\documentclass[reqno]{amsart} \usepackage{amssymb} \AtBeginDocument{{\noindent\small {\em Electronic Journal of Differential Equations}, Vol. 2004(2004), No. 39, pp. 1--14.\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 2004 Texas State University - San Marcos.} \vspace{7mm}} \begin{document} \title[\hfilneg EJDE-2004/39\hfil Structure of group invariants] {Structure of group invariants \\ of a quasiperiodic flow} \author[Lennard F. Bakker\hfil EJDE-2004/39\hfilneg] {Lennard F. Bakker} \address{Department of Mathematics\\ Brigham Young University\\ 292 TMCB \\ Provo, UT 84602 USA} \email{bakker@math.byu.edu} \date{} \thanks{Submitted July 02 2002. Published March 22, 2004.} \subjclass[2000]{37C55, 37C80, 20E34, 11R04} \keywords{Generalized symmetry, quasiperiodic flow, semidirect product} \begin{abstract} It is shown that the multiplier representation of the generalized symmetry group of a quasiperiodic flow induces a semidirect product structure on certain group invariants (including the generalized symmetry group) of the flow's smooth conjugacy class. \end{abstract} \maketitle \newtheorem{theorem}{Theorem}[section] \newtheorem{corollary}[theorem]{Corollary} \newtheorem{lemma}[theorem]{Lemma} \newtheorem{example}[theorem]{Example} \numberwithin{equation}{section} \section{Introduction} The generalized symmetry group, $S_\phi$, of a smooth flow $\phi:\mathbb{R}\times T^n\to T^n$ is the collection of all diffeomorphisms of $T^n$ that map the generating vector field of $\phi$ to a uniformly scaled copy of itself (see next section for definitions). The multiplier representation of $S_\phi$ is the one-dimensional linear representation \[ \rho_\phi:S_\phi\to\mathbb{R}^*\equiv\mathop{\rm GL}(\mathbb{R}) \] that takes a generalized symmetry $R\in S_\phi$ to its unique multiplier $\rho_\phi(R)$ (Theorem 2.8 in \cite{BC}), the multiplier being the scalar by which the generating vector field of $\phi$ is uniformly scaled by $R$. For each subgroup $\Lambda$ of the multiplier group $\rho_\phi(S_\phi)$, the multiplier representation induces the short exact sequence of groups, \[ \mathop{\rm id}{}_{T^n}\to \ker \rho_\phi\to \rho_\phi^{-1}(\Lambda) \stackrel{j_\Lambda}{\to } \Lambda\to 1, \] in which $\mathop{\rm id}_{T^n}$ is the identity diffeomorphism of $T^n$, $\ker \rho_\phi\to\rho_\phi^{-1}(\Lambda)$ is the canonical monomorphism, and $j_\Lambda:\rho_\phi^{-1}(\Lambda)\to\Lambda \cong\rho_\phi^{-1}(\Lambda)/\ker \rho_\phi$ is $\rho_\phi\vert\rho_\phi^{-1}(\Lambda)$. This short exact sequence indicates that $\rho_\phi^{-1}(\Lambda)$ is a group extension of $\ker \rho_\phi$ by the Abelian group $\Lambda$. When $\phi$ is a quasiperiodic flow on $T^n$, it will be shown that \begin{enumerate} \item[(i)] every element of $\rho_\phi(S_\phi)$ is a real algebraic integer of degree at most $n$ (Corollary \ref{solve4}), \item[(ii)] $\ker \rho_\phi\cong T^n$ (Corollary \ref{solve7}), \item[(iii)] every $R\in S_\phi$ with $\rho_\phi(R)=-1$ is an involution (Corollary \ref{solve8}), \item[(iv)] $\rho_\phi(S_\phi)$ is isomorphic to an Abelian subgroup of $\mathop{\rm GL}(n,\mathbb{Z})$ (Theorem \ref{rep2}), and \item[(v)] for each subgroup $\Lambda<\rho_\phi(S_\phi)$ there is a splitting map $h_\Lambda:\Lambda\to\rho_\phi^{-1}(\Lambda)$ for the extension (Theorem \ref{rep4}). \end{enumerate} The main result (Theorem \ref{rep5}) is that \[ \rho_\phi^{-1}(\Lambda)=\ker \rho_\phi \rtimes_\Gamma h_{\Lambda}(\Lambda) \] for every $\Lambda<\rho_\phi(S_\phi)$; that is, $\rho_\phi^{-1}(\Lambda)$ is the semidirect product of $\ker \rho_\phi$ by $h_{\Lambda}(\Lambda)$ corresponding to the conjugating homomorphism $\Gamma:h_\Lambda(\Lambda)\to\mathop{\rm Aut}(\ker \rho_\phi)$. \section{Multipliers and Quasiperiodic Flows} A generalized symmetry of a (smooth, i.e.\ $C^\infty$) flow $\phi$ on the $n$-torus $T^n$ ($n\geq 2$) is an $R\in\mathop{\rm Diff}(T^n)$ (the group of smooth diffeomorphisms on $T^n$) for which there exists an $\alpha\in\mathbb{R}^*$ such that \[ R\phi(t,\theta)=\phi\big(\alpha t,R(\theta)\big)\quad\mbox{for all $t\in \mathbb R$ and all }\theta\in T^n. \] This condition is $R\phi_t=\phi_{\alpha t}R$ for all $t\in\mathbb{R}$, where $\phi_t$ is the diffeomorphism of $T^n$ defined by $\phi_t(\theta)=\phi(t,\theta)$. A generalized symmetry of $\phi$ is characterized by its action on the generating vector field $X$ of $\phi$, which vector field is defined by \[ X(\theta)=\frac{d}{dt} \phi_t (\theta)\Big|_{t=0},\quad\theta\in T^n. \] (In what follows, $\mathbf{T}$ is the tangent functor, and $R_*X=\mathbf{T}RXR^{-1}$ is the push-forward of $X$ by $R$.) \begin{theorem}\label{char} An $R\in\mathop{\rm Diff}(T^n)$ is a generalized symmetry of a flow $\phi$ on $T^n$ if and only if there exists a unique $\alpha\in \mathbb{R}^*$ such that $R_*X=\alpha X$. \end{theorem} For the proof of this theorem, see Proposition 1.4 and Lemma 2.7 in \cite{BC}. The generalized symmetry group, $S_\phi$, of a flow $\phi$ on $T^n$ is the collection of all the generalized symmetries of $\phi$. The Abelian group $F_\phi=\{\phi_t:t\in\mathbb{R}\}\subset\mathop{\rm Diff}(T^n)$ generated by $\phi$ is a subgroup of the normal subgroup $\ker \rho_\phi$ of $S_\phi$. On the other hand, $S_\phi$ is the group theoretic normalizer of $F_\phi$ in $\mathop{\rm Diff}(T^n)$ (Theorem 2.5 \cite{BC}). The unique $\alpha$ attached to an $R\in S_\phi$ in Theorem \ref{char} is $\rho_\phi(R)$, the multiplier of $R$. An $R\in S_\phi$ with $\rho_\phi(R)=1$ is known as a (classical) symmetry of $\phi$ (p.8 \cite{LR}); the symmetry group of $\phi$ is $\ker \rho_\phi=\rho_\phi^{-1}(\{1\})$. An $R\in S_\phi$ with $\rho_\phi(R)=-1$ is called a reversing symmetry (p.4 \cite{LR}); if $R^2=\mathop{\rm id}_{T^n}$, then $R$ is a reversing involution or a classical time-reversing symmetry of $\phi$; the reversing symmetry group of $\phi$ is $\rho_\phi^{-1}(\{1,-1\})$ (p.8 \cite{LR}). An $R\in