1: %Fix a mistake Hitrik found in wave-eq section.
2:
3: %Currently this is the version sent to Zelditch and Hitrik for comments
4: %4 - rearrange a few lemmas, etc.
5: %10/28/05 Fix some edits from Zworski
6: %Fix as many small edits as possible.
7: %Fix e.v. issues and normal form props
8: %Add section on damped wave equation
9: %Fix proof of ``guillemin'' type lemma
10: %Note: on 6/22/05: ``good enough for government work''. First version actually distributed to
11: %anyone.
12: %first distribution revision: remove some fluff and polish for preprint distribution
13: %without boundary business. Based on first part of Version 7. Will continue
14: %to build with boundary results when I get there...
15: %NOTE: Don't forget to edit this as we go!!!!!!
16: %seventh revision: adding sections for characteristic reflecting off
17: %manifold boundary.
18: %sixth revision: reorganize to put all geometry together.
19: %Fix proof of technical lemma.
20: %fifth revision, including second main theorem and proof
21: %fourth revision, better version of st/unst manifold thm, much more background
22: %third version, fixing errors and including more details/lemmas, etc.
23: %second version, removed elliptic components
24:
25:
26: %01/02/06
27:
28: %Normal formatting
29: \documentclass{amsart}
30:
31: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
32: %This is thesis formatting header stuff between lines
33: %\documentclass[12pt]{amsart}
34: %\renewcommand{\baselinestretch}{2}
35: %\setlength{\oddsidemargin}{.4in}
36: %\setlength{\evensidemargin}{0in}
37: %\setlength{\topmargin}{0in}
38: %\setlength{\textwidth}{6in}
39: %\setlength{\textheight}{9in}
40: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
41:
42:
43:
44:
45: \usepackage{epsfig}
46: \usepackage{amscd}
47: \usepackage{amsmath, amssymb, amsthm}
48: \usepackage{latexsym}
49: \usepackage{graphics}
50: \usepackage{color}
51: %\usepackage{mathdots}
52: %\pagestyle{headings}
53:
54: \newtheorem*{main-theorem}{Main Theorem}
55: \newtheorem{proposition}{Proposition}[section]
56: \newtheorem{theorem}{Theorem}
57: \newtheorem*{old-thm}{Theorem}
58: \newtheorem{lemma}[proposition]{Lemma}
59: \newtheorem{corollary}[theorem]{Corollary}
60: \newtheorem{conjecture}[theorem]{Conjecture}
61: \newtheorem{question}[proposition]{Question}
62: \theoremstyle{definition}
63: \newtheorem{definition}[proposition]{Definition}
64: \newtheorem*{remark}{Remark}
65: \newtheorem{claim}[proposition]{Claim}
66: %\newtheorem{exercise}{Exercise}
67: \newtheorem{example}[proposition]{Example}
68: \numberwithin{equation}{section}
69:
70:
71:
72:
73: \def\ZZ{{\mathbb Z}}
74: \def\reals{{\mathbb R}}
75: \def\cx{{\mathbb C}}
76: \def\LL{{\mathcal L}}
77: \def\JJ{{\mathcal J}}
78: \def\Ci{{\mathcal C}^\infty}
79: \def\Re{\,\mathrm{Re}\,}
80: \def\Im{\,\mathrm{Im}\,}
81: \def\tom{\widetilde{\omega}}
82: \def\diam{\mathrm{diam}\,}
83: \def\sgn{\mathrm{sgn}\,}
84: \def\tr{\mathrm{tr}\,}
85: \def\WF{\mathrm{WF}_h\,}
86: \def\div{\mathrm{div}\,}
87: \def\supp{\mathrm{supp}\,}
88: \def\Arg{\mathrm{Arg}\,}
89: \def\id{\,\mathrm{id}\,}
90: \def\Ph{\widehat{\varphi}}
91: \def\Phd{\widehat{\varphi}_{\gamma,d}}
92: \def\P{\varphi}
93: \def\Pd{\varphi_{\gamma,d}}
94: \def\O{{\mathcal O}}
95: \def\SS{{\mathbb S}}
96: \def\s{{\mathcal S}}
97: \def\OPS{{\mathrm OP} {\mathcal S}}
98: \def\cchar{\mathrm{char}\,}
99: \def\nbhd{\mathrm{neigh}\,}
100: \def\neigh{\mathrm{neigh}\,}
101: \def\Op{\mathrm{Op}\,}
102: \def\Opb{\mathrm{Op}_{h,b}^w}
103: \def\Psib{\Psi_{h,b}}
104: \def\esssupp{\text{ess-supp}\,}
105: \def\ad{\mathrm{ad}\,}
106: \def\neg{\mathcal{N}}
107: \def\hsc{{\left( \tilde{h}/h \right)}}
108: \def\hscnp{\frac{\tilde{h}}{h}}
109: \def\csh{{\left( h/\tilde{h} \right)}}
110: \def\phi{\varphi}
111: \def\half{{\frac{1}{2}}}
112: \def\dist{\text{dist}\,}
113: \def\diag{\text{diag}\,}
114: \def\l{{L^2(X)}}
115: \def\be{\begin{eqnarray*}}
116: \def\ben{\begin{eqnarray}}
117: \def\ee{\end{eqnarray*}}
118: \def\een{\end{eqnarray}}
119: \def\lll{\left\langle}
120: \def\rrr{\right\rangle}
121: \def\contraction{\Bigg\lrcorner}
122:
123:
124: %\newcommand{\subfrac}[2]{\genfrac{}{}{0}{}{#1}{#2}}
125: \begin{document}
126: \title[Non-concentration]{Semiclassical Non-concentration near
127: Hyperbolic Orbits}
128: \author{Hans Christianson}
129: \address{Department of Mathematics, University of California, Berkeley, CA 94720 USA}
130: \email{hans@math.berkeley.edu}
131: \keywords{loxodromic orbit, complex hyperbolic orbit, Hamiltonian flow, semiclassical estimates, non-concentration}
132:
133: \begin{abstract}
134: For a large class of semiclassical pseudodifferential operators, including Schr\"odinger operators,
135: $ P ( h ) = -h^2 \Delta_g + V (x) $, on compact Riemannian manifolds, we give logarithmic lower bounds on
136: the mass of eigenfunctions outside neighbourhoods of generic closed hyperbolic orbits. More precisely we show that
137: if $ A $ is a pseudodifferential operator which is microlocally equal to the identity near the
138: hyperbolic orbit and microlocally zero away from the orbit, then
139: \[ \| u \| \leq C (\sqrt{\log(1/h)}/ h ) \| P (h)u \| + C \sqrt {\log(1/h )} \| ( I - A ) u \| \,. \]
140: This generalizes earlier estimates of Colin de Verdi\`ere-Parisse \cite{CVP} obtained for a special case, and
141: of Burq-Zworski \cite{BZ} for real hyperbolic orbits.
142: \end{abstract}
143: \maketitle
144: %\setcounter{section}{-1}
145:
146: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
147:
148: \section{Introduction}
149: \label{intro}
150: \indent
151: To motivate the general result, we first present two applications. If
152: $(X,g)$ is a Riemannian manifold with Laplacian $\Delta_g$, we consider the eigenvalue problem
153: \be
154: -\Delta_g u = \lambda^2 u, \ \ \
155: \|u\|_{L^2(X)} = 1.
156: \ee
157: If $U$ is a small neighbourhood of a closed {\em hyperbolic} geodesic $\gamma$,
158: we show that
159: \be
160: \int_{X \setminus U} |u|^2 dx \geq \frac{c}{ \log | \lambda
161: |},
162: \ee
163: that is, if $u$
164: concentrates near $\gamma$, the rate is logarithmic. This generalizes
165: results of Colin de Verdi\`ere-Parisse \cite{CVP} and Burq-Zworski
166: \cite{BZ}.
167:
168: As another application of our main results we consider the damped wave equation
169: \begin{eqnarray*}
170: \left\{ \begin{array}{l}
171: \left( \partial_t^2 - \Delta + 2a(x) \partial_t \right) u(x,t) = 0, \quad (x,t) \in X \times (0, \infty) \\
172: u(x,0) = 0, \quad \partial_t u(x,0) = f(x).
173: \end{array} \right.
174: \end{eqnarray*}
175: We prove in \S \ref{wave-section} that if $a(x)>0$ outside a
176: neighbourhood of a closed hyperbolic geodesic $\gamma$, we have the following energy estimate:
177: \be
178: \| \partial_t u \|_{L^2(X)}^2 + \left\| \nabla u
179: \right\|_{L^2(X)}^2 \leq C e^{-t/C} \|f\|_{H^\epsilon(X)}^2,
180: \ee
181: for all $\epsilon>0$. (In \S \ref{wave-section} a weaker
182: geometric control condition in the spirit of Rauch-Taylor \cite{RT} is
183: considered.) This application was suggested to us by M. Hitrik,
184: and it generalizes an example of Lebeau \cite{Leb}.
