\documentclass[reqno]{amsart} \usepackage{hyperref} \AtBeginDocument{{\noindent\small \emph{Electronic Journal of Differential Equations}, Vol. 2008(2008), No. 09, pp. 1--10.\newline ISSN: 1072-6691. URL: http://ejde.math.txstate.edu or http://ejde.math.unt.edu \newline ftp ejde.math.txstate.edu (login: ftp)} \thanks{\copyright 2008 Texas State University - San Marcos.} \vspace{9mm}} \begin{document} \title[\hfilneg EJDE-2008/09\hfil Maslov index for Hamiltonian systems] {Maslov index for Hamiltonian systems} \author[A. Portaluri\hfil EJDE-2008/09\hfilneg] {Alessandro Portaluri} \address{Alessandro Portaluri\newline Department of Mathematics and Applications\\ Universit\`a di Milano-Bicocca\\ Street R. Cozzi, 53 city Milano, Italy} \email{alessandro.portaluri@unimib.it} \thanks{Submitted September 10, 2007. Published January 17, 2008.} \subjclass[2000]{53D12, 37J05} \keywords{Maslov index; Kashiwara index; Hamiltonian systems} \begin{abstract} The aim of this article is to give an explicit formula for computing the Maslov index of the fundamental solutions of linear autonomous Hamiltonian systems in terms of the Conley-Zehnder index and the map time one flow. \end{abstract} \maketitle \numberwithin{equation}{section} \newtheorem{theorem}{Theorem}[section] \newtheorem{corollary}[theorem]{Corollary} \newtheorem{lemma}[theorem]{Lemma} \newtheorem{proposition}[theorem]{Proposition} \newtheorem{definition}[theorem]{Definition} \newtheorem{remark}[theorem]{Remark} \section{Introduction}\label{sec:intro} The Maslov index is a semi-integer homotopy invariant of paths $l$ of Lagrangian subspaces of a symplectic vector space $(V, \omega)$ which gives the algebraic counts of non transverse intersections of the family $\{l(t)\}_{t \in [0,1]}$ with a given Lagrangian subspace $l_{*}$. To be more precise, let us denote by $\Lambda(V):=\Lambda(V, \omega)$ the set of all Lagrangian subspaces of the symplectic space $V$ and let $\Sigma(l_*) = \{ l \in \Lambda: l \cap l_* \not= (0)\}$ be the \emph{train\/} or the \emph{ Maslov cycle\/} of $l_*$. Then, it can be proven that $\Sigma(l_*)$ is a co-oriented one codimensional algebraic subvariety of the Lagrangian Grassmannian $\Lambda(V)$ and the Maslov index counts algebraically the number of intersections of $l$ with $\Sigma(l_*)$. This is the basic invariant out of which many others are defined. For example, if $\phi \colon [a,b] \to \mathop{\rm Sp}(V)$ is a path of symplectic automorphisms of $V$ and $l_*$ is a fixed Lagrangian subspace, then the Maslov index of $\phi$ is by definition the number of intersections of the path $[a,b]\ni t \mapsto \phi_{t}(l_*)\in \Lambda(V)$ with the train of $l_*$. The aim of this paper is to explicitly compute the Maslov index of the fundamental solution associated to \[ w'(x) = H w(x) \] where $H$ is a (real) constant Hamiltonian matrix. The idea in order to perform our computation is to relate the Maslov index with the Conley-Zehnder index and then to compute an arising correction term which is written in terms of an invariant of a triple of Lagrangian subspaces also known in literature as Kashiwara index. We remark that the result is not new and it was already proven in \cite{MarPicTau01}. However the contribution of this paper is to provide a different and we hope a simpler proof of this formula. \section{Preliminaries}\label{sec:symplecticgroup} The purpose of this section is to recall some well-known facts about the geometry of the Lagrangian Grassmannian and the Maslov index needed in our computation. For further details see for instance \cite{Abbo1, Dui76, GiaPicPor, MarPicTau01,RobSal}. \begin{definition} \label{def1.1} \rm Let $V$ be a finite dimensional real vector space. A symplectic form $\omega$ is a non degenerate anti-symmetric bilinear form on $V$. A symplectic vector space is a pair $(V, \omega)$. \end{definition} The archetypical example of symplectic space is $(\mathbb{R}^{2n}, \omega_0)$ where the symplectic structure $\omega_0$ is defined as follows. Given the splitting $\mathbb{R}^{2n}= \mathbb{R}^n \oplus \mathbb{R}^n$ and the scalar product of $\mathbb{R}^n$ $\langle \cdot, \cdot \rangle$ then for each $z_k=(x_k, y_k) \in \mathbb{R}^n \oplus \mathbb{R}^n$ for $k=1,2$ we have \[ \omega_0(z_1, z_2) = \langle x_1, y_2 \rangle- \langle x_2, y_1 \rangle. \] This symplectic structure $\omega_0$ can be represented against the scalar product by setting $ \omega_0 (z_1, z_2) = \langle J z_1, z_2 \rangle$ for all $z_i \in \mathbb{R}^{2n}$ with $i=1,2$ where we denoted by $J$ the \emph{standard complex structure\/} of $\mathbb{R}^{2n}$ which can be written with respect to the canonical basis of $\mathbb{R}^{2n}$ as \begin{equation}\label{eq:strutcomplessa} J = \begin{pmatrix} 0 & -I_n \\ I_n & 0 \end{pmatrix} \end{equation} where $I_n$ is the $n$ by $n$ identity matrix. Given a linear subspace of the symplectic vector space $(V, \omega)$, we define the orthogonal of $W$ with respect to the symplectic form $\omega $ as the linear subspace $W^{\sharp}$ given by $ W^{\sharp}=\{v \in V : \omega(u,v)=0\; \forall u\, \in W\}$. \begin{definition} \label{def1.2} \rm Let $W$ be a linear subspace of $V$. Then \begin{itemize} \item[(i)] $W$ is \emph{isotropic\/} if $W\subset W^{\sharp}$; \item[(ii)] $W$ is \emph{symplectic\/} if $ W^{\sharp}\cap W= {0}$. \item[(iii)] $W$ is \emph{Lagrangian\/} if $W = W^\sharp$. \end{itemize} \end{definition} \begin{definition} \label{def1.3} \rm Let $(V_1, \omega_1)$ and $(V_2, \omega_2)$ be symplectic vector spaces. A symplectic isomorphism from $(V_1, \omega_1)$ to $(V_2, \omega_2)$ is a bijective linear map $\varphi \colon V_1 \to V_2$ such that $\varphi^* \omega_2 = \omega_1$, meaning that \[ \omega_2(\varphi(u), \varphi(v))= \omega_1(u,v), \quad \forall \, u, v \in V_1. \] In the case $(V_1, \omega_1)=(V_2, \omega_2)$, $\varphi$ is called a symplectic automorphism or symplectomorphism. \end{definition} The matrices which correspond to symplectic automorphisms of the standard symplectic space $(\mathbb{R}^{2n}, \omega_0)$ are called symplectic and they are characterized by the equation \[ A^T JA =J \] where $A^T$ denotes the adjoint of $A$. The set of all symplectic automorphisms of $(V, \omega)$ forms a group, denoted by $\mathop{\rm Sp}(V, \omega)$. The set of the symplectic matrices is a Lie group, denoted by $\mathop{\rm Sp}(2n)$. Since each symplectic vector space of dimension $2n$ is symplectically isomorphic to $(\mathbb{R}^{2n}, \omega_0)$, then $\mathop{\rm Sp}(V, \omega)$ is isomorphic to $\mathop{\rm Sp}(2n)$. The Lie algebra of $\mathop{\rm Sp}(2n)$ is \[ \mathop{\rm sp} (2n):= \{H \in L(2n): H^TJ + JH=0\} \] where $L(2n)$ is the vector space of all real matrices of order $2n$. The matrices in $\mathop{\rm sp}(2n)$ are called \emph{infinitesimally symplectic\/} or \emph{Hamiltonian\/}. \subsection{The Krein signature on $\mathop{\rm Sp}(2n)$} Following the argument given in \cite[Chapter 1, Section 1.3]{Abbo1} we briefly recall the definition of Krein signature of the eigenvalues of a symplectic matrix. In order to define the Krein signature of a symplectic matrix $A$ we shall consider $A$ as acting on $\mathbb{C}^{2n}$ in the usual way \[ A(\xi + i \eta) := A \xi + i A \eta, \quad \forall \, \xi, \eta \in \mathbb{R}^{2n} \] and we define the Hermitian form $g(\xi,\eta) := \langle G \xi, \eta \rangle$ where $G:=-i J$. The complex symplectic group $\mathop{\rm Sp}(2n, \mathbb{C})$ consists of the complex matrices $A$ such that \begin{equation}\label{eq:sullacomplessadisp} A^* GA = G \end{equation} where as usually $ A^*=\bar A^T$ denotes the transposed conjugate of $A$. \begin{definition}\label{def:signaturadokrein} \rm Let $\lambda$ be an eigenvalue on the unit circle of a complex symplectic matrix. The \emph{Krein signature\/} of $\lambda$ is the signature of the restriction of the Hermitian form $g$ to the generalized eigenspace $E_\lambda$. \end{definition} If the real symplectic matrix $A$ has an eigenvalue $\lambda$ on the unit circle of Krein signature $(p,q)$, it is often convenient to say that $A$ has $p+q$ eigenvalues $\lambda$, and that $p$ of them are Krein-positive and $q$ which are Krein-negative. Let $A$ be a semisimple symplectic matrix, meaning that the algebraic and