,\m \subseteq \widetilde{\NN}_c$, \ie that any orbit in ${\rm supp}\,\m$ is a $c$-minimizer, see (\ref{maneset}). To this purpose, let us consider $(x,v)\in{\rm supp}\,\m$ and let $\g(t)\=\p(\Phi_t(x,v))$, where $\Phi_t$ is the Euler-Lagrange flow and $\p$ the canonical projection on $M$. Given any interval $[a,b]\subset\RRR$, let us consider the difference $u(\g(b))-u(\g(a))$ and rewrite it as: % \be u(\g(b))-u(\g(a))=\int_a^b d_{\g(s)} u(\g(s)) \dot\g(s)\, ds= \int_a^b \big[L_\eta(\g(s),\dot\g(s))+H_\eta(\g(s),d_{\g(s)} u) \big] ds\,, \label{4.12}\ee % where the second equality follows from the definition of the Hamiltonian as the Legendre-Fenchel transform of the Lagrangian and the fact that $\g(s)$ is an orbit of the Euler-Lagrange flow. Note that along the orbit $H_\eta(\g(s),d_{\g(s)} u)=\a(c)$, because ${\rm supp}\,\m$ is invariant and ${\rm supp}\,\m^*$ is in the critical part of $\Lambda$. Then % \be \int_a^b L_\eta(\g(s),\dot\g(s)) ds = u(\g(b))-u(\g(a)) -\a(c) (b-a)\;.\ee % On the other hand, any other curve $\g_1:[a,b]\to M$ such that $\g_1(a)=\g(a)$ and $\g_1(b)=\g(b)$ satisfies: % \be u(\g(b))-u(\g(a))=\int_a^b d_{\g_1(s)} u(\g_1(s)) \dot\g(s)\, ds\le \int_a^b \big[L_\eta(\g_1(s),\dot\g_1(s))+H_\eta(\g_1(s),d_{\g_1(s)} u) \big] ds\label{4.14}\ee % where the second inequality follows again by the duality between Hamiltonian and Lagrangian. Note that now $H_\eta(\g_1(s),d_{\g_1(s)} u)\le \a(c)$, because $\Lambda=\{(x,\eta(x)+du(x))\}$ is subcritical. Then % \be \int_a^b L_\eta(\g_1(s),\dot\g_1(s)) ds \ge u(\g(b))-u(\g(a)) -\a(c) (b-a) \ee % and this proves that $\g$ is a $c$-minimizer. Let us finally observe that the Hamilton function on any invariant Lagrangian graph $\Lambda=\{(x,\eta+du)\}$ is a constant: $H(x,\eta+du)=k$ (see Appendix \ref{app2}). Then $u$ is a classical solution of the Hamilton-Jacobi equation with cohomology class $c$. As discussed in Section II, there is only one possible value of $k$ for which such solutions can exist, namely $k=\a(c)$, and this shows that $\Lambda$ coincides with its critical part. By the result proved in item ({\it ii}), if $\m^*$ is supported on $\Lambda$, then $\m$ is a $c$-action minimizing measure and this proves the last claim in the statement of the Lemma. \end{Proof} We can now prove the well-known uniqueness result for Lagrangian graphs supporting invariant measures of full support, in a fixed cohomology class, stated in the Introduction. \begin{Teo}\label{uniqcom} \it If $\Lambda \subset {\rm T}^*M$ is a Lagrangian graph on which the Hamiltonian dynamics admits an invariant measure $\m^*$ with full support, then $\Lambda = \cL \big(\widetilde{\cM}_c\big)= \cA_c^*$, where $c$ is the cohomology class of $\Lambda$.\ Therefore, if $\Lambda_1$ and $\Lambda_2$ are two Lagrangian graphs as above, with the same cohomology class, then $\Lambda_1=\Lambda_2$. \end{Teo} \begin{Proof} By Lemma \ref{lemmadellacontesa}, the measure $\mu=\cL^{-1}_*\m^*$ is $c$-minimizing. This means that $\LL^{-1}(\Lambda)={\rm supp}\,\m\subseteq \widetilde\cM_c\subseteq \widetilde\cA_c$, where the last inclusion follows from Theorem \ref{emmeinenne}. Note however that, by Theorems \ref{Theograph} and \ref{Mane}, $\widetilde\cM_c$ and $\widetilde\cA_c$ are graphs over their bases and, since ${\rm supp}\,\m$ is a graph over the whole $M$, it follows that % \be \LL^{-1}(\Lambda)={\rm supp}\,\m = \widetilde\cM_c = \widetilde\cA_c\,.\ee % \end{Proof} One can deduce something more from the above proof. \begin{Teo}\label{diffbeta} If $\Lambda$ and $\mu$ are as in Theorem \ref{uniqcom} and $\r$ is the rotation vector of $\m=\cL^{-1}\m^*$, then $\Lambda=\cL \big(\widetilde{\cM}^\r\big)$. Therefore, if $\Lambda_1$ and $\Lambda_2$ are two Lagrangian graphs supporting measures of full support and the same rotation vector $\rho$, then $\Lambda_1 = \Lambda_2$. Moreover, Mather's $\beta$-function is differentiable at $\r$ with $\partial \beta(\rho) = c$, where $c$ is the cohomology class of $\L$.\\ \end{Teo} \begin{Proof} The first claim follows from the fact that $\widetilde{\cM}^\rho$ is a graph over $M$ and that by definition $\widetilde{\cM}^\rho \supseteq {\rm supp}\,\m = \LL^{-1}(\Lambda).$ As far as the differentiability of $\beta$ at $\rho$ is concerned, suppose that $c'\in {\rm H}^1(M;\R)$ is a subderivative of $\beta$ at $\rho$. Therefore, $\b(\r) = \langle c', \rho \rangle - \a(c')$ and this implies that $\calM^{\rho} \subseteq \calM_{c'}$; in fact, for any $\m \in \calM^{\rho}$: $$ \int_{{\rm T}M} (L-\hat{\eta}')\, d\m = \int_{{\rm T}M} L\,d\m - \int_{{\rm T}M} \hat{\eta}'\,d\m = \b(\rho) - \langle c',\rho\rangle = -\a(c')\,, $$ where $\eta'$ is a closed $1$-form of cohomology $c'$. As a result, $\widetilde{\cM}^{\rho} = \cL^{-1}\left(\Lambda\right) \subseteq \widetilde{\cM}_{c'}$. The graph property of $\widetilde{\cM}_{c'}$ and of $\widetilde{\cA}_{c'}$ implies that $\widetilde{\cA}_{c'} = \widetilde{\cM}_{c'} = \cL^{-1} \left( \Lambda \right)$ and $\cA^*_{c'} = \Lambda $. As a consequence, $c'=c$. In fact, by Theorem \ref{astar} and Eq.(\ref{2.8}), there exists a $C^1$ function $v:M\longrightarrow \R$, such that $\Lambda = \{ (x, \eta' + d v):\; x \in M \}$, where $\eta'$ is a closed $1$-form with $[\eta']=c'$. This, by definition, means that $c'$ is the cohomology class of $\Lambda$ and therefore $c'=c$. \end{Proof} We are now in the position of proving the main uniqueness result in a homology class, stated in the Introduction. \begin{Teo}\label{uniqschw} Let $\Lambda$ be a Schwartzman strictly ergodic invariant Lagrangian graph with homology class $\rho$. The following properties are satisfied: \begin{itemize} \item[{\rm (i)}] if $\Lambda \cap \cA^*_c \neq \emptyset$, then $\Lambda=\cA^*_c$ and $c=c_{\Lambda}$, where $c_{\Lambda}$ is the cohomology class of $\Lambda$. % In particular, the set $\Lambda$ does not intersect any other %invariant Lagrangian graph. \item[{\rm (ii)}] the Mather function $\alpha $ is differentiable at $c_{\Lambda}$ and $\partial \alpha (c_{\Lambda}) = \rho$. \end{itemize} Therefore, \begin{itemize} \item[{\rm (iii)}] any invariant Lagrangian graph that carries a measure with rotation vector $\rho$ is equal to the graph $\Lambda$; \item[{\rm (iv)}] any invariant Lagrangian graph is either disjoint from $\Lambda$ or equal to $\Lambda$. \end{itemize} \end{Teo} We shall need the following Lemma. \begin{Lem}\label{lemmetto} Let $\rho,c$ be respectively an arbitrary homology class in $\rH_1(M;\R)$ and an arbitrary cohomology class $ \rH^1(M;\R)$. We have $$ {\rm (1)}\; \widetilde{\cM}^{\rho} \cap \widetilde{\cA_c} \neq \emptyset \quad \Longleftrightarrow \quad {\rm (2)}\; \widetilde{\cM}^{\rho} \subseteq \widetilde{\cA_c} \quad \Longleftrightarrow \quad {\rm (3)}\;\rho \in \partial \a(c)\,. $$ \end{Lem} {\bf Proof of Lemma \ref{lemmetto}.} The implication ${\rm (2) }\Longrightarrow {\rm (1)}$ is trivial. Let us prove that ${\rm (1)} \Longrightarrow {\rm (3)}$. If $\widetilde{\cM}^{\rho} \cap \widetilde{\cA_c}\neq 0$, then there exists a $c$-minimizing invariant measure $\m$ with rotation vector $\rho$. Let $\eta$ be a closed $1$-form with $[\eta]=c$; from the definition of $\a$ and $\b$: \beqano -\a(c) = \int_{{\rm T}M} (L-\hat{\eta})\,d\m = \int_{{\rm T}M} L\,d\m - \langle c,\rho\rangle = \b(\rho) - \langle c,\rho\rangle\,; \eeqano since $\b$ and $\a$ are convex conjugated, then $\rho$ is a subderivative of $\a$ at $c$.\\ Finally, in order to show $(3) \Longrightarrow (2)$, let us prove that any action minimizing measure with rotation vector $\rho$ is $c$-minimizing. In fact, if $\rho \in \partial \a(c)$ then $\a(c)= \langle c,\rho\rangle - \b(\rho)$; therefore for any $\m \in \calM^{\rho}$ and $\eta$ as above: $$ -\a(c) = \b(\rho) - \langle c,\rho\rangle = \int_{{\rm T}M}(L-\hat{\eta})\, d\m. $$ This proves that $\m \in \calM_c$ and concludes the proof. \qed\\ {\bf Proof of Theorem \ref{uniqschw}} ({\it i}) From Theorem \ref{uniqcom}, it follows that $\Lambda=\cA^*_{c_{\Lambda}}$. Let us show that it does not intersect any other Aubry set. Suppose by contradiction that $\Lambda$ intersects another Aubry set $\cA_{c}^*$. By Theorem \ref{diffbeta}, $\Lambda = \cL^{-1}\left(\widetilde{\cM}^{\rho} \right)$, then $\widetilde{\cM}^{\rho} \cap \widetilde{\cA}_c \neq \emptyset$ and, because of the previous lemma and the graph property of $\widetilde{\cA}_c$, we can conclude that $\cA^*_c = \Lambda$. The same argument used in the proof of Theorem \ref{diffbeta} allows us to conclude that $c=c_{\Lambda}$.