185:
186: We now turn to the general case. Let $X$ be a compact $n$-dimensional manifold without boundary. We consider a selfadjoint
187: pseudodifferential operator, $P(h)$, with real principal symbol $p$.
188: %and belonging to an appropriate class $P(h) \in \Psi_h^{k,m}(X, \Omega_X^\half)$. Ostensibly $P(h)$ is the Schr\"{o}dinger operator
189: %\begin{eqnarray*}
190: %P(h) = -h^2 \Delta_g + V(x)
191: %\end{eqnarray*}
192: %with $\Delta_g$ the Laplace-Beltrami operator on a compact Riemannian manifold, or on $ \reals^n $, with
193: %a confining $ V ( x ) $.
194: We assume throughout if $p=0$ then $dp \neq 0$, and that $p$ is elliptic outside
195: of a compact subset of $T^*X$.
196: %\be
197: %p \geq \lll \xi \rrr^m / C \text{ for } |\xi| \geq C.
198: %\ee
199: %For the example of the Schr\"{o}dinger operator, the principal symbol is
200: %\begin{eqnarray*}
201: %p(x, \xi) = |\xi|^2 + V(x).
202: %\end{eqnarray*}
203: Assume that
204: $$ \gamma \subset p^{-1} ( 0 ) $$
205: is a closed loxodromic orbit of the Hamiltonian flow of $p$.
206: Let $N \subset \{p=0\}$ be a Poincar\'{e} section for $\gamma$ and let $S$ be the Poincar\'e map.
207: The assumption that $\gamma$ be loxodromic means that no eigenvalue of $dS(0,0)$ lies on the unit circle.
208: We assume also that $dS(0,0)$ has no real negative eigenvalues.
209:
210: \begin{main-theorem}
211: Let $A \in \Psi_h^{0,0}$ be a pseudodifferential operator whose principal symbol is
212: $1$ near $\gamma$ and $0$ away from $\gamma$. Then, there exist constants $ h_0>0
213: $ and $0 < C < \infty$ so that we have uniformly in $ 0 < h < h_0 $,
214: \begin{eqnarray}
215: \label{main-theorem-3-est}
216: \| u \| \leq C \frac{ \sqrt{\log ( 1/h )}} h \| P ( h ) u \| + C
217: \sqrt{ \log (1/h ) } \| ( I - A ) u \| \,,
218: \end{eqnarray}
219: where the norms are $ L^2 $ norms on $ X$.
220: In particular if a family, $u = u ( h ) $ satisfies
221: \begin{eqnarray*}
222: P(h) u = \O_{L^2}(h^\infty), \ \ \
223: \| u \|_{L^2(X)} = 1\, ,
224: \end{eqnarray*}
225: then
226: \begin{eqnarray}
227: \label{main-theorem-3-est-2}
228: \left\| (I - A) u \right\|_{L^2(X)} \geq \frac{1}{C} \log \left(\left(1/h\right)\right)^{-\half} \,, \ \ 0 < h < h_0 \,.
229: \end{eqnarray}
230: \end{main-theorem}
231: We note that the assumptions on $ A $ imply that
232: $\WF (A) $ is contained in a neighbourhood of $\gamma$, while $\WF(I - A)$ is away from $\gamma$, see
233: \S \ref{preliminaries} for definitions.
234:
235:
236: Colin de Verdi\`ere and Parisse \cite{CVP} have shown that the
237: estimates (\ref{main-theorem-3-est}-\ref{main-theorem-3-est-2}) are sharp in the case where
238: $ X $ is a segment of a hyperbolic cylinder and $ P ( h ) = -h^2 \Delta_g $ is its Dirichlet Laplacian.
239: Even though the closed orbit at the ``neck'' of the cylinder is hyperbolic, the flow is completely
240: integrable in that case. This shows that eliminating the $\log (h^{-1})$ factor requires global
241: conditions on the classical flow.
242:
243: The assumption that the Poincar{\'e} map has no negative eigenvalues
244: is standard in the literature on quantum Birkhoff normal forms (see, for example, \cite{IaSj},
245: \cite{ISZ}, and \cite{Ze}), and in the present work serves to
246: eliminate cases in which current techniques seem to break down. It is
247: important to note that this case does arise,
248: as in the example in \cite{Kl} \S $3.4$.
249:
250: There are many examples in which the hypotheses of the theorem are
251: satisfied, the simplest of which is the case in which $p = |\xi|^2- E(h)$ for $E(h)>0$.
252: Then the Hamiltonian flow of $p$ is the geodesic flow, so if the
253: geodesic flow has a closed hyperbolic orbit, there is
254: non-concentration of eigenfunctions, $u(h)$, for the equation
255: \be
256: -h^2 \Delta u(h) = E(h) u(h).
257: \ee
258: Another example of such a $p$ is the case $p = |\xi|^2 + V(x)$, where
259: $V(x)$ is a confining potential with three ``bumps'' or ``obstacles''
260: in the lowest energy level (see Figure \ref{fig:fig13}). In the
261: appendix to \cite{Sjo2a} it is shown that for an interval of energies
262: $V(x) \sim 0$, there is a closed hyperblic orbit $\gamma$ of the Hamiltonian
263: flow which ``reflects'' off the bumps (see Figure \ref{fig:fig14}).
264: Loxodromic orbits may be constructed by considering $3$-dimensional
265: hyperbolic billiard problems (see, for example, \cite{AuMa}), although
266: in the present work we are assuming the orbit does not intersect
267: the boundary of the manifold. In addition, Proposition
268: \ref{normal-prop-1} gives a somewhat artificial means of constructing a manifold diffeomorphic to a neighbourhood in $T^*\SS_{(t, \tau)}^1
269: \times T^* \reals_{(x, \xi)}^{n-1}$ which contains a loxodromic orbit $\gamma$
270: by starting with the Poincar\'{e} map $\gamma$ is to have.
271:
272: %\begin{figure}
273: %\caption{\label{fig:fig13} A confining potential $V(x)$ with three bumps
274: % at the lowest energy $E <0$, and the energy
275: % level $V(x) = 0$.}
276: %\include{fig13}
277: %\end{figure}
278:
279: \begin{figure}
280: \centerline{\epsfig{figure=three-bump, height=2.5in}}
281: \caption{\label{fig:fig13} A confining potential $V(x)$ with three
282: bumps at the lowest energy level $E <0$.}
283: \end{figure}
284:
285:
286: \begin{figure}
287: \include{fig14}
288: \caption{\label{fig:fig14} The level set $V(x) = 0$ and the closed
289: hyperbolic orbit $\gamma$.}
290: \end{figure}
291:
292: \indent In order to prove the Main Theorem, we will first prove that
293: the principal symbol of $P(h)$ can be put into a normal form near $\gamma$. This
294: will allow analysis of small complex perturbations of $P(h)$. These are defined as follows:
295: let $a \in \Ci(T^*X,[0,1])$ be equal to $0$ in a neighbourhood of $\gamma$ and $1$ outside of a larger
296: neighbourhood of $ \gamma $. For $z \in [-1,1] + i[-\delta, \delta]$, define
297: \begin{eqnarray}
298: \label{Q(z)}
299: Q(z):= P(h) - z - ih Ca^w,
300: \end{eqnarray}
301: for a constant $C$ to be fixed later. The following theorem states that by perturbing $P(h)$ into $Q(z)$
302: we are able to push the spectrum of $P(h)$ into the lower half plane.
303: \begin{theorem}
304: \label{main-theorem-1}
305: There exist constants $c_0>0$, $h_0>0$, and $N_0$ such that for $u$ with $\WF(u)$ in a sufficiently small
306: neighbourhood of $\gamma$, $z \in [-1,1] + i(-c_0h, + \infty)$, and $0 < h < h_0$ we have
307: \begin{eqnarray}
308: \label{main-theorem-1-est}
309: \left\| Q(z) u \right\|_\l \geq C^{-1} h^{N_0} \left\| u \right\|_\l
310: \end{eqnarray}
311: for some constant $C$.
312: \end{theorem}
313: Using Theorem \ref{main-theorem-1} and a semiclassical adaptation of the ``three-lines'' theorem from complex analysis,
314: we will be able to deduce the following estimate.
315: \begin{theorem}
316: \label{main-theorem-2}
317: Suppose $Q(z)$ is given by \eqref{Q(z)}, and $z \in I \Subset (-\infty, \infty)$. Then there
318: is $h_0 >0$ and $0 < C < \infty$ such that for $0 < h < h_0 $,
319: \begin{eqnarray}
320: \label{main-theorem-2-est-1}
321: \left\| Q(z)^{-1} \right\|_{L^2(X) \to L^2(X)} \leq C \frac{\log (1/h) }{h}.