geometric multiplicity of its eigenvalues coincides. \begin{definition}\label{def: normalisemisemplicisimplectiic3} \rm We say that $A$ is in \emph{normal form\/} if $A = A_1 \oplus \dots \oplus A_p$, where $A_i$ has one of the forms listed below: \begin{itemize} \item[(i)] $A_1 =\begin{pmatrix} \cos \alpha & -\sin \alpha \\ \sin \alpha & \cos \alpha \end{pmatrix}$, for $\alpha \in \mathbb{R}$. \item[(ii)] $A_2=\begin{pmatrix} \mu & 0 \\ 0 & \mu^{-1} \end{pmatrix}$, for $\mu \in \mathbb{R}$ and $|\mu|>1$. \item[(iii)] \[ A_3 =\begin{pmatrix} \lambda \cos \alpha &-\lambda \sin \alpha& 0& 0 \\ \lambda \sin \alpha & \lambda\cos \alpha & 0 & 0 \\ 0 & 0 &\lambda^{-1} \cos \alpha & -\lambda^{-1} \sin \alpha \\ 0 & 0& \lambda^{-1} \sin \alpha & \lambda^{- 1}\cos \alpha \end{pmatrix}, \] for $\alpha \in \mathbb{R} \backslash\pi \mathbb{Z}$, $\mu \in \mathbb{R}$, $| \mu|>1$. \end{itemize} \end{definition} \subsection{The Maslov index} In this section we define the Maslov index for Lagrangian and symplectic paths. Our basic reference is \cite{RobSal}. Given a symplectic space $(V, \omega)$, let us consider the set of its Lagrangian subspaces $\Lambda(V, \omega)$. For any $L_0\in \Lambda(V, \omega)$ fixed and for all $k=0,1,\ldots,n$ we set \[ \Lambda_k(L_0)=\big\{L\in\Lambda:\mathrm{dim}(L\cap L_0)=k\big\},\quad \Sigma(L_0)=\cup_{k=1}^n\Lambda_k(L_0). \] It can be proven that each stratum $\Lambda_k(L_0)$ is connected of codimension $\frac12k(k+1)$ in $\Lambda$. If $l \colon [a,b] \to \Lambda$ is a $C^1$-curve of Lagrangian subspaces, we say that $l$ has a \emph{crossing\/} with the {\em train\/} $\Sigma(L_0)$ of $L_0$ at the instant $t=t_0$ if $l(t_0)\in\Sigma(L_0)$. At each non transverse crossing time $t_0 \in [a,b]$ we define the \emph{crossing form\/} $\Gamma$ as the quadratic form \[ \Gamma (l, L_0, t_0)= l'(t_0)|_{l(t_0)\cap L_0} \] and we say that a crossing $t$ is called \emph{regular\/} if the crossing form is nonsingular. It is called \emph{simple\/} if it is regular and in addition $l(t_0) \in \Lambda_1(L_0)$. \begin{definition}\label{def:RobSalindex} \rm Let $l \colon [a,b] \to \Lambda$ be a smooth curve having only \emph{ regular crossings\/} we define the \emph{Maslov index\/} \begin{equation}\label{eq: MaslovdiRobSal} \mu(l, L_0):=\frac{1}{2}\mathop{\rm sign} \Gamma (l, L_0, a)+ \sum_{t\in\left]a,b\right[}\mathop{\rm sign} \Gamma (l, L_0, t)+ \frac{1}{2}\mathop{\rm sign} \Gamma (l, L_0, b) \end{equation} where the summation runs over all crossings $t$. \end{definition} For the properties of this number we refer to \cite{RobSal}. Now let $\psi \colon [a,b] \to \mathop{\rm Sp}(2n)$ be a continuous path of symplectic matrices and $L \in \Lambda(n)$ where $\Lambda(n)$ denotes the set of all Lagrangian subspaces of the symplectic space $(\mathbb{R}^{2n}, \omega_0)$. Then we define the \emph{Maslov index\/} of the $\psi$ as \[ \mu_{L}(\psi):= \mu(\psi L, L). \] Given the vertical Lagrangian subspace $L_0 = \{0\}\oplus \mathbb{R}^n$ and assuming that $\psi$ has the block decomposition \begin{equation}\label{eq: blockdecomposition} \psi(t)=\begin{pmatrix} a(t) & b(t) \\ b(t) & d(t) \end{pmatrix}, \end{equation} then the crossing form of the path of Lagrangian subspaces $\psi L_0$ at the crossing instant $t=t_0$ is the quadratic form $\Gamma(\psi, t_0)\colon \ker\, b(t_0) \to \mathbb{R}$ given by \begin{equation}\label{eq: calcolocrossingsimplettici} \Gamma(\psi, t_0)(v) = - \langle d(t_0)v , b'(t_0) v \rangle \end{equation} where $b(t_0)$ and $d(t_0)$ are the block matrices defined in \eqref{eq: blockdecomposition}. Lemma below will be crucial in our final computation. \begin{lemma}\label{thm:lagrangianiprodotto} Consider the symplectic vector space $\mathbb{R}^{2n} \times \mathbb{R}^{2n}$ equipped by the symplectic form $\overline\omega = - \omega_0 \times \omega_0$. Then \begin{equation}\label{eq:unadellaultime} \mu(\psi L, L_1) = \mu( \mathop{\rm Gr}(\psi),L \times L_1) \end{equation} where $\mathop{\rm Gr}$ denotes the graph and where $L, L_1 \in \Lambda(n)$. \end{lemma} For the proof of this result see \cite[Theorem 3.2]{RobSal}. \begin{definition}\label{def:conleyzehnder} \rm