% File: geometry.tex % Section: MATH % Title: Basic geometry % Last modified: 06.06.2002 % \documentclass[a4paper,12pt]{article} \usepackage{amsmath,amssymb} \textwidth 16cm \textheight 25cm \oddsidemargin 0cm \topmargin -1.5cm \pagestyle{empty} \begin{document} \begin{center} \large\textbf{BASIC GEOMETRY} \end{center} \vspace{.0cm} \begin{center} \large\textbf{Manifolds} \end{center} \vspace{.1cm} An $n$-dimensional differentiable manifold $M$ is a set (actually, topological space) which locally (in a vicinity of each point) behaves like ordinary Euclidean space $\mathbb{R}^n$\,. Namely, $M$ is supplied with an atlas consisting of local coordinate maps $\varphi_\alpha:U_\alpha\rightarrow \mathbb{R}^n$\,, where subsets $U_\alpha\subset M$ are open and cover the whole of $M$. On intersections $U_\alpha\cap U_\beta$\,, the two coordinate systems should be related by smooth (infinitely differentiable, or $C^\infty$) transformations, i.e., both $\varphi_\alpha\circ\varphi_\beta^{-1}$ and $\varphi_\beta\circ\varphi_\alpha^{-1}\in C^\infty$\,. In other words, any point $x\in M$ is characterized by (possibly, several) sets of local coordinates (i.e., sets of $n$ real numbers $x^i_\alpha=\varphi^i_\alpha(x)\,, \ i=1,\ldots,n$)\,, one set for each map $\varphi_\alpha$ such that $x\in U_\alpha$\,, different sets being transformed to each other by $n$ smooth invertible functions of $n$ arguments $x^{i'}_\beta\,(x^i_\alpha)$\,. Speaking about $x$, we (explicitly or implicitly) do it in terms of local coordinates provided by one of the maps $\varphi_\alpha$\,. In what follows, we tacitly use one of existent coordinate maps, omitting index $\alpha$ everywhere. A smooth map $\mathbb{R}\rightarrow M$ defines a (smooth) curve $x(t)$ on $M$ parametrized by $t\in\mathbb{R}$ with the corresponding coordinate dependence $x^i(t)$\,. We may imagine the motion along this curve with $t$ being time. The corresponding velocity appears under the name of \emph{tangent vector}, which can be defined at each point of the curve. Let, say, $x(0)=x$\,. Then \begin{equation} \label{tang} \xi^i = \frac{dx^i(t)}{dt}|_{t=0} \ \ \ \ \ \ \ \Leftrightarrow \ \ \ \ \ \ \ x^i(t) = x^i + t\xi^i + \mathcal{O}(t^2) \end{equation} are components of the tangent vector to the curve $x(t)$ at a point $x$\,. Of course, $\xi$ is also tangent (at $x$) to any other curve $x_1(t)$ such that $x_1^i(t)-x^i(t)=\mathcal{O}(t^2)$\,. Tangent vectors to $M$ at a point $x$ form an $n$-dimensional linear space denoted by $T_x M$\,. Under a change of coordinates (or, rather, of the coordinate map) tangent vectors transform as follows: \begin{equation} \label{tang'} \xi_\beta^{i'} = \frac{\partial x_\beta^{i'}} {\partial x_\alpha^i}\,\xi_\alpha^i\,, \ \ \ \ \ \ \xi_\alpha^{i} = \frac{\partial x_\alpha^i} {\partial x_\beta^{i'}}\,\xi_\beta^{i'}\,, \ \ \ \ \ \ \frac{\partial x_\alpha^i}{\partial x_\beta^{i'}}\, \frac{\partial x_\beta^{i'}}{\partial x_\alpha^j} = \delta^i_j\ . \end{equation} The right equality in (\ref{tang}) can be written only for coordinates, not for the points $x(t)$ themselves, because $M$ is not a linear space and we cannot add its points. However, we can do it with functions on $M$\,: one can write an analog of (\ref{tang}) in the coordinate-free form, \begin{equation} \label{tang1} f(x(t)) = f(x) + t\,(\xi f)_x + \mathcal{O}(t^2)\,, \ \ \ \ \ \ \ \ \ \ (\xi f)_x \doteq \frac{df(x(t))}{dt}|_{t=0} \ . \end{equation} Because of $f(x(t))=f(\{x^i(t)\})$\,, we find that \begin{equation} \label{action} (\xi f)_x = \frac{\partial f}{\partial x^i}\,\frac{dx^i(t)}{dt}|_{t=0} = \xi^i\,\partial_i f(x)\,, \ \ \ \ \ \ \xi\,x^i = \xi^i\,, \ \ \ \ \ \xi = \xi^i \partial_i\,, \ \ \ \ \ (\partial_i)^j = \delta_i^j\,, \end{equation} so that $\partial_i\equiv\partial^x_i\equiv\partial/\partial x^i$ form the basis of $T_x M$\,. Thus, the action of a tangent vector $\xi\in T_x M$ on a function $f\in C^\infty(M)$ results in a number $(\xi f)_x$ which is nothing but the rate of change of $f$ along any curve having $\xi$ as its tangent vector (at $x$)\,, or, for short, along the vector $\xi$ itself. A smooth map $\varphi:M\rightarrow N\,, \ \varphi(x)=y$\,, \ to some manifold $N$ naturally induces two other maps. The first of them, $\varphi_*:T_x M\rightarrow T_y N$\,, sends a vector $\xi\in T_x M$ tangent to a curve on $M$ to a vector $\eta=\varphi_*\xi\in T_y N$ tangent to the image (on $N$) of this curve. The second map, a `pullback' $\varphi^*:C^\infty(N)\rightarrow C^\infty(M)$\,, simply re-reads a function $f$ on $N$ as a new function $g$ on $M$ according to $g=\varphi^*f=f\circ\varphi$\,, or $g(x)=f(y(x))$\,. Since, on any curve, $g(x(t))$ and $f(y(x(t)))$ are identical functions of $t$, having thus the same small-$t$ expansion, we conclude that \begin{equation} \label{pull} (\eta f)_y = (\xi g)_x \ \ \ \ \ \ \ \text{or} \ \ \ \ \ \ \ (\varphi_*\xi)f = \xi(\varphi^* f)\,. \end{equation} In essence, these formulas describe the change of variables: \begin{equation} \label{change} x\rightarrow\varphi(x) = y(x) \ \ \ \ \ \ \Rightarrow \ \ \ \ \ \ \eta^{i'}(y) = \frac{\partial y^{i'}}{\partial x^i}\,\xi^i(x)\,, \ \ \ \ \ \varphi_*\,\partial^x_i = \frac{\partial}{\partial x^i(y)} = \frac{\partial y^{i'}}{\partial x^i}\,\partial^y_{i'}\,. \end{equation} For a composition of maps we have \begin{equation} \label{compos} M\xrightarrow{\varphi}N\xrightarrow{\psi}K\,: \ \ \ \ \ \ \ (\psi\circ\varphi)^* = \varphi^*\circ\psi^*\,, \ \ \ \ \ \ (\psi\circ\varphi)_* = \psi_*\circ\varphi_*\ . \end{equation} \vspace{.3cm} \begin{center} \large\textbf{Tensor fields} \end{center} \vspace{.1cm} Tangent vectors are elements of $T_x M$; their components $\xi^i$ carry upper indices and transform according to (\ref{tang'}) or (\ref{change}). Elements of the dual space $T^*_x M\doteq(T_x M)^*$\,, i.e., linear functions on $T_x M$, are called 1-forms; their components are labelled by lower indices. More generally, \emph{tensors} $T^r_s\in(T_x M)^{\otimes r}(T^*_x M)^{\otimes s}$ are characterized by sets of components $T^{i_1\ldots i_r}_{j_1\ldots j_s}(x)$ transforming (under $x\rightarrow y$) as \begin{equation} \label{tens} (T')^{i'_1\ldots i'_r}_{j'_1\ldots j'_s}(y) = T^{i_1\ldots i_r}_{j_1\ldots j_s}(x)\, \frac{\partial y^{i'_1}}{\partial x^{i_1}} \cdots\frac{\partial y^{i'_r}}{\partial x^{i_r}}\, \frac{\partial x^{j_1}}{\partial y^{j'_1}}\cdots \frac{\partial x^{j_s}}{\partial y^{j'_s}}\,. \end{equation} We define a \emph{tensor field} $T^r_s(x)$ on $M$ by specifying a tensor of this type at each point $x\in M$. Explicit $x$-dependence of a tensor field $T(x)$ should not be confused with its transformation properties under change of coordinates, or mappings to other manifolds; the latter are described by $T(x)\rightarrow T'(y)$ according to (\ref{tens})\,. Tensors of the type $T^0_k$ can be considered as $k$-linear functions on $(T_x M)^{\otimes k}$ through $<\!T,\xi_1\otimes\ldots\otimes\xi_k\!>= T(\xi_1,\ldots,\xi_k)$\,. Let us look more closely at the totally antisymmetric functions $\omega^k(\xi_1,\ldots,\xi_k)$ ($k$-forms)\,. They form an \emph{exterior algebra} with respect to the \emph{wedge product} operation which produces a $(k+m)$-form $\omega^k\wedge\omega^m$ via \begin{equation} \label{shuffle} (\omega^k\wedge\omega^m)(\xi_1,\ldots,\xi_{k+m}) \doteq \sum_{\text{shuffles}}\, (-)^\sigma\,\omega^k(\xi_{i_1},\ldots,\xi_{i_k})\, \omega^m(\xi_{j_1},\ldots,\xi_{j_m})\,. \end{equation} In particular, the wedge product of $n$ 1-forms is given explicitly by \begin{equation} \label{det} (\omega_1\wedge\ldots\wedge\omega_n)\,(\xi_1,\ldots,\xi_n) = \text{det}\,\omega_i(\xi_j)\,. \end{equation} Wedge product is associative and skew-commutative: \begin{equation} \label{skew} (\omega^k\wedge\omega^l)\wedge\omega^m = \omega^k\wedge(\omega^l\wedge\omega^m)\,, \ \ \ \ \ \ \ \ \ \omega^k\wedge\omega^m = (-)^{km}\,\omega^m\wedge\omega^k\,. \end{equation} While a tensor field $T^0_0(x)$ is simply a function on $M$, \, $T^1_0(x)$ is called a \emph{vector field}, and $T^0_1(x)$ a \emph{differential form}\,. The last name is justified by the following definition: \begin{equation} \label{diff} <\!df,\xi\!>_x \ \doteq\ (\xi f)_x \,=\, \xi^i\,\partial_i f(x) \ \ \ \ \Rightarrow \ \ \ \ \ <\!dx^i,\xi\!> = \xi^i\,, \ \ \ \ \ <\!dx^i,\partial_j\!> = \delta^i_j\ , \end{equation} which agrees with the conventional rule $df(x) = \partial_i f\,dx^i$\,. 1-forms $dx^i$ constitute a basis in $T^*_xM$ dual to $\{\partial_j\}$\,. Evidently, they transform like components of vectors, as it should be. So, any differential form on $M$ (a tensor field $T^0_1$) can be written down as $\omega=\omega_i(x)\,dx^i$\,. The operation $d$ here is the $k=0$ case of \emph{exterior differentiation} which generally sends $k$-forms to $(k+1)$-forms, and is uniquely determined by the following conditions: \begin{equation} \label{d} d(f\omega) = df\wedge\omega + f\,d\omega\,, \ \ \ \ d(\omega^k\wedge\omega^m) = d\omega^k\wedge\omega^m + (-)^k\omega^k\wedge d\omega^m, \ \ \ \ d\circ d = 0\,. \end{equation} Since $\{dx^{i_1}\!\wedge\ldots\wedge dx^{i_k}\}$ is the basis in the space of $k$-forms at $x$, one explicitly obtains \begin{equation} \label{} \omega^k(x) = \sum\,\omega_{i_1\ldots i_k}(x)\,dx^{i_1}\!\wedge\ldots\wedge dx^{i_k} \ \ \ \ \ \ \Rightarrow \ \ \ \ \ \ d\omega^k = \sum\,d\omega_{i_1\ldots i_k}\wedge dx^{i_1}\!\wedge\ldots\wedge dx^{i_k}\,. \end{equation} There is another general formula for $d\omega^k$ which involves the notion of the commutator of vector fields (see (\ref{commut}) below)\,: \begin{multline} \label{d2} d\omega\,(\xi_0,\xi_1,\ldots,\xi_k) = \sum_{i=0}^{k}\,(-)^i\,\xi_i (\omega\,(\xi_0,\ldots,\xi_{i-1},\xi_{i+1},\ldots,\xi_k)) \\ + \sum_{i_x \ \doteq \ <\!\omega\,,\,\varphi_*\xi\!>_y \ \ \ \ \ \ \ \ \text{or} \ \ \ \ \ \ \ \ <\!\varphi^*\omega\,,\,\xi\!> \ = \ \varphi^*\!<\!\omega\,,\,\varphi_*\xi\!>\,. \end{equation} As a simple corollary, we deduce the relation \ $\varphi^*(df)=d(\varphi^*f)$\,: \begin{equation} \label{} <\!\varphi^*(df)\,,\,\xi\!> \ = \,\varphi^*\!<\!df\,,\,\varphi_*\xi\!> \ = \,\varphi^*((\varphi_*\xi)f) \, = \,\xi(\varphi^*f) \, = \ <\!d(\varphi^*f)\,,\,\xi\!