\\ ({\it ii}) Suppose that $ h \in \partial \a(c_{\Lambda}) $. The previous lemma implies that $\widetilde{\cM}^{h} \subseteq \Lambda $; the Schwartzman unique ergodicity property of $\Lambda$ implies $h=\rho$. Therefore $ \alpha $ is differentiable at $c_{\Lambda}$ and $\partial \alpha (c_{\Lambda}) = \rho$. To prove ({\it iv}), let $\Lambda_1$ be an invariant Lagrangian graph, and call $c_1$ its cohomology class. If the compact invariant set $\Lambda\cap\Lambda_1$ is not empty, then we can find a probability measure $\mu^*$ invariant under the flow and whose support is contained in this intersection. Since $\mu^*$ is contained in the Lagrangian graph $\Lambda_1$, by Lemma \ref{lemmadellacontesa}, it is $c_1$-minimizing. Hence, the support of $\mu^*$ is contained in $\cA^*_{c_1}$. This shows that the intersection $\Lambda \cap \cA^*_{c_1}$ contains the support of $\mu^*$ and is therefore not empty. By ({\it i}), $\Lambda=\cA^*_{c_1}$. Moreover, note that $\cA^*_{c_1}\subseteq \L_1$, because, on the one hand, $\L_1={\rm Graph}(\eta_1+du_1)$, with $[\eta_1]=c_1$ and $u_1$ a classical solution to the Hamilton-Jacobi equation (see proof of Lemma \ref{lemmadellacontesa}), and, on the other hand, $\cA^*_{c_1}=\cap_{u\in\cS^1_\eta} {\rm Graph} (\eta_1+du_1)$, see (\ref{2.8}). Therefore, $\Lambda= \Lambda_1$, since they are both graphs over the base. To prove ({\it iii}), consider an invariant Lagrangian graph $\Lambda_1$, with cohomology class $c_1$, which carries an invariant measure $\mu^*$ whose rotation vector is $\rho$. By Lemma \ref{lemmadellacontesa}, the measure $\mu^*$ is $c_1$-minimizing. Therefore, we have $\widetilde{\cM}^\rho\cap \widetilde\cA_{c_1}\neq \emptyset$. By Lemma \ref{lemmetto}, it follows that ${\cal L}^{-1}(\Lambda)=\widetilde{\cM}^\rho\subseteq \widetilde\cA_{c_1}\subseteq {\cal L}^{-1}(\Lambda_1)$. Again, this forces the equality $\Lambda=\Lambda_1$ by the graph property.\qed \\ Finally, observe that Lemma \ref{CaracMinim}, Theorem \ref{emmeinenne} and Lemma \ref{lemmetto} imply the following property. \begin{Cor} The Mather function $\a$ is differentiable at $c$ if and only if the restriction of the Euler-Lagrange flow to $\widetilde{\cA}_c$ is Schwartzman uniquely ergodic, \ie if and only if all invariant measures supported on $\widetilde{\cA}_c$ have the same rotation vector (see Appendix \ref{app1} for the definition and a discussion of Schwartzman ergodic flows). \end{Cor} %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% %%%%%%%%%%%%%%%%%%%%%%%% SECTION 4%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% \section{Global uniqueness of KAM tori}\label{sec4} %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% In this section we motivate more precisely the problem of uniqueness of KAM tori and we prove Corollary \ref{corollary1}. We also show how to generalize Corollary \ref{corollary1} to cover the case of invariant tori belonging to the closure of the set of KAM tori. KAM theory concerns the study of existence of KAM tori (see Definition \ref{kamtorus}) in quasi-integrable Hamiltonian systems of the form $H(x,p)=H_0(p)+\e f(x,p)$, where: $(x,p)$ are local coordinates on $\TTT^n \times\RRR^n$, $\e$ is a ``small'' parameter and $f(x,p)$ a smooth function. If $\e=0$ the system is integrable, in the sense that the dynamics can be explictly solved: in particular each torus $\TTT^n\times\{p_0\}$ is invariant and the motion on it corresponds to a rotation with frequency $\r(p_0)=\frac{\partial H_0}{\partial p}(p_0)$. The question addressed by KAM theory is whether this foliation of phase space into invariant tori, on which the motion is (quasi-)periodic, persists even if $\e\neq 0$. In 1954 Kolmogorov stated (and Arnol'd and Moser proved it later in different contexts) that, in spite of the generic disappearence of the invariant submanifolds filled by periodic orbits, pointed out by Poincar\'e, for small $\e$ it is always possible to find KAM tori corresponding to ``{\it strongly non-resonant}'', \ie {\it Diophantine}, rotation vectors. Let us recall here the definition and some properties of Diophantine vectors: given $C,\t>0$, we say that $\r\in\RRR^n$ is a $(C,\t)$-Diophantine vector if and only if $C|\media{\r,\n}|\ge |\n|^{-\t}$, $\forall \n\in\ZZZ^n\setminus \{0\}$. The set of $(C,\t)$-Diophantine vectors will be denoted by $\DD(C,\t)$. Note that, if $\tn-1$, the Diophantine vectors have full measure in $\RRR^n$, that is $\lim_{R\to\io}\m_0(\cup_{C>0} \DD(C,\t)\cap B_R)/\m_0(B_R)=1$, where $\m_0$ is the Lesbegue measure and $B_R$ is the ball of radius $R$ centered at $0$; for $\t=n-1$, $\cup_{C>0} \DD(C,\t)$ has measure zero but Haussdorf dimension $n$. The celebrated KAM Theorem (in one of its several versions) not only shows the existence of such tori, but also provides an explicit method to construct them. \\ {\bf Theorem} (Kolmogorov--Arnol'd--Moser) \cite{Salamon}. {\it Let $n\ge 2$, $\t>n-1$, $C>0$, $\ell>2\t+2$, $M>0$ and $r>0$ be given. Let $B_r\in\RRR^n$ be the open ball of radius $r$ centered at the origin. Let $H\in C^\ell(\TTT^n\times B_r)$ be of the form % \be H(x,p)=H_0(p)+\e f(x,p)\label{1.2}\ee % with $|H_0|_{C^\ell}\le M$, $|f|_{C^\ell}\le M$, $\left|\frac{\partial^2 H_0} {\partial p^2}\right|\ge M^{-1}$ and $\r= \frac{\partial H_0}{\partial p}(0)\in \DD(C,\t)$. Then, for any $s<\ell- 2\t-1$, there exists $\e_0>0$ such that for any $\e\le\e_0$ the Hamiltonian {\rm(\ref{1.2})} admits a $C^{s,s+\t}$ KAM torus with rotation vector $\r$, \ie a $C^{s+\t}$ invariant torus such that the Hamiltonian flow on it is $C^s$-conjugated with a rotation with frequency $\r$.} \\ {\bf Remarks.} \\ 1) If $H\in C^\io$ then the KAM torus mentioned in the theorem above is $C^\io$. If $H$ is real analytic then the KAM torus is real analytic.\\ 2) As already mentioned, the proof of this theorem is constructive and it actually contains much more informations than those summarized in the above statement. For instance, in the analytic case, the proof consists of an iterative method allowing one to construct order by order the series defining the conjugation function (and to prove convergence of the formal series). In the differentiable case the proof is based on the idea of iteratively approximating differentiable functions by analytic ones and then using the inductive approximation scheme of the analytic case. In the differentiable case the proof provides an explicit construction of a KAM torus (however the construction {\it a priori} depends on a number of arbitrary choices one has to make along the proof -- \eg the choice of cutoffs one needs to introduce in the iterative approximation scheme).\\ 3) The invariant torus constructed in the proof of the KAM Theorem is locally unique, in the sense that for any prescribed (and admissible) $s$ there is at most one $C^{s,s+\t}$ KAM torus with rotation vector $\r$ within a $C^s$-distance $\d(n,s,C,\t)$ to the one constructed in the proof of the KAM Theorem, see \cite{BroerTakens, salamon-zehnder, Salamon}. Note that the $C^s$-distance $\d$ within which one can prove uniqueness of the KAM torus in a prescribed regularity class depends both on the irrationality properties of $\r$ and on the regularity class $s$ itself. It is then {\it a priori} possible that even for small $\e$ there exist different KAM tori {\it within a prescribed $C^1$-distance} from the one constructed in the proof of the Theorem, possibly less regular than that torus. Quite surprisingly, even in the analytic case, we are not aware of any proof of ``global'' uniqueness of the invariant analytic KAM torus with rotation vector $\r$ (of course in the analytic case the analytic torus one manages to construct is unique within the class of analytic tori -- however nothing {\it a priori} guarantees that less regular invariant tori with the same rotation vector exist). \\ The questions arisen in remark (3) is our main motivation for the study of the problem of global uniqueness of KAM tori. Our result, in the form stated in the Corollary in Section \ref{sec1}, settles the question and shows that, at least in the case of Tonelli Hamiltonians, it is not possible to have two different KAM tori with the same rotation vector. Note that the assumption of strict convexity of the Hamiltonian is necessary to exclude trivial sources of non-uniqueness: for instance, in the context of quasi-integrable Hamiltonians, global uniqueness could be lost simply because the unperturbed Hamiltonian induces a map $p\to \partial_p H_0(p)$ from actions to frequencies that is not one to one. Let us also remark that, apparently, the Hamiltonian considered in KAM Theorem is not a Tonelli Hamiltonian, since the latter, by definition, is defined globally on the whole $\TTT^n\times\RRR^n$. However any $C^\ell$ strictly convex Hamiltonian defined on $\TTT^n\times B_r$ for some $r>0$ can be extended to a global $C^\ell$ Tonelli Hamiltonian. Then in the statement of the KAM Theorem above it is actually enough to assume $H$ to be a $C^\ell$ Tonelli Hamiltonian, locally satisfying the (in)equalities listed after (\ref{1.2}).