322: \end{eqnarray}
323: If $\phi \in \Ci_c(X)$ is supported away from $\gamma$, then
324: \begin{eqnarray}
325: \label{main-theorem-2-est-2}
326: \left\| Q(z)^{-1} \phi \right\|_{L^2(X) \to L^2(X)} \leq C \frac{\sqrt{ \log (1/h)}}{h}.
327: \end{eqnarray}
328: \end{theorem}
329: In order to apply the results of Theorems \ref{main-theorem-1} and \ref{main-theorem-2} to the Main Theorem,
330: we observe that for $A$ as in the statement of the Main Theorem we
331: have $Q(0)A = P(h)A$ microlocally and apply a commutator argument.
332:
333: This note is organized as follows. \S \ref{preliminaries} recalls basic facts about
334: $h$-pseudodifferential operators on manifolds. This is followed in \S \ref{FIO} with a review
335: of some standard results from the theory of $h$-Fourier Integral Operators. In \S \ref{symplectic-geometry} we present
336: some symplectic geometry and prove the principal symbol can be put
337: into a normal form in the case all the eigenvalues of $dS(0)$ are distinct.
338: \S \ref{main-theorem-1-proof} contains the proof of Theorem
339: \ref{main-theorem-1} in the case of distinct eigenvalues, then re-examines the normal
340: form of the principal symbol to show how it may be extended to the
341: case when the eigenvalues are not distinct, and contains the details
342: of the more general case of Theorem \ref{main-theorem-1}. Finally, in \S \ref{main-theorem-2-proof} we prove Theorem
343: \ref{main-theorem-2} and the Main Theorem. In \S
344: \ref{wave-section} we follow a suggestion of M. Hitrik to apply the
345: techniques of \S \ref{symplectic-geometry}-\ref{main-theorem-2-proof}
346: to the damped wave equation.
347:
348: The impetus for this paper came when M. Zworski suggested generalizing results from the appendix of \cite{BZ},
349: as well as correcting a mistake which was discovered by J.-F. Bony, S. Fujiie, T. Ramond,
350: and M. Zerzeri (see \cite{BFRZ} for their closely related work). This paper generalizes the statements of the theorems
351: from the case of real hyperbolic trajectories to complex hyperbolic or loxodromic trajectories as well as
352: correcting the mistake.
353:
354: {\bf Acknowledgements:} The author would like to thank Maciej Zworski for much help
355: and support during the writing of this paper, Alan Weinstein and Fr\'ed\'eric Naud for helpful
356: conversations, as well as the NSF for partial support. He would
357: like to thank Michael Hitrik for comments on an early draft and suggesting Section
358: \ref{wave-section}, and Laurent Thomann and Steve Zelditch for careful reading of an
359: early draft of this paper.
360:
361:
362: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
363:
364: \section{Preliminaries}
365: \label{preliminaries}
366: This section contains some basic definitions and results from semiclassical and microlocal
367: analysis which we will be using throughout the paper. This is essentially standard, but we
368: include it for completeness. We will follow the presentation in \cite{BZ}, \S $2$.
369: Let $X$ be a smooth, compact manifold. We will be operating on half-densities,
370: \begin{eqnarray*}
371: u(x)|dx|^{\frac{1}{2}} \in \Ci\left(X, \Omega_X^{\frac{1}{2}}\right),
372: \end{eqnarray*}
373: with the informal change of variables formula
374: \begin{eqnarray*}
375: u(x)|dx|^{\frac{1}{2}} = v(y)|dy|^\half, \,\, \text{for}\,\, y = \kappa(x)
376: \Leftrightarrow v(\kappa(x))|\kappa'(x)|^\half = u(x).
377: \end{eqnarray*}
378: By symbols on $X$ we mean
379: \begin{eqnarray*}
380: \lefteqn{\s^{k,m} \left(T^*X, \Omega_{T^*X}^\half \right):= } \\
381: & = &\left\{ a \in \Ci(T^*X \times (0,1], \Omega_{T^*X}^\half): \left| \partial_x^\alpha
382: \partial_\xi^\beta a(x, \xi; h) \right| \leq C_{\alpha \beta}h^{-m} \langle \xi \rangle^{k - |\beta|} \right\}.
383: \end{eqnarray*}
384: There is a corresponding class of pseudodifferential operators $\Psi_h^{k,m}(X, \Omega_X^\half)$
385: acting on half-densities defined by the local formula (Weyl calculus) in $\reals^n$:
386: \begin{eqnarray*}
387: \Op_h^w(a)u(x) = \frac{1}{(2 \pi h)^n} \int \int a \left( \frac{x + y}{2}, \xi; h \right)
388: e^{i \langle x-y, \xi \rangle / h }u(y) dy d\xi.
389: \end{eqnarray*}
390: We will occasionally use the shorthand notations $a^w := \Op_h^w(a)$ and $A:=\Op_h^w(a)$ when
391: there is no ambiguity in doing so. \\
392: \indent We have the principal symbol map
393: \begin{eqnarray*}
394: \sigma_h : \Psi_h^{k,m} \left( X, \Omega_X^\half \right) \to \s^{k,m} \left/ \s^{k, m-1}
395: \left(T^*X, \Omega_{T^*X}^\half \right) \right.,
396: \end{eqnarray*}
397: which gives the left inverse of $\Op_h^w$ in the sense that
398: \begin{eqnarray*}
399: \sigma_h \circ \Op_h^w: \s^{k,m} \to \s^{k,m}/\s^{k, m-1}
400: \end{eqnarray*}
401: is the natural projection. Acting on half-densities in the Weyl calculus, the principal
402: symbol is actually well-defined in $\s^{k,m} / \s^{k, m-2}$, that is, up to $\O(h^2)$ in
403: $h$ (see, for example \cite{EvZw} Appendix D).
404:
405: We will use the notion of wave front sets for pseudodifferential operators on manifolds.
406: If $a \in \s^{k,m}(T^*X, \Omega_{T^*X}^\half)$, we define the singular support or essential support for $a$:
407: \begin{eqnarray*}
408: \esssupp_h a \subset T^*X \bigsqcup \SS^*X,
409: \end{eqnarray*}
410: where $\SS^*X = (T^*X \setminus \{0\}) / \reals_+$ is the cosphere bundle (quotient taken with
411: respect to the usual multiplication in the fibers), and the union is disjoint. $\esssupp_h a$ is defined using complements:
412: \begin{eqnarray*}
413: \lefteqn{\esssupp_h a := } \\
414: & = & \complement \left\{ (x, \xi) \in T^*X : \exists \epsilon >0, \,\,\, \partial_x^\alpha
415: \partial_\xi^\beta a(x', \xi') = \O(h^\infty), \,\,\, d(x, x') + |\xi - \xi'| < \epsilon \right\} \\
416: && \bigcup \complement \{ (x, \xi) \in T^*X \setminus 0 : \exists \epsilon > 0, \,\,\, \partial_x^\alpha
417: \partial_\xi^\beta a(x', \xi') = \O (h^\infty \langle \xi \rangle^{-\infty}), \\
418: && \quad \quad \quad d(x, x') + 1 / |\xi'| + | \xi/ |\xi| - \xi' /
419: |\xi'| | < \epsilon \} / \reals_+.
420: \end{eqnarray*}
421: We then define the wave front set of a pseudodifferential operator $A \in \Psi_h^{k,m}( X, \Omega_X^\half )$:
422: \begin{eqnarray*}
423: \WF(A) : = \esssupp_h(a), \,\,\, \text{for} \,\,\, A = \Op_h^w(a).
424: \end{eqnarray*}
425: Finally for distributional half-densities $u \in \Ci( (0, 1]_h, \mathcal{D}'(X, \Omega_X^\half))$
426: such that there is $N_0$ so that $h^{N_0}u$ is bounded in $\mathcal{D}'(X, \Omega_X^\half)$, we can
427: define the semiclassical wave front set of $u$, again by complement:
428: \begin{eqnarray*}
429: \lefteqn{\WF (u) := } \\
430: &= & \complement \{(x, \xi) : \exists A \in \Psi_h^{0,0}, \,\, \text{with} \,\, \sigma_h(A)(x,\xi) \neq 0, \,\, \\
431: && \quad \text{and} \,\,Au \in h^\infty \Ci((0,1]_h, \Ci(X, \Omega_X^\half)) \}.