Given a continuous path of symplectic matrices $\psi$, we define the \emph{Conley-Zehnder index\/} $\mu_{CZ}(\psi)$ as \[ \mu_{CZ}(\psi):= \mu( \mathop{\rm Gr}(\psi),\Delta), \] where $\Delta \subset \mathbb{R}^{2n}\times \mathbb{R}^{2n}$ denotes the diagonal in the product space. \end{definition} Let us consider the path \begin{equation}\label{eq:dlpirmotipoformnorm} \psi_1(x) =\begin{pmatrix} \cos \alpha x & -\sin \alpha x \\ \sin \alpha x & \cos \alpha x \end{pmatrix}. \end{equation} Given the Lagrangian $L_0= \{0\} \oplus \mathbb{R}$ in $\mathbb{R}^2$ let us consider the path of Lagrangian subspaces of $\mathbb{R}^2$ given by $l_1:=\psi_1 \,L_0$. It is easy to check that the crossing points are of the form $x \in \pi\mathbb{Z}/\alpha$ so by formula \eqref{eq: calcolocrossingsimplettici} if $x_0$ is a crossing instant then we have \[ \Gamma(\psi_1, x_0)(k)= \alpha k\ \cos(\alpha x_0)k \cos(\alpha x_0)= \alpha k^2 \cos^2(\alpha x_0). \] Thus if $\alpha \not=0$, we have \[ \mathop{\rm sign}\Gamma(\psi_1, x_0)= \begin{cases} 1 & \text{if } \alpha >0 \\ -1 & \text{if } \alpha < 0. \end{cases} \] Summing up we have the following lemma. \begin{lemma}\label{thm: Maslov1normale} Let $\psi_1:[0,1] \to \mathop{\rm Sp} (2)$ be the path of symplectic matrices given in \eqref{eq:dlpirmotipoformnorm}. Then the Maslov index is given by \begin{enumerate} \item non transverse end-point $ \mu(\psi_1)=\alpha/\pi$; \item transverse end-point $\mu(\psi_1)= \big[\frac{\alpha}{\pi}\big] + 1/2$. \end{enumerate} where we have denoted by $[\cdot]$ the integer part. \end{lemma} Let $\psi(x)= e^{xH}$ be the fundamental solution of the linear system \[ z'(x)=Hz(x) \quad x \in [0,1] \] where $H$ is a semi-simple infinitesimally symplectic matrix and let us denote by $L_0'$ the vertical Lagrangian $\oplus_{j=1}^p L_0^j$ of the symplectic space $(V, \oplus_{j=1}^{p}\omega_j)$ where $\omega_j$ is the standard symplectic form in $\mathbb{R}^{2m}$ for $m =1,2$ corresponding to the decomposition of $V$ into $2$ and $4$ dimensional $\psi(1)$-invariant symplectic subspaces. Then as a direct consequence of the product property of the Maslov index the following holds. \begin{proposition}\label{thm:Teoremafinalemaslovnormaleloop} Let $e^{i\alpha_1}, \dots, e^{i\alpha_k}$ be the Krein positive eigenvalues of $\psi$. Then the Maslov index with respect to the Lagrangian $L_0'$ is given by: \begin{equation}\label{eq:formuladelmaslovnormaleperloop} \mu_{L'_0}(\psi)= \sum_{j=1}^{k}f\Big(\frac{\alpha_j}{\pi}\Big), \end{equation} where $f$ be the function which holds identity on semi-integer and is the closest semi-integer not integer otherwise. \end{proposition} \begin{remark} \label{rmk1.11} \rm By using the zero property for the Maslov index (see for instance \cite{RobSal}) a direct computation shows that the (ii) and (iii) of Definition \ref{def: normalisemisemplicisimplectiic3} do not give any non null contribution to the Maslov index. \end{remark} \subsection{The Kashiwara and H\"ormander index} The aim of this section is to discuss a different notion of Maslov index. Our basic references are \cite{Dui76}, \cite{LioVer}, \cite[Section~8]{CapLeeMil} and \cite[Section 3]{GiaPicPor}. The H\"ormander index, or four-fold index has been introduced in \cite[Chapter 10, Sect. 3.3]{Hor} who also gave an explicit formula in terms of a triple of Lagrangian subspaces, which is known in literature with the name of \emph{Kashiwara index\/} and which we now describe. Given the Lagrangians $L_1,L_2,L_3\in\Lambda(V,\omega)$, the \emph{ Kashiwara index\/} $\tau_V(L_1,L_2,L_3)$ is defined as the signature of the (symmetric bilinear form associated to the) quadratic form $Q:L_1\oplus L_2\oplus L_3\to\mathbb{R}$ given by: \begin{equation}\label{eq:formulapercalcolodiKashiwara} Q(x_1,x_2,x_3)=\omega(x_1,x_2)+\omega(x_2,x_3)+\omega(x_3,x_1). \end{equation} It is proven in \cite[Section~8]{CapLeeMil} that $\tau_V$ is the unique integer valued map on $\Lambda\times\Lambda\times\Lambda$ satisfying the following properties: \begin{itemize} \item[(P1)] (skew symmetry) If $\sigma$ is a permutation