S_\phi$ with $\rho_\phi(R)\ne\pm1$, if it exists, is another type of symmetry of $\phi$. Two flows $\phi$ and $\psi$ are smoothly conjugate if and only if there is a $V\in\mathop{\rm Diff}(T^n)$ such that $V\phi_t=\psi_tV$ for all $t\in\mathbb{R}$. (This is equivalent to $V_*X=Y$ where $X$ is the generating vector field for $\phi$, and $Y$ is the generating vector field for $\psi$.) A flow $\phi$ on $T^n$ with generating vector field $X$ is quasiperiodic if and only if there exists a $V\in\mathop{\rm Diff}(T^n)$ such that $V_*X$ is a constant vector field whose coefficients are independent over $\mathbb{Q}$ (see pp.79-80 \cite{HB}). (Recall that real numbers $a_1,a_2,\dots ,a_n$ are independent over $\mathbb{Q}$ if for $m=(m_1,m_2,\dots ,m_n)\in\mathbb{Z}^n$, the equation $\sum^n_{j=1}m_ja_j=0$ implies that $m_j=0$ for all $j=1,2,\dots ,n$.) The frequencies of a quasiperiodic flow $\phi$ generated by a constant vector field $X$ are the components of $X$. \begin{example}\label{example1} \rm Identify $T^3$ with $S^1\times S^1\times S^1$ where $S^1=\mathbb{R}/\mathbb{Z}$. Let $\theta=(\theta_1,\theta_2,\theta_3)$ be global coordinates on $T^3$. The quasiperiodic flow $\phi$ on $T^3$ generated by vector field \[ X=\frac{\partial}{\partial\theta_1}+7^{1/3}\frac{\partial}{\partial\theta_2}+ 7^{2/3}\frac{\partial}{\partial\theta_3} \] is \[ \phi_t(\theta)=\phi(t,\theta_1,\theta_2,\theta_3)=\big(\theta_1+t,\theta_2+7^{1/3}t,\theta_3+7^{2/3}t\big), \] where the addition in the components of $\phi$ is mod $1$. For each $c=(c_1,c_2,c_3)\in T^3$, the translation \[ R_c(\theta_1,\theta_2,\theta_3)=\big(\theta_1+c_1,\theta_2+c_2,\theta_3+c_3\big) \] of $T^3$ is a symmetry of $\phi$ because \[ R_c\phi(t,\theta_1,\theta_2,\theta_3) = \big(\theta_1+c_1+t,\theta_2+c_2+7^{1/3}t,\theta_3+c_3+7^{2/3}t\big) = \theta(t, R_c(\theta_1,\theta_2,\theta_3). \] The involution $N(\theta_1,\theta_2,\theta_3)=(-\theta_1,\theta_2,\theta_3)$ of $T^3$ is a reversing symmetry of $\phi$ because \[ N\phi(t,\theta_1,\theta_2,\theta_3) = (-\theta_1-t,-\theta_2-7^{1/3}t,-\theta_3-7^{2/3}t\big) = \phi\big(-t,N(\theta_1,\theta_2,\theta_3)\big). \] \end{example} \begin{theorem}\label{mult} If $\phi$ is a quasiperiodic, then $\{1,-1\}<\rho_\phi(S_\phi)$. \end{theorem} \begin{proof} Suppose $\phi$ is quasiperiodic. Then there is a $V\in\mathop{\rm Diff}(T^n)$ such that $Y=V_*X$ is a constant vector field. Let $\psi$ be the flow generated by $Y$. For any $t\in\mathbb{R}$, the diffeomorphism $\psi_t$ satisfies $(\psi_t)_*Y=Y$, so that $1\in\rho_\psi(S_\psi)$. On the other hand, the map $N:T^n\to T^n$ defined by $N(\theta)=-\theta$ satisfies $N_*Y=-Y$, so that $-1\in\rho_\psi(S_\psi)$. The flows $\phi$ and $\psi$ are smoothly conjugate because $Y=V_*X$. This implies that $\rho_\phi(S_\phi)=\rho_\psi(S_\psi)$ (Theorem 4.2 \cite{BC}), and so $\{1,-1\}<\rho_\phi(S_\phi)$.\end{proof} \begin{theorem}\label{nonabelian} If $\phi$ is quasiperiodic and $\Lambda$ is a nontrivial subgroup of $\rho_\phi(S_\phi)$, then $\rho_\phi^{-1}(\Lambda)$ is non-Abelian, and hence the generalized symmetry group of $\phi$ and the reversing symmetry group of $\phi$ are non-Abelian. \end{theorem} \begin{proof} Suppose $\phi$ is quasiperiodic and $\Lambda$ is a nontrivial subgroup of $\rho_\phi(S_\phi)$. Then there is an $R\in S_\phi$ such that $\alpha=\rho_\phi(R)\ne1$. Thus $R\phi_1=\phi_{\alpha}R$. If $\phi_1=\phi_{\alpha}$, then $\phi$ would be periodic. Thus, $\rho_\phi^{-1}(\Lambda)$ is non-Abelian. By Theorem \ref{mult}, both $\rho_\phi(S_\phi)$ and $\rho_\phi\big(\rho_\phi^{-1}(\{1,-1\})\big)$ contain $-1$, so that $S_\phi=\rho_\phi^{-1}(\rho_\phi(S_\phi))$ and $\rho_\phi^{-1}(\{1,-1\})$ are both non-Abelian. \end{proof} For any $\Lambda<\rho_\phi(S_\phi)$, $\rho_\phi^{-1}(\Lambda)$ is an invariant of the smooth conjugacy class of $\phi$ in the sense that if $\phi$ and $\psi$ are smoothly conjugate, then $\rho_\phi^{-1}(\Lambda)$ and $\rho_\psi^{-1}(\Lambda)$ are conjugate subgroups of $\mathop{\rm Diff}(T^n)$ (Theorem 4.3 \cite{BC}). Because a quasiperiodic flow $\phi$ is smoothly conjugate to a quasiperiodic flow $\psi$ generated by a constant vector field, the group structure of $\mathop{\rm id}_{T_n}\to\ker \rho_\phi\to\rho_\phi^{-1}(\Lambda)\to\Lambda\to 1$ is determined by that of $\mathop{\rm id}_{T^n}\to\ker \rho_\psi\to\rho_\psi^{-1}(\Lambda)\to\Lambda\to 1$. Attention is therefore restricted to a quasiperiodic flow $\phi$ generated by a constant vector field $X$. \section{Lifting the Generalized Symmetry Equation} The generalized symmetry equation of a flow $\phi$ on $T^n$ is the equation $R_*X=\alpha X$ that appears in Theorem \ref{char}. Lifting it from $\mathbf{T}T^n$ to $\mathbf{T}\mathbb{R}^n$, the universal cover of $\mathbf{T}T^n$, requires lifting the diffeomorphism $R$ of $T^n$ to a diffeomorphism of $\mathbb{R}^n$, and lifting the vector field $X$ on $T^n$ to a vector field on $\mathbb{R}^n$. The covering map $\pi:\mathbb{R}^n\to T^n$ is a local diffeomorphism for which \[ \pi(x+m)=\pi(x) \] for any $x\in\mathbb{R}^n$ and any $m\in\mathbb{Z}^n$. Let $R:T^n\to T^n$ be a continuous map. A {\it lift} of $R\pi:\mathbb{R}^n\to T^n$ is a continuous map $Q:\mathbb{R}^n\to\mathbb{R}^n$ for which $R\pi=\pi Q$. Since $\pi$ is a fixed map, $Q$ is also said to be a lift of $R$. Any two lifts of $R$ differ by a deck transformation of $\pi$, which is a translation of $\mathbb{R}^n$ by an $m\in\mathbb{Z}^n$. \begin{theorem}\label{lift3} Let $R:T^n\to T^n$ and $Q:\mathbb{R}^n\to\mathbb{R}^n$. Then $Q$ is a lift of a diffeomorphism $R$ of $T^n$ if and only if $Q$ is a diffeomorphism of $\mathbb{R}^n$ such that {\rm a)} for any $m\in\mathbb{Z}^n$, $Q(x+m)-Q(x)$ is independent of $x\in\mathbb{R}^n$, and {\rm b)} the map $l_Q(m)=Q(x+m)-Q(x)$ is an isomorphism of $\mathbb{Z}^n$. \end{theorem} The proof of this theorem uses standard arguments in topology, we omit it. The canonical projections $\tau_{\mathbb{R}^n}:\mathbf{T}\mathbb{R}^n\to\mathbb{R}^n$ and $\tau_{T^n}:\mathbf{T}T^n\to T^n$ are smooth. The former is a lift of the latter, \[ \tau_{T^n}\mathbf{T}\pi=\pi\tau_{\mathbb{R}^n}, \] which lift sends $w\in\mathbf{T}_x\mathbb{R}^n$ to $x\in\mathbb{R}^n$. The covering map $\mathbf{T}\pi:\mathbf{T}\mathbb{R}^n \to \mathbf{T}T^n$ is a local diffeomorphism.A vector field on $T^n$ is a smooth map $Y:T^n\to\mathbf{T}T^n$ such that $\tau_{T^n}Y=\mathop{\rm id}_{T^n}$. A vector field on $\mathbb{R}^n$ is a smooth map $Z:\mathbb{R}^n\to\mathbf{T}\mathbb{R}^n$ such that $\tau_{\mathbb{R}^n}Z=\mathop{\rm id}_{\mathbb{R}^n}$. \begin{lemma}\label{lift4} If $Y$ is a vector field on $T^n$, then there is only one lift of $Y$ that is a vector field on $\mathbb{R}^n$. \end{lemma} \begin{proof} Let $x_0\in\mathbb{R}^n$, $\theta_0\in T^n$ be such that $Y\pi(x_0)=Y(\theta_0)$. Let $w_{x_0}\in\mathbf{T}_{x_0}\mathbb{R}^n$ be the only vector such that $\mathbf{T}\pi(w_{x_0})=Y(\theta_0)$. By the Lifting Theorem (Theorem 4.1, p.143 \cite{BR}), there exists a unique lift $Z:\mathbb{R}^n\to\mathbf{T}\mathbb{R}^n$ such that $Y\pi=\mathbf{T}\pi Z$ and $Z(x_0)=w_{x_0}$. It needs only be checked that this $Z$ is a vector field. Because $Y$ is a vector field on $T^n$, $Z$ is a lift of $Y\pi$, and $\tau_{\mathbb{R}^n}$ is a lift of $\tau_{T^n}$, it follows that \[ \pi(x)=\tau_{T_n}Y\pi(x)=\tau_{T^n}\mathbf{T}\pi Z(x)=\pi\tau_{\mathbb{R}^n}Z(x). \] So the difference $x-\tau_{\mathbb{R}^n}Z(x)$ is a discrete valued map. Because $\mathbb{R}^n$ is connected, this difference is a constant (see Proposition 4.5, p.10 \cite{BR}). This constant is zero because $\tau_{\mathbb{R}^n}Z(x_0)=x_0$, and so $\tau_{\mathbb{R}^n}Z=\mathop{\rm id}_{\mathbb{R}^n}$. The equation $Y\pi=\mathbf{T}\pi Z$ implies that $Z$ is smooth because $\pi$ and $\mathbf{T}\pi$ are local diffeomorphisms and because $Y$ is smooth. The choice of the only vector $w\in\mathbf{T}_{x_0+m}\mathbb{R}^n$ for any $0\ne m\in\mathbb{Z}^n$ such that $\mathbf{T}\pi(w)=Y(\theta_0)$ would lead to a lift $Z_m$ of $Y$ that is not a vector field on $\mathbb{R}^n$ because $\tau_{\mathbb{R}^n}Z_m(x)=x+m$. The collection $\{Z_m:m\in\mathbb{Z}\}$, with $Z_0=Z$, accounts for all the lifts of $Y$ by the uniqueness of the lift and the uniqueness of the vector $w$. Therefore $Z$ is the only lift of $Y$ that is a vector field on $\mathbb{R}^n$. \end{proof} For a vector field $X$ on $T^n$, let $\hat X$ denote the only lift of $X$ that is a vector field on $\mathbb{R}^n$ as described in Lemma \ref{lift4}; $\hat X$ satisfies $X\pi=\mathbf{T}\pi \hat X$. For a diffeomorphism $R$ of $T^n$, let $\hat R$ be a lift of $R$; the lift $\hat R$ is a diffeomorphism of $\mathbb{R}^n$ (by Theorem \ref{lift3}) for which $R\pi=\pi\hat R$. \begin{lemma}\label{lift5} The only lift of the vector field $R_*X$ on $T^n$ that is a vector field on $\mathbb{R}^n$ is $\hat R_*\hat X$.\end{lemma} \begin{proof}A lift of $R_*X$ is $\hat R_*\hat X$ because \begin{align*} \mathbf{T}\pi\hat R_*\hat X & =\mathbf{T}\pi\mathbf{T}\hat R\hat X\hat R^{-1} = \mathbf{T}(\pi \hat R)\hat X\hat R^{-1} = \mathbf{T}(R\pi)\hat X\hat R^{-1} \\ & = \mathbf{T}R\mathbf{T}\pi\hat X\hat R^{-1} = \mathbf{T}RX\pi\hat R^{-1} = \mathbf{T}RXR^{-1}\pi = R_*X\pi. \end{align*} By definition, $\hat R_*\hat X$ is a vector field on $\mathbb{R}^n$. By Lemma \ref{lift4}, it is the only lift of $R_*X$ that is a vector field on $\mathbb{R}^n$.\end{proof} \begin{lemma}\label{lift6} For any $\alpha\in\mathbb{R}^*$, the only lift of the vector field $\alpha X$ on $T^n$ that is a vector field on $\mathbb{R}^n$ is $\alpha\hat X$.\end{lemma} \begin{proof}A lift of $\alpha X$ is $\alpha\hat X$ because $\mathbf{T}\pi(\alpha\hat X)=\alpha\mathbf{T}\pi\hat X=\alpha X\pi$. Only one lift of $\alpha X$ is a vector field (Lemma \ref{lift4}), and $\alpha\hat X$ is this lift. \end{proof} \begin{theorem}\label{lift7} Let $X$ be a vector field on $T^n$, $\hat X$ the lift of $X$ that is a vector field on $\mathbb{R}^n$, $R$ a diffeomorphism of $T^n$, $\hat R$ a lift of $R$, and $\alpha$ a nonzero real number. Then $R_*X=\alpha X$ if and only if $\hat R_*\hat X=\alpha\hat X$. \end{theorem} \begin{proof} Suppose that $R_*X=\alpha X$. By Lemma \ref{lift5}, $\hat R_*\hat X$ is a lift of $R_*X$: $\mathbf{T}\pi\hat R_*\hat X=R_*X\pi$. By Lemma \ref{lift6}, $\alpha \hat X$ is a lift of $\alpha X$: $\mathbf{T}\pi(\alpha \hat X)=\alpha X\pi$. Then \[ \mathbf{T}\pi\big(\hat R_*\hat X-\alpha\hat X\big)=\big(R_*X-\alpha X\big)\pi=\mathbf{0}_{T^n}\pi, \] where $\mathbf{0}_{T^n}$ is the zero vector field on $T^n$. So $\hat R_*\hat X-\alpha \hat X$ is a lift of $\mathbf{0}_{T^n}$. The only lift of $\mathbf{0}_{T^n}$ that is a vector field on $\mathbb{R}^n$ is $\mathbf{0}_{\mathbb{R}^n}$, the zero vector field on $\mathbb{R}^n$. By Lemma \ref{lift5} and Lemma \ref{lift6}, the difference $\hat R_*\hat X-\alpha\hat X$ is a vector field on $\mathbb{R}^n$. By Lemma \ref{lift4}, $\hat R_*\hat X-\alpha\hat X=\mathbf{0}_{\mathbb{R}^n}$. Thus, $\hat R_*\hat X=\alpha\hat X$.Suppose that $\hat R_*\hat X=\alpha\hat X$. Then \begin{align*} R_*X\pi& = \mathbf{T}RXR^{-1}\pi= \mathbf{T}RX\pi\hat R^{-1} = \mathbf{T}R\mathbf{T}\pi\hat X\hat R^{-1} \\ & = \mathbf{T}(R\pi)\hat X\hat R^{-1} = \mathbf{T}(\pi\hat R)\hat X\hat R^{-1}= \mathbf{T}\pi\mathbf{T}\hat R\hat X\hat R^{-1} \\ & = \mathbf{T}\pi\hat R_*\hat X = \mathbf{T}\pi(\alpha\hat X)= \alpha\mathbf{T}\pi\hat X= \alpha X\pi. \end{align*} The surjectivity of $\pi$ implies that $R_*X=\alpha X$. \end{proof} \section{Solving the Lifted Generalized Symmetry Equation} The lift of $R_*X=\alpha X$ is an equation on $\mathbf{T}\mathbb{R}^n$ of the form $Q_*\hat X=\alpha\hat X$ for $Q\in\mathop{\rm Diff}(\mathbb{R}^n)$. With global coordinates $x=(x_1,x_2,\dots ,x_n)$ on $\mathbb{R}^n$, the diffeomorphism $Q$ has the form \[ Q(x_1,x_2,\dots ,x_n)=(f_1(x_1,x_2,\dots ,x_n),\dots ,f_n(x_1,x_2,\dots ,x_n)) \] for smooth functions $f_i:\mathbb{R}^n\to\mathbb{R}$, $i=1,\dots ,n$. Let $\theta=(\theta_1,\theta_2,\dots ,\theta_n)$ be global coordinates on $T^n$ such that $\theta_i=x_i$ mod 1, $i=1,2,\dots ,n$. If \[ X(\theta)=a_1\frac{\partial}{\partial \theta_1}+ a_2\frac{\partial}{\partial \theta_2}+\dots+a_n\frac{\partial}{\partial \theta_n} \] for constants $a_i\in\mathbb{R}$, $i=1,\dots ,n$, then \[ \hat X(x)= a_1\frac{\partial}{\partial x_1}+ a_2\frac{\partial}{\partial x_2} +\dots +a_n\frac{\partial}{\partial x_n}, \] so that $Q_*\hat X=\alpha\hat X$ has the form \[ \sum^n_{j=1}a_j\frac{\partial f_i}{\partial x_j}=\alpha a_i,\ i=1,\dots ,n. \] This is an uncoupled system of linear, first order equations which is readily solved for its general solution. \begin{lemma}\label{solve1} For real numbers $a_1,a_2,\dots ,a_n$ and $\alpha$ with $a_n\ne0$, the general solution of the system of $n$ linear partial differential equations \[ \sum^n_{j=1}a_j\frac{\partial f_i}{\partial x_j}=\alpha a_i,\ i=1,\dots ,n \] is \[f_i(x)=\alpha\frac{a_i}{a_n}x_n + h_i\Big(x_1-\frac{a_1}{a_n}x_n,x_2 -\frac{a_2}{a_n}x_n,\dots ,x_{n-1}-\frac{a_{n-1}}{a_n}x_n\Big), \] for arbitrary smooth functions $h_i:\mathbb{R}^{n-1}\to\mathbb{R}$, $i=1,\dots ,n$. \end{lemma} \begin{proof} For each $i=1,\dots ,n$, consider the initial value problem \begin{gather*} \sum^n_{j=1}a_j\frac{\partial f_i}{\partial x_j}=\alpha a_i \\ x_j(0,s_1,s_2,\dots ,s_{n-1}) = s_j \hbox{\rm\ for\ }j=1,\dots ,n-1 \\ x_n(0,s_1,s_2,\dots ,s_{n-1}) = 0 \\ f_i(0,s_1,s_2,\dots ,s_{n-1}) = h_i(s_1,s_2,\dots ,s_{n-1}) \end{gather*} for parameters $(s_1,s_2,\dots ,s_{n-1})\in\mathbb{R}^{n-1}$ and initial data $h_i:\mathbb{R}^{n-1}\to\mathbb{R}$. Using the method of characteristics (see \cite{JO} for example), the solution of the initial value problem in parametric form is \begin{gather*} x_j(t,s_1,s_2,\dots ,s_{n-1}) = a_jt+s_j\hbox{ for }j=1,\dots ,n-1 \\ x_n(t,s_1,s_2,\dots ,s_{n-1}) = a_nt \\ f_i(t,s_1,s_2,\dots ,s_{n-1}) = \alpha a_it +h_i(s_1,s_2,\dots ,s_{n-1}). \end{gather*} The coordinates $(x_1,x_2,\dots ,x_n)$ and the parameters $(t,s_1,s_2,\dots ,s_{n-1})$ are related by \[ \begin{bmatrix} x_1 \\ x_2 \\ x_3 \\ \vdots \\ x_{n-1} \\ x_n \end{bmatrix} =\begin{bmatrix}a_1 & 1 & 0 & 0 & \dots & 0 \\a_2 & 0 & 1 & 0 & \dots & 0 \\ a_3 & 0 & 0 & 1 & \ldots & 0 \\\vdots & \vdots & \vdots & \vdots & \ddots & \vdots \\ a_{n-1} & 0 & 0 & 0 & \ldots & 1 \\ a_n & 0 & 0 & 0 & \ldots & 0\end{bmatrix} \begin{bmatrix} t \\ s_1 \\ s_2 \\ \vdots \\ s_{n-2} \\ s_{n-1} \end{bmatrix} \] The determinant of the $n\times n$ matrix is $(-1)^na_n$, which is nonzero by hypothesis. Inverting the matrix equation gives \[ \begin{bmatrix} t \\ s_1 \\ s_2 \\ \vdots \\ s_{n-2} \\ s_{n-1} \end{bmatrix} =\begin{bmatrix}0 & 0 & \ldots & 0 & 0 & 1/a_n \\1 & 0 & \ldots & 0 & 0 & -a_1/a_n \\0 & 1 & \ldots & 0 & 0 & -a_2/a_n \\ \vdots & \vdots & \ddots & \vdots & \vdots & \vdots \\ 0 & 0 & \ldots & 1 & 0 & -a_{n-2}/a_n \\0 & 0 & \ldots & 0 & 1 & -a_{n-1}/a_n \end{bmatrix} \begin{bmatrix}x_1 \\ x_2 \\ x_3 \\ \vdots \\ x_{n-1} \\ x_n\end{bmatrix} \] Substitution of the expressions for $t$ and the $s_i$'s in terms of the $x_i$'s into \[ f_i(x_1,x_2,\dots ,x_n)=\alpha a_it+h_i(s_1,s_2,\dots ,s_{n-1}) \] gives the desired form of the general solution. \end{proof} \begin{lemma}\label{solve2} If $a_1,a_2,\dots ,a_n$ are independent over $\mathbb{Q}$, then \[ J=\Big\{\Big(m_1-\frac{a_1}{a_n}m_n,\dots ,m_{n-1}-\frac{a_{n-1}}{a_n}m_n\Big): m_1,dots ,m_n\in\mathbb{Z}\Big\} \] is a dense subset of $\mathbb{R}^{n-1}$.\end{lemma} \begin{proof}Suppose $a_1,a_2,\dots ,a_n$ are independent over $\mathbb{Q}$. This implies that none of the $a_i$'s are zero. In particular, $a_n\ne0$. Consider the flow \[ \psi_t(\theta_1,\dots ,\theta_{n-1},\theta_n)=(\theta_1-(a_1/a_n)t,\dots , \theta_{n-1}-(a_{n-1}/a_n)t,\theta_n-t) \] on $T^n$ which is generated by the vector field \[ Y=-\frac{a_1}{a_n}\frac{\partial}{\partial \theta_1}-\frac{a_2}{a_n} \frac{\partial}{\partial \theta_2}-\dots -\frac{a_{n-1}}{a_n} \frac{\partial}{\partial \theta_{n-1}}-\frac{\partial}{\partial \theta_n}. \] The coefficients of $Y$ are independent over $\mathbb{Q}$ because $a_1,a_2,\dots ,a_n$ are independent over $\mathbb{Q}$ and \[ m_1a_1+\dots +m_na_n=0\Leftrightarrow -m_1\frac{a_1}{a_n}-\dots -m_{n-1} \frac{a_{n-1}}{a_n}-m_n=0. \] So the orbit of $\psi$ through any point $\theta_0\in T^n$, \[ \gamma_\psi(\theta_0)=\{\psi_t(\theta_0):t\in\mathbb{R}\}, \] is dense in $T^n$ (Corollary 1, p. 287 \cite{AR}).The submanifold \[ P=\{(\theta_1,\dots ,\theta_{n-1},\theta_n):\theta_n=0\} \] of $T^n$, which is diffeomorphic to $T^{n-1}$, is a global Poincar\'e section for $\psi$ because $X(\theta)\not\in\mathbf{T}_\theta P$ for every $\theta\in P$ and because $\gamma_\psi(\theta_0)\cap P\ne\emptyset$ for every $\theta_0\in T^n$. Define the projection $\wp:T^n \to T^{n-1}$ by \[ \wp(\theta_1,\theta_2,\dots ,\theta_{n-1},\theta_n)=(\theta_1,\theta_2,\dots , \theta_{n-1}) \] and the injection $\imath:T^{n-1}\to T^n$ by\[\imath(\theta_1,\theta_2,\dots , \theta_{n-1})=(\theta_1,\theta_2,\dots ,\theta_{n-1},0). \] The Poincar\'e map induced on $\wp(P)$ by $\psi$ is given by $\bar\psi=\wp\psi_1\imath$ because $\psi_1(\theta_0)\in P$ when $\theta_0\in P$. For any $\kappa\in\mathbb{Z}$, $\bar\psi^\kappa=\wp\psi_{\kappa}\imath$. So, for instance, with $0=(0,0,\dots ,0)\in T^n$ and $\bar 0=\wp(0)$, \[ \wp\big(\gamma_\psi(0)\cap P\big) = \{\bar\psi^\kappa(\bar 0):\kappa\in\mathbb{Z}\} = \Big\{\Big(-\frac{a_1}{a_n}\kappa,-\frac{a_2}{a_n}\kappa,\dots , -\frac{a_{n-1}}{a_n}\kappa\Big):\kappa\in\mathbb{Z}\Big\}, \] where for each $i=1,\dots ,n-1$, the quantity $-(a_i/a_n)\kappa$ is taken mod 1. With $\bar\pi:\mathbb{R}^{n-1}\to T^{n-1}$ as the covering map, \[ J=\bar\pi^{-1}\big(\wp(\gamma_\psi(0)\cap P)\big). \] If $\wp(\gamma_\psi(0)\cap P)$ were dense in $\wp(P)$, then $J$ would be dense in $R^{n-1}$ because $\bar\pi$ is a covering map. (That is, if $\wp(\gamma_\psi(0)\cap P)\cap[0,1)^{n-1}$ is dense in the fundamental domain $[0,1)^{n-1}$ of the covering map $\bar\pi$, then by translation, it is dense in $\mathbb{R}^{n-1}$.)Define $\chi:\mathbb{R}\times T^{n-1}\to T^n$ by \[ \chi(t,\theta_1,\theta_2,\dots ,\theta_{n-1})=\psi\big(t,\imath(\theta_1,\theta_2, \dots ,\theta_{n-1})\big). \] The map $\chi$ is a local diffeomorphism by the Inverse Function Theorem because \[ \mathbf{T}\chi=\begin{bmatrix}-a_1/a_n & 1 & 0 & \ldots & 0 \\-a_2/a_n & 0 & 1 & \ldots & 0 \\\vdots & \vdots & \vdots & \ddots & \vdots \\-a_{n-1}/a_n & 0 & 0 &\ldots & 1 \\-1 & 0 & 0 & \ldots & 0\end{bmatrix} \] has determinant of $(-1)^{n+1}$. Let $O$ be a small open subset of $\wp(P)$. For $\epsilon>0$, the set $O_\epsilon=(-\epsilon,\epsilon)\times O$ is an open subset in the domain of $\chi$. For $\epsilon$ small enough, the image $\chi(O_\epsilon)$ is open in $T^n$ because $\chi$ is a local diffeomorphism. By the denseness of $\gamma_\psi(0)$ in $T^n$, there is a point $\theta_0$ in $\chi(O_\epsilon)\cap\gamma_\psi(0)$. By the definition of $\chi(O_\epsilon)$, there is an $\bar\epsilon\in(-\epsilon,\epsilon)$ and a $\bar\theta_0\in O$ such that $\chi(\bar\epsilon,\bar\theta_0)=\theta_0$. Thus $\imath(\bar \theta_0)\in\gamma_\psi(0)$, and so $\wp(\gamma_\psi(0)\cap P)$ intersects $O$ at $\bar\theta_0$. Since $O$ is any small open subset of $\wp(P)$, the set $\wp(\gamma_\psi(0)\cap P)$ is dense in $\wp(P)$. \end{proof} \begin{theorem}\label{solve3} If $\alpha\in\mathbb{R}^*$ and the coefficients of $X=\sum_{i=1}^n a_i\partial/\partial\theta_i$ are independent over $\mathbb{Q}$, then for each $R\in\mathop{\rm Diff}(T^n)$ that satisfies $R_*X=\alpha X$ there exist $B=(b_{ij})\in\mathop{\rm GL}(n,\mathbb{Z})$ and $c\in\mathbb{R}^n$ such that \[\hat R(x)=Bx+c \] for $x=(x_1,x_2,\dots ,x_n)$, in which\[b_{in}= \alpha\frac{a_i}{a_n}-\sum_{j=1}^{n-1}b_{ij}\frac{a_j}{a_n},\ i=1,\dots ,n. \]\end{theorem} \begin{proof} Suppose that the $a_1,a_2,\dots ,a_n$ are independent over $\mathbb{Q}$. For $\alpha\in\mathbb{R}^*$, suppose that $R\in\mathop{\rm Diff}(T^n)$ is a solution of $R_*X=\alpha X$. A lift $\hat R$ of $R$ is a diffeomorphism of $\mathbb{R}^n$ by Theorem \ref{lift3}. The lift of $X$ that is a vector field on $\mathbb{R}^n$ is $\hat X=\sum^n_{i=1}a_i(\partial/\partial x_i)$. By Theorem \ref{lift7}, $\hat R$ is a solution of $\hat R_*\hat X=\alpha\hat X$. With global coordinates $(x_1,x_2,\dots ,x_n)$ on ${\mathbb R }^n$ write \[ \hat R(x)=(f_1(x_1,\dots ,x_n),\dots ,f_n(x_1,\dots ,x_n)). \] In terms of this coordinate description, the equation $\hat R_*\hat X=\alpha X$ written out is \[ \sum_{j=1}^n a_j\frac{\partial f_i}{\partial x_j}=\alpha a_i,\ i=1,\dots ,n. \] The independence of the coefficients of $\hat X$ over $\mathbb{Q}$ implies that $a_n\ne0$. By Lemma \ref{solve1}, there are smooth functions $h_i:\mathbb{R}^{n-1}\to\mathbb{R}$, $i=1,\dots ,n$, such that \[ f_i(x_1,\dots ,x_n)=\alpha\frac{a_i}{a_n}x_n+h_i(s_1,s_2,\dots ,s_{n-1}) \] where \[ s_i=x_i-\frac{a_i}{a_n}x_n,\ i=1,\dots ,n-1. \] By Theorem \ref{lift3}, $\hat R(x+m)-\hat R(x)$ is independent of $x$ for each $m\in\mathbb{R}^n$. This implies for each $i=1,\dots ,n$ that \begin{align*} & f_i(x+m)-f_i(x)\\ & =f_i(x_1+m_1,x_2+m_2,\dots ,x_n+m_n)-f_i(x_1,x_2,\dots ,x_n) \\ & =\alpha\frac{a_i}{a_n}m_n + h_i\big(s_1+m_1-\frac{a_1}{a_n}m_n,\dots ,s_{n-1} +m_{n-1}-\frac{a_{n-1}}{a_n}m_n\big) - h_i(s_1,\dots ,s_{n-1}) \end{align*} is independent of $x$ for every $m=(m_1,m_2,\dots ,m_n)\in\mathbb{Z}^n$. This independence means that $f_i(x+m)-f_i(x)$ is a function of $m$ only. So for each $j=1,\dots ,n-1$, \begin{align*} 0 & = \frac{\partial}{\partial x_j}\big[f_i(x_1+m_1,x_2+m_2,\dots,x_n+m_n) -f_i(x_1,x_2,\dots ,x_n)\big] \\ & = \frac{\partial h_i}{\partial s_j}\Big(s_1+m_1-\frac{a_1}{a_n}m_n,\dots , s_{n-1}+m_{n-1}-\frac{a_{n-1}}{a_n}m_n\Big) -\frac{\partial h_i}{\partial s_j}\big(s_1,\dots ,s_{n-1}\big). \end{align*} So, in particular \[ \frac{\partial h_i}{\partial s_j}\Big(m_1-\frac{a_1}{a_n}m_n,\dots ,m_{n-1} -\frac{a_{n-1}}{a_n}m_n\Big)= \frac{\partial h_i}{\partial s_j}\big(0,\dots ,0\big) \] for all $(m_1,\dots ,m_n)\in\mathbb{Z}^n$. By Lemma \ref{solve2}, the set \[ \Big\{\Big(m_1-\frac{a_1}{a_n}m_n,\dots ,m_{n-1}-\frac{a_{n-1}}{a_n}m_n\Big): m_1,\dots ,m_n\in\mathbb{Z}\Big\} \] is dense in $\mathbb{R}^{n-1}$, which together with the smoothness of $h_i$ implies that $\partial h_i/\partial s_j$ is a constant. Let this constant be $b_{ij}$ for $i=1,\dots ,n$, $j=1,\dots ,n-1$. By Taylor's Theorem, \[ h_i(s_1,\dots ,s_{n-1})=c_i + \sum_{j=1}^{n-1} b_{ij}s_j \] for constants $c_i\in\mathbb{R}$. Thus, \begin{align*} f_i(x_1,\dots ,x_n) & = c_i + \alpha\frac{a_i}{a_n}x_n+\sum_{j=1}^{n-1} b_{ij} \Big(x_j-\frac{a_j}{a_n}x_n\Big) \\ & = c_i + \sum_{j=1}^{n-1}b_{ij}x_j+\Big(\alpha\frac{a_i}{a_n} -\sum_{j=1}^{n-1}b_{ij}\frac{a_j}{a_n}\Big)x_n. \end{align*} For each $i=1,2,\dots ,n$, set \[ b_{in}=\alpha\frac{a_i}{a_n}-\sum_{j=1}^{n-1}b_{ij}\frac{a_j}{a_n}\Big. \] Then for each $i=1,2,\dots ,n$, \[ f_i(x_1,x_2,\dots ,x_n)=c_i + \sum_{j=1}^n b_{ij}x_j. \] So $\hat R$ has the form $\hat R(x)=Bx+c$ where $B=(b_{ij})$ is an $n\times n$ matrix, and $c\in\mathbb{R}^n$.By Theorem \ref{lift3}, the map $l_{\hat R}(m)=\hat R(x+m)-\hat R(x)$ is an isomorphism of $\mathbb{Z}^n$. By the formula for $f_i$ derived above, \[ f_i(x_1+m_1,\dots ,x_n+m_n)-f_i(x_1,x_2,\dots ,x_m)=\sum_{j=1}^n b_{ij}m_j \] for each $i=1,2,\dots ,n$. This implies that $l_{\hat R}(m)=Bm$. Since $l_{\hat R}$ is an isomorphism of $\mathbb{Z}^n$, it follows that $B\in\mathop{\rm GL}(n,\mathbb{Z})$. \end{proof} Theorem \ref{solve3} restricts the search for lifts of generalized symmetries of a quasiperiodic flow on $T^n$ to affine maps on $\mathbb{R}^n$ of the form $Q(x)=Bx+c$ for $B\in{\rm GL}(n,\mathbb{Z})$ and $c\in\mathbb{R}^n$. For an affine map of this form, the difference \[ Q(x+m)-Q(x)=B(x+m)+c-(Bx+c)=Bm \] is independent of $x$, and the map $l_Q(m)=Q(x+m)-Q(x)$ is an isomorphism of $\mathbb{Z}^n$, so that $Q$ is a lift of a diffeomorphism $R$ on $T^n$ by Theorem \ref{lift3}. If $Q$ is a solution of $Q_*\hat X=\alpha\hat X$, then by Theorem \ref{lift7}, $R$ is a solution of $R_*X=\alpha X$, so that by Theorem \ref{char}, $R\in S_\phi$.The following two corollaries of Theorem \ref{solve3} restrict the possibilities for the multipliers of the generalized symmetries of a quasiperiodic flow on $T^n$. One restriction employs the notion of an {\it algebraic integer}, which is a complex number that is a root of a monic polynomial in the polynomial ring $\mathbb{Z}[z]$. If $m$ is the smallest degree of a monic polynomial in $\mathbb{Z}[z]$ for which an algebraic integer is a root, then $m$ is the {\it degree} of that algebraic integer (Definition 1.1, p.1 \cite{RM}). \begin{corollary}\label{solve4} If $\phi$ is a quasiperiodic flow on $T^n$ with generating vector field $X=\sum^n_{i=1}a_i\partial/\partial\theta_i$, then each $\alpha\in\rho_\phi(S_\phi)$ is a real algebraic integer of degree at most $n$, and $\rho_\phi(S_\phi)\cap\mathbb{Q}=\{1,-1\}$. \end{corollary} \begin{proof} For each $\alpha\in\rho_\phi(S_\phi)$ (which is real) there is an $R\in S_\phi$ such that $\rho_\phi(R)=\alpha$. By Theorem \ref{solve3} there is a $B\in\mathop{\rm GL}(n,\mathbb{Z})$ such that $\mathbf{T}\hat R=B$. Then by Theorem \ref{char} and Theorem \ref{lift7}, \[ B\hat X=\hat R_*\hat X=\alpha\hat X. \] So, $\alpha$ is an eigenvalue of $B$ (and $\hat X$ is an eigenvector of $B$.) The characteristic polynomial of $B$ is an n-degree monic polynomial in $\mathbb{Z}[z]$: \[ z^n+d_{n-1}z^{n-1}+\dots +d_1z+d_0. \] Thus $\alpha$ is a real algebraic integer of degree at most $n$. The value of $d_0$ is $\hbox{\rm det}(B)$, which is a unit in $\mathbb{Z}$ (Theorem 3.5, p.351 \cite{HU}). The only units in $\mathbb{Z}$ are $\pm 1$. So the only possible rational roots of the characteristic polynomial of $B$ are $\pm1$ (Proposition 6.8, p.160 \cite{HU}). This means that $\rho_\phi(S_\phi)\cap\mathbb{Q}\subset\{1,-1\}$. But $\rho_\phi(S_\phi)\cap\mathbb{Q}\supset\{1,-1\}$ by Theorem \ref{mult}. Thus, $\rho_\phi(S_\phi)\cap\mathbb{Q}=\{1,-1\}$. \end{proof} The other restriction on the possibilities for the multipliers of any generalized symmetries of $\phi$ employs linear combinations over $\mathbb{Z}$ of pair wise ratios of the entries of the ``eigenvector'' $\hat X$ (which entries are the frequencies of $\phi$).\begin{corollary}\label{solve5}If $\phi$ is a quasiperiodic flow on $T^n$ with generating vector field $X=\sum^n_{i=1}a_i\partial/\partial\theta_i$, then for any $\alpha\in\rho_\phi(S_\phi)$ there exists a $B=(b_{ij})\in\mathop{\rm GL}(n,\mathbb{Z})$ such that \[ \alpha=\sum_{j=1}^n b_{ij}\frac{a_j}{a_i},\ \ i=1,\dots ,n. \] \end{corollary} \begin{proof} Suppose that $\alpha\in\rho_\phi(S_\phi)$. Then there is an $R\in S_\phi$ such that $\alpha=\rho_\phi(R)$. By Theorem \ref{solve3}, there is a $B=(b_{ij})\in{\rm GL}(n,\mathbb{Z})$ such that $\mathbf{T}\hat R=B$ with \[ b_{in}= \alpha\frac{a_i}{a_n}-\sum_{j=1}^{n-1}b_{ij}\frac{a_j}{a_n},\quad i=1,\dots ,n. \] Solving this equation for $\alpha$ gives \[ \alpha=\sum_{j=1}^n b_{ij}\frac{a_j}{a_i},\ \ i=1,\dots ,n. \] \end{proof} The multiplier group of any quasiperiodic flow $\phi$ always contains $\{1,-1\}$ as stated in Theorem \ref{mult}. For each $t\in\mathbb{R}$, the diffeomorphism $\phi_t$ is in $S_\phi$ by definition. A lift of $\phi_t$ is $\hat \phi_t(x)=Ix+t\hat X$, where $I=\delta_{ij}$ is the $n\times n$ identity matrix, so that by Corollary \ref{solve5}, \[ \alpha=\sum^n_{j=1}\delta_{ij}\frac{a_j}{a_i}=\frac{a_i}{a_i}=1 \] for each $i=1,\dots ,n$. A lift of the reversing involution $N$ defined in the proof of Theorem \ref{mult} is $\hat N(x)=-Ix$, so that by Corollary \ref{solve5}, \[ \alpha=-\sum_{j=1}^n \delta_{ij}\frac{a_j}{a_i}=-\frac{a_i}{a_i}=-1 \] for each $i=1,\dots ,n$. Corollary \ref{solve5} enables a complete description of all symmetries and reversing symmetries of $\phi$. \begin{theorem}\label{solve6} Suppose that $\phi$ is a quasiperiodic flow on $T^n$ with generating vector field $X=\sum^n_{i=1}a_i\partial/\partial\theta_i$. If $\rho_\phi(R)=\pm1$ for an $R\in S_\phi$, then there is $c\in\mathbb{R}^n$ such that $\hat R(x)=\rho_\phi(R)Ix+c$. \end{theorem} \begin{proof} Let $R\in S_\phi$. By Theorem \ref{solve3} there exists a $B=(b_{ij})\in\mathop{\rm GL}(n,\mathbb{Z})$ and a $c\in\mathbb{R}^n$ such that $\hat R(x)=Bx+c$. By Corollary \ref{solve5}, the entries of $B$ satisfy \[ \rho_\phi(R)=\sum_{j=1}^n b_{ij}\frac{a_j}{a_i} \] for each $i=1,2,\dots ,n$. By hypothesis, $\rho_\phi(R)=\pm 1$. Then for each $i=1,2,\dots ,n$, \[ b_{i1}a_1+\dots +(b_{ii}\mp1)a_i+\dots +b_{in}a_n=0. \] By the independence of $a_1,a_2,\dots ,a_n$ over $\mathbb{Q}$, $b_{ij}=0$ when $i\ne j$ and $b_{ii}=\rho_\phi(R)$ for all $i=1,2,\dots ,n$. Therefore, $\hat R(x)=\rho_\phi(R)Ix+c$.\end{proof} \begin{corollary}\label{solve7} If $\phi$ is a quasiperiodic flow on $T^n$, then $\ker \rho_\phi\cong T^n$. \end{corollary} \begin{proof} Let $R\in S_\phi$ such that $\rho_\phi(R)=1$. By Theorem \ref{solve6}, $\hat R(x)=Ix+c$ for some $c\in\mathbb{R}^n$. Now, for any $c\in\mathbb{R}^n$, the $Q\in\mathop{\rm Diff}(T^n)$ induced by $\hat Q(x)=Ix+c$ satisfies $Q_*X=X$ by Theorem \ref{lift7} because $\hat Q_*\hat X=\hat X$. So, by Theorem \ref{char}, $Q\in\ker \rho_\phi$. Since $c$ is arbitrary, $Q\pi=\pi\hat Q$, and $\pi(\mathbb{R}^n)=T^n$, it follows that $\ker \rho_\phi\cong T^n$. \end{proof} \begin{corollary}\label{solve8} If $\phi$ is a quasiperiodic flow on $T^n$, then every reversing symmetry of $\phi$ is an involution.