>. \end{equation} Generalization to higher tensors is straightforward: \begin{equation} \label{higher} \varphi_*(\xi\otimes\xi') = \varphi_*\xi\otimes\varphi_*\xi'\,, \ \ \ \ \ \ \varphi^*(\omega\wedge\omega') = \varphi^*\omega\wedge\varphi^*\omega'\,, \ \ \ \ \ \ \varphi^*(d\omega)=d(\varphi^*\omega)\,. \end{equation} In components, (\ref{pull1}) gives, in perfect agreement with (\ref{tens})\,,\ $(\varphi^*\omega)_i(x)=\omega_j(y)\frac{\partial y^j}{\partial x^i}$\,. Of course, it is nothing more than a mere change of variables: \begin{equation} \label{} \omega(y) = \omega_i(y)\,dy^i \ \ \Rightarrow \ \ (\varphi^*\omega)(x) = \omega(y(x)) = \omega_j(y(x))\,dy^j(x) = \omega_j(y(x))\,\frac{\partial y^j(x)}{\partial x^i}\,dx^i. \end{equation} It is instructive to recognize that underlying transformations of bases and components are accomplished by the same matrix \ $\partial y^j/\partial x^i$\,: \begin{equation} \label{} \varphi_*\partial^x_i = \frac{\partial y^j}{\partial x^i}\,\partial^y_j\ , \ \ \ \ \ (\varphi_*\xi)^j = \frac{\partial y^j}{\partial x^i}\,\xi^i\,, \ \ \ \ \ \varphi^*dy^j = \frac{\partial y^j}{\partial x^i}\,dx^i\,, \ \ \ \ \ (\varphi^*\omega)_i = \frac{\partial y^j}{\partial x^i}\,\omega_j\ . \end{equation} \vspace{.3cm} \begin{center} \large\textbf{Lie derivative} \end{center} \vspace{.1cm} By means of the action (\ref{action})\,, vector fields send functions to functions, $\xi:f(x)\rightarrow(\xi f)_x$\,. This fact can be used to endow the space of vector fields on $M$ with a Lie algebra structure. The appropriate commutator is defined by \begin{equation} \label{commut} [\xi\,,\,\eta]f \doteq \xi(\eta f) - \eta(\xi f)\,, \ \ \ \ \ \ \ \ [\xi\,,\,\eta]^i = \xi^j\,\partial_j \eta^i - \eta^j\,\partial_j \xi^i\,, \end{equation} proves to be a true vector field (a first order differential operator on functions)\,, \begin{equation} \label{dif} \xi(fg)=(\xi f)\,g+f\,\xi g \ \ \ \ \ \ \Rightarrow \ \ \ \ \ \ [\xi,\eta](fg)=([\xi,\eta]f)\,g+f\,[\xi,\eta]g \end{equation} obeys the Jacobi identity, and is respected by the map $\varphi_*$\,, \ $\varphi_*[\xi\,,\,\eta]=[\varphi_*\xi\,,\,\varphi_*\eta]$\,, as can be shown by the following computation: \begin{multline} \label{} \varphi^*((\varphi_*[\xi\,,\,\eta])f) = [\xi\,,\,\eta](\varphi^*f) = \xi(\eta(\varphi^*f)) - \eta(\xi(\varphi^*f)) = \xi\varphi^*((\varphi_*\eta)f) - \eta\varphi^*((\varphi_*\xi)f) \\ = \varphi^*((\varphi_*\xi)(\varphi_*\eta)f) - \varphi^*((\varphi_*\eta)(\varphi_*\xi)f) = \varphi^*([\varphi_*\xi\,,\,\varphi_*\eta]f)\,. \end{multline} It turns out that both (\ref{action}),\,(\ref{commut}) can be interpreted as particular cases of a certain universal action of vector fields on tensors of arbitrary rank, called a Lie derivative: \begin{equation} \label{Lie} L_\xi f \doteq \xi f\,, \ \ \ \ \ \ \ \ L_\xi\,\eta \doteq [\xi,\eta]\,. \end{equation} Generally, $L_\xi$ is a \emph{derivation} of the corresponding tensor algebra: it respects the type of a tensor, $L_\xi : T^r_s\rightarrow T^r_s$\,, satisfies the tensorial Leibniz rule, \begin{equation} \label{deriv} L_\xi\,(T\otimes T') = (L_\xi T)\otimes T' + T\otimes L_\xi T'\,, \ \ \ \ \ \ L_\xi\,(\omega\wedge\omega') = (L_\xi\omega)\wedge\omega' + \omega\wedge L_\xi\omega'\,, \end{equation} and also commutes with contraction of tensors, e.g., \begin{equation} \label{conv} L_\xi\!<\!\omega\,,\,\eta\!> \ = \ <\!L_\xi\omega\,,\,\eta\!> + <\!\omega\,,\,L_\xi\eta\!>\,, \end{equation} whence we obtain, for example, an explicit recipe of the $L_\xi$-action on $k$-forms\,: \begin{equation} \label{Lk} (L_\xi\omega^k)\,(\eta_1,\ldots,\eta_k) = \xi(\omega^k(\eta_1,\ldots,\eta_k)) - \sum_{i=1}^{k}\,\omega^k(\eta_1,\ldots,[\xi,\eta_i],\ldots,\eta_k)\,. \end{equation} Postulates (\ref{Lie})--(\ref{conv}) unambiguously define the operation $L_\xi$ on arbitrary tensors $T^r_s$\,: \begin{equation} \label{Lcomp} (L_\xi T)^{i_1\ldots i_r}_{j_1\ldots j_s} = \xi^n\partial_n T^{i_1\ldots i_r}_{j_1\ldots j_s} -\sum_{m=1}^{r}\,\partial_n\xi^{i_m}\,T^{i_1\ldots n\ldots i_r}_{j_1\ldots j_s} +\sum_{m=1}^{s}\,\partial_{j_m}\xi^n\,T^{i_1\ldots i_r}_{j_1\ldots n\ldots j_s}\,. \end{equation} The following important properties can be verified: \begin{equation} \label{Liehom} L_\xi L_\eta - L_\eta L_\xi \equiv [L_\xi\,,\,L_\eta] = L_{[\xi,\,\eta]}\ , \ \ \ \ \ \ \ L_\xi\circ d = d\circ L_\xi\,. \end{equation} At last, the \emph{interior product} \,$\iota_\xi$\,, \,a map which sends forms $\omega^{k+1}$ to $\omega^k$\,, is defined by \begin{equation} \label{inner} \iota_\xi f = 0\,, \ \ \ \ \ \ \ (\iota_\xi \omega)\,(\eta_1,\ldots,\eta_k) = \omega\,(\xi,\eta_1,\ldots,\eta_k)\,. \end{equation} It exhibits the following properties: \begin{gather} \label{iota1} \iota_\xi(\omega^k\wedge\omega^m) = \iota_\xi\omega^k\wedge\omega^m + (-)^k\omega^k\wedge \iota_\xi\omega^m, \ \ \ \ \ \ \ \ \iota_\xi\circ\iota_\eta = -\iota_\eta\circ\iota_\xi\,, \\ \label{iota2} [L_\xi\,,\,\iota_\eta] = \iota_{[\xi,\,\eta]}\,, \ \ \ \ \ \ d\circ\iota_\xi + \iota_\xi\circ d = L_\xi\,, \ \ \ \ \ \ \varphi^*\circ\iota_{\varphi_*\xi} = \iota_\xi\circ\varphi^*\,. \end{gather} To consider the (group) flows on manifolds, we have to examine the case of \emph{invertible} maps $\varphi : M\rightarrow N$ in more details. In this case, we may study both $\varphi_*$ and $\varphi^*$ uniformly, as if directed from $M$ to $N$, by introducing a map $\tilde\varphi : T^r_s(M)\longrightarrow T^r_s(N)$ as follows: \begin{equation} \label{phitild} \tilde\varphi f = (\varphi^*)^{-1}f\,, \ \ \ \ \ \tilde\varphi\,\xi = \varphi_*\,\xi\,, \ \ \ \ \ \tilde\varphi\,\omega = (\varphi^*)^{-1}\omega\,, \ \ \ \ \ \tilde\varphi\,(a\otimes b) = (\tilde\varphi\,a\otimes\tilde\varphi\,b)\,. \end{equation} Actually, the component form of $\tilde\varphi$ is given by (\ref{tens})\,, being merely a change of coordinates: \begin{equation} \label{} (\tilde\varphi f)(y) = f(x(y))\,, \ \ \ \ \ \tilde\varphi\,dx^j = \frac{\partial x^j}{\partial y^i}\,dy^i\,, \ \ \ \ \ (\tilde\varphi\,\omega)_i = \frac{\partial x^j}{\partial y^i}\,\omega_j\ . \end{equation} It follows from (\ref{compos}) that $\widetilde{\varphi\circ\rho} = \tilde\varphi\circ\tilde\rho$, and from (\ref{pull1}) that $\tilde\varphi$ commutes with the contraction: \begin{equation} \label{} \tilde\varphi<\!