\\ Given the proof of our main results in Section \ref{sec3}, the proof of Corollary \ref{corollary1} is very simple.\\ {\bf Proof of Corollary \ref{corollary1}.} Since the Lagrangian KAM torus $\TT$ admits an invariant measure $\m^*$ of full support, which is the image via the conjugation $\f$ of the uniform measure on $\TTT^n$, then the claims follow from Theorem \ref{diffbeta}. Note that for rationally independent rotation vectors, a classical remark, generally attributed to Herman \cite{Herman} (see Appendix \ref{app2} for a proof), implies that $\TT$ is automatically Lagrangian.\qed\\ An interesting generalization of the result of Corollary \ref{corollary1} concerns the invariant tori belonging to the $C^0$-closure $\overline\Upsilon$ of the set $\Upsilon$ of all Lagrangian KAM tori. Note that, for quasi-integrable systems, $\Upsilon$ is not empty. The set $\Upsilon$ can be seen as a subset of ${\rm Lip}(\TTT^n,\RRR^n)$. This follows from Theorem \ref{diffbeta}, and from Mather's graph theorem, see Theorems \ref{Theograph}, \ref{matherrotation}, and the results in \cite{Mather91}. Moreover, any family of invariant Lagrangians graphs on which the function $\alpha$ (or $H$) is bounded gives rise to a family of functions in ${\rm Lip}(\TTT^n,\RRR^n)$ with uniformly bounded Lipschitz constant. This is because, given $\L$ in such a family and denoting by $(\eta+du)$ its graph, for any pair of points $x,y\in\TTT^n$ and any smooth curve $\g(t)$ on $\L$ connecting $x$ to $y$ with unit speed, we have that $u(x)-u(y)\le \int_0^{|x-y|}L_\eta(\g(t),\dot\g(t))+\a(c)|x-y|$, where $c=[\eta]$, see (\ref{4.14}). By Ascoli-Arzel\`a theorem, it follows that $\overline\Upsilon$ is also a subset of ${\rm Lip}(\TTT^n,\RRR^n)$, consisting of functions whose graphs are invariant $C^0$-Lagrangian tori. Herman \cite{HermanTori} showed that, for a generic Hamiltonian $H$ close enough to an integrable Hamiltonian $H_0$, the dynamics on the generic tori in $\overline \Upsilon$ is not conjugated to a rotation. These ``new'' tori therefore represent the majority, in the sense of topology, and hence most invariant tori cannot be obtained by the KAM algorithm. More precisely, Herman showed that in $\overline\Upsilon$ there exists a dense $G_\d$ set (\ie a dense countable intersection of open sets) of invariant Lagrangian graphs on which the dynamics is strictly ergodic and weakly mixing, and for which the rotation vector, in the sense of Section \ref{sec2}, is not Diophantine. These invariant graphs are therefore not obtained by the KAM theorem, however our uniqueness result do still apply to these graphs since strict ergodicity implies Schwartzman strict ergodicity. More generally, given any Tonelli Lagrangian on $\T^n$, we consider the set $\tilde \Upsilon$ of invariant Lagrangian graphs on which the dynamics of the flow is topologically conjugated to an {\it ergodic} linear flow on $\T^n$ (of course, far from the canonical integrable Lagrangian the set $\tilde \Upsilon$ may be empty). The dynamics on anyone of the invariant graph in $\tilde \Upsilon$ is strictly ergodic. Since the set of strictly ergodic flows on a compact set is a $G_\delta$ in the $C^0$ topology, see for example \cite[Corollaire 4.5] {FathiHerman}, it follows that there exists a dense $G_\delta$ subset ${\cal G}$ of the $C^0$ closure of $\tilde \Upsilon$ in ${\rm Lip}(\TTT^n,\RRR^n)$, such that the dynamics on any $\Lambda\in {\cal G}$ is strictly ergodic. Therefore we have the following proposition. \begin{Prop} There exists a dense $G_\d$ set $\cal G$ in the $C^0$ closure of $\tilde \Upsilon$ consisting of strictly ergodic invariant Lagrangian graphs. Any $\Lambda\in {\cal G}$ satisfies the following properties: \begin{itemize} \item[{\rm (i)}] the invariant graph $\Lambda$ has a well-defined rotation vector $\rho(\Lambda)$. \item[{\rm (ii)}] Any invariant Lagrangian graph that intersects $\Lambda$ coincides with $\Lambda$. \item[{\rm (iii)}] Any Lagrangian invariant graph that carries an invariant measure whose rotation is $\rho(\Lambda)$ coincides with $\Lambda$. \end{itemize} \end{Prop} %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% \appendix \appendixpage \section{Schwartzman unique and strict ergodicity}\label{app1} %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% In Section \ref{sec2} we have introduced the concept of rotation vector of a measure. This is closely related to the notion of Schwartzman asymptotic cycle of a flow, introduced by Sol Schwartzman in \cite{Schwartzman}, as a first attempt to develop an algebraic topological approach to the study of dynamics. In particular, we would like to provide some examples and investigate some properties of what we call {\it Schwartzman uniquely ergodic flows}. One can give a different description of the Schwartzman asymptotic cycle of a flow. This is also known as the flux homomorphism in volume preserving and symplectic geometry, see \cite{Banyaga}[Chapter 3]. We will use the description given in \cite{Fathi80}[pages 67-70]. This definition has the technical advantadge of not relying on the Krylov-Bogolioubov theory of generic orbits in a dynamical system, although a more geometrical definition showing that ``averaged'' pieces of long orbits converge almost everywhere in the first homology group for any invariant measure is certainly more heuristic and intuitive. Let us start with some standard facts. As usual we set $\T=\R/\Z$. The space $\T$ is a topological group for the addition. An important feature of $\T$ is that the canonical projection $\pi:\R\to\T$ is a covering map. Therefore given any continuous path $\gamma:[a,b]\to\T$, with $a\leq b$, we can find a continuous lift $\bar\gamma: [a,b]\to\T$ such that $\gamma=\pi\bar\gamma$. Any two such lifts differ by an integer. It follows that the quantity $\bar\gamma(b)-\bar\gamma(a)$ does not depend on the lift. We will set $${\cal V}(\gamma)=\bar\gamma(b)-\bar\gamma(a)\in\R.$$ This quantity remains constant on the homotopy class, with fixed end points, of the path $\gamma$. Morever, if $\gamma$ is a closed path, \ie we have $\gamma(a)= \gamma(b)$ then ${\cal V}(\gamma)\in\Z$, and, if such a closed path is homotopic to $0$ (with fixed endpoint) then ${\cal V}(\gamma)=0$. It is also clear that for $c\in[a,b]$, we have $${\cal V}(\gamma|[a,b])= {\cal V}(\gamma|[a,c])+{\cal V}(\gamma|[c,b]).$$ Note also that two continuous paths $\gamma_1,\gamma_2:[a,b]\to \T$ can be added by the formula $$(\gamma_1+\gamma_2)(t)= \gamma_1(t)+\gamma_2(t).$$ For this addition we have $${\cal V}(\gamma_1+\gamma_2)={\cal V}(\gamma_1)+{\cal V}(\gamma_2).$$ Another important property of the map ${\cal V}$ is its continuity on the functional space $C^0([a,b],\T)$, endowed with the topology of uniform convergence. If we call $\theta$ the closed 1-form on $\T$ whose lift to $\R$ is the usual differential form $dt$ on $\R$, where $dt$ is the differential of the identity map $\R\to\R,t\mapsto t$. It is well-known that when $\gamma:[a,b]\to \T$ is $C^1$, we have $${\cal V}(\gamma)=\int_\gamma\,\theta= \int_a^b\theta_{\gamma(t)}(\dot\gamma(t))\,dt.