432: \end{eqnarray*}
433: For $A = \Op_h^w(a)$ and $B = \Op_h^w(b)$, $a \in \s^{k,m}$, $b \in \s^{k',m'}$ we have the composition
434: formula (see, for example, \cite{DiSj})
435: \begin{eqnarray}
436: \label{Weyl-comp}
437: A \circ B = \Op_h^w \left( a \# b \right),
438: \end{eqnarray}
439: where
440: \begin{eqnarray}
441: \label{a-pound-b}
442: \s^{k + k', m+m'} \ni a \# b (x, \xi) := \left. e^{\frac{ih}{2} \omega(Dx, D_\xi; D_y, D_\eta)}
443: \left( a(x, \xi) b(y, \eta) \right) \right|_{{x = y} \atop {\xi = \eta}} ,
444: \end{eqnarray}
445: with $\omega$ the standard symplectic form. \\
446: \indent We will need the definition of microlocal equivalence of operators. Suppose
447: $T: \Ci(X, \Omega_X^\half) \to \Ci(X, \Omega_X^\half)$ and that for any seminorm $\| \cdot \|_1$ on
448: $\Ci(X, \Omega_X^\half)$ there is a second seminorm $\| \cdot \|_2$ on $\Ci(X, \Omega_X^\half)$ such that
449: \begin{eqnarray*}
450: \| Tu\|_1 = \O(h^{-M_0})\|u \|_2
451: \end{eqnarray*}
452: for some $M_0$ fixed. Then we say $T$ is {\it semiclassically tempered}. We assume for the rest of
453: this paper that all operators satisfy this condition. Let $U,V \subset T^*X$ be open precompact sets.
454: We think of operators defined microlocally near $V \times U$ as equivalence classes of tempered operators.
455: The equivalence relation is
456: \begin{eqnarray*}
457: T \sim T' \Longleftrightarrow A(T-T')B = \O(h^\infty): \mathcal{D}'\left( X, \Omega_X^\half \right) \to \Ci
458: \left(X, \Omega_X^\half \right)
459: \end{eqnarray*}
460: for any $A,B \in \Psi_h^{0,0}(X, \Omega_X^\half)$ such that
461: \begin{eqnarray*}
462: && \WF (A) \subset \widetilde{V}, \quad \WF (B) \subset \widetilde{U}, \,\, \text{with} \,\, \widetilde{V},
463: \widetilde{U} \,\, \text{open and } \\
464: && \quad \quad \overline{V} \Subset \widetilde{V} \Subset T^*X, \quad \overline{U} \Subset
465: \widetilde{U} \Subset T^*X.
466: \end{eqnarray*}
467: In the course of this paper, when we say $P=Q$ {\it microlocally} near $U \times V$, we mean for any $A$, $B$ as above,
468: \begin{eqnarray*}
469: APB - AQB = \O_{L^2 \to L^2}\left( h^\infty \right),
470: \end{eqnarray*}
471: or in any other norm by the assumed precompactness of $U$ and $V$. Similarly, we say $B = T^{-1}$ on $V \times V$ if
472: $BT = I$ microlocally near $U \times U$ and $TB = I$ microlocally near $V \times U$. \\
473: \indent For this paper, we will need the following semiclassical version of Beals's Theorem (see \cite{DiSj} for a proof).
474: Recall for operators $A$ and $B$, the notation $\ad_BA$ is defined as
475: \begin{eqnarray*}
476: \ad_B A = \left[B, A \right].
477: \end{eqnarray*}
478: \begin{old-thm}[Beals's Theorem]
479: Let $A: \s \to \s'$ be a continuous linear operator. Then $A = \Op_h^w(a)$ for a symbol $a \in \s^{0,0}$ if and only
480: if for all $N \in \mathbb{N}$ and all linear symbols $l_1, \ldots l_N$,
481: \begin{eqnarray*}
482: \ad_{\Op_h^w(l_1)} \circ \ad_{\Op_h^w(l_2)} \circ \cdots \circ \ad_{\Op_h^w(l_N)} A = \O(h^N)_{L^2 \to L^2}.
483: \end{eqnarray*}
484: \end{old-thm}
485: The following lemma (given more generally in \cite{BoCh}) will be used in the proof of Theorem \ref{main-theorem-1}.
486: We include a sketch of the proof from \cite{SjZw2} here for completeness. It is easiest to phrase in terms of
487: order functions.
488: A smooth function $m \in \Ci(T^*X; \reals)$ is called an order function if it satisfies
489: \begin{eqnarray*}
490: m(x, \xi) \leq C m(y, \eta) \left\langle \dist (x-y) + |\xi - \eta| \right\rangle^N
491: \end{eqnarray*}
492: for some $N \in \mathbb{N}$. We say $a \in \s^l(m)$ if
493: \begin{eqnarray*}
494: \left| \partial^\alpha a \right| \leq C_{\alpha }h^{-l} m.
495: \end{eqnarray*}
496: If $l = 0$, we write $\s(m):= \s^0(m)$.
497: \begin{lemma}
498: \label{etG-lemma}
499: Let $m$ be an order function, and suppose $G \in \Ci (T^*X ; \reals)$ satisfies
500: \begin{eqnarray}
501: \label{G-cond-1}
502: G(x, \xi) - \log \left( m(x, \xi) \right) = \O (1),
503: \end{eqnarray}
504: and
505: \begin{eqnarray}
506: \label{G-cond-2}
507: \partial_x^\alpha \partial_\xi^\beta G(x,\xi) = \O(1) \,\,\, \text{for} \,\,\, (\alpha, \beta) \neq (0,0).
508: \end{eqnarray}
509: Then for $G^w = \Op_h^w(G)$ and $|t|$ sufficiently small,
510: \begin{eqnarray*}
511: \exp (tG^w) = \Op_h^w(b_t)
512: \end{eqnarray*}
513: for $b_t \in \s(m^t)$. Here $e^{tG^w}$ is defined as the unique solution to the ordinary
514: differential equation
515: \be
516: \left\{ \begin{array}{l} \partial_t \left( U(t) \right) - G^w
517: U(t) = 0 \\
518: U(0) = \id. \end{array} \right.
519: \ee
520: \end{lemma}
521: \begin{proof}[Sketch of Proof]
522: The conditions on $G$ \eqref{G-cond-1} and \eqref{G-cond-2} are equivalent to saying $e^{tG} \in \s(m^t)$. We will
523: compare $\exp tG^w$ and $\Op_h^w (\exp t G)$.
524: \begin{claim}
525: \label{U-inv-claim}
526: Set $U(t):= \Op_h^w (e^{tG}): \s \to \s$. For $|t| < \epsilon_0$, $U(t)$ is invertible and $U(t)^{-1} = \Op_h^w (b_t)$ for
527: $b_t \in \s(m^{-t})$, where $\epsilon_0$ depends only on $G$.
528: \end{claim}
529: \begin{proof}[Proof of Claim]
530: Using the composition law, we see $U(-t) U(t) = \id + \Op_h^w(E_t)$, with $E_t = \O(t)$. Hence $\id + \Op_h^w(E_t)$ is
531: invertible and using Beals's Theorem, we get $(\id + \Op_h^w(E_t))^{-1} = \Op_h^w(c_t)$ for $c_t \in \s(1)$.
532: Thus $\Op_h^w(c_t)U(-t)U(t) = \id$, so
533: \begin{eqnarray*}
534: U(t)^{-1} = \Op_h^w \left( c_t \# \exp (-tG) \right),
535: \end{eqnarray*}
536: and subsequently $b_t \in \s(m^{-t})$.
537: \end{proof}
538: Now observe that
539: \begin{eqnarray*}
540: \frac{d}{dt} U(-t) = - \Op_h^w \left( G \exp(-tG) \right), \,\,\, \text{and} \,\,\, U(-t) G^w = \Op_h^w \left( e^{-tG} \# G \right),
541: \end{eqnarray*}
542: so that
543: \begin{eqnarray}
544: \lefteqn{\frac{d}{dt} \left( U(-t) e^{tG^w} \right) =} \label{Op(A_t)}
545: \\ & = & -\Op_h^w \left( G \exp(-tG) \right) e^{tG^w} +
546: \Op_h^w \left( e^{-tG} \# G \right) e^{tG^w} \nonumber \\
547: & = & \Op_h^w (A_t) e^{tG^w}, \nonumber
548: \end{eqnarray}
549: for $A_t \in \s(m^{-t})$. To see \eqref{Op(A_t)}, recall that by the composition law,
550: \begin{eqnarray*}
551: e^{-tG} \# G = e^{-tG} G + \left( \text{terms with }\,G \,\,\, \text{derivatives} \right).
552: \end{eqnarray*}
553: Then the first terms in \eqref{Op(A_t)} will cancel and the remaining terms will all involve at least one
554: derivative of $G$, which is then bounded by \eqref{G-cond-2}. \\
555: \indent Set $C(t):= -\Op_h^w(A_t)U(-t)^{-1}$. Claim \ref{U-inv-claim} implies $C(t) = \Op_h^w (c_t)$ for a family
556: $c_t \in \s(1)$. The composition law implies $c_t$ depends smoothly on $t$. Then
557: \begin{eqnarray*}
558: \left( \frac{\partial}{\partial t} + C(t) \right) \left( U(-t) e^{tG^w} \right) = \Op_h^w (A_t) e^{tG^w} -
559: \Op_h^w (A_t) e^{tG^w} = 0,
560: \end{eqnarray*}
561: so we have reduced the problem to proving the following claim.