\end{corollary} \begin{proof}Suppose $R\in S_\phi$ is a reversing symmetry. By Theorem \ref{solve6}, $\hat R(x)=-Ix+c$ for some $c\in\mathbb{R}^n$, and so $\hat R^2(x)=Ix$. This implies that $R^2=\mathop{\rm id}_{T^n}$. \end{proof} \begin{example}\label{example2} \rm Recall the quasiperiodic flow $\phi$ on $T^3$ and its generating vector field \[ X=\frac{\partial}{\partial\theta_1}+7^{1/3}\frac{\partial}{\partial\theta_2}+ 7^{2/3}\frac{\partial}{\partial\theta_3} \] from Example \ref{example1}. By Corollary \ref{solve7}, the symmetry group of $\phi$ is exactly the group of translations $\{R_c:c\in T^n\}$ on $T^n$, where $R_c(\theta)=\theta+c$. By Corollary \ref{solve8}, every reversing symmetry of $\phi$ is an involution. In particular, this implies that the reversing symmetry group of $\phi$ is a semidirect product of the symmetry group of $\phi$ by the $\mathbb{Z}_2$ subgroup generated by reversing involution $N(\theta)=-\theta$ (see p.8 in \cite{LR}). Are there symmetries of $\phi$ with multipliers other than $\pm 1$? The $\mathop{\rm GL}(3,\mathbb{Z})$ matrix \[ B=(b_{ij})=\begin{bmatrix}-2 & 1 & 0 \\ 0 & -2 & 1 \\ 7 & 0 & -2\end{bmatrix} \] induces a $Q\in\mathop{\rm Diff}(T^3)$ by Theorem \ref{lift3}. Since \[ \hat Q_*\hat X=\mathbf{T}\hat Q\hat X=B\hat X=\big( -2+7^{1/3}\big)\hat X, \] Theorem \ref{lift7} implies that $Q_*X=(-2+7^{1/3})X$. Hence, by Theorem \ref{char}, $Q\in S_\phi$. The number $-2+7^{1/3}$ is $\rho_\phi(Q)$, the multiplier of $Q$, is an algebraic integer of degree at most $3$ by Corollary \ref{solve4}, and satisfies \[ -2+7^{1/3}=\sum_{j=1}^3 b_{ij}\frac{a_j}{a_i}, \quad i=1,2,3, \] by Corollary \ref{solve5}. (The matrix $B$ was found by using Theorem 3.1 in \cite{BA1}, a result which characterizes the matrices in $\mathop{\rm GL}(3,\mathbb{Z})$ inducing generalized symmetries of a quasiperiodic flow generated by a vector field of a certain type, of which $X$ above is.) Since $S_\phi$ is a group and $\rho_\phi:S_\phi\to\mathbb{R}^*$ is a homomorphism, it follows for each $k\in\mathbb{Z}$ that $Q^k\in S_\phi$ with $\rho_\phi(Q^k)=\big(\rho_\phi(Q)\big)^k=(-2+7^{1/3})^k$, and that $NQ^k\in S_\phi$ with $\rho_\phi(NQ^k)=-(-2+7^{1/3})^k$. \end{example} \section{A Splitting Map for the Extension} For a quasiperiodic flow $\phi$ on $T^n$, Theorem \ref{solve3} implies that $\mathbf{T}\hat R\in\mathop{\rm GL}(n,\mathbb{Z})$ for every $R\in S_\phi$. Set \[ \Pi_\phi=\{B\in\mathop{\rm GL}(n,\mathbb{Z}): \hbox{ there is }R\in S_\phi\hbox{ for which }B=\mathbf{T}\hat R\}, \] and define a map $\nu_\phi:\Pi_\phi\to\rho_\phi(S_\phi)$ by $\nu_\phi(B)=\rho_\phi(R)$ where $R\in S_\phi$ with $\mathbf{T}\hat R=B$. \begin{lemma}\label{rep0} If $\phi$ is a quasiperiodic flow on $T^n$ with generating vector field $X$, then $\nu_\phi$ is well-defined. \end{lemma} \begin{proof} Let $B\in\Pi_\phi$, and suppose there are $R,Q\in S_\phi$ with $\mathbf{T}\hat R=B=\mathbf{T}\hat Q$. Then $RQ^{-1}\in S_\phi$ and $\hat R\hat Q^{-1}$ is a lift of $RQ^{-1}$ for which $\mathbf{T}(\hat R\hat Q^{-1})=BB^{-1}=I$. Hence $\hat R\hat Q^{-1}(x)=Ix+c$ for some $c\in\mathbb{R}^n$. This implies that $(\hat R\hat Q^{-1})_*\hat X=\hat X$, so that by Theorem \ref{lift7}, $(RQ^{-1})_*X=X$. By Theorem \ref{char}, $\rho_\phi(RQ^{-1})=1$. Because $\rho_\phi$ is a homomorphism, $\rho_\phi(R)=\rho_\phi(Q)$.\end{proof} \begin{lemma}\label{rep1} If $\phi$ is a quasiperiodic flow on $T^n$ with generating vector field $X$, then $\Pi_\phi$ is a subgroup of $\mathop{\rm GL}(n,\mathbb{Z})$.\end{lemma} \begin{proof} Let $B,C\in\Pi_\phi$. Then there are $R,Q\in S_\phi$ such that $\mathbf{T}\hat R=B$ and $\mathbf{T}\hat Q=C$. The latter implies that $\mathbf{T}\hat Q^{-1}=(\mathbf{T}\hat Q)^{-1}=C^{-1}$. Then $BC^{-1}=\mathbf{T}\hat R\mathbf{T}\hat Q^{-1}=\mathbf{T}(\hat R\hat Q^{-1})$. The diffeomorphism $x\to \hat R\hat Q^{-1}x$ of $\mathbb{R}^n$ satisfies conditions a) and b) of Theorem \ref{lift3}, and so is a lift of a diffeomorphism $V$ of $T^n$. Let $\alpha=\rho_\phi(R)$ and $\beta=\rho_\phi(Q)$. Then $\rho_\phi(Q^{-1})=\beta^{-1}$ because $\rho_\phi$ is a homomorphism, and so $(\hat Q^{-1})_*\hat X=\beta^{-1}\hat X$. Thus, $\mathbf{T}(\hat R\hat Q^{-1})\hat X=(\hat R\hat Q^{-1})_*\hat X =\alpha\beta^{-1}\hat X$. By Theorem \ref{lift7}, $V_*X=\alpha\beta^{-1}X$, so that by Theorem \ref{char}, $V\in S_\phi$. The lifts $\hat R\hat Q^{-1}$ and $\hat V$ of $V$ differ by a deck transformation of $\pi$, so that $BC^{-1}=\mathbf{T}(\hat R\hat Q^{-1})=\mathbf{T}\hat V$. Therefore, $BC^{-1}\in \Pi_\phi$.\end{proof} \begin{theorem}\label{rep2} If $\phi$ is a quasiperiodic flow on $T^n$ with generating vector field $X$, then $\nu_\phi$ is an isomorphism and $\Pi_\phi$ is an Abelian subgroup of $\mathop{\rm GL}(n,\mathbb{Z})$.\end{theorem} \begin{proof} Let $B,C\in\Pi_\phi$. Then there are $R,Q\in S_\phi$ such that $\mathbf{T}\hat R=B$ and $\mathbf{T}\hat Q=C$. Let $\alpha=\rho_\phi(R)$ and $\beta=\rho_\phi(Q)$. By Theorem \ref{char} and Theorem \ref{lift7}, $\mathbf{T}\hat R\hat X=\alpha\hat X$ and $\mathbf{T}\hat Q\hat X=\beta\hat X$. By Lemma \ref{rep1}, $BC\in \Pi_\phi$, so that there is a $V\in S_\phi$ such that $\mathbf{T}\hat V=BC$. Hence, $\hat V_*\hat X=\mathbf{T}\hat V\hat X=BC\hat X=\alpha\beta\hat X$. By Theorem \ref{lift7} and Theorem \ref{char}, $\rho_\phi(V)=\alpha\beta$. Thus, $\nu_\phi(BC)=\alpha\beta=\nu_\phi(B)\nu_\phi(C)$. By definition, $\nu_\phi$ is surjective, and by Theorem \ref{solve6}, $\ker \nu_\phi=\{I\}$. Therefore, $\nu_\phi$ is an isomorphism. The multiplier group $\rho_\phi(S_\phi)$ is Abelian because it is a subgroup of the Abelian group $\mathbb{R}^*$. Thus $\Pi_\phi$ is Abelian.