$$ If $X$ is a topological space, and $F:X\times [a,b]\to \T$ is a given map, we will define ${\cal V}(F):X\to \R$ by $$\forall x\in X,\ {\cal V}(F)(x)={\cal V}(F_x),$$ where $F_x:[a,b]\to\T$ is defined by $F_x(t)=F(x,t)$. The continuity of ${\cal V}$ on $C^0([a,b],\T)$ implies that ${\cal V}(F)$ is continuous. Furthermore, the continuity of ${\cal V}$ on $C^0([a,b],\T)$ also implies that the map $C^0(X\times[a,b],\T)\to C^0(X,\T),F\mapsto {\cal V}(F)$ is continuous , when we provide spaces of continuous maps with the compact open topology. If $F$ can be lifted to a continuous map $\bar F:X\times [a,b]\to \R$ with $F=\pi\bar F$, then $${\cal V}(F)(x)=\bar F(x,b)-\bar F(x,a).$$ Suppose now that $X$ is a topological space, and that $(\phi_t)_{t\in\R}$ is a continuous flow on $X$. We will define $\Phi:X\times [0,1]\to X$ by $\Phi(x,t) =\phi_t(x)$. If $f:X\to \T$ is continuous, we set $${\cal V}(f,\phi_t)={\cal V}(f\circ \Phi):X\to\R.$$ There is another another way to define ${\cal V}(f,\phi_t)$ which is used in \cite{Fathi80}. The function $F(f,\Phi):X\times[0,1]\to\T$, defined by $$F(f,\Phi)(x,t)= f(\phi_t(x))-f(x)$$ is continuous and identically $0$ on $X\times\{0\}$, it is therefore homotopic to a constant and can be lifted to a continuous map $F(f,\Phi):X\times[0,1]\to \R$, with $F(f,\Phi)|X\times\{0\}$ identically $0$. We have $${\cal V}(f,\phi_t)(x)=F(f,\Phi)(x,1).$$ Note that is $f$ is homotopic to $0$ then it can be lifted continuously to $\bar f:X\to \R$. In that case $\bar F(f,\Phi)=\bar f\Phi-\bar f$, and $${\cal V}(f,\phi_t)(x)=\bar f(\phi_1(x))-\bar f.$$ If $\mu$ is a measure with compact support invariant under the flow $\phi_t$, for a continuous $f:X\to\T$, we define ${\cal S}(\mu,\phi_t)(f)$, or simply ${\cal S}(\mu)(f)$ when $\phi_t$ is fixed, by $${\cal S}(\mu)(f)=\int_X{\cal V}(f,\phi_t)(x)\,d\mu(x).$$ It is not difficult to verify that for $f_1,f_2:X\to\T$, then $${\cal S}(\mu)(f_1+f_2)={\cal S}(\mu)(f_1)+{\cal S}(\mu)(f_2).$$ Moreover, if $f:X\to\T$ is homotopic to $0$ it can be lifted to $\bar f:X\to\R$ and $${\cal S}(\mu)(f)=\int_X[\bar f(\ph_1(x))-\bar f(x)]\,d\mu(x)=\int_X \bar f(\ph_1(x))\,d\mu(x)-\int_X\bar f(x)\,d\mu(x)=0$$ since $\mu$ is invariant by $\phi_1$. Therefore, if we denote by $[X,\T]$ the set of homotopy classes of continuous maps from $X$ to $\T$, which is an additive group, the map ${\cal S}(\mu)$ is a well-defined additive homomorphism from the additive group $[X,\T]$ to $\R$. When $X$ is a good space (like a manifold or a locally finite polyhedron), it is well-known that $[X,\T]$ is canonically identified with the first cohomology group $\rH^1(X;\Z)$. In that case ${\cal S}(\mu)$ is in $\operatorname{Hom}(\rH^1(X;\Z),\R)$. Since the first cohomology group with real coefficients $\rH^1(X;\R)$ is $\rH^1(X;\Z) \otimes \R$, we can view ${\cal S}(\mu)$ as an element of the dual $\rH^1(X;\R)^*$ of the $\R$-vector space $\rH^1(X;\R)$. When $\rH^1(X;\R)$ is finite-dimensional then $\rH^1(X;\R)^*$ is in fact equal to the first homology group $\rH_1(X;\R)$, and therefore ${\cal S}(\mu)$ defines an element of $\rH_1(X;\R)$, \ie a 1-cycle. This 1-cycle ${\cal S}(\mu)$ is called the {\it Schwartzman asymptotic cycle} of $\mu$. Note that $\rH^1(X;\R)$ is finite dimensional when $X$ is a finite poilyhedron or a compact manifold. It should be also noted that for a manifold $M$ the projection $TM\to M$ is a homotopy equivalence. Therefore $\rH^1(TM;\R)=\rH^1(M;\R)$ is finite dimensional when $M$ is a compact manifold. We now study the behavior of Schwartzman asymptotic cycles under semi-conjugacy. \begin{Prop}\label{SchSemiConj} Suppose $\phi^i_t:X_i\to X_i, i=1,2$ are two continuous flows. Suppose also that $\psi:X_1\to X_2$ is a continuous semi-conjugation between the flows, \ie $\psi\circ \phi^1_t=\phi^2_t\circ \psi$, for every $t\in\R$. Given a probability measure $\mu$ with compact support on $N_1$ invariant under $\phi^1_t$, then, for every continuous map $f:X_2\to\T$, we have $${\cal S}(\psi_*\mu,\phi^2_t)([f])={\cal S}(\mu,\phi^1_t)([f\circ \psi]),$$ where $\psi_*\mu$ is the image of $\mu$ under $\psi$. In particular, if we are in the situation where $\operatorname{Hom}([X_i,\T])\equiv \rH_1(X_i;\R), i=1,2$, we obtain $${\cal S}(\psi_*\mu,\phi^2_t)=H_1(\psi)({\cal S}(\mu,\phi^1_t)).$$ \end{Prop} \begin{Proof} Notice that $f\psi\phi^1_t(x)-f\psi(x)=f\phi^2_t(\psi(x))- f(\psi(x))$. Therefore by uniqueness of liftings ${\cal V}(f\psi,\phi^1_t)(x) ={\cal V}(f,\phi^2_t)(\psi(x))$. An integration with respect to $\mu$ finishes the proof. \end{Proof} We would like now to relate the Schwartzman asymptotic cycles to the rotation vectors $\rho(\mu)$ defined for Lagrangian flows. We first consider the case of a $C^1$ flow $\phi_t$ on the manifold $N$. We call $X$ the continuous vector field on $N$ generating $\phi_t$, \ie $$\forall x\in N, \quad X(x)={\frac {d\phi_t(x)}{dt}}\Big|_{t=0}.$$ By the flow property $\phi_{t+t'}=\phi_t\circ\phi_{t'}$, this implies $$\forall x\in N,\ \forall t\in \R, \quad \frac {d\phi_t(x)}{dt}=X(\phi_t(x)).$$ In the case of a manifold $N$, the identification of $[N,\T]$ with $\rH^1(N;\Z)$ is best described with the de Rham cohomology. We consider the natural map $I_N:[X,\T]\to \rH^1(X;\R)$ defined by $$I_N([f])=[f^*\theta],$$ where $[f]$ on the left hand side denotes the homotopy class of the $C^\infty$ map $f:N\to\T$, and $[f^*\theta]$ on the right hand side is the cohomology class of the pullback by $f$ of the closed 1-form on $\T$ whose lift to $\R$ is $dt$. Note that any homotopy class in $[N,\T]$ contains smooth maps because $C^\infty$ maps are dense in $C^0$ maps (for the Whitney topology). Therefore the map $I_N$ is indeed defined on the whole of $[N,\T]$. As it is well-known, this map $I_N$ induces an isomorphism of $[N,\T]$ on $\rH^1(N;\Z)\subset \rH^1(N;\R)=\rH^1(N;\Z)\otimes \R$. Given a $C^\infty$ map $f:N\to \T$, the $C^1$ flow $\phi_t$ on $N$, and $x\in N$, we compute ${\cal V}(f,\phi_t)(x)$. If $\gamma_x:[0,1]\to N$ is the path $t\mapsto\phi_t(x)$, by definition, we have ${\cal V}(f,\phi_t)(x)={\cal V} (f\circ\gamma_x)$. Since $\gamma_x$ is $C^1$, we get $${\cal V}(f,\phi_t)(x)=\int_{f\circ \gamma_x}\theta=\int_{\gamma_x}f^*\theta.$$ Since $\gamma_x(t)=\phi_t(x)$, we have $\dot\gamma_x(t)=X(\phi_t(x))$. It follows that $${\cal V}(f,\phi_t)(x)=\int_0^1(f^*\theta)_{\phi_t(x)}(X[\phi_t(x)])\,dt.$$ Recall that the interior product $i_X\omega$ of a differential form $\omega$ with $X$ is given by $$(i_X\omega)_x(\cdots)=\omega_x(X(x),\cdots).$$ When $\omega$ is a differential 1-form then $i_X\omega$ is a function. With this notation, we get $${\cal V}(f,\phi_t)(x)=\int_0^1(i_Xf^*\theta)(\phi_t(x))\,dt.$$ Therefore if $\mu$ is an invariant measure for $\phi_t$, which we will assume to have a compact support, we obtain $${\cal S}(\mu)=\int_N\int_0^1(i_Xf^*\theta)(\phi_t(x))\,dtd\mu(x).$$ Since $i_Xf^*\theta$ is continuous, and we are assuming that $\mu$ has a compact support, we have $${\cal S}(\mu)=\int_0^1\int_N(i_Xf^*\theta)(\phi_t(x))\,d\mu(x)dt.$$ By invariance of $\mu$ under $\phi_t$, we get $\int_N(i_Xf^*\theta)(\phi_t(x)) \,d\mu(x)= \int_N(i_Xf^*\theta)(x)\, d\mu(x)$, and therefore $${\cal S}(\mu)=\int_0^1\int_N(i_Xf^*\theta)(x)\,d\mu(x)dt=\int_N(i_Xf^*\theta) \,d\mu.$$ This shows that as an element of $\rH^1(M;\R)^*$, the Schwarztman asymptotic cycle ${\cal S}(\mu)$ is given by $${\cal S}(\mu)([\omega])=\int_Ni_X\omega\,d\mu.