562: \begin{claim}
563: Suppose $C(t) = \Op_h^w (c_t)$ with $c_t \in \s(1)$ depending smoothly on $t \in (-\epsilon_0, \epsilon_0)$. If $Q(t)$ solves
564: \begin{eqnarray*}
565: \left\{ \begin{array}{c}
566: \left( \frac{\displaystyle \partial}{\displaystyle \partial t} + C(t) \right) Q(t) = 0, \\
567: Q(0) = \Op_h^w(q), \,\,\, \text{with} \,\,\, q \in \s(1),
568: \end{array} \right.
569: \end{eqnarray*}
570: then $Q(t) = \Op_h^w (q_t)$ with $q_t \in \s(1)$ depending smoothly on $t \in (-\epsilon_0, \epsilon_0)$.
571: \end{claim}
572: \begin{proof}[Proof of Claim]
573: The Picard existence theorem for ODEs implies $Q(t)$ exists and is bounded on $L^2$. We want to use Beals's
574: Theorem to show $Q(t)$ is actually a quantized family of symbols. Let $l_1, \ldots, l_N$ be linear symbols.
575: We will use induction to show that for any $N$ and any choice of the $l_j$, $\ad_{\Op_h^w (l_1)} \circ \cdots \circ
576: \ad_{\Op_h^w(l_N)}Q(t) = \O(h^N)_{L^2\to L^2}$. Since we are dealing with linear symbols, we take $h = 1$ for
577: convenience. First note
578: \begin{eqnarray*}
579: &&\frac{d}{dt} \ad_{\Op_h^w (l_1)} \circ \cdots \circ \ad_{\Op_h^w(l_N)} Q(t) + \ad_{\Op_h^w (l_1)} \circ \cdots \circ
580: \ad_{\Op_h^w(l_N)}\\
581: && \quad \quad \quad \quad \quad \quad \quad \quad \quad \cdot \left( C(t) Q(t) \right) = 0
582: \end{eqnarray*}
583: For the induction step, assume $ \ad_{\Op_h^w (l_1)} \circ \cdots \circ \ad_{\Op_h^w(l_k)}Q(t) = \O(1)$ is known
584: for $k<N$ and observe
585: \begin{eqnarray*}
586: \lefteqn{ \ad_{\Op_h^w (l_1)} \circ \cdots \circ \ad_{\Op_h^w(l_N)} \left( C(t) Q(t) \right) = } \\
587: & = & C(t) \ad_{\Op_h^w (l_1)} \circ \cdots \circ \ad_{\Op_h^w(l_N)} Q(t) + R(t),
588: \end{eqnarray*}
589: where $R(t)$ is a sum of terms of the form $A_k(t) \ad_{\Op_h^w (l_1)} \circ \cdots \circ \ad_{\Op_h^w(l_k)}Q(t)$ for
590: each $k<N$ and $A_k(t) = \Op_h^w (a_k(t))$ with $a_k(t) \in \s(1)$. Set $\tilde{Q}(t) = \ad_{\Op_h^w (l_1)} \circ
591: \cdots \circ \ad_{\Op_h^w(l_N)}Q(t)$, and note that $\tilde{Q}$ solves
592: \begin{eqnarray*}
593: \left\{ \begin{array}{c}
594: \left( \frac{\displaystyle \partial}{\displaystyle \partial t} + C(t) \right) \tilde{Q}(t) = -R(t), \\
595: \tilde{Q}(0) = \O (1)_{L^2 \to L^2}.
596: \end{array} \right.
597: \end{eqnarray*}
598: Since $R(t) = \O(1)_{L^2 \to L^2}$ by the induction hypothesis, Picard's theorem implies $\tilde{Q}(t):L^2 \to L^2$
599: as desired.
600: \end{proof}
601: \end{proof}
602:
603: We will need to review some basic facts about the calculus of symbols with two parameters. We will only use symbol spaces with two
604: parameters in the context of microlocal estimates, in which case we
605: may assume we are working in an open subset of $\reals^{2n}$. We define the
606: following spaces of symbols with two parameters:
607: \begin{eqnarray*}
608: \lefteqn{\s^{k,m, \widetilde{m}} \left( \reals^{2n} \right) := } \\
609: & = &\Big\{ a \in \Ci \left( \reals^{2n} \times (0,1]^2 \right): \\
610: && \quad \quad \left| \partial_x^\alpha \partial_\xi^\beta a(x, \xi; h, \tilde{h}) \right|
611: \leq C_{\alpha \beta}h^{-m}\tilde{h}^{-\widetilde{m}} \langle \xi \rangle^{k - |\beta|} \Big\}.
612: \end{eqnarray*}
613: For the applications in this paper, we assume $\tilde{h} >h$ and
614: define the scaled spaces:
615: \begin{eqnarray*}
616: \lefteqn{\s_{\delta}^{k,m, \widetilde{m}} \left(\reals^{2n} \right):= } \\
617: & = & \Bigg\{ a \in \Ci \left(\reals^{2n} \times (0,1]^2 \right): \\
618: && \quad \quad \left| \partial_x^\alpha \partial_\xi^\beta a(x, \xi; h, \tilde{h}) \right|
619: \leq C_{\alpha \beta}h^{-m}\tilde{h}^{-\widetilde{m}} \left( \frac{\tilde{h}}{h} \right)^{\delta(|\alpha| + |\beta|)}
620: \langle \xi \rangle^{k - |\beta|} \Bigg\}.
621: \end{eqnarray*}
622: As before, we have the corresponding spaces of semiclassical pseudodifferential operators $\Psi^{k, m, \widetilde{m}}$
623: and $\Psi_{\delta}^{k,m, \widetilde{m}}$, where we will usually add a subscript of $h$ or $\tilde{h}$ to indicate which
624: parameter is used in the quantization. The relationship between $\Psi_h$ and
625: $\Psi_{\tilde{h}}$ is given in the following lemma.
626: \begin{lemma}
627: \label{U-hsc-lemma}
628: Let $a \in \s_{0}^{k,m,\tilde{m}}$, and set
629: \be
630: b(X, \Xi) = a\left(\csh^{\half}X, \csh^{\half} \Xi
631: \right) \in \s_{-\half}^{k,m,\tilde{m}}.
632: \ee
633: There is a linear operator $T_{h, \tilde{h}}$, unitary on $L^2$, and an operator
634: such that
635: \be
636: \Op_{\tilde{h}}^w(b) T_{h, \tilde{h}} u = T_{h, \tilde{h}} \Op_h^w(a) u.
637: \ee
638: \end{lemma}
639: \begin{proof}
640: For $u \in L^2(\reals^{n})$, define $T_{h, \tilde{h}}$ by
641: \begin{eqnarray}
642: \label{U-hsc}
643: T_{h, \tilde{h}} u (X) := \csh^{\frac{n}{4}}u\left( \csh^{\half} X \right).
644: \end{eqnarray}
645: We see immediately that $T_{h, \tilde{h}}$ conjugates operators $a^w(x, hD_x)$ and $b^w(X, \tilde{h} D_X)$.
646: \end{proof}
647: We have the following microlocal commutator lemma.
648: \begin{lemma}
649: \label{2-param-lemma}
650: Suppose $a \in \s_{0}^{-\infty,0,0}$, $b \in
651: \s_{-\half}^{-\infty,m,\widetilde{m}}$, and $\tilde{h}>h$.
652:
653: (a) If $A = \Op_{\tilde{h}}^w (a)$ and $B = \Op_{\tilde{h}}^w(b)$,
654: \be
655: [A,B] &=& \frac{\tilde{h}}{i}\Op_{\tilde{h}}^w (\{a,b\}) + \O\left(
656: h^{3/2} \tilde{h}^{3/2} \right).
657: \ee
658:
659: (b) More generally, for each $l>1$,
660: \be
661: \ad_A^l B = \O_{L^2 \to L^2} \left( h \tilde{h}^{l-1} \right) .
662: \ee
663: \end{lemma}
664: \begin{proof}
665: Without loss of generality, $m = \widetilde{m} = 0$, so for (a) we have from the Weyl
666: calculus:
667: \be
668: [A,B] = \frac{\tilde{h}}{i} \Op_{\tilde{h}}^w (\{a,b\}) + \tilde{h}^3 \O
669: \left( \sum_{|\alpha| = |\beta| = 3} \partial^\alpha a \partial^\beta b \right),
670: \ee
671: since the second order term vanishes in the Weyl expansion of the
672: commutator. Note $\partial^\alpha a$
673: is bounded for all $\alpha$, and observe for $|\beta|=3$,
674: \be
675: \tilde{h}^3 \partial^\beta b & = & \tilde{h}^3 \O \left(h^{3/2}
676: \tilde{h}^{-3/2} \right).