\omega\,,\,\xi\!> \ = \ (\varphi^*)^{-1}\!<\!\omega\,,\,\xi\!> \ = \ <\!(\varphi^*)^{-1}\omega\,,\,\varphi_*\xi\!> \ = \ <\!\tilde\varphi\omega\,,\,\tilde\varphi\xi\!>\,. \end{equation} Let now $\xi$ be a vector field on a manifold $M$. By solving differential equations, we can find its \emph{integral curves} which at each point $x$ have $\xi(x)$ as tangent vector. A \emph{flow} $\xi^t$ generated by $\xi$ is a one-parameter Abelian group of transformations on $M$ which describes the travelling along these integral curves: \begin{equation} \label{flow} \xi^t : \ x(s)\,\rightarrow \ \xi^t x(s) = x(s+t) \ \ \ \ \ \Longrightarrow \ \ \ \ \ \ \xi^s\xi^t = \xi^{s+t}, \ \ \ \ \xi^{-t} = (\xi^t)^{-1}\,. \end{equation} This definition implies $\xi^t_*\xi=\xi$ and \begin{equation} \label{fxit} \frac{d}{dt}f(\xi^t x) = (\xi f)(\xi^t x) \ \ \ \ \ \ \ \Longrightarrow \ \ \ \ \ \ \ \frac{d^n}{dt^n}f(\xi^t x) = (\xi^n\!f)(\xi^t x)\,, \end{equation} which results in the Taylor expansion characteristic to an exponential: \begin{equation} \label{expon} f(\xi^t x) = \sum_{n=0}^{\infty}\frac{t^n}{n!}\ (\xi^n\!f)_x \ \ \ \ \ \ \ \text{or} \ \ \ \ \ \ \ (\xi^t)^* f = \sum_{n=0}^{\infty}\frac{t^n}{n!}\ \xi^n\!f\,. \end{equation} These expressions for $\xi^t$ generalize the well-known exponential relations between elements of matrix groups and their Lie algebras. As a corollary, we find that \begin{equation} \label{comflows} [\xi\,,\,\eta] = 0 \ \ \ \ \ \ \Longleftrightarrow \ \ \ \ \ \ \xi^t\circ\eta^s = \eta^s\circ\xi^t\,. \end{equation} Now we can give a universal definition of the Lie derivative $L_\xi$ in terms of the flow $\xi^t$: \begin{equation} \label{Lie2} L_\xi = -\frac{d}{dt}\,\tilde{\xi^t}|_{t=0}\ . \end{equation} In particular, \,$\tilde{\xi^t} T=T \ \Leftrightarrow \ L_\xi T=0$ \,for any tensor field $T$. Owing to the properties of $L_\xi$ and $\tilde\varphi$, it suffices to verify (\ref{Lie2}) on functions and vectors. From (\ref{expon}) one obtains \begin{equation} \label{Lie2f} -\frac{d}{dt}\,\tilde{\xi^t}f|_{t=0} = -\frac{d}{dt}\,(\xi^{-t})^*f|_{t=0} = \xi f = L_\xi f\,, \end{equation} and, invoking also (\ref{pull'})\,, \begin{multline} \label{Lie2v} (-\frac{d}{dt}\,\tilde{\xi^t}\eta)f|_{t=0} = -\frac{d}{dt}\,(\xi^t_*\eta)f|_{t=0} = -\frac{d}{dt}\,(\xi^{-t})^*\,(\eta\,(\xi^t)^*f)|_{t=0} \\ = -\frac{d}{dt}\,(1-t\xi+\mathcal{O}(t^2))\,\eta\, (1+t\xi+\mathcal{O}(t^2))f|_{t=0} = \xi\eta f - \eta\xi f = (L_\xi \eta)f\,. \end{multline} At last, consider a map $\varphi : M\rightarrow N$\,. By definition, it sends integral curves (on $M$) of a vector field $\xi$ to those of $\varphi_*\xi$ (on $N$)\,. This entails \begin{equation} \label{phiflow} \varphi\circ\xi^t = (\varphi_*\xi)^t\circ\varphi\,, \ \ \ \ \ \tilde\varphi\circ L_\xi = L_{\varphi_*\xi}\circ\tilde\varphi\,, \ \ \ \ \ \tilde{\xi^t}\circ L_\xi = L_\xi\circ\tilde{\xi^t}\,. \end{equation} \vspace{.3cm} \begin{center} \large\textbf{Lie groups} \end{center} \vspace{.1cm} A Lie group $G$ \,is a manifold supplied with a differentiable group multiplication law. We denote a unity in a group by $e$, left and right shifts by $L_h$ and $R_h$\,, \begin{equation} \label{shifts} L_h : g\rightarrow hg\,, \ \ \ \ R_h : g\rightarrow gh\,, \ \ \ \ L_{gh} = L_g\circ L_h\,, \ \ \ \ R_{gh} = R_h\circ R_g\,, \ \ \ \ L_g\circ R_h = R_h\circ L_g\,. \end{equation} A vector field $A$ on $G$ is called left-invariant if $(L_g)_*A=A$\,. Such fields constitute a subalgebra of the Lie algebra of all vector fields on $G$\,. This subalgebra $\mathfrak{g}$ is called a Lie algebra of the Lie group $G$\,. Equally well we may think of $\mathfrak{g}$ as a tangent space $T_e G$ at unity: translating the vectors of this space by left shifts just recovers left-invariant fields. Let $\varphi : G\rightarrow H$ be a homomorphism of Lie groups. Then $\varphi_* : \mathfrak{g}\rightarrow\mathfrak{h}$ is a Lie algebra homomorphism. Indeed, $\varphi_*$ commutes with Lie bracket and sends $A\in\mathfrak{g}$ to a left-invariant field also: \begin{multline} \label{hom} \varphi(gh) = \varphi(g)\,\varphi(h) \ \ \ \Longrightarrow \ \ \ \varphi\circ L_g = L_{\varphi(g)}\circ\varphi \ \ \ \Longrightarrow \ \ \ \varphi_*\circ (L_g)_* = (L_{\varphi(g)})_*\circ\varphi_* \\ \ \ \ \Longrightarrow \ \ \ (L_{\varphi(g)})_*\,\varphi_*A = \varphi_*(L_g)_* A = \varphi_*A \ \ \ \Longrightarrow \ \ \ \varphi_*A \in\mathfrak{h}\,. \end{multline} In particular, an automorphism of $G$ (isomorphism $\varphi:G\rightarrow G$) induces an automorphism (invertible endomorphism) $\varphi_* : \mathfrak{g}\rightarrow\mathfrak{g}$\,. Consider a flow $A^t$ on $G$ generated by $A\in\mathfrak{g}$\,. Due to (\ref{phiflow}) it commutes with $L_g$\,. Denoting $A^t e\doteq e^{tA}$, we find that \begin{equation} \label{etA} A^t g = A^t L_g e = L_g A^t e = g