$$ We can now easily compute Schwartzman asypmtotic cycles for linear flows on $\T^n$. Such a flow is determined by a constant vector field $\alpha\in \R^n$ on $\T^n$ (here we use the canonical trivialisation of the tangent bundle of $\T^n$), the associated flow $R^\alpha_t: \T^n\to\T^n$ is defined by $R^\alpha_t(x)=x+[t\alpha]$, where $[t\alpha]$ is the class in $\T^n=\R^n/\Z^n$ of the vector $t\alpha\in\R^n$. If $\omega$ is a 1-form with constant coefficients, \ie $\omega=\sum_{i=1}^na_idx_i$, with $a_i\in\R$, the interior product $i_\alpha\omega$ is the constant function $\sum_{i=1}^n\alpha_ia_i$. Therefore, it follows that ${\cal S}(\mu)=\alpha\in \R^n\equiv \rH_1(\T^n;\R)$. We now compute Schwartzman asymptotic cycles for Euler-Lagrange flows. In this case $N=TM$ and $\phi_t$ is an Euler-Lagrange flow $\phi_t^L$ of some Lagrangian $L$. If we call $X_L$ the vector field generating $\phi_t^L$, since this flow is obtained from a second order ODE on $M$, we get $$\forall x\in M, \forall v\in T_xM, \quad T\pi(X_L(x,v))=v,$$ where $T\pi:T(TM)\to TM$ denotes the canonical projection. Since this projection $\pi$ is a homotopy equivalence, to compute ${\cal S}(\mu)$ we only need to consider forms of the type $\pi^*\omega$ where $\omega$ is a closed 1-form on the base $M$. In this case $(i_{X_L}\pi^*\omega) (x,v)=\omega_x(T\pi(X_L(x,v))=\omega_x(v)$. Therefore, for any measure probability measure $\mu$ on $TM$ with compact support and invariant under $\phi_t^L$, we obtain $${\cal S }(\mu)[\pi^*\omega]=\int_{TM}\omega_x(v)\,d\mu(x,v) =\int_{TM}\hat\omega\,d\mu.$$ This is precisely $\rho(\mu)$ as it was defined above in section \ref{sec2}. Note that the only property we have used is the fact that $\phi_t$ is the flow of a second order ODE on the base $M$. To simplify things, in the remainder of this appendix, we will assume that $X$ is a compact space, for which we have $[X,\T]=\rH^1(X;\Z)$, and $\rH_1(X;\Z)$ is finitely generated. In that case, the dual space $\rH^1(X;\R)^*$ is $\rH_1(X;\R)$, and for every flow $\phi_t$ on $X$ and every probability measure $\mu$ on $X$ invariant under $\phi_t$, the Schwartzman asymptotic cycle is an element of the finite dimensional-vector space $\rH_1(X;\R)$. Suppose that $x$ is a periodic point of $\phi_t$ or period $T>0$. One can define an invariant probability measure $\mu_{x,t_0}$ for $\phi_t$ by $$\int_Xg(x)\,d\mu_{x,t_0}=\frac1{t_0}\int_0^{t_0}g((\phi_t(x))\,dt,$$ where $g:X\to \R$ is a measurable function. We let the reader verify that ${\cal S}(\mu_{x,t_0})$ is equal in $\rH_1(X;\R)$ to the homology class $[\gamma_{x,t_0}]/t_0$, where $\gamma_{x,t_0}$ is the loop $t\mapsto \phi_t(x), t\in[0,t_0]$. When $x$ is a fixed point of $\phi_t$, then the Dirac mass $\delta_x$ at $x$ is invariant under $\phi_t$, and in that case ${\cal S} (\delta_x)=0$. \begin{Def}\rm For a flow $\phi_t$ on $X$, we denote by ${\cal S}(\phi_t)$ the set of all Scwartzman asymptotic cycles ${\cal S}(\mu)$, where $\mu$ is an arbitrary probability measure on $X$ invariant under $\phi_t$. \end{Def} Since $X$ is compact, note that for the weak topology the set $\calM(X)$ of probability Borel measures on $X$ is compact and convex. It is even metrizable, since we are assuming $X$ metrizable. Furthermore the subset ${\calM}(X,\phi_t)\subseteq {\calM}(X)$ of probability measures invariant under $\phi_t$ is, as it is well-known, compact convex and non empty. Therefore ${\cal S}(\phi_t)$ is a compact convex non-empty subset of $\rH_1(X;\R)$. For the case of a linear flow $R^\alpha$ on $\T^n$, we have shown above that ${\cal S}(R^\alpha_t)=\{\alpha\}\subset \R^n\equiv \rH_1(\T^n;\R)$. The following corollary is an easy consequence of Proposition \ref{SchSemiConj}. \begin{Cor}\label{SchCon} For $i=1,2$, suppose that $\phi^i_t$ is a continous flow on the compact space $X_i$, which satisfies $\operatorname{Hom}([X_i,\T], \R)\equiv \rH_1(X_i;\R)$. If $\psi:X_1\to X_2$ is a topological conjugacy between $\ph^1_t$ and $\phi^2_t$ (\ie the map $\psi$ is a homeomorphism that satisfies $\psi\phi^1_t=\phi^2_t \psi$, for all $t\in \R$), then we have $${\cal S}(\phi^2_t)=H_1(\psi)[{\cal S}(\phi^1_t)].$$ \end{Cor} We denote by ${\mathfrak F}(X)$ the set of continuous flows on $X$. We can embed ${\mathfrak F}(X)$ in $C^0(X\times [0,1],X)$ by the map $\phi_t\mapsto F^{\phi_t}\in C^0(X\times [0,1],X)$, where $$F^{\phi_t}(x,t)=\phi_t(x).$$ The topology on $C^0(X\times [0,1],X)$ is the compact open (or uniform) topology, and we endow ${\mathfrak F}(X)$ with the topology inherited from the embedding given above. \begin{Lem}The map $\phi_t\mapsto {\cal S}(\phi_t)$ is upper semi-continuous on ${\mathfrak F}(X)$. This means that for each open subset $U\subseteq \rH_1(X;\R)$, the set $\{\phi_t\in {\mathfrak F}(X)\mid {\cal S}(\phi_t)\subset U\}$ is open in ${\mathfrak F}(X)$. \end{Lem} \begin{Proof} Since the topology on $C^0(X\times [0,1],X)$ is metrizable, if this where not true we could find an open set $U\subset \rH_1(X;\R)$ and a sequence $\phi^n_t$ of continuous flows on $X$ converging uniformly to a flow $\phi_t$, with ${\cal S}(\phi_t)\subset U$, and $ {\cal S}(\phi^n_t)$ is not contained in $ U$. this means that for each $n$ we can find a probability measure $\mu_n$ on $X$ invariant under $\phi^n_t$ and such that its Schwartzman asymptotic cycle ${\cal S}(\mu_n,\phi^n_t)$ for $\phi_n^t$ is not in the open set $U$. Since ${\calM}(X)$ is compact for the weak topology, extracting a subsequence if necessary, we can assume that $\mu_n\to\mu$. It is not difficult to show that $\mu$ is invariant under the flow $\phi_t$. We now show that ${\cal S}(\mu_n, \phi^n_t)\to {\cal S}(\mu,\phi_t)$. This will yield a contradiction and finish the proof because ${\cal S}(\mu_n,\phi^n_t)$ is in the closed set $\rH_1(X;\R) \setminus U$, for every $n$, and ${\cal S}(\mu,\phi_t)\in U$. To show that the linear maps ${\cal S}(\mu_n,\phi^n_t)\in \rH_1(X;\R)= \rH^1(X;\R)^*$ converge to the linear map ${\cal S}(\mu,\phi_t)$, it suffices to show that ${\cal S}(\mu_n,\phi^n_t)([f])\to {\cal S}(\mu,\phi_t)([f])$, for every $[f]\in[X,\T]=\rH^1(X;\Z)\subset \rH^1(X;\R)=\rH^1(X;\Z)\otimes\R$. Fix now a continuous map $f:X\to \T$. Denote by $F_n, F:X\times [0,1]\to \T$ the maps defined by $$F_n(x,t)=f(\phi^n_t(x))-f(x)\text { and}\; F(x,t)=f(\phi_t(x))-f(x).$$ By the uniform continuity of $f$ on the compact metric space $X$, the sequence $F_n$ converges uniformly to $F$. Since $F_n|X\times \{0\}\equiv 0$, if we call $\tilde F_n:X\times [0,1]\to \R$ the lift of $F_n$ such that $\tilde F_n|X\times \{0\}\equiv 0$, then the sequence $\tilde F_n$ also converges uniformly to $\tilde F$, that is the lift of $F$ such that $\tilde F |X\times \{0\}\equiv 0$. Since the $\mu_n$ are probability measures, we have $$\left| \int_X\tilde F_n(x,1)\mu_n(x)-\int_X\tilde F(x,1)\mu_n(x)\right| \leq \lVert \tilde F_n-\tilde F\rVert_\infty\longrightarrow 0.$$ Since $\mu_n\to \m$ weakly, we also have $$\left|\int_X\tilde F(x,1)\mu_n(x)-\int_X\tilde F(x,1)\mu(x)\right| \longrightarrow 0.$$ Therefore ${\cal S}(\mu_n,\phi^n_t)([f])=\int_X\tilde F_n(x,1)\mu_n(x)\to {\cal S}(\mu,\phi_t)([f])=\int_X\tilde F(x,1)\mu_0(x)$. \end{Proof} \begin{Def}{\bf[Schwartzman unique ergodicity]} \rm We say that a flow $\phi_t$ is Schwartzman uniquely ergodic if ${\cal S}(\phi_t)$ is reduced to one point. \end{Def} By the computation done above linear flows on the torus $\T^n$ are Schwartzman uniquely ergodic. Of course, all uniquely ergodic flows (\ie flows having exactly one invariant probability measure) are also Schwartzman uniquely ergodic. Moreover, by Corollary \ref{SchCon}, any flow topologically conjugate to a Schwartzman uniquely ergodic flow is itself Schwartzman uniquely ergodic. \begin{Teo} The set ${\mathfrak S}(N)$ of Schwartzman uniquely ergodic flows is a $G_\delta$ in ${\mathfrak F}(X)$. \end{Teo} \begin{Proof}. Fix some norm on $\rH^1(X;\R)$. We will measure diameters of subsets of $\rH^1(X;\R)$. with respect to that norm. Fix $\epsilon>0$. Call ${\cal U}_\epsilon$ the set of flows $\phi_t$ such that the diameter of ${\cal S}(\phi_t)\subset \rH_1(X;\R)$ is $<\epsilon$. If $\phi^0_t\in {\cal U}_\epsilon$, we can find $U$ an open subset of $\rH_1(X;\R)$ of diameter $<\epsilon$ and containing ${\cal S}(\phi^0_t)$. By the lemma above the set $\{\phi_t\in {\mathfrak F}(X)\mid {\cal S}(\phi_t) \subset U\}$ is open in ${\mathfrak F}(X)$ contains $\phi^0_t$ and is contained in ${\cal U}_\epsilon$. The set of Schwartzman uniquely ergodic flows is $\cap_{n\geq 1}{\cal U}_{1/n}$. \end{Proof} \begin{Prop} Let $\f_t:X\longrightarrow X$ be a continuous flow on the compact path connected space $X$. Suppose that there exist $t_i\uparrow +\infty$ such that $\f_{t_i}{\longrightarrow} \f$ in $C(N,N)$ (with the $C^0$-topology). Then, $\f_t$ is Schwartzman uniquely ergodic. In particular, periodic flows and (uniformly) recurrent flows are Schwartzman uniquely ergodic (in both cases $\f={\rm Id}$). \end{Prop} \begin{Proof} Fix a continuous map $f:X\to\T$. Consider the function $F:X\times [0,+\infty)\to \T, (x,t)\mapsto f(\phi_t(x))-f(x)$. We have $F(x,0)=0$, for every $x\in X$. Call $\bar F:X\times [0,+\infty)\to \R$ the (unique) continuous lift of $F$ such that $\bar F(x,0)=0$, for every $x\in X$. The definition of the Schwartzman asymptotic cycle gives $${\cal S}(\mu)([f])=\int_X\bar F(x,1)\,d\mu(x),$$ for every probability measure invariant under $\phi_t$. We claim that we have $$\forall t,t'\geq 0,\forall x\in X, \quad \bar F(x,t+t')=\bar F(\phi_t(x),t') +F(x,t).