677: \ee
678: For part (b) we again assume $m = \widetilde{m} = 0$, and we observe that for $l>1$ we no longer have the
679: same gain in powers of $h$ as in part (a). This follows from the fact
680: that the $\tilde{h}$-principal symbol for the commutator $[A,[A,B]]$, $-i
681: \tilde{h} \{a, -i \tilde{h}\{a,b\} \}$, satisfies
682: \ben
683: -i \tilde{h}\{a, -i \tilde{h}\{a,b\}\} & = & -\tilde{h}^2 \Big( \partial_\Xi a
684: \partial_X \left(\partial_\Xi a \partial_X b - \partial_X a
685: \partial_\Xi b \right) \label{comm-calc-1} \\
686: && \quad - \partial_X a \partial_\Xi \left( \partial_\Xi a \partial_X b - \partial_X a
687: \partial_\Xi b \right) \Big) \nonumber \\
688: & \in & \s_{0}^{-\infty,-1 ,-1}, \label{comm-calc-2}
689: \een
690: since $\{a,b\}$ involves products of derivatives of both $a$ and $b$.
691:
692: For general $l>1$, assume
693: \be
694: \sigma_{\tilde{h}} \left( \ad_A^l B \right) \in \s_{0}^{0, -1, 1-l}
695: \ee
696: and a calculation similar to (\ref{comm-calc-1}-\ref{comm-calc-2})
697: finishes the induction.
698: \end{proof}
699:
700:
701:
702:
703: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
704:
705: \section{$h$-Fourier Integral Operators}
706: \label{FIO}
707: In this section we review some facts about $h$-Fourier Integral Operators ($h$-FIOs). See \cite{duistermaat} for a
708: comprehensive introduction to general FIOs without $h$, or
709: \cite{EvZw}, \S $10.1$ with the addition of the $h$ parameter.
710: For this note, we are only interested in a special class of $h$-FIOs, namely those associated to a symplectomorphism.
711: In order to motivate this, suppose $f:X \to Y$ is a diffeomorphism. Then we write
712: \begin{eqnarray*}
713: f^*u(x) = u(f(x))= \frac{1}{(2 \pi h)^n} \int e^{i \langle f(x) -y, \xi \rangle /h} u(y) dy d\xi,
714: \end{eqnarray*}
715: and $f^*: \Ci(Y) \to \Ci(X)$ is an $h$-FIO associated to the nondegenerate phase function $\phi = \langle f(x) -y,
716: \xi \rangle$. We recall the notation from \cite{duistermaat}: if $A:\Ci_c(Y) \to \mathcal{D}'(X)$ is a continuous mapping with
717: distributional kernel $K_A \in \mathcal{D}'(X \times Y)$,
718: \begin{eqnarray*}
719: \WF'(A) & = & \{((x, \xi),(y, \eta)) \in (T^*X \times T^*Y) \setminus 0 :
720: \\
721: && \quad \quad (x, y ; \xi, - \eta) \in \WF (K_A) \}.
722: \end{eqnarray*}
723: In this notation, we note
724: \begin{eqnarray*}
725: \WF' f^* \subset \left\{ ((x, \xi),(y, \eta)): y = f(x), \,\, \xi = \,^t D_x f \cdot \eta \right\},
726: \end{eqnarray*}
727: which is the graph of the induced symplectomorphism
728: \begin{eqnarray*}
729: \kappa (x, \xi) = (f(x), (\,^tD_x f)^{-1}(\xi)).
730: \end{eqnarray*}
731: \indent To continue, we follow \cite{SjZw}, and let $A(t)$ be a smooth family of pseudodifferential operators:
732: $A(t) = \Op_h^w(a(t))$ with
733: \begin{eqnarray*}
734: a(t) \in \Ci \left( [-1,1]_t; \s^{-\infty,0} \left( T^*X \right) \right),
735: \end{eqnarray*}
736: such that for each $t$, $\WF(A(t)) \Subset T^*X$. Let $U(t): L^2(X) \to L^2(X)$ be defined by
737: \begin{eqnarray}
738: \left\{ \begin{array}{c}
739: hD_tU(t) + U(t)A(t) = 0, \\
740: U(0) = U_0 \in \Psi_h^{0,0}(X),
741: \end{array} \right. \label{U(t)}
742: \end{eqnarray}
743: where $D_t = -i \partial / \partial t$ as usual. If we let $a_0(t)$ be the real-valued $h$-principal symbol of
744: $A(t)$ and let $\kappa(t)$ be the family of symplectomorphisms defined by
745: \begin{eqnarray*}
746: \left\{ \begin{array}{c}
747: \frac{\displaystyle d}{\displaystyle dt} \kappa(t)(x, \xi) = \left( \kappa(t) \right)_* \left( H_{a_0(t)}(x, \xi) \right), \\
748: \kappa(0)(x, \xi) = (x, \xi),
749: \end{array} \right.
750: \end{eqnarray*}
751: for $(x, \xi) \in T^*X$, then $U(t)$ is a family of $h$-FIOs associated to $\kappa(t)$. We have the following
752: well-known theorem of Egorov (see, for example \cite{EvZw}, \S $10.1$).
753: \begin{old-thm}[Egorov's Theorem]
754: Suppose $B \in \Psi_h^{k,m}(X)$, and $U(t)$ defined as above. Suppose further that $U_0$ in \eqref{U(t)} is
755: elliptic ($\sigma_h(U_0) \geq c >0$). Then there exists a smooth family of pseudodifferential operators $V(t)$
756: such that
757: \begin{eqnarray}
758: \left\{ \begin{array}{c}
759: \sigma_h \left( V(t) B U(t) \right) = \left( \kappa(t) \right)^* \sigma_h(B), \\
760: V(t)U(t) -I, \,\, U(t)V(t) -I \in \Psi_h^{-\infty, -\infty}(X). \label{egorov1}
761: \end{array} \right.
762: \end{eqnarray}
763: \end{old-thm}
764: \begin{proof} As $U_0$ is elliptic, there exists an approximate inverse $V_0$, such that $U_0 V_0 -I, \,\, V_0
765: U_0 - I \in \Psi_h^{-\infty, -\infty}$. Let $V(t)$ solve
766: \begin{eqnarray*}
767: \left\{ \begin{array}{c}
768: h D_t V(t) - A(t) V(t) = 0, \\
769: V(0) = V_0.
770: \end{array} \right.
771: \end{eqnarray*}
772: Write $B(t) = V(t)BU(t)$, so that
773: \begin{eqnarray*}
774: hD_t B(t) = A(t)V(t)BU(t) - V(t)BU(t)A(t) = [A(t), B(t)]
775: \end{eqnarray*}
776: modulo $\Psi_h^{-\infty, -\infty}$. But the principal symbol of $[A(t), B(t)]$ is
777: \begin{eqnarray*}
778: \sigma_h \left( [A(t), B(t)] \right) = \frac{h}{i} \left\{ \sigma_h ( A(t)), \sigma_h (B(t)) \right\} =
779: \frac{h}{i} H_{a_0(t)} \sigma_h(B(t)),
780: \end{eqnarray*}
781: so \eqref{egorov1} follows from the definition of $\kappa(t)$.
782: \end{proof}
783: Let $U:= U(1)$, and suppose the graph of $\kappa$ is denoted by $C$. Then we introduce the standard notation
784: \begin{eqnarray*}
785: U \in I_h^0(X \times X; C'), \,\,\, \text{with}\,\,\, C' = \left\{ (x, \xi; y, -\eta) : (x, \xi) =
786: \kappa(y, \eta) \right\},
787: \end{eqnarray*}
788: meaning $U$ is the $h$-FIO associated to the graph of $\kappa$. The next few results when taken together
789: will say that locally all $h$-FIOs associated to symplectic graphs are of the same form as $U(1)$. First
790: a well-known lemma.
791: \begin{lemma}
792: \label{deform-lemma}
793: Suppose $\kappa : \nbhd (0,0) \to \nbhd (0,0)$ is a symplectomorphism fixing $(0,0)$. Then there exists
794: a smooth family of symplectomorphisms $\kappa_t$ fixing $(0,0)$ such that $\kappa_0 = \id$ and $\kappa_1
795: = \kappa$. Further, there is a smooth family of functions $g_t$ such that
796: \begin{eqnarray*}
797: \frac{d}{dt} \kappa_t = (\kappa_t)_* H_{g_t}.
798: \end{eqnarray*}
799: \end{lemma}
800: The proof of Lemma \ref{deform-lemma} is standard, but we include a
801: sketch here, as it will be used in the proof of Proposition
802: \ref{normal-prop-1} (see \cite{EvZw} \S $10.1$ for details).