e^{tA} = R_{e^{tA}}g\,, \ \ \ \ \ \ \ e^{tA} e^{sA} = A^s e^{tA} = A^s A^t e = e^{(t+s)A}. \end{equation} So, an integral curve $e^{tA}$ (an orbit of unity $e$ under the flow $A^t$) is a 1-parameter Abelian subgroup of $G$\,. In matrix groups, it turns out to be a true exponential. For any group homomorphism $\varphi:G\rightarrow H$ one obtains \begin{equation} \label{phietA} \varphi(e^{tA}) = \varphi\circ A^t e = (\varphi_* A)^t\varphi(e) = (\varphi_* A)^t e = e^{t\varphi_* A}\,. \end{equation} A homomorphism $\rho$ of a Lie group $G$ to the group of automorphisms of a vector space $V$ is called a \emph{representation} of $G$ in $V$, or, in other words, $G$ acts on $V$ ($G\triangleright V$)\,: \begin{equation} \label{reprG} \rho(gh) = \rho(g)\circ\rho(h) \ \ \ \ \ \ \text{or} \ \ \ \ \ \ (gh)\triangleright x = g\triangleright(h\triangleright x)\,, \ \ \ \ \ x\in V\,. \end{equation} Due to (\ref{hom})\,, $\rho_*$ is then a \emph{representation} of $\mathfrak{g}$ in $V$\,. Let us consider \emph{adjoint} representations of both $G$ and $\mathfrak{g}$ \,in $\mathfrak{g}$\,. We begin with a homomorphism A\!D\,:\,$G\rightarrow\text{Aut}G$ which associates with each $h\in G$ an \emph{inner automorphism} \begin{equation} \label{int} \text{A\!D}_{h} : g\rightarrow hgh^{-1} \ \ \ \ \ \ \Longrightarrow \ \ \ \ \ \ \text{A\!D}_{\,h} = R_{h^{-1}}\circ L_h\,, \ \ \ \ \ \text{A\!D}_{gh} = \text{A\!D}_g\circ\text{A\!D}_{\,h}\,. \end{equation} The corresponding maps \begin{equation} \label{Adj} (\text{A\!D}_{h})_*\doteq\text{Ad}_{\,h} : \mathfrak{g}\rightarrow\mathfrak{g}\,, \ \ \ \ \ \text{Ad}_{\,h} = (R_{h^{-1}})_*\circ(L_h)_* = (R_{h^{-1}})_*\,, \ \ \ \ \ \text{Ad}_{\,gh} = \text{Ad}_{\,g}\circ\text{Ad}_{\,h} \end{equation} establish the adjoint representation of $G$ in $\mathfrak{g}$ as a homomorphism Ad\,: $G\rightarrow\text{Aut}\mathfrak{g}$ given explicitly by $g\rightarrow\text{Ad}_g$ (in the matrix case, $\text{Ad}_{\,g}\,A$ is simply $gAg^{-1}$)\,. From (\ref{phietA}) one finds \begin{equation} \label{getA} g\,e^{tA} g^{-1} \equiv \text{A\!D}_g\,e^{tA} = e^{t\text{Ad}_g A}\,. \end{equation} Further, $\text{Ad}_*\doteq\text{ad} :\mathfrak{g}\rightarrow\text{End}\mathfrak{g}$ \,is nothing but the adjoint representation of $\mathfrak{g}$ given by $\text{ad}_A B=[A,B]$\,. Indeed, \begin{equation} \label{adj} \text{ad}_A = \frac{d}{dt}\,\text{Ad}_{e^{tA}}|_{t=0} = \frac{d}{dt}\,(R_{e^{-tA}})_*|_{t=0} = \frac{d}{dt}\,A^{-t}_*|_{t=0} = -\frac{d}{dt}\,\tilde{A^t}|_{t=0} = L_A = [A,\cdot\,]\,. \end{equation} We differentiate directly in $\text{End}\mathfrak{g}$ because it is a vector space where addition is allowed (and tangent spaces are naturally identified with the underlying space)\,. Owing to the property $\text{Ad}_{e^{tA}}\circ\text{Ad}_{e^{sA}}=\text{Ad}_{e^{(t+s)A}}$\,, linear transformations $\text{Ad}_{e^{tA}}$ with fixed $A$ form a 1-parameter matrix group (an orbit of $\mathbf{1}$ under the corresponding flow on $\text{Aut}\mathfrak{g}$) which can be written down as (ordinary) exponential of its tangent vector $\text{ad}_A$\,: \begin{equation} \label{adjexp} \text{Ad}_{e^{tA}} = e^{t\,\text{ad}_A} = \mathbf{1} + t[A,\cdot\,] + \frac{t^2}{2}[A,[A,\cdot\,]] + \ldots\ . \end{equation} In a dual space $\mathfrak{g}^*$ of left-invariant 1-forms $\omega$ such that $(L_g)^*\omega=\omega$\,, the group $G$ acts (from the right) via \emph{coadjoint} (anti)representation $\text{Ad}^*:G\rightarrow\text{Aut}\mathfrak{g}^*$, according to \begin{equation} \label{Ad*} <\!\text{Ad}^*_g\,\omega\,,\,A\!> \ \doteq \ <\!\omega\,,\,\text{Ad}_g\,A\!>\,, \ \ \ \ \ \ \ \ \ \text{Ad}^*_{gh} = \text{Ad}^*_h\circ\text{Ad}^*_g\ . \end{equation} Of course, $(\text{Ad}^*)_*\doteq\text{ad}^* :\mathfrak{g}\rightarrow\text{End}\mathfrak{g}^*$ is a coadjoint (anti)representation of $\mathfrak{g}$\,: \begin{equation} \label{ad*} \text{ad}^*_A = \frac{d}{dt}\,\text{Ad}^*_{e^{tA}}|_{t=0} \ \ \ \ \ \ \Longrightarrow \ \ \ \ \ \ <\!\text{ad}^*_A\,\omega\,,\,B\!> \ = \ <\!\omega\,,\,\text{ad}_A B\!>\,. \end{equation} A Lie group $G$ \,acts on a manifold $M$ from the right, $M\triangleleft G$\,, if \begin{equation} \label{act} x(gh) = (xg)h \ \ \ \ \ \text{or} \ \ \ \ \ R_{gh} = R_h\circ R_g \ \ \ \ \ \ \ \ \ (x\triangleleft g \equiv xg \equiv R_g\,x\,, \ \ x\in M)\,. \end{equation} Each $A\in\mathfrak{g}$ induces a \emph{fundamental vector field} $A^*$ on $M$ via \begin{equation} \label{fund} (A^*)^t x = x e^{tA} \equiv R_{e^{tA}}\,x\ . \end{equation} In other words, the flow $(A^*)^t$ produces the same $x$-orbit as the right action of $e^{tA}$ (equivalently, the vectors $A^*$ are tangent to orbits $x e^{tA}$ generated by $A$)\,. Hence, $A\rightarrow A^*$ is a homomorphism of Lie algebras. If $M=G$ (when $G$ acts on itself by right shifts), $A^*\equiv A$\,. E.g., for $\text{Ad}^*$ (it acts from the right!) we can find the value of the field $A^*$ at some $\omega\in\mathfrak{g}^*$ simply by differentiating the orbit: \begin{equation} \label{ad*om} A^*(\omega) = \frac{d}{dt}\,\text{Ad}^*_{e^{tA}}\,\omega\,|_{t=0} = \text{ad}^*_A\,\omega\ , \ \ \ \ \ \ \ \ A^*_i(\omega) = -C^k_{ij}\,\omega_k A^j\,. \end{equation} $G$-invariant functions on $M$ (they are constant on orbits, $f(xg)=f(x)$) are annihilated by the fundamental vector fields, $A^*f\equiv 0 \ \ \forall A\in\mathfrak{g}$\,, because, due