of the set $\{1,2,3\}$, \[ \tau_V(L_{\sigma(1)},L_{\sigma(2)},L_{\sigma(3)}) =\mathop{\rm sign}(\sigma)\tau_V(L_1,L_2,L_3); \] \item[(P2)] (symplectic additivity) given the symplectic spaces $(V,\omega)$, $(\widetilde V,\widetilde\omega)$, and the Lagrangians $L_1,L_2,L_3\in\Lambda(V,\omega)$, $\widetilde L_1, \widetilde L_2, \widetilde L_3 \in\Lambda(\widetilde V,\widetilde \omega)$, we have \[ \tau_{V\oplus \widetilde V}(L_1\oplus \widetilde L_1,L_2\oplus \widetilde L_2, L_3\oplus \widetilde L_3)=\tau_V(L_1,L_2,L_3)+\tau_{\widetilde V}(\widetilde L_1, \widetilde L_2,\widetilde L_3); \] \item[(P3)] (symplectic invariance) if $\phi:(V,\omega)\to(\widetilde V,\widetilde\omega)$ is a symplectomorphism, then \[ \tau_V(L_1,L_2,L_3)=\tau_{\widetilde V}(\phi(L_1),\phi(L_2),\phi(L_3)); \] \item[(P4)] (normalization) if $V=\mathbb{R}^2$ is endowed with the canonical symplectic form, and $L_1=\mathbb{R}(1,0)$, $L_2=\mathbb{R}(1,1)$, $L_3=\mathbb{R}(0,1)$, then $\tau_V(L_1,L_2,L_3)=1$. \end{itemize} Let $(V, \omega)$ be a $2n$-dimensional symplectic vector space and let $L_1, L_2, L_3$ be three Lagrangians and let us assume that $L_3$ is transversal both to $L_1$ and $L_2$. If $L_1$ and $L_2$ are transversal we can choose coordinates $z =(x,y) \in V$ in such a way that $L_1$ is defined by the equation $y=0$, $L_2$ by the equation $x=0$ and consequently $L_3$ is defined by $ y=Ax$ for some symmetric non-singular matrix $A$. We claim that \[ \tau_V(L_1, L_2, L_3)= \mathop{\rm sign}A. \] In fact, since every symplectic vector space of dimension $2n$ is symplectically isomorphic to $(\mathbb{R}^{2n}, \omega_0)$, by property [P3] on the symplectic invariance of the Kashiwara index and by equation \eqref{eq:formulapercalcolodiKashiwara}, it is enough to compute the signature of the quadratic form $Q$ for \[ x_1=(x,0), \quad x_2=(z, Az), \quad x_3=(0, y), \quad \textrm{where } x,y,z \in \mathbb{R}^n, \] and $\omega =\omega_0$. Thus, $\omega_0(x_1, x_2)= \langle z, Az \rangle$, $\omega_0(x_3, x_1)= -\langle x, y \rangle$ and $\omega_0(x_2, x_3)= \langle x, y \rangle$ and by this we conclude that \[ Q(x_1, x_2, x_3)= \langle z, Az \rangle. \] In the general case, let $K = L_1 \cap L_2$. Then $K$ is an isotropic linear subspace of the symplectic space $(V,\omega)$ and $K^\#/K:=V^K$ is a symplectic vector space with the symplectic form induced by $(V, \omega)$. If $L$ is any Lagrangian subspace in $(V, \omega)$ then $L^K = L \cap K^\# \mod K$, is a Lagrangian subspace in $V^K$. \begin{lemma}\label{thm:checipossiamoridurre} For an arbitrary subspace $K$ of $L_1 \cap L_2 + L_2 \cap L_3 + L_3 \cap L_1$, \[ \tau_V(L_1, L_2, L_3)=\tau_{V^K}(L^K_1, L^K_2, L^K_3). \] where for $i = 1,2,3$ the Lagrangian subspaces $L_i^K$ are the image of $L_i$ under the symplectic reduction \[ (K+ K^{\#}) \to V^K :=(K+ K^{\#})/(K\cap K^{\#}). \] \end{lemma} For the proof of the above lemma, see \cite[Proposition 1.5.10]{LioVer}. We will now proceed to a geometrical description of $\tau_V$ using the Maslov index for paths. To this aim we will introduce the \emph{H\"ormander index\/}. \begin{lemma}\label{thm:perdefq} Given four Lagrangians $L_0,L_1,L_0',L_1'\in\Lambda$ and any continuous curve $l:[a,b]\to\Lambda$ such that $l(a)=L_0'$ and $l(b)=L_1'$, then the value of the quantity $\mu(l,L_1)-\mu(l, L_0)$ does \emph{not\/} depend on the choice of $l$. \end{lemma} The proof of the above lemma can be found in \cite[Theorem 3.5]{RobSal}. We are now ready for defining the map $ s:\Lambda\times\Lambda\times\Lambda\times\Lambda\to \frac12\mathbb{Z}$. \begin{definition}\label{thm:deffourfold} \rm Given $L_0,L_1,L_0',L_1'\in\Lambda$, the \emph{H\"ormander index\/} $s(L_0,L_1;L_0',L_1')$ is the half-integer $\mu(l,L_1)-\mu(l, L_0)$, where $l:[a,b]\to\Lambda$ is any continuous curve joining $l(a)=L_0'$ with $l(b)=L_1'$. \end{definition} The H\"ormander's index, satisfies the following symmetries. (See, for instance \cite[Proposition 3.23]{GiaPicPor}). We can now establish the relation between the H\"ormander index $s$ and the Kashiwara index $\tau_V$.This will be made in the same way as in \cite[Section 