\end{proof} A splitting map for the short exact sequence, \[ \mathop{\rm id}{}_{T^n}\to \ker \rho_\phi\to \rho_\phi^{-1}(\Lambda) \stackrel{j_\Lambda}{\to } \Lambda\to 1, \] is a homomorphism $h_\Lambda:\Lambda\to\rho_\phi^{-1}(\Lambda)$ such that $j_\Lambda h_\Lambda$ is the identity isomorphism on $\Lambda$. Take for $h_\Lambda$ the map where for each $\alpha\in\Lambda$, the image $h_\Lambda(\alpha)$ is the diffeomorphism in $\rho_\phi^{-1}(\Lambda)$ induced by the $\mathop{\rm GL}(n,\mathbb{Z})$ matrix $\nu_\phi^{-1}(\alpha)$. \begin{theorem}\label{rep4} If $\phi$ is a quasiperiodic flow on $T^n$, then $h_\Lambda$ is a splitting map for the extension $\mathop{\rm id}_{T^n}\to\ker \rho_\phi\to\rho_\phi^{-1}(\Lambda) \to\Lambda\to 1$ for each $\Lambda<\rho_\phi(S_\phi)$.\end{theorem} \begin{proof} For arbitrary $\alpha,\beta\in\Lambda$, set $R=h_\Lambda(\alpha)$, $Q=h_\Lambda(\beta)$, and $V=h_\Lambda(\alpha\beta)$. Then $\hat R(x)=\nu_\phi^{-1}(\alpha)x$, $\hat Q(x)=\nu_\phi^{-1}(\beta)x$, and $\hat V(x)=\nu_\phi^{-1}(\alpha\beta)x$. By Theorem \ref{rep2}, $\nu_\phi^{-1}$ is an isomorphism, so that $\hat V(x)=\nu_\phi^{-1}(\alpha)\nu_\phi^{-1}(\beta)x$. Because \begin{align*} h_\Lambda(\alpha)h_\Lambda(\beta)\pi(x) & = RQ\pi(x)= \pi \hat R\hat Q(x)= \pi\nu_\phi^{-1}(\alpha)\nu_\phi^{-1}(\beta)x \\ & = \pi\nu_\phi^{-1}(\alpha\beta)x= \pi \hat V(x)= V\pi(x) = h_\Lambda(\alpha\beta)\pi(x), \end{align*} and because $\pi$ is surjective, $h_\Lambda(\alpha)h_\Lambda(\beta)=h_\Lambda(\alpha\beta)$. Let $B=\mathbf{T}\hat R=\nu_\phi^{-1}(\alpha)$. Then $\nu_\phi(B)=\rho_\phi(R)$, so that \[ j_\Lambda h_\Lambda(\alpha)=j_\Lambda(R)=\rho_\phi(R)=\nu_\phi(B) =\nu_\phi(\nu_\phi^{-1}(\alpha))=\alpha. \] Therefore, $h_\Lambda$ is a splitting map for the extension.\end{proof} \begin{theorem}\label{rep5} If $\phi$ is a quasiperiodic flow on $T^n$, then \[ \rho_\phi^{-1}(\Lambda)=\ker \rho_\phi \rtimes_\Gamma h_{\Lambda}(\Lambda) \] for each $\Lambda<\rho_\phi(S_\phi)$, where $\Gamma:h_\Lambda(\Lambda)\to\mathop{\rm Aut}(\ker \rho_\phi)$ is the conjugating homomorphism. Moreover, if $\Lambda$ is a nontrivial subgroup of $\rho_\phi(S_\phi)$, then $\Gamma$ is nontrivial. \end{theorem} \begin{proof} By Theorem \ref{rep4}, $h_\Lambda$ is a splitting map for the extension \[ \mathop{\rm id}{}_{T^n}\to \ker \rho_\phi\to \rho_\phi^{-1}(\Lambda) \stackrel{j_\Lambda}{\to } \Lambda\to 1. \] Thus, $\rho_\phi^{-1}(\Lambda)=\big(\ker \rho_\phi\big)\big(h_\Lambda(\Lambda)\big)$ and $\ker \rho_\phi\cap h_\Lambda(\Lambda)={\rm id}_{T^n}$ (Theorem 9.5.1, p.240 \cite{SC}). Since $\ker \rho_\phi$ is a normal subgroup of $\rho_\phi^{-1}(\Lambda)$, then $\rho_\phi^{-1}(\Lambda)=\ker \rho_\phi\rtimes_\Gamma h_\Lambda(\Lambda)$ where $\Gamma:h_\Lambda(\Lambda)\to\mathop{\rm Aut}(\ker \rho_\phi)$ is the conjugating homomorphism (see p.21 in \cite{AB}). If $\Gamma$ is the trivial homomorphism, then $\rho_\phi^{-1}(\Lambda)$ is Abelian since $\ker \rho_\phi$ is Abelian by Corollary \ref{solve7} and $h_\Lambda(\Lambda)$ is Abelian by Theorem \ref{rep2} (see p.21 in \cite{AB}). But $\rho_\phi^{-1}(\Lambda)$ is non-Abelian by Theorem \ref{nonabelian} whenever $\Lambda$ is a nontrivial subgroup of $\rho_\phi(S_\phi)$. \end{proof} \begin{example}\rm For the quasiperiodic flow $\phi$ on $T^3$ with frequencies $1$, $7^{1/3}$, and $7^{2/3}$, it was shown in Example \ref{example2} that $\alpha=-2+7^{1/3}\in\rho_\phi(S_\phi)$. The set \[ \Lambda=\{(-1)^j\alpha^k:j\in\{0,1\}, k\in\mathbb{Z}\} \] is a nontrivial subgroup of $\rho_\phi(S_\phi)$ that is isomorphic to $\mathbb{Z}_2\times\mathbb{Z}$. By Theorem \ref{rep5} and Corollary \ref{solve7}, \[ \rho_\phi^{-1}(\Lambda)\cong T^3\rtimes_\Gamma \big(\mathbb{Z}_2\times\mathbb{Z}\big), \] where $\Gamma$ is the (nontrivial) conjugating homomorphism. In particular, every element of $\rho_\phi^{-1}(\Lambda)$ can be written uniquely as $R_c N^jQ^k$ where $R_c\in\hbox{\rm ker}\rho_\phi$ is a translation by $c$ on $T^n$ (as defined in Example \ref{example1}), $N$ is the reversing involution (as defined Example \ref{example1}), and $Q$ is the generalized symmetry of $\phi$ whose multiplier is $\alpha$ (as defined in Example \ref{example2}). Thus \[ \rho_\phi^{-1}(\Lambda) = \{R_cN^jQ^k:c\in T^n, j\in\{0,1\},k\in\mathbb{Z} \}. \] \end{example} \begin{thebibliography}{00} \bibitem{AB} J.L. Alperin and R.B. Bell, {\it Groups and Representations}, Graduate Texts in Mathematics Vol. 162, Springer-Verlag, New York, 1995. \bibitem{AR} V.I. Arnold, {\it Mathematical Methods of Classical Mechanics}, Second Edition, Graduate Texts in Mathematics Vol. 60, Springer-Verlag, New York, 1989. \bibitem{BA1} L. F. Bakker, {\it A Reducible Representation of the Generalized Symmetry Group of a Quasiperiodic Flow}, in ``Dynamical Systems and Differential Equations,'' (eds. W. Feng, S. Hu, and X. Lu), AIMS Press (2003). \bibitem{BA3} L. F. Bakker, {\it Quasiperiodic Flows and Algebraic Number Fields}, submitted to the Proceedings of the Fourth International Conference on Dynamic Systems and Applications, June 2003; listed on arxiv.org as preprint math.DS/0307389. \bibitem{BC} L. F. Bakker and G. Conner, {\it A Class of Generalized Symmetries of Smooth Flows}, to appear in Communications on Pure and Applied Analysis. \bibitem{BR} G. E. Bredon, {\it Topology and Geometry}, Graduate Texts in Mathematics Vol. 139, Springer-Verlag, New York, 1993. \bibitem{HB} H.W. Broer, {\it A family of quasiperiodic attractors}, in {\it Structures in Dynamics, Finite Dimensional Deterministic Studies}, Studies in Mathematical Physics Vol. 2, North-Holland, Elsevier Science Publishers B.V., New York, 1991. \bibitem{HU} T. W. Hungerford, {\it Algebra}, Graduate Texts in Mathematics Vol. 73, Springer-Verlag, New York, 1974. \bibitem{JO} F. John, {\it Partial Differential Equations}, Fourth Edition, Applied Mathematical Sciences Vol. 1, Springer-Verlag, New York, 1982. \bibitem{LR} J. S. W. Lamb and J. A. G. Roberts, {\it Time-reversal symmetry in dynamical systems: A survey}, Physica D 112 (1998), 1-39. \bibitem{RM} R. A. Mollin, {\it Algebraic Number Theory}, Chapman and Hall/CRC, Boca Raton, 1999. \bibitem{SC} W. R. Scott, {\it Group Theory}, Dover Publications, Inc., New York, 1987. \end{thebibliography}\end{document}