$$ In fact, if we fix $t$ and we consider each side of the equality above as a (continuous) function of $(x,t')$ with values in $\R$, we see that the two sides are equal for $t'=0$, and that they both lift the function $$(x,t')\mapsto f(\phi_{t+t'}(x))-f(x)=f(\phi_t'(\phi_t(x))-f(\phi_t(x))+ f(\phi_t(x))-f(x)$$ with values in $\T$. By induction, it follows easily that $$\forall k\in\N, \quad \bar F(x, k)=\sum_{j=0}^{k-1}\bar F(\phi_j(x),1).$$ Therefore, if $t\geq 0$ and $[t]$ is its integer part, we also obtain \begin{equation*} \bar F(x,t)=\bar F(\phi_{[t]}(x),t-[t])+\sum_{j=0}^{[t]-1}\bar F(\phi_j(x),1). \tag{$*$} \end{equation*} It follows that \begin{equation*} \forall t\geq 0,\forall x\in X, \quad \lvert \bar F(x,t)\rvert\leq ([t]+1)\lVert \bar F|X\times [0,1]\rVert_\infty.\tag{$**$} \end{equation*} By compactness $\lVert \bar F|X\times [0,1]\rVert_\infty$ is finite. If we integrate equality ($*$) with respect to a probability measure $\mu$ on $X$ invariant under the flow $\phi_t$, we obtain $$\int_X\bar F(x,t)\,d\mu(x)=\int_X\bar F(x,t-[t])\,d\mu(x)+[t]\int_X\bar F(x,1)\,d\mu(x).$$ Therefore we have \begin{equation*} {\cal S}(\mu)([f])=\lim_{t\to+\infty}\int_X\frac{\bar F(x,t)}t\,d\mu(x). \tag{$***$} \end{equation*} Suppose now that we set $\gamma_x(s)=\phi_s(x)$; we have $\bar F(x,t)={\cal V}(f\gamma_x|[0,t])$. Fix now some point $x_0\in X$, and consider $t_i\to +\infty$ such that $\phi_{t_i}\to \phi$ in the $C^0$ topology. Since $\bar F(x_0,t)/t$ is bounded in absolute value by $2 \lVert \bar F|X\times [0,1]\rVert_\infty$, for $t\geq 1$, extracting a subsequence if necessary, we can assume that $\bar F(x_0,t_i)/t_i\to c\in \R$. If $x\in X$, we can find a continuous path $\gamma:[0,1]\to M$ with $\gamma(0)=x_0$ and $\gamma(1)=x$. The map $\Gamma:[0,1]\times [0,t]\to \T,(s,s')\to \phi_{s'}(\gamma(s))$ is continuous, therefore we can lift it to a continuous function with values in $\R$, and this implies the equality $${\cal V}(\Gamma|[0,1]\times\{0\})+{\cal V}(\Gamma|\{1\}\times[0,t]) -{\cal V}(\Gamma|[0,1]\times\{1\})-{\cal V}(\Gamma|\{0\}\times[0,t])=0.$$ This can be rewritten as $${\cal V}(f\gamma_x|[0,t])-{\cal V}(f\gamma_{x_0}|[0,t])={\cal V} (f\phi_t\gamma)-{\cal V}(f\gamma),$$ which translates to $$\bar F(x,t)-\bar F(x_0,t)={\cal V}(f\phi_t\gamma)-{\cal V}(f\gamma).$$ Since $\phi_{t_i}\to \phi$ uniformly, by continuity of ${\cal V}$, the left hand-side remains bounded as $t=t_i\to +\infty$. It follows that $(\bar F(x,t_i)-\bar F(x_0,t_i))/t_i\to 0$. Hence for every $x\in X$, we also have that $\bar F(x,t_i)/t_i$ tends to the same limit $c$ as $\bar F(x_0,t_i)/t_i$. Since $\bar F(x,t)/t$ is uniformly bounded for $t\geq 1$, by $(**)$, by Lebesgue's dominated convergence we obtain from ($***$) that ${\cal S}(\mu)([f])=c$, where $c$ is independent of the invariant measure $\mu$. This is of course true for any $f:X\to \T$. Therefore ${\cal S}(\mu)$ does not depend on the invariant measure $\mu$. \end{Proof} An interesting property of Schwartzman uniquely ergodic flows (which also shows that they have some kind of rigidity) is the following proposition, that follows immediately from the definition of Schwartzman unique ergodicity and what we remarked above about the asymptotic cycles of fixed and periodic points (see also \cite{Schwartzman}). \begin{Prop} Suppose that $\phi_t $ is a Schwartzman uniquely ergodic flow on $X$. If there exists either a fixed point or a closed orbit homologous to zero, then all closed orbits are homologous to zero. In the remaining case, if $C_1$ and $C_2$ are closed orbits with periods $\t_1$ and $\t_2$, then $\frac{C_1} {\t_1}$ and $\frac{C_2}{\t_2}$ are homologous. Since $[C_1]$ and $[C_2]$ are in $\rH_1(X;\Z)$, it follows in this case that the ratio of the periods of any two closed orbits must be rational. Consequently, for any continuous family of periodic orbits of $\f_t$, all orbits have the same period. \end{Prop} \begin{Def}{\bf[Schwartzman strict ergodicity]} \rm We say that a flow $\phi_t$ is Schwartzman strictly ergodic if it is Schwartzman uniquely ergodic and it has an invariant measure $\mu$ of full support (\ie $\mu(U)>0$ for every non-empty open subset $U$ of $X$). \end{Def} Linear flows on the torus $\T^n$ are Schwartzman strictly ergodic (they preserve Lebesgue measure). Of course, all strictly ergodic flows (\ie flows having exactly one invariant probability measure, and the support of this measure is full) are also Schwartzman strictly ergodic. A minimal flow which is Schwartzman uniquely ergodic is in fact Schwartzman strictly ergodic (because all invariant measures have full support). Moreover, any flow topologically conjugate to a Schwartzman strictly ergodic flow is also Schwartzman strictly ergodic. %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% \section{Lagrangian graphs and KAM tori}\label{app2} %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% In this Appendix we recall some very well-known properties of Lagrangian graphs and KAM tori used in the previous sections. We provide the standard proofs for the convenience of the reader. First of all, let us observe that, as remarked in Section \ref{sec1}, the canonical symplectic structure is intrinsically defined, \ie it does not depend on the choice of the local coordinates. In fact, it is easy to check that the Liouville form $\l$ can be equivalently defined as $$ \lambda {(x,p)} = (d\pi{(x,p)})^*p \;\in {\rm T}^*_{(x,p)}{\rm T}^*M,$$ where $\pi: {\rm T}^*M \,\longrightarrow\, M$ is the canonical projection on $M$. Now, let us prove that smooth Lagrangian graphs correspond to closed $1$-forms. \begin{Prop} Let $\Lambda=\{(x,\eta(x)),\; x\in M\}$ a smooth section of ${\rm T}^*M$. $\Lambda$ is Lagrangian if and only if $\eta$ is a closed $1$-form. \end{Prop} \begin{Proof} Let us consider: \beqano s_{\eta}: M &\longrightarrow & {\rm T}^*M\\ x &\longmapsto& (x,\eta(x))\,. \eeqano We want to prove first that $s_{\eta}^*\l = \eta$, where $\l$ is the tautological form introduced above and $s_{\eta}^*\l$ denotes its pull-back. Recalling that $\lambda{(x,p)} = (d\pi {(x,p)})^*p$, we get: \beqano (s_\eta^*\lambda){(x)} &=& (ds_{\eta}(x))^*\l{(x,p)} = (ds_{\eta}(x))^*(d \pi{(x,p)})^*\eta(x) = \\ &=& d(\pi \circ s_{\eta}{(x,p)})^* \eta(x) = \eta(x)\,, \eeqano where in the last equality we used that $\pi \circ s_{\eta}$ is the identity map. Using this property, the claim follows immediately. In fact: \beqano \Lambda\;\mbox{is Lagrangian} &\Longleftrightarrow& \omega \big|_{{\rm T}\Lambda} = 0 \quad \Longleftrightarrow \quad s_{\eta}^* \omega = 0 \quad \Longleftrightarrow \quad s_{\eta}^* d\l = 0 \\ &\Longleftrightarrow& d s_{\eta}^* \l = 0 \quad \Longleftrightarrow \quad d\eta = 0 \quad \Longleftrightarrow \quad {\eta}\; \mbox{is closed}. \eeqano \end{Proof} Let us now consider a Hamiltonian $H: {\rm T}^*M \longrightarrow \R$. %and see the relation betweenthe property of being Lagrangian and the %dynamics. \\ \begin{Prop} Let $\Lambda$ be a Lagrangian submanifold. Then $\L$ is invariant if and only if $H\big|_{\Lambda} \equiv {\rm const.