3]{GiaPicPor}. We define $\overline{s}:\Lambda\times\Lambda\times\Lambda\to\mathbb{Z}$ by: \begin{equation}\label{eq:defqbarra} \overline{s}(L_0,L_1,L_2):=2s(L_0,L_1;L_2,L_0). \end{equation} Observe that the function $s$ is completely determined by $\overline{s}$, because of the following identity \begin{equation}\label{eq:relqqbar} \begin{aligned} &2s(L_0,L_1;L_0',L_1')=2s(L_0,L_1;L_0',L_0)+2s(L_0,L_1;L_0,L_1')\\ &=\overline{s}(L_0,L_1,L_0')-\overline{s} (L_0,L_1,L_1'). \end{aligned} \end{equation} \begin{proposition}\label{thm:q=tauV} The map $\overline{s}$ defined in \eqref{eq:defqbarra} coincides with the Kashiwara index $\tau_V$. \end{proposition} \begin{proof} By uniqueness, it suffices to prove that $\overline{s}$ satisfies the properties (P1), (P2), (P3) and (P4). See \cite{GiaPicPor} for further details. \end{proof} As a direct consequence of Proposition \ref{thm:q=tauV} and formula \eqref{eq:relqqbar} we have \begin{equation}\label{eq:relstau} s(L_0,L_1;L_0',L_1')=\frac12[\tau_V(L_0,L_1,L_0')-\tau_V (L_0,L_1,L_1')]. \end{equation} \section{The main result}\label{sec:Maslovconleyzehnder} Let $\psi$ be the fundamental solution of the linear Hamiltonian system \[ w'(x)= H w(x), \quad x \in [0,1]. \] By Lemma \ref{thm:lagrangianiprodotto} we have $\mu_{L_0}(\psi) = \mu(\mathop{\rm Gr}(\psi), L_0 \times L_0)$; hence \begin{equation}\label{eq: passaggiopsifi} \begin{aligned} \mu(\mathop{\rm Gr}(\psi), L_0 \times L_0) &= \mu(\mathop{\rm Gr}(\psi), \Delta) + s(\Delta,L_0 \times L_0, \mathop{\rm Gr}(I), \mathop{\rm Gr}(\psi(1)))\\ &= \mu_{CZ}(\psi)-\frac12\tau_V(\Delta, L_0 \times L_0, \mathop{\rm Gr}(\psi(1))) \end{aligned} \end{equation} where the last equality follows by \eqref{eq:relstau}. For one periodic loop the last term in formula \eqref{eq: passaggiopsifi} vanishes identically because of the anti-symmetry of the Kashiwara index and by the fact that $\mathop{\rm Gr}(\psi(1))= \Delta$. Thus in this case we conclude that \[ \mu_{L_0}(\psi)=\mu_{CZ}(\psi). \] From now on we assume the following transversality condition: \begin{enumerate} \item[(H)] $\psi(1)L_0 \cap L_0 = \{0\}$. \end{enumerate} Let $L= L_0 \times L_0$ and $L_2= \mathop{\rm Gr}\,(\psi(1))$. Thus we only need to compute the last term in formula \eqref{eq: passaggiopsifi} which is $-\frac12\tau_V(\Delta, L, L_2)$ where the product form can be represented with respect to the scalar product in $\mathbb{R}^{4n}$ by the matrix \[ \widetilde J=\begin{pmatrix} -J & 0 \\ 0 & J \end{pmatrix}. \] for $J$ defined in \eqref{eq:strutcomplessa}. We denote by $K$ the isotropic subspace $\Delta \cap L$; it is the set of all vectors of the form $(0, u, 0, u)$ for $u \in \mathbb{R}^n$. Moreover $K^{\#}$ is \begin{align*} K^{\#}&=\{(x, y, z, v)\in \mathbb{R}^{4n}: \overline \omega[(x, y, z, v), (0,u,0,u)^T] =0 \}\\ &=\{(x, y, x, v) : x, y, v \in \mathbb{R}^n\}. \end{align*} Identifying the quotient space $K^\#/K$ with the orthogonal complement $S_K$ of $K$ in $K^\#$ we have $ S_K= \{(t, w, t, -w) \colon t, w \in \mathbb{R}^n\};$ moreover $\Delta \cap K^{\#}=\Delta$, $L\cap K^{\#}= L$. Now if $\psi(1)$ has the following block decomposition \[ \psi(1)= \begin{pmatrix} A&B\\C&D \end{pmatrix}, \] then $ L_2 \cap K^{\#}= \{\big(r, s ,Ar + Bs, Cr + Ds\big)\colon Ar+Bs=r; r , s \in \mathbb{R}^n\}$. Since $K^{\#} = S_K \oplus K$ then the image in $S_K$ of an arbitrary point in $K^\#$ is represented by the point $[\epsilon, \eta, \epsilon, - \eta]$ where $[\cdot]$ denotes the equivalence class in the quotient space. Thus we have \begin{itemize} \item[(i)] $\Delta^K = \{[\alpha,0,\alpha,0]: \alpha \in \mathbb{R}^n\}$; \item[(ii)] $L^K=\{[0,u,0,-u]; u \in \mathbb{R}^n\}$: \item[(iii)] $L_2^K=\{[r,s,r, Cr+Ds]: Ar + Bs =r\} = \{[r,s-Ds,r, Cr+Ds]; Ar + Bs =r\}$. \end{itemize} Then we have \begin{gather*} \bar \omega(x_1, x_2)= \bar \omega\big([\alpha,0,\alpha,0],[0,u,0,-u] \big)= -2\langle \alpha, u\rangle.\\ \bar \omega(x_2,x_3)= \bar\omega\left([0,u,0,-u], \left[r,s-Ds,r,Cr\right]\right)= 2\langle u, r\rangle.