803: \begin{proof}[Sketch of Proof]
804: First suppose $K: \reals^{2n} \to \reals^{2n}$ is a linear symplectic
805: transformation. Write the polar decomposition of $K$, $K = QP$ with
806: $Q$ orthogonal and $P$ positive definite. It is standard that $K$
807: symplectic implies $Q$ and $P$ are both symplectic as well. Identify
808: $\reals^{2n}$ with $\cx^n$ on which $Q$ is unitary. Write $Q = \exp
809: iB$ for $B$ Hermitian and $P = \exp A$ for $A$ real symmetric and $JA
810: + AJ = 0$, where
811: \begin{eqnarray*}
812: J := \left( \begin{array}{cc} 0 & -I \\ I & 0 \end{array} \right)
813: \end{eqnarray*}
814: is the standard matrix of symplectic structure on $\reals^{2n}$.
815: Then $K_t = \exp (itB) \exp(tA)$ satisfies $K_0 = \id$ and
816: $K_1 = K$.
817:
818: In the case $\kappa$ is nonlinear, set $K = \partial \kappa(0,0)$ and
819: choose $K_t$ such that $K_0 = \id$ and $K_\half = K$. Then set
820: \begin{eqnarray*}
821: \tilde{\kappa}_t(x, \xi) = \frac{1}{t} \kappa(t(x, \xi)),
822: \end{eqnarray*}
823: and note that $\tilde{\kappa}_t$ satisfies $\tilde{\kappa}_0 = K$,
824: $\tilde{\kappa}_1 = \kappa$. Rescale $\tilde{\kappa}_t$ in $t$, so
825: that $\tilde{\kappa}_t \equiv K$ near $1/2$ and $\tilde{\kappa}_1 =
826: \kappa$. Rescale $K_t$ so that $K_0 = \id$ and $K_t \equiv K$ near
827: $1/2$. Then $\kappa_t$ is defined for $0 \leq t \leq 1$ by taking
828: $K_t$ for $0 \leq t \leq 1/2$ and $\tilde{\kappa}_t$ for $1/2 \leq t
829: \leq 1$.
830:
831: To show $\frac{d}{dt} \kappa_t = (\kappa_t)_* H_{g_t}$, set $V_t =\frac{d}{dt}
832: \kappa_t$. Cartan's formula then gives for $\omega$ the symplectic form
833: \begin{eqnarray*}
834: \mathcal{L}_{V_t} \omega = d\omega \contraction V_t + d( \omega \contraction
835: V_t),
836: \end{eqnarray*}
837: but $\mathcal{L}_{V_t} \omega = \frac{d}{dt} \kappa_t^* \omega = 0$ since
838: $\kappa_t$ is symplectic for each $t$. Hence $\omega \contraction V_t =
839: dg_t$ for some smooth function $g_t$ by the Poincar\'{e} lemma, in
840: other words, $V_t = (\kappa_t)_* H_{g_t}$.
841: \end{proof}
842:
843:
844:
845: We have the following version of Egorov's theorem.
846: \begin{proposition}
847: \label{AF=FB}
848: Suppose $U$ is an open neighbourhood of $(0,0)$ and $\kappa: U \to U$ is a symplectomorphism fixing $(0,0)$.
849: Then there is a bounded operator $F : L^2 \to L^2$ such that for all $A = \Op_h^w(a)$,
850: \begin{eqnarray*}
851: AF = FB \,\, \text{microlocally on}\,\, U \times U,
852: \end{eqnarray*}
853: where $B = \Op_h^w(b)$ for a Weyl symbol $b$ satisfying
854: \begin{eqnarray*}
855: b = \kappa^* a + \O(h^2).
856: \end{eqnarray*}
857: $F$ is microlocally invertible in $U \times U$ and $F^{-1} A F = B$ microlocally in $U \times U$.
858: \end{proposition}
859: Proposition \ref{AF=FB} is a standard result, however we include a proof as we will be using it for the proof
860: of Theorem \ref{gamma-egorov}.
861: \begin{proof}
862: For $0 \leq t \leq 1$ let $\kappa_t$ be a smooth family of symplectomorphisms satisfying $\kappa_0 = \id$,
863: $\kappa_1 = \kappa$, and let $g_t$ satisfy $\frac{d}{dt} \kappa_t =
864: (\kappa_t)_* H_{g_t}$. Let $G_t = \Op_h^w(g_t)$,
865: and solve the following equations
866: \begin{eqnarray*}
867: && \left\{ \begin{array}{c}
868: h D_tF(t) + F(t) G(t) = 0, \,\, (0 \leq t \leq 1) \\
869: F(0) = I,
870: \end{array} \right. \\
871: && \left\{ \begin{array}{c}
872: hD_t \tilde{F}(t) - G(t) \tilde{F}(t) = 0,\,\, (0 \leq t \leq 1) \\
873: \tilde{F}(0) = I.
874: \end{array} \right.
875: \end{eqnarray*}
876: Then $F(t), \tilde{F}(t) = \O(1) : L^2 \to L^2$ and
877: \begin{eqnarray*}
878: hD_t\left( F(t) \tilde{F}(t) \right) = -F(t)G(t) \tilde{F}(t) + F(t) G(t) \tilde{F}(t) = 0,
879: \end{eqnarray*}
880: so $F(t) \tilde{F}(t) = I$ for $0 \leq t \leq 1$. Similarly,
881: $E(t) = \tilde{F} F - I$ satisfies
882: \ben
883: \label{ode-eg}
884: hD_t E(t) = G(t) \tilde{F}(t) F(t) - \tilde{F}(t) F(t) G(t) = [G(t), E(t) ]
885: \een
886: with $E(0) = 0$. But equation \eqref{ode-eg} has
887: unique solution $E(t) \equiv 0$ for the initial condition $E(0) = 0$.
888: Hence $\tilde{F}(t) F(t) =I$ microlocally. \\
889: \indent Now set $B(t) = \tilde{F}(t) A F(t)$. We would like to show $B(t) = \Op_h^w(b_t)$, for $b_t =
890: \kappa_t^*a + \O(h^2)$. Set $\tilde{B}(t) = \Op_h^w (\kappa_t^*a)$. Then
891: \begin{eqnarray*}
892: hD_t \tilde{B}(t) & = & \frac{h}{i} \Op_h^w \left( \frac{d}{dt}\kappa_t^*a \right) \\
893: & = & \frac{h}{i} \Op_h^w \left(\{ g_t, \kappa_t^*a \}\right) \\
894: & = & \left[ G(t), \tilde{B}(t) \right] + E_1(t),
895: \end{eqnarray*}
896: where $E_1(t) = \Op_h^w(e_1(t))$ for $e_1(t)$ a smooth family of symbols.
897: Note if we take $g_t \# (\kappa_t^* a) - (\kappa_t^* a) \# g_t$, the composition formula \eqref{a-pound-b}
898: implies the $h^2$ term vanishes for the Weyl calculus since $\omega^2$ is symmetric while
899: \begin{eqnarray*}
900: g_t(x,\xi) \kappa_t^* a(y, \eta) - \kappa_t^* a(x, \xi) g_t(y, \eta)
901: \end{eqnarray*}
902: is antisymmetric. Thus $E_1(t) \in \Psi_h^{0,-3}$, since we are working
903: microlocally. We calculate
904: \begin{eqnarray}
905: \lefteqn{hD_t \left( F(t) \tilde{B}(t) \tilde{F}(t) \right) =} \label{B-de1}\\
906: & = & -F(t) G(t) \tilde{B}(t) \tilde{F}(t) + F(t)
907: \left( \left[ G(t), \tilde{B}(t) \right] + E_1(t) \right) \tilde{F}(t) \label{B-de2}\\
908: && \quad + F(t)\tilde{B}(t) G(t) \tilde{F}(t) \nonumber \\
909: & = & F(t) E_1(t) \tilde{F}(t) \label{B-de3}\\
910: & = & \O(h^3).\nonumber
911: \end{eqnarray}
912: Integrating in $t$ and dividing by $h$ we get
913: \begin{eqnarray}
914: \label{B-int}
915: F(t) \tilde{B}(t) \tilde{F}(t) = A + \frac{i}{h} \int_0^t F(s)E_1(s)\tilde{F}(s) ds = A + \O(h^2),
916: \end{eqnarray}
917: so that $\tilde{B}(t) - B(t) = \O(h^2)$.