to (\ref{fxit})\,, \begin{equation} \label{A*1} (A^*f)_x = \frac{d}{dt}\,f((A^*)^t x)|_{t=0} = \frac{d}{dt}\,f(x e^{tA})|_{t=0} = 0\,. \end{equation} Under the (right) group action, the fields $A^*$ transform as follows: \begin{equation} \label{RgA*} (R_g)_*\,A^* = ((R_g)_*A)^* = (\text{Ad}_{g^{-1}}A)^*\,. \end{equation} Specifically, if $A^*$ is tangent to the orbit $x e^{tA}$ at a point $x$, then $(R_g)_*\,A^*$ is tangent at $xg$ to the orbit $(xe^{tA})g=(xg)(g^{-1}e^{tA}g)=(xg)e^{t\text{Ad}_{g^{-1}}A}$ generated by $\text{Ad}_{g^{-1}}A$\,. %\newpage \vspace{.3cm} \begin{center} \large\textbf{Hamiltonian dynamics} \end{center} \vspace{.1cm} Hamiltonian dynamics on manifolds is formulated in terms of flows generated by \emph{Hamiltonian vector fields}. The corresponding \emph{Hamilton equations} are written as follows: \begin{equation} \label{dyn} \dot{F}(x) \doteq \frac{d}{dt}\,F(\bar{H}^t x)|_{t=0} = (\bar{H}F)_x \equiv \{H,F\}_x\,, \ \ \ \ \ \ \dot{x} \doteq \frac{d}{dt}\,\bar{H}^t x|_{t=0} = \bar{H}(x)\,. \end{equation} Here $x$ is a point of some manifold $M$ (phase space), $H$ and $F$ are functions on $M$, and $\{\,,\,\}$ is a bilinear operation (\emph{Poisson brackets}) producing a Hamiltonian vector field $\bar{H}$ from a given function $H(x)$. In the last relation of (\ref{dyn}) we slightly abuse notation by differentiating points instead of their coordinates. A natural setting for this construction is provided by \emph{symplectic manifolds} carrying a $T_2^0$ tensor field (2-form) $\Omega$ which is closed ($d\Omega=0$) and non-degenerate: \begin{equation} \label{Omega} \Omega = \frac{1}{2}\,\Omega_{ij}(x)\,dx^i\wedge dx^j\,, \ \ \ \ \ \Omega_{ij} = -\Omega_{ji}\,, \ \ \ \ \ \exists\,\Lambda^{ij}(x) : \ \Lambda^{ik}\Omega_{kj} = \delta^i_j\,, \ \ \ \ \Lambda^{ij} = -\Lambda^{ji}\,. \end{equation} We define $\bar{H}$ and $\{\,,\,\}$ by \begin{equation} \label{barH} (dH)_i = \Omega_{ij}\bar{H}^j \ \ \ \ \ \text{or} \ \ \ \ \ \bar{H}^i = \Lambda^{ij}\,\partial_j H\,, \ \ \ \ \ \ \ \{H,F\} = \Lambda^{ij}\,\partial_i F\,\partial_j H\,, \end{equation} which can be cast in coordinate-free form through the use of interior product $\iota$ (\ref{inner})\,: \begin{equation} \label{barH2} dH = -\iota_{\bar H}\Omega\,, \ \ \ \ \ \ \ \{H,F\} = \bar{H}F = \ <\!dF,\bar{H}\!> \ = -<\!\iota_{\bar F}\Omega,\bar{H}\!> \ = \ \Omega(\bar{H},\bar{F})\,. \end{equation} It is non-degeneracy of the symplectic form $\Omega$ (invertibility of $\Omega_{ij}$) that enables us to associate with $dH$ a vector field $\bar{H}$\,. Moreover, it establishes isomorphism between tangent and cotangent spaces on $M$ relating (bijectively) 1-forms $\iota_\xi\,\Omega$ with vectors $\xi$\,: \begin{equation} \label{iso} (\iota_\xi\,\Omega)_i = (\iota_\xi\,\Omega)(e_i) = \Omega(\xi,e_i) = -\Omega_{ij}\xi^j\,, \ \ \ \ \ \ \xi^j = -\Lambda^{ji}(\iota_\xi\,\Omega)_i\,. \end{equation} Also, it follows from $d\Omega=0$ and (\ref{iota2}) that Hamiltonian flows respect symplectic form, \begin{equation} \label{resp} L_{\bar H}\Omega = (d\circ\iota_{\bar H} + \iota_{\bar H}\circ d)\,\Omega = -d^2 H = 0 \ \ \ \ \ \ \Longrightarrow \ \ \ \ \ \ (\bar{H}^t)^*\Omega = \Omega\,. \end{equation} Any function $H(x)$ on a symplectic manifold may serve as a Hamiltonian and generate the corresponding flow. Thus, the equality $\{H,F\} = 0$ means that $F$ is invariant under the flow $\bar{H}^t$\,, and vice versa. Poisson brackets evidently obey \begin{equation} \label{Leib} \{H,F\} = -\{F,H\}\,, \ \ \ \{H,FG\} = \{H,F\}G + F\{H,G\}\,, \ \ \ \{H,\Phi(F)\} = \{H,F\}\frac{d\Phi}{dF}\,. \end{equation} The Jacobi property of Poisson brackets stems from $d\,\Omega=0$ via (\ref{barH2}) and (\ref{d2})\,, \begin{multline} \label{Jacobi} \{F,\{H,G\}\} + \{G,\{F,H\}\} +\{H,\{G,F\}\} \equiv \{F,\{H,G\}\} + \text{cycle} = \bar{F}\bar{H}G + \text{cycle} \\ \!\! = \bar{H}\bar{F}G - [\bar{H},\bar{F}]G + \text{cycle} = \bar{H}\Omega(\bar{F},\bar{G}) - \Omega([\bar{H},\bar{F}],\bar{G}) + \text{cycle} = d\Omega(\bar{H},\bar{F},\bar{G}) = 0 \end{multline} and additionally shows that the Poisson structure is respected by Hamiltonian flows, \begin{equation} \label{resp2} L_{\bar H}\{F,G\} = \{L_{\bar H}F,G\} + \{F,L_{\bar H}G\}\,, \end{equation} and that the map $H\rightarrow\bar H$ is a homomorphism from a Poisson algebra to a Lie algebra of Hamiltonian vector fields, \begin{equation} \label{homo} \overline{\{H,F\}} = [\bar{H},\bar{F}] \end{equation} (acting by both parts of (\ref{homo}) on any function $G$ reproduces the Jacobi identity)\,. \vspace{.1cm} A canonical example of symplectic manifold is given by a cotangent bundle $T^*M$\,, the set of all cotangent spaces (spaces of 1-forms) over a manifold $M$. Its points are denoted by $(p,q)$\,, with $q$ a point on $M$ and $p=p_i\,dq^i$ a 1-form $\in T^*_qM$, and coordinates by $(p_i,q^j)$\,. A symplectic form here is defined by \begin{equation} \label{can} \Omega = dp_i\wedge dq^i\,. \end{equation} Obviously, it is closed and non-degenerate. Its component presentation generalizes (\ref{Omega})\,, \begin{gather} \Omega = \frac{1}{2}\,(\Omega^{ij}\,dp_i\wedge dp_j + \Omega^i{}_j\,dp_i\wedge dq^j + \Omega_i{}^j\,dq^i\wedge dp_j + \Omega_{ij}\,dq^i\wedge dq^j) = dp_i\wedge dq^i\,, \\ \Omega^{ij} = \Omega_{ij} = \Lambda^{ij} = \Lambda_{ij} = 