\\ \bar \omega(x_3, x_1) = \bar \omega\left(\left[r,s-Ds,r,Cr\right], [\alpha,0,\alpha,0]\right)=\langle s-Ds-Cr, \alpha \rangle . \end{gather*} Hence the quadratic form $Q$ is given by \[ Q(x_1,x_2, x_3)= -2\langle\alpha, u\rangle+2\langle u, r\rangle+\langle s-Ds-Cr, \alpha\rangle, \] where $\alpha, u, r, s, \in \mathbb{R}^n$ and $Ar + Bs =r$. Due to the transversality condition $(H)$ we have $s= B^{-1}(I_n-A) r$ and by setting $2X = (I_n-D)B^{-1}(I_n -A) -C$ the quadratic form $Q$ can be written as follows \[ Q(x_1,x_2, x_3)= -2\langle\alpha, u\rangle+2\langle u, r\rangle+ 2 \langle X r, \alpha\rangle = \langle Y w, w \rangle \] for $w =(\alpha, u, r)$ and $Y$ given by \[ Y= \begin{pmatrix} 0_n & - I_n & X\\ -I_n & 0_n & I_n\\ X^T & I_n & 0_n \end{pmatrix}. \] The Cayley-Hamilton polynomial of $A$ is given by \[ p_Y(\lambda)= \lambda^3 I_n -(2I_n + XX^T)\lambda + (X + X^T). \] In order to compute the spectrum of $Y$ we prove the following result. \begin{lemma}\label{lem:lemmafava} For any symplectic block matrix of the form $ \begin{pmatrix} A&B\\C&D \end{pmatrix}$, the $n$ by $n$ matrix $2X = (I_n-D)B^{-1}(I_n -A) -C$ is symmetric. \end{lemma} \begin{proof} In fact \begin{gather*} 2X = (B^{-1}-DB^{-1})(I_n-A)-C= B^{-1} - B^{-1}A - DB^{-1} + DB^{-1}A -C \\ 2X^T = [B^T]^{-1}-[B^T]^{-1}D^T - A^T[B^T]^{-1}+ A^T[B^T]^{-1}D^T-C^T. \end{gather*} Moreover by multiplying this last equation on the right by $BB^{-1}$ we have \begin{align*} 2X^T &= ([B^T]^{-1}B-[B^T]^{-1}D^TB-A^T[B^T]^{-1}B+A^T[B^T]^{-1}D^TB-C^TB)B^{-1}\\ &= ([B^T]^{-1}B-[B^T]^{-1}B^TD-A^T[B^T]^{-1}B+A^T[B^T]^{-1}B^TD-C^TB)B^{-1}\\ &= ([B^T]^{-1}B-D-A^T[B^T]^{-1}B+A^TD-C^TB)B^{-1}\\ &= ([B^T]^{-1}B-D-A^T[B^T]^{-1}B+I_n)B^{-1}\\ &= [B^T]^{-1}-DB^{-1}- A^T[B^T]^{-1}+ B^{-1}, \end{align*} where we used the relations $A^TD-C^TB=I_n$, $A^TC=C^TA$ and finally $D^TB=B^TD$. Thus by the expression for $2X$ and this last equality it follows that in order to prove the thesis it is enough to show that \[ -B^{-1}A + DB^{-1}A -C =[B^T]^{-1}-A^T[B^T]^{-1}. \] Now we observe that by multiplying on the left the first member of the above equality by $[B^T]^{-1}B^T$, we have \begin{align*} ([B^T]^{-1}B^T)(-B^{-1}A + DB^{-1}A -C) &= [B^T]^{-1}(-B^TB^{-1}A + B^TDB^{-1}A -B^TC)\\ &= [B^T]^{-1}(-B^TB^{-1}A + D^TBB^{-1}A -B^TC)\\ &= [B^T]^{-1}(-B^TB^{-1}A + D^TA -B^TC)\\ &= -B^{-1}A+[B^T]^{-1}. \end{align*} Thus we reduced to show that $-B^{-1}A+[B^T]^{-1}=[B^T]^{-1}-A^T[B^T]^{-1}$ or which is the same to $B^{-1}A=A^T[B^T]^{-1}$. Otherwise stated since $A^T[B^T]^{-1}= (B^{-1}A)^T$, it is enough to check that the $n$ by $n$ matrix $U=B^{-1}A$ is symmetric. In fact \begin{align*} U&=I_n\cdot U = (A^TD-C^TB)B^{-1}A\\ &=A^TDB^{-1}A-C^TA\\ &=A^TDB^{-1}A-A^TC\\ &= A^T(DB^{-1}A-C); \end{align*} moreover $U^T=A^T[B^T]^{-1}$. Thus the condition $U^T=U$ reduced to show that $A^T[B^T]^{-1}=A^T(DB^{-1}A-C)$ and then the only thing to prove is that $[B^T]^{-1}=DB^{-1}A-C$. In fact by multiplying the second member of this last equality on the left by $[B^T]^{-1}B^T$, it then follows that \begin{align*} [B^T]^{-1}(B^TDB^{-1}A-B^TC) &=[B^T]^{-1}(D^TBB^{-1}A-B^TC)\\ &=[B^T]^{-1}(D^TA-B^TC)\\ &=[B^T]^{-1}I_n \end{align*} and this completes the proof of the Lemma. \end{proof} Now since by Lemma \ref{lem:lemmafava} $X$ is a symmetric matrix, there exists an $n$ by $n$ orthogonal matrix such that $M^T X M = \mathop{\rm diag}(\lambda_1, \dots, \lambda_k)$ where the eigenvalues $\lambda_j$ are counted with multiplicity. Hence in order to compute the solutions of $p_Y(\lambda)=0$ it is enough to compute the solutions of $M^T p_Y(\lambda) M=0$ which is the same to solve \[ 0=\lambda^3 -(2 + \lambda_j^2)\lambda + 2\lambda_j=(\lambda-\lambda_j)(\lambda^2 + \lambda_j \lambda -2), \ \ \textrm{for}\ \ j=1, \dots, k. \] Now the solutions of the equation $\lambda^2 + \lambda_j \lambda -2=0$ are one positive and one negative and therefore they do not give any contribution to the signature of $Y$. Thus we proved that $ \mathop{\rm sign}(Y) = \mathop{\rm sign}(X)$ and therefore \[ -\frac12\tau_{\mathbb{R}^{4n}}(\Delta, L, L_2)=-\frac12 \,\tau_{\mathbb{R}^{2n}}(\Delta^K, L^K, L_2^K)= -\frac12\, \mathop{\rm sign}X. \] Summing up the previous calculation we proved