}$ \end{Prop} \begin{Proof} ({\it i}) The Hamiltonian vector field $X_{H}$ is defined by $\omega(X_H,\cdot)=dH$. Since $\Lambda$ is invariant, $X_{H}\big|_{\Lambda}$ is tangent to $\Lambda$. But $\Lambda$ is Lagrangian, therefore $0=\omega(X_H,V)=dH\cdot V$ for any $V \in T\Lambda$, and this implies that $H$ is constant on $\Lambda$. ({\it ii}) Since $H$ is constant on $\L$, we have that, for every $V\in {\rm T}\L$, $0=dH\cdot V= \omega(X_H,V)$. Since $\L$ is Lagrangian, $X_H$ belongs to ${\rm T}\L$ itself and therefore $\L$ is invariant. \end{Proof} \begin{Rem} As observed in the proof of Lemma \ref{lemmadellacontesa}, if $\Lambda$ is an invariant Lagrangian graph, then the value of the constant is given by $\a(c_{\Lambda})$, where $\a$ is the $\a$-function associated to $H$ and $c_{\Lambda}$ the cohomology class of $\Lambda$. \end{Rem} To conclude, let us show that in the case of KAM tori with rationally independent rotation vectors, the condition of being Lagrangian is automatically satisfied. Actually, the same result holds in a slightly more general setting, as was observed in \cite{Herman}[page 52, Proposition 3.2], from where we borrowed the proof. \begin{Prop} \label{Lagrangiantori} Given a Hamiltonian $H$, let $\TT\subset\TTT^n\times\RRR^n$ be an invariant graph over $\TTT^n$ such that the Hamiltonian flow on $\TT$ is conjugated to a flow $R_t$ on $\TTT^n$, which is transitive, \ie with a dense orbit. Then $\TT$ is Lagrangian. \end{Prop} %$\f: \TTT^n\to \TT$ %such that $\f^{-1}\circ\Phi_t\circ\f= R_\r^t$, $\forall t\in\RRR$, %where $R_\r^t: x \to x+\r t$. \begin{Proof} Let $\f: \TTT^n\to \TT$ be the conjugation: $\f^{-1}\circ\Phi_t\circ\f= R_t$, $\forall t\in\RRR$. Consider the inclusion $i_{\TT}$ of $\TT$ into $\T^n\times \R^n$. We want to prove that $\omega\big|_{\TT} = i_{\TT}^*\omega \= 0$. Let us start by proving that the restriction of the symplectic form $i_{\TT}^*\omega$ is invariant under the Hamiltonian flow $\Phi_t$. In fact, \beqano \Phi_t^*\left(i_{\TT}^*\omega\right) = \left(i_{\TT}\circ\Phi_t\right)^* \omega = \left(\Phi_t \circ i_{\TT}\right)^*\omega = i_{\TT}^* \left(\Phi_t^*\omega \right) = i_{\TT}^*\omega\,, \eeqano where we used that $\TT$ is invariant ($\Phi_t \circ i_{\TT} = i_{\TT}\circ \Phi_t $) and $\Phi_t$ is a symplectomorphism for any $t\in \R$ (\ie $\Phi_t^*\omega = \omega$). Consider now the $1$-form on $\T^n$ given by $\omega_1 = \f^* \left(i_{\TT}^*\omega\right)$. Let us show that $\omega_1$ is invariant under $R_t$; in fact, \beqano (R_t)^*\omega_1 &=& ({R_t})^* \left(\f^*\left(i_{\TT}^*\omega\right) \right) = \left( \f \circ R_t \right)^*i_{\TT}^*\omega = \left( \Phi_t \circ \f \right)^*i_{\TT}^*\omega = \\ &=&\f^* \left( \Phi_t^* \left(i_{\TT}^*\omega \right)\right) = \f^*\left(i_{\TT}^*\omega \right) = \omega_1\,, \eeqano where we used that $\f^{-1}\circ\Phi_t\circ\f= R_t$. Since $R_t$ is transitive, then $\omega_1$ invariant implies $\omega_1$ constant: $\omega_1 = \sum_{in$ such that $f^t(\cal{U})\cap \cal{U} \neq \emptyset$. \end{Def} We will denote the set of {\it non-wandering} points for $\Phi_t$ by $\Omega(\Phi_t)$. Note that, if $\m$ is an invariant measure, then ${\rm supp}\ \m\subseteq \Omega(\Phi_t)$. In fact, by the ergodic decomposition theorem, every point $x\in{\rm supp}\ \m$ is in the support of an ergodic invariant measure $\m_1$: therefore, $x$ is non-wandering, by the ergodicity of $\m_1$. Given this remark, (\ref{app.1}) is a simple consequence of the following Proposition. \begin{Prop}\label{propnonwandering} If $M$ is a compact manifold and $L$ a Tonelli Lagrangian on ${\rm T} M$, then $\Omega \left( \Phi^L_t\big|\widetilde{\cN}_c\right) \subseteq \widetilde{\cA}_c$ for each $c\in \rH^1(M;\R)$. \end{Prop} \noindent{\bf{Remarks.}}\\ 1) Proposition \ref{propnonwandering} also shows that the Aubry set is non-empty. In fact, any continuous flow on a compact space possesses non-wandering points.\\ 2) As remarked above, every point in the support of an invariant measure $\m$ is non-wandering. Therefore, if $x\in{\rm supp}\ \m\subseteq \widetilde{\cN}_c$, then $x\in \Omega \left( \Phi^L_t\big|\widetilde{\cN}_c \right)$ and, by the above proposition, $x\in \widetilde{\cA}_c$. This proves (\ref{app.1}).\\ Before proving the above statement, we need some preliminary results. Let $u:M\longrightarrow \R$ be an $\eta$-critical subsolution, with $[\eta]=c$ (see Definition \ref{etacriticalsub}), and $\g:[a,b]\longrightarrow M$ a curve. It is easy to check (see for instance \cite{Fathibook}), that for all $a\leq t \leq t'\leq b$: \beqa{numero3} u(\g(t')) - u(\g(t)) \leq \int_t^{t'} L(\g(s),\dot{\g}(s))\,ds + \a(c)(t'-t)\,, \eeqa % see also (\ref{4.12}) and (\ref{4.14}). We will say that $\g:[a,b] \longrightarrow M$ is $(u,L,\a(c))$-{\it calibrated} on $[a,b]$ if for all $a\leq t \leq t' \leq b$ the above inequality is an equality: $$u(\g(t')) - u(\g(t)) = \int_t^{t'} L(\g(s),\dot{\g}(s))\,ds + \a(c)(t'-t)\,.$$ In \cite{Fathibook} many properties of such curves have been studied. In particular, they provide a useful characterization of the Aubry and Ma\~n\'e sets in terms of critical subsolutions. Let us denote by $\widetilde{ {\cal I}}_c(u)$ the set of points $(x,v)\in {\rm T}M$ such that the curve $\g(t)=\pi \Phi^L_t(x,v)$ is $(u,L,\a(c))$-calibrated on $\R$. In \cite{Fathibook} it is proven that: \beqa{numero2} \widetilde{\cA}_c = \bigcap_{u\in\cS^1_{\eta}} \widetilde{{\cal I}}_c(u) \quad {\rm and} \quad \widetilde{\cN}_c = \bigcup_{u\in\cS^1_{\eta}} \widetilde{{\cal I}}_c(u). \eeqa \begin{Proof}[{\bf Proposition \ref{propnonwandering}}] Without any loss of generality we can assume that $c=0$ and $\a(0)=0$. Let $(x,v) \in \Omega \left( \Phi^L_t\big|\widetilde{\cN}_0\right)$. By the definition of non-wandering point, there exist a sequence $(x_k,v_k) \in \widetilde{\cN_0}$ and $t_k \rightarrow +\infty$, such that $(x_k,v_k) \rightarrow (x,v)$ and $\Phi^L_{t_k}(x_k,v_k) \rightarrow (x,v)$ as $k\rightarrow +\infty$. From (\ref{numero2}), for each $(x_k,v_k)$ there exists a critical subsolution $u_k$, such that the curve $\g_k (t) = \pi \Phi^L_{t} (x_k,v_k)$ is $(u_k,L,0)$-calibrated. Moreover, up to extracting a subsequence, we can assume that, on any compact interval, $\g_k$ converge in the $C^1$-topology to $\g(t)=\Phi^L_t(x,v)$. Pick now any critical subsolution $u$. If we show that $\g$ is $(u,L,0)$-calibrated, using (\ref{numero2}) we can conclude that $(x,v)\in \widetilde{\cA}_0$. First of all, observe that, by the continuity of $u$, $$u(\g_k(t_k)) - u(x_k)\stackrel{k\rightarrow \infty}{\longrightarrow} 0\;.$$ Using that $\eta$-critical subsolutions are equi-Lipschitz \cite{Fathibook}, we can also conclude that $$u_k(\g_k(t_k)) - u_k(x_k)\stackrel{k\rightarrow \infty}{\longrightarrow} 0$$ % and, therefore, % \be \int_0^{t_k} L(\g_k(s),\dot{\g}_k(s))\,ds = u_k(\g_k(t_k)) - u_k(x_k) \stackrel{k\rightarrow \infty}{\longrightarrow} 0\;. \ee % Let $0\leq a\leq b$ and choose $t_k \geq b$. Observe now that \beqano u(\g_k(b)) - u(\g_k(a)) &=& u(\g_k(t_k)) - u(x_k) - \left[u(\g_k(t_k))- u(\g_k(b)) \right] - \left[u(\g_k(a))- u(x_k) \right] \geq\\ &\geq &u(\g_k(t_k)) - u(x_k) - \int_b^{t_k} L(\g_k(s),\dot{\g}_k(s))\,ds - \int_0^{a} L(\g_k(s),\dot{\g}_k(s))\,ds = \\ &=& u(\g_k(t_k)) - u(x_k) + \int_a^b L(\g_k(s),\dot{\g}_k(s))\,ds - \int_0^{t_k} L(\g_k(s),\dot{\g}_k(s))\,ds\,; \eeqano taking the limit as $k\rightarrow \infty$ on both sides, one can conclude: $$ u(\g(b)) - u(\g(a)) \geq \int_a^b L(\g(s),\dot{\g}(s))\,ds $$ and therefore, from (\ref{numero3}), it follows the equality. This shows that $\g$ is $(u,L,0)$-calibrated on $[0,\infty)$. To show that it is indeed calibrated on all $\R$, one can make a symmetric argument, letting $(y_k,w_k)= \Phi^L_{t_k}(x_k,v_k)$ play the role of $(x_k,v_k)$ in the previous argument. In fact, one has $(y_k,w_k) \rightarrow (x,v)$ and $\Phi^L_{-t_k}(y_k,w_k) \rightarrow (x,v)$ as $k\rightarrow +\infty$ and the very same argument works. \end{Proof} %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% %%%%%%%%%%%%%%%%%%%REFERENCES %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% \def\cprime{$'$} \begin{thebibliography}{10} \expandafter\ifx\csname natexlab\endcsname\relax\def\natexlab#1{#1}\fi \expandafter\ifx\csname bibnamefont\endcsname\relax \def\bibnamefont#1{#1}\fi \expandafter\ifx\csname bibfnamefont\endcsname\relax \def\bibfnamefont#1{#1}\fi \expandafter\ifx\csname citenamefont\endcsname\relax \def\citenamefont#1{#1}\fi \expandafter\ifx\csname url\endcsname\relax \def\url#1{\texttt{#1}}\fi \expandafter\ifx\csname urlprefix\endcsname\relax\def\urlprefix{URL }\fi \providecommand{\bibinfo}[2]{#2} \providecommand{\eprint}[2][]{\url{#2}} \bibitem{Arnold} Vladimir~I. Arnol'd. \newblock Proof of a theorem of {A}. {N}. {K}olmogorov on the preservation of conditionally periodic motions under a small perturbation of the {H}amiltonian. \newblock {\em Uspehi Mat. Nauk}, 18\;(5 (113)):\;13--40, 1963. \bibitem{Banyaga} Augustin Banyaga. \newblock The structure of classical diffeomorphism groups. \newblock{\em Mathematics and its Applications}, 400. Kluwer Academic Publishers Group, Dordrecht, 1997. \bibitem{Bernard} Patrick Bernard. \newblock Existence of ${C}^{1,1}$ critical sub-solutions of the {H}amilton-{J}acobi equation on compact manifolds. \newblock {\em Ann. Sci. ENS} (To appear). \bibitem{Bernardsymp} Patrick Bernard. \newblock Symplectic aspects of {A}ubry-{M}ather theory. \newblock {\em Duke Math. J.