0\,, \ \ \ \Omega^i{}_j = -\Omega_j{}^i = \Lambda^i{}_j = -\Lambda_j{}^i = \delta^i_j\ , \end{gather} with matrix indices being summed over like $(MN)^i_j=M^{ik}N_{kj}+M^i{}_k N^k{}_j$\ . We can now obtain explicit formulas for the Hamiltonian vector field $\bar{H}=(\bar{H}_i,\bar{H}^j)$ (like any vector tangent to $T^*M$, it is parametrized by doubled number of components)\,, \begin{equation} \label{vect} \bar{H}_i = \Lambda_i{}^j\,\partial_j H = -\partial_i H \equiv -\frac{\partial H}{\partial q^i}\,, \ \ \ \ \ \ \ \ \bar{H}^j = \Lambda^j{}_i\,\partial^i H = \partial^j H \equiv \frac{\partial H}{\partial p_j}\ , \end{equation} and for the Poisson brackets: \begin{equation} \label{Pois} \{H,F\} = \bar{H}F = \bar{H}_i\,\partial^i\!F + \bar{H}^i\,\partial_i F = \frac{\partial H}{\partial p_i}\,\frac{\partial F}{\partial q^i} - \frac{\partial H}{\partial q^i}\,\frac{\partial F}{\partial p_i}\ . \end{equation} Now eq.\,(\ref{dyn}) produces the Hamilton equations in canonical form: \begin{equation} \label{Ham} \dot{p}_i = -\frac{\partial H}{\partial q^i}\,, \ \ \ \ \dot{q}^i = \frac{\partial H}{\partial p_i}\,, \ \ \ \ \ \ \dot{F} = \{H,F\}\,. \end{equation} Let us now consider a more general situation when the tensor $\Lambda^{ij}(x)$ is not invertible (Poisson brackets are degenerate)\,. A symplectic form $\Omega$ does not exist, but the Poisson structure does, and Hamiltonian vector fields can be constructed from differentials of functions, as before ($\Lambda^{ij}$ is assumed to guarantee the properties (\ref{Leib})--(\ref{homo}) of Poisson brackets)\,. It turns out in this case, unlike the symplectic one, that Hamiltonian vectors span not the entire tangent bundle $TM$\,, but only certain subspaces of each $T_xM$, being therefore tangent to certain submanifolds (\emph{symplectic leaves}) where the induced Poisson structure is no longer degenerate, and its inverse is symplectic. These submanifolds constitute a symplectic foliation on $M$. Hamiltonian vector fields span entire tangent bundles of each symplectic leaf, and only those functions $F$ on $M$ that are constant on each leaf result in $\{F,\cdot\}\equiv0$ (it is these functions that made the original Poisson structure degenerate: they annihilated Poisson brackets without being constant throughout)\,. The whole picture is consistent owing to (\ref{homo})\,: a commutator of two Hamiltonian vector fields tangent to a leaf is also tangent to this leaf (Hamiltonian flows do not take out things from the leaf)\,. Therefore, symplectic leaves prove to be common integral submanifolds for all Hamiltonian flows. They carry an induced symplectic structure given by $\Omega(\bar{H},\bar{F})=\{H,F\}$\,. Invertibility of the corresponding tensors when restricted to the leaves is the result of quotienting by kernels of the relevant Poisson brackets. A famous example of a degenerate Poisson structure is provided by the (Lie-Berezin-Kostant-)\,Kirillov bracket on a dual space $\mathfrak{g}^*$ to a Lie algebra $\mathfrak{g}$\,. Any element $A\in\mathfrak{g}$ by definition specifies on $\mathfrak{g}^*$ a linear function \begin{equation} \label{A()} A(\omega)= \ <\!\omega\,,A\!> \ = A^i\omega_i\,, \ \ \ \ \ \ \ e_i(\omega) = \omega_i\,. \end{equation} Let us define Poisson brackets of such linear functions as their Lie algebra commutator, \begin{equation} \label{Kir0} \{A(\omega),B(\omega)\} \equiv \{A,B\}(\omega) \doteq [A,B]\,(\omega) = \ <\!\omega\,,[A,B]\!>\,. \end{equation} Then the coordinate functions $\omega_i$ have the following brackets: \begin{equation} \label{Kir1} \{\omega_i,\omega_j\}\doteq \ <\!\omega,[e_i,e_j]\!> \ =C^k_{ij}\,\omega_k\,. \end{equation} This (Kirillov) bracket extends to arbitrary functions via the last equality in (\ref{Leib})\,, \begin{equation} \label{Kirf} \{H,F\}(\omega) = C^k_{ij}\,\omega_k\,\frac{\partial H}{\partial\omega_i} \frac{\partial F}{\partial\omega_j}\,, \end{equation} whence a Hamiltonian vector field $\bar H$ is \begin{equation} \label{KirH} \bar{H}_i(\omega)=-C^k_{ij}\,\omega_k\,\frac{\partial H}{\partial\omega_j}\,. \end{equation} The last two formulas can be also rewritten in coordinate-free form. Identifying $T_\omega\mathfrak{g}^*$ with $\mathfrak{g}^*$, we note the following simple equality for the differential of a linear function $A(\omega)$\,: \begin{equation} \label{} <\!d(A(\omega))\,,\lambda\!> \ = A^i<\!d\omega_i\,,\lambda\!> \ = A^i\lambda_i = \ <\!\lambda\,,A\!> \ \ \ \ \ \ \ (\lambda\in\mathfrak{g}^* \simeq T_\omega\mathfrak{g}^*) \end{equation} Similarly, differential of any function $H$ on $\mathfrak{g}^*$ can be treated as a $\mathfrak{g}$-valued field $H'(\omega)$\,, \begin{equation} \label{F'} <\!dH\,,\lambda\!>_\omega \ \doteq \ <\!\lambda\,,H'(\omega)\!> \ \ \ \ \ \ \Longrightarrow \ \ \ \ \ \ H'(\omega) = \frac{\partial H}{\partial\omega_i}\,e_i\,, \ \ \ \ \omega'_i = e_i\,, \ \ \ \ A'(\omega) = A\,. \end{equation} In these terms, the Kirillov bracket and Hamiltonian vector field are written as \begin{equation} \label{Kirf2} \{H,F\}(\omega) = \ <\!\omega\,,[H'(\omega),F'(\omega)]\!>\,, \ \ \ \ \ \ \ \ \bar{H}(\omega) = \text{ad}^*_{H'(\omega)}\,\omega\,. \end{equation} The last relation is verified as follows: \begin{equation} \notag <\!\bar{H},F'\!>_\omega \ = \ <\!dF,\bar{H}\!>_\omega \ = (\bar{H}F)_\omega = \{H,F\}(\omega) = \ <\!