the following result. \begin{theorem}\label{thm:Teoremafinalemaslov} Let $\psi\colon [0,1] \to \mathop{\rm Sp}(2n)$ be the fundamental solution of the Hamiltonian system \[ w'(x)= Hw(x), \quad x \in [0,1] \] and let us assume that condition $(H)$ holds. Then the Maslov index of $\psi$ is \begin{equation}\label{eq:finale} \mu_{L_0}(\psi)= \mu_{CZ}(\psi) +\frac12\,\mathop{\rm sign} \widetilde X \end{equation} for $\widetilde X = C+ (D-I_n)B^{-1}(I_n -A) $. \end{theorem} \begin{corollary} \label{coro2.3} Let $e^{i \alpha_1}, \dots, e^{i \alpha_k}$ be the Krein positive purely imaginary eigenvalues of the fundamental solution $\psi(x)=e^{xH}$ counted with algebraic multiplicity and we assume that $(H)$ and $\det(\psi(1) - I_{2n})\not=0$ hold. Then the Maslov index of $\psi$ is given by \begin{equation} \mu_{L_0}(\psi)= \sum_{j=1}^{k}g\Big(\frac{\alpha_j}{\pi}\Big)+ \frac12\,\mathop{\rm sign} \widetilde X, \end{equation} where we denoted by $g$ the double integer part function which holds the identity on integers and it is the closest odd integer otherwise. \end{corollary} \begin{thebibliography}{99} \bibitem{Abbo1} A. Abbondandolo; \emph{Morse Theory for Hamiltonian Systems}, Pitman Research Notes in Mathematics, vol.\ 425, Chapman \& Hall, London, 2001. \bibitem{CapLeeMil} S. E. Cappell, R. Lee, E. Y. Miller; \emph{On the Maslov index}, Comm.\ Pure Appl.\ Math.\ {\bf 47} (1994), no.\ 2, 121--186. \bibitem{Dui76} J. J. Duistermaat; \emph{One the Morse index in Variational Calculus\/}, Adv. Math., 173-195 (1976). \bibitem{GiaPicPor} R. Giamb\`o, P.Piccione, A. Portaluri; \emph{On the Maslov index of Lagrangian paths that are not transversal to the Maslov cycle. Semi-Riemannian index theorems in the degenerate case.\/} (math.DG/0306187). \bibitem{Hor} L. H\"ormander; \emph{Fourier integral operators}, Acta Math.\ {\bf 127} (1971), 79--183. \bibitem{JavPic} M. A. Javaloyes, P. Piccione \emph{Conjugate points and Maslov index in locally symmetric semi-Riemannian manifolds.\/} Differential Geom. Appl. {\bf 24} (2006), no. 5 521-541. \bibitem{LioVer} G. Lions, M. Vergne; \emph{The Weil representation, Maslov index and theta series}, Progress in Mathematics No.\ 6, Birk\"auser, Boston--Basel, 1980. \bibitem{MarPicTau01} R. C. Nostre Marques, Paolo Piccione, Daniel V. Tausk; \emph{On the Morse and Maslov index for periodic geodesics of arbitrary causal character.\/} Differential Geom. Appl. Proc. Conf., Opava (Czech Republic), August 27-31 2001 Silesian University, Opava, 2001, 343-358. \bibitem{RobSal} J.\ Robbin, D.\ Salamon; \emph{The Maslov Index for Paths}, Topology {\bf 32}, No.\ 4 (1993), 827--844. \end{thebibliography} \section*{Addendum posted on July 19, 2008} The main contribution of the above article \cite{Por} (as stated in its introduction) is to provide a different proof of a result already proven in \cite{Nostre}. After the publication of \cite{Por}, the author was informed by Prof. Maurice de Gosson that the main result stated in \cite[Theorem 3.2]{Por} basically can be found (up to minor details) in the proofs of \cite[Proposition 3]{deGoPiccio} and \cite[Proposition 5.7]{deGo12}. \begin{thebibliography}{99} \bibitem[A1]{deGoPiccio} M. de Gosson, S. de Gosson {\em An extension of the Conley--Zehnder Index, a product formula and an application to the Weyl representation of metaplectic operators.\/} J. Math. Phys., 47 (available on line 13. December 2006), 2006. \bibitem[A2]{deGo12} M. de Gosson, S. de Gosson, P. Piccione {\em On a product formula for the Conley-Zehnder index of symplectic paths and its applications.\/} Annals of Global Analysis and Geometry (online 14 March 2008). \bibitem[A3]{Nostre} Nostre Marques, Regina C.; Piccione, Paolo; Tausk, Daniel V. {\em On the Morse and the Maslov index for periodic geodesics of arbitrary causal character.\/} Differential geometry and its applications (Opava, 2001), 343--358, Math. Publ., 3, Silesian Univ. Opava, Opava, 2001. \bibitem[A4]{Por} A. Portaluri {\em Maslov index for Hamiltonian systems\/}. Electron. J. Differential Equations, 2008 (2008), No. 09, 1-10. \end{thebibliography} \end{document}