} (To appear). \bibitem{BroerTakens} Henk Broer and Floris Takens. \newblock Unicity of KAM tori. \newblock {\em Ergodic Theory Dynam. Systems}, 27(3):713--724, 2007. %\newblock Preprint, Groningen, 2005, available at %\verb+http://math.rug.nl/~broer/pdf/btun.pdf+ %http://www.math.rug.nl/$\sim$broer/pdf/btun.pdf \bibitem{Cannas} Ana Cannas~da~Silva. \newblock Lectures on symplectic geometry. \newblock {\em Lecture Notes in Mathematics}, 1764. Springer-Verlag, Berlin, 2001. \bibitem{Carneiro} Mario Jorge Dias Carneiro. \newblock On minimizing measures of the action of autonomous {L}agrangians. \newblock {\em Nonlinearity}, 8(6):1077--1085, 1995. \bibitem{CGGG} Ovidiu Costin, Guido Gentile, Alessandro Giuliani and Giovanni Gallavotti. \newblock Borel Summability and Lindstedt Series. \newblock {\em Communications in Mathematical Physics}, 269(1):175--193, 2007. \bibitem{ManeII} Gonzalo Contreras, Jorge Delgado and Renato Iturriaga. \newblock Lagrangian flows: the dynamics of globally minimizing orbits. {II}. \newblock {\em Bol. Soc. Brasil. Mat. (N.S.)}, 28(2):155--196, 1997. \bibitem{ContrerasIturriaga} Gonzalo Contreras and Renato Iturriaga. \newblock {\em Global minimizers of autonomous {L}agrangians}. \newblock 22$\sp {\rm o}$ Col\'oquio Brasileiro de Matem\'atica. [22nd Brazilian Mathematics Colloquium]. Instituto de Matem\'atica Pura e Aplicada (IMPA), Rio de Janeiro, 1999. \bibitem{Fathi80} Albert Fathi. \newblock Structure of the group of homeomorphisms preserving a good measure on a compact manifold. \newblock {\em Ann. Sci. \'Ecole Norm. Sup. (4)} 13 no. 1, 45--93, 1980. \bibitem{Fathibook} Albert Fathi. \newblock The Weak {K}{A}{M} theorem in {L}agrangian dynamics. \newblock Cambridge University Press. (To appear). \bibitem{FathiHerman} Albert Fathi and Michael~R. Herman. \newblock Existence de diff\'eomorphismes minimaux. \newblock {\em Asterisque} 49, 37--59, 1977. \bibitem{FathiSiconolfi} Albert Fathi and Antonio Siconolfi. \newblock Existence of {$C\sp 1$} critical subsolutions of the {H}amilton-{J}acobi equation. \newblock {\em Invent. Math.}, 155(2):363--388, 2004. \bibitem{GG1} Guido Gentile and Giovanni Gallavotti. \newblock Hyperbolic Low-Dimensional Invariant Tori and Summations of Divergent Series. \newblock {\em Communications in Mathematical Physics}, 227(3):421--460, 2002. \bibitem{GG2} Guido Gentile and Giovanni Gallavotti. \newblock Degenerate Elliptic Resonances. \newblock {\em Communications in Mathematical Physics}, 257(2):319--362, 2005. \bibitem{Hedlund} Gustav~A. Hedlund. \newblock Geodesics on a two-dimensional {R}iemannian manifold with periodic coefficients. \newblock {\em Ann. of Math. (2)}, 33(4):719--739, 1932. \bibitem{Herman} Michael~R. Herman. \newblock In\'egalit\'es ``a priori'' pour des tores lagrangiens invariants par des diff\'eomorphismes symplectiques. \newblock {\em Inst. Hautes \'Etudes Sci. Publ. Math.}, (70):47--101 (1990), 1989. \bibitem{HermanTori} Michael~R. Herman. \newblock On the dynamics of {L}agrangian tori invariant by symplectic diffeomorphisms. \newblock {\em Progress in variational methods in Hamiltonian systems and elliptic equations} (L'Aquila, 1990), 92--112. Pitman Res. Notes Math. Ser., 243, Longman Sci. Tech., Harlow, 1992. \bibitem{Kolmogorov} Andrei~N. Kolmogorov. \newblock On conservation of conditionally periodic motions for a small change in {H}amilton's function. \newblock {\em Dokl. Akad. Nauk SSSR (N.S.)}, 98\;:\;527--530, 1954. \bibitem{LPV} P.-L.~Lions, G.~Papanicolau and S.R.S.~Varadhan. \newblock Homogenization of Hamilton-Jacobi equation. \newblock Unpublished preprint, 1987. \bibitem{ManeI} Ricardo Ma{\~n}{\'e}. \newblock Lagrangian flows: the dynamics of globally minimizing orbits. \newblock {\em Bol. Soc. Brasil. Mat. (N.S.)}, 28(2):141--153, 1997. \bibitem{Massart} Daniel Massart. \newblock Vertices of {M}ather's {B}eta function, {I}{I}. \newblock Preprint on ArXiv, December 2006. \bibitem{Mather90} John~N. Mather. \newblock Differentiability of the minimal average action as a function of the rotation number. \newblock {\em Bol. Soc. Brasil. Mat. (N.S.)}, 21(1):59--70, 1990. \bibitem{Mather91} John~N. Mather. \newblock Action minimizing invariant measures for positive definite {L}agrangian systems. \newblock {\em Math. Z.}, 207(2):169--207, 1991. \bibitem{Mather93} John~N. Mather. \newblock Variational construction of connecting orbits. \newblock {\em Ann. Inst. Fourier (Grenoble)}, 43(5):1349--1386, 1993. \bibitem{Moser} J\"urgen Moser. \newblock On invariant curves of area-preserving mappings of an annulus. \newblock {\em Nachr. Akad. Wiss. G\"ottingen Math.-Phys. Kl. II}, 1962:1--20, 1962. \bibitem{Paternain-Siburg} Gabriel~P. Paternain, Leonid Polterovich and Karl~Friedrich Siburg. \newblock Boundary rigidity for {L}agrangian submanifolds, non-removable intersections, and {A}ubry-{M}ather theory. \newblock {\em Mosc. Math. J.}, 3(2):593--619, 745, 2003. %\newblock Dedicated to Vladimir I. Arnol'd on the occasion of his 65th %birthday. \bibitem{salamon-zehnder} D.~Salamon and E~Zehnder. \newblock K{AM} theory in configuration space. \newblock {\em Comment. Math. Helv.}, 64\;(1):\;84--132, 1989. \bibitem{Salamon} Dietmar~A. Salamon. \newblock The {K}olmogorov-{A}rnold-{M}oser theorem. \newblock {\em Math. Phys. Electron. J.}, 10:Paper 3, 37 pp. (electronic), 2004. \bibitem{Schwartzman} Sol Schwartzman. \newblock Asymptotic cycles. \newblock {\em Ann. of Math. (2)}, 66(2):270--284, 1957. \end{thebibliography} \vspace{1.truecm} {\sc Albert Fathi}\\ {\it\'Ecole normale sup\'erieure de Lyon,\\ Unit\'e de Math\'ematiques Pures et Appliqu\'ees, UMR CNRS 5669,\\ 46, all\'ee d'Italie,\\ 69364 Lyon Cedex 07 France}\\ {\sc Alessandro Giuliani}\\ {\it Dipartimento di Matematica,\\ Universit\`a degli Studi di Roma Tre,\\ L.go S. Leonardo Murialdo 1\\ 00146 Roma Italy}\\ {\sc Alfonso Sorrentino}\\ {\it Department of Mathematics,\\ Princeton University,\\ Washington Road,\\ Princeton NJ 08544 USA} \end{document} \begin{Rem} From a geometric point of view, if we consider the space ${\rm T}^*M$ equipped with the canonical symplectic form, the graph of the differential of a $C^1$ $\eta$-critical subsolution (plus the $1$-form $\eta$) is nothing else than a $c$-Lagrangian graph (\ie a Lagrangian graph with cohomology class $c$, as introduced in Section \ref{sec1}). In particular, Ma\~n\'e's $c$-critical energy level $\cE^*_{c}$ corresponds to a $(2n-1)$-dimensional hypersurface, such that the region it bounds is convex and does not contain in its interior any $c$-Lagrangian graph. In other words, all $c$-Lagrangian graphs that are contained in the region bounded by $\cE^*_c$ must intersect $\cE^*_c$ and all these intersections contain the Aubry set $\cA^*_c$ (that constitutes a sort of ``{\it non-removable intersection set}'' - see \cite{Paternain-Siburg}). It follows that the Aubry set (and also Mather set) only depends on the geometry of the critical energy level (that encodes all information about dynamics) and it follows easily from this definition that they are invariant by exact symplectomorphisms (see also \cite{Bernardsymp}). \end{Rem} \bibitem{Paternain-Siburg} Gabriel~P. Paternain, Leonid Polterovich, and Karl~Friedrich Siburg. \newblock Boundary rigidity for {L}agrangian submanifolds, non-removable intersections, and {A}ubry-{M}ather theory. \newblock {\em Mosc. Math. J.}, 3(2):593--619, 745, 2003. \newblock Dedicated to Vladimir I. Arnol'd on the occasion of his 65th birthday. \setcounter{equation}{0} \renewcommand{\theequation}{\arabic{section}.\arabic{equation}} ---------------0801230946404--