\omega\,,[H',F']\!> \ = \ <\!\text{ad}^*_{H'}\omega\,,F'\!>. \end{equation} Comparing (\ref{Kirf2}) to (\ref{ad*om})\,, we notice that $\bar{H}(\omega)$ at a point $\omega\in\mathfrak{g}^*$ coincides with the value of fundamental vector field of coadjoint action induced by the element $H'(\omega)\in\mathfrak{g}$\,. For linear Hamiltonians $A(\omega)$\,, when $A'(\omega)$ is independent of $\omega$\,, the picture becomes even more impressive: Hamiltonian vector fields $\bar{A}(\omega)$ produced from $A(\omega)$ by the Kirillov bracket, and fundamental vector fields $A^*(\omega)$ generated by the coadjoint action, coincide identically on $\mathfrak{g}^*$, \begin{equation} \label{A'} A\in\mathfrak{g} \ \ \ \ \Longrightarrow \ \ \ \ A(\omega)= \ <\!\omega\,,A\!>\,, \ \ \ \ \ A'(\omega) = A\,, \ \ \ \ \ \bar{A}(\omega) = A^*(\omega) = \text{ad}^*_A \omega\,, \end{equation} spanning the tangent spaces to coadjoint orbits. The $\text{Ad}^*$-\,invariant functions (they are constant on the orbits) are thus annihilated by any $\bar H$, have zero Poisson brackets with any function, and produce zero Hamiltonian vector fields: \begin{equation} \label{} F(\text{Ad}^*_g\,\omega) = F(\omega) \ \ \ \ \Longrightarrow \ \ \ \ \bar{H}F = \{H,F\} = - \bar{F}H = 0\,, \ \ \ \bar{F}(\omega) = \text{ad}^*_{F'(\omega)}\,\omega = 0\,. \end{equation} Due to these functions, the Kirillov bracket is degenerate on the whole $\mathfrak{g}^*$. However, restricting to the orbits removes the degeneracy, and a symplectic (Kirillov) form can be defined on symplectic leaves (which are exactly coadjoint orbits) in a standard fashion: \begin{equation} \label{Kirom} \Omega^{\text K}(\bar{H},\bar{F})_\omega = \{H,F\}(\omega) = \ <\!\omega\,,[H',F']\!>\,. \end{equation} Coadjoint orbits provide an important (actually, the basic) example of \emph{Hamiltonian action} of a group on a symplectic manifold: the flows induced by the group action prove to be the Hamiltonian ones. Namely, there exists a Lie algebra homomorphism $A\rightarrow\tilde{A}(x)$ from $\mathfrak{g}$ to the Poisson algebra of functions on a manifold $M$, and the Hamiltonian $\tilde{A}(x)$ generates exactly the fundamental vector field $A^*(x)$\,: \begin{equation} \label{Hamact} \widetilde{[A,B]}\,(x) = \{\tilde{A},\tilde{B}\}(x)\,, \ \ \ \ \ \ A^*(x) = \bar{\tilde{A}}(x)\,. \end{equation} Unlike the $\text{Ad}^*$ example above, where explicit formulas (\ref{A'}) for $\tilde{A}$ and $A^*$ were simple enough, in the general case it may not be so. However, for a given $x\!\in\! M$, \ $\tilde{A}(x)$ can be viewed as a linear function on $\mathfrak{g}$\,, or an element of $\mathfrak{g}^*$. This motivates the following definition of the \emph{moment map} \ $\mu:M\rightarrow\mathfrak{g}^*$\,: \begin{equation} \label{mu} \tilde{A}(x) \doteq \ <\!\mu(x)\,,A\!> \ = A(\mu(x)) \ \ \ \ \ \ \Longrightarrow \ \ \ \ \ \tilde{A} = A\circ\mu = \mu^*\!A \end{equation} (in the last equality we denote the function $A(\omega)$ simply by $A$)\,. By (\ref{A'}), on coadjoint orbits $\mu(\omega)=\omega$\,, \, or \,$\mu=\text{id}$\,. Consider some properties of the moment map. By definition, it relates the Hamiltonians $A(\omega)$ and $\tilde{A}(x)$ for each $A\in\mathfrak{g}$\,, as well as the corresponding Poisson structures\,: \begin{equation} \label{muP} \mu^*\{A,B\} = \mu^*[A,B] = \widetilde{[A,B]} = \{\tilde{A},\tilde{B}\} = \{\mu^*\!A\,,\,\mu^*\!B\}\,. \end{equation} Also, it is \emph{equivariant} (commutes with the group action)\,, \begin{equation} \label{equivar} \mu(xg)=\text{Ad}^*_g\ \mu(x) \ \ \ \ \ \ \text{or} \ \ \ \ \ \ \mu\circ g = \text{Ad}^*_g\circ\mu\,, \end{equation} because it sends fundamental vector fields (and thus the corresponding flows) acting on $M$ to those acting on $\mathfrak{g}^*$ via \, $\mu_*:A^*(x)\rightarrow A^*(\mu(x))=\bar{A}(\mu(x))$\,. \,Indeed, for any $B\in\mathfrak{g}$\,, \begin{equation} \notag ((\mu_*A^*)B)(\mu(x)) = (A^*(\mu^*\!B))(x) = (A^* \tilde{B})(x) = \{\tilde{A},\tilde{B}\}(x) = [A,B](\mu(x)) = (\bar{A}B)(\mu(x)) \end{equation} Another simple corollary of (\ref{muP}) shows that $\mu^*$ relates symplectic forms $\Omega$ and $\Omega^{\text K}$ restricted to the corresponding orbits on $M$ and $\mathfrak{g}^*$: \begin{multline} \label{muom} \Omega\,(A^*,B^*)_x = \Omega\,(\bar{\tilde A},\bar{\tilde B})_x = \{\tilde{A},\tilde{B}\}\,(x) = [A,B]\,(\mu(x)) = \Omega^{\text K}(\bar{A},\bar{B})_{\mu(x)} \\ = \Omega^{\text K}(\mu_* A^*,\mu_* B^*)_{\mu(x)} = (\mu^*\Omega^{\text K})(A^*,B^*)_x \ \ \ \ \ \Longrightarrow \ \ \ \ \mu^*\Omega^{\text K}=\Omega \ \ \text{on orbits.} \end{multline} Thus, the moment map $\mu$ sends $G$-\,orbits on $M$ to those on $\mathfrak{g}^*$, together with all the Hamiltonian dynamics defined on them. The following statement is a Hamiltonian analog of Noether's theorem. If the dynamics on $M$ is governed by some $G$-invariant Hamiltonian $H(x)$\,, then the moment $\mu$\,, taken as a $\mathfrak{g}^*$-\,valued function on $M$, is conserved: $\dot\mu=0$\,. Indeed, (\ref{A*1}) means that $A^*\!H=0$ for any $A\in\mathfrak{g}$\,. Therefore, \begin{equation} \label{Noether} <\!\bar{H}\mu,A\!> \ = \bar{H}<\!\mu,A\!> \ = \bar{H}\tilde{A} =-\bar{\tilde A}H = -A^* H =0 \ \ \ \ \Longrightarrow \ \ \ \ \dot\mu = \bar{H}\mu = 0\,. \end{equation} \end{document}