918:
919: We will construct families of pseudodifferential operators $B_k(t)$ so
920: that for each $m$
921: \ben
922: B(t) = \tilde{B}(t) + B_1(t) + \cdots + B_m(t) + \O(h^{m+2}). \label{B-ind}
923: \een
924: Let
925: \be
926: \tilde{e}_1(t) = (\kappa_t)^* \int_0^t (\kappa_s^{-1})^* e_1(s) ds,
927: \ee
928: and set $\tilde{E}_1(t) = \Op_h^w( \tilde{e}_1(t))$. Observe
929: \be
930: hD_t \tilde{E}_1 = \left[ G(t), \tilde{E}_1 \right] + \frac{h}{i}\left(
931: E_1(t) + E_2(t)\right) ,
932: \ee
933: where $E_2(t) \in \Psi_h^{0,-4}$ by the Weyl calculus, since
934: $[G,\tilde{E}_1] = \O(h^4)$. Then as in (\ref{B-de1}-\ref{B-de3})
935: \be
936: hD_t \left( F(t)\tilde{E}_1(t) \tilde{F}(t) \right) & = & - F(t)
937: \left[ G(t), \tilde{E}_1(t) \right] \tilde{F}(t) + F(t) hD_t\left(
938: \tilde{E}_1(t) \right) \tilde{F}(t) \\
939: & = & \frac{h}{i}\left( F(t) E_1(t) \tilde{F}(t) + F(t) E_2(t)
940: \tilde{F}(t)\right) .
941: \ee
942: Integrating in $t$ gives
943: \be
944: F(t) \tilde{E}_1(t) \tilde{F}(t) = \int_0^t F(s)E_1(s) \tilde{F}(s) ds
945: + \frac{i}{h}\int_0^t F(s)E_2(s) \tilde{F}(s) ds,
946: \ee
947: and substituting in \eqref{B-int} gives
948: \be
949: \tilde{B}(t) - B(t) & = & \frac{i}{h} \tilde{E}_1(t) - \tilde{F}(t) \left(
950: \frac{i}{h} \int_0^t F(s)E_2(s)\tilde{F}(s)ds \right) F(t) \\
951: & = & \frac{i}{h} \tilde{E}_1(t) + \O(h^3).
952: \ee
953: Setting $B_1(t) = i \tilde{E}_1(t)/h$ and continuing inductively gives
954: $B_k(t)$ satisfying \eqref{B-ind}.
955:
956: \indent Let $l$ be a linear symbol, and $L = \Op_h^w(l)$. Then
957: \begin{eqnarray*}
958: \ad_L (\tilde{B} - B) = \left[ \tilde{B} - B, L \right] = \O(h^2).
959: \end{eqnarray*}
960: Fix $N$. From \eqref{B-ind} we can choose $B_1, \ldots , B_N$ so that
961: replacing $\tilde{B}$ with $\tilde{B}+ B_1 + \cdots + B_N$, we have for $l_1, \ldots, l_N$ linear symbols, $L_k = \Op_h^w(l_k)$,
962: \begin{eqnarray*}
963: \ad_{L_1} \circ \cdots \circ \ad_{L_N} (\tilde{B} - B) = \O(h^{N+2}),
964: \end{eqnarray*}
965: so Beals's Theorem implies $B(t) = \Op_h^w(b(t))$ for $b(t) = \kappa_t^* a + \O(h^2)$.
966: \end{proof}
967: The next proposition is essentially a converse to Proposition \ref{AF=FB}.
968: \begin{proposition}
969: \label{U-FIO}
970: Suppose $U = \O(1): L^2 \to L^2$ and for all pseudodifferential operators $A ,B \in \Psi_h^{0,0}(X)$
971: such that $\sigma_h(B) = \kappa^* \sigma_h(A)$, $AU = UB$ microlocally near $(\rho_0, \rho_0)$,
972: where $\kappa : \nbhd (\rho_0, \rho_0) \to \nbhd (\rho_0 , \rho_0)$ is a symplectomorphism fixing
973: $(\rho_0, \rho_0)$. Then $U \in I_h^0(X \times X; C')$ microlocally near $(\rho_0, \rho_0)$.
974: \end{proposition}
975: \begin{proof}
976: Choose $\kappa_t$ a smooth family of symplectomorphisms such that $\kappa_0 = \id$, $\kappa_1 = \kappa$,
977: and $\kappa_t( \rho_0) = \rho_0$. Choose $a(t)$ a smooth family of functions satisfying $\frac{d}{dt}
978: \kappa_t = (\kappa_t)_* H_{a(t)}$, and let $A(t) = \Op_h^w(a(t))$. Let $U(t)$ be a solution to
979: \begin{eqnarray*}
980: \left\{ \begin{array}{c}
981: hD_t U(t) - U(t)A(t) = 0, \\
982: U(1) = U,
983: \end{array} \right.
984: \end{eqnarray*}
985: for $0 \leq t \leq 1$. Next let $A$ and $B$ satisfy the assumptions of the proposition. Since $AU = UB$,
986: we can find $V(t)$ satisfying
987: \begin{eqnarray}
988: \label{V(t)-1}
989: \left\{ \begin{array}{c}
990: A U(t)V(t) = U(t)BV(t), \\
991: V(0) = \id.
992: \end{array} \right.
993: \end{eqnarray}
994: By Egorov's theorem, the right hand side of \eqref{V(t)-1} is equal to
995: \begin{eqnarray*}
996: U(t) V(t) \left( V(t)^{-1} B V(t) \right) = U(t) V(t) A + \O(h).
997: \end{eqnarray*}
998: Setting $t = 0$, we see $[U(0), A ] = \O(h)$. Applying the same argument to $[U(t), A]$ and another choice
999: of $\tilde{A}, \tilde{B}$ satisfying the hypotheses of the proposition yields by induction,
1000: \begin{eqnarray}
1001: \label{beals-00}
1002: \ad_{A_1} \circ \cdots \circ \ad_{A_N} U(0) = \O(h^N)
1003: \end{eqnarray}
1004: for any choice of $A_1, \ldots, A_N \in \Psi_h^{0,0}(X)$. Since we are only interested in what $U(t)$ looks
1005: like microlocally, \eqref{beals-00} is sufficient to apply Beals's Theorem and conclude that $U(0) \in
1006: \Psi_h^{0,0}(X)$. Thus $U(t)$ and hence $U(1) = U$ is in $I_h^0(X \times X; C')$ for the twisted graph
1007: \begin{eqnarray*}
1008: C' = \left\{ (x,\xi, y, -\eta): (y, \eta) = \kappa(x, \xi) \right\}.
1009: \end{eqnarray*}
1010: \end{proof}
1011: Using the following more general version of the Poincar{\'e} lemma
1012: from \cite{Wei2}, we will be able to generalize Proposition \ref{AF=FB} to a
1013: neighbourhood of a periodic orbit.
1014: \begin{lemma}
1015: \label{poincare-lemma}
1016: Let $N \subset T^*X$ be a closed submanifold, and assume $(x, \xi) \in
1017: N$ implies $(x,0) \in N$. Then if $\omega$ is a closed $k$-form such that $\left. \omega \right|_N = 0$, then there is a $(k-1)$-form $I( \omega)$ in a neighbourhood of $N$ such that $\omega = d I(\omega)$.
1018: \end{lemma}
1019: \begin{proof}
1020: Let $m_s : T^*X \to T^*X$, $m_s: (x, \xi) \mapsto (x, s \xi)$, be multiplication by $s$ in the fibres, and define
1021: \begin{eqnarray*}
1022: X_s = \left. \left( \frac{d}{dr} m_r \right) \right|_{r=s}.
1023: \end{eqnarray*}
1024: That is, in coordinates,
1025: \begin{eqnarray*}
1026: X_s = \frac{1}{s} \sum_j \xi_j \frac{\partial}{\partial_{\xi_j}}
1027: \end{eqnarray*}
1028: is just $1/s$ times the radial vector field. Then
1029: \begin{eqnarray*}
1030: \left. \frac{d}{dr} (m_r^* \omega) \right|_{r=s} = m_s^* \left( X_s \contraction d \omega \right)
1031: + d \left( m_s^*( X \contraction \omega) \right),
1032: \end{eqnarray*}
1033: and integrating in $r$ gives
1034: \begin{eqnarray*}
1035: \omega - m_0^* \omega = I ( d\omega) + d I(\omega)
1036: \end{eqnarray*}
1037: for
1038: \begin{eqnarray*}
1039: I (\omega) = \int_0^1 m_r^* (X_r \contraction \omega ) dr.
1040: \end{eqnarray*}
1041: Now $\left. \omega \right|_N = 0$ and $ d \omega = 0$ finishes the proof.
1042: \end{proof}
1043:
1044: \begin{theorem}
1045: \label{gamma-egorov}
1046: Suppose $N \subset T^*X$ is a closed submanifold such that $(x, \xi)
1047: \in N$ implies $(x,0) \in N$, and assume
1048: $\kappa: \neigh (N) \to \kappa(\neigh(N))$ is a symplectomorphism which is smoothly homotopic in the symplectic
1049: group to identity on $N$. Then there is a bounded linear operator $F: L^2(\neigh(N)) \to L^2(\kappa(\neigh(N)))$
1050: such that for all $A = \Op_h^w(a)$,
1051: \begin{eqnarray*}
1052: AF = FB \,\,\, \text{microlocally on } \neigh(N) \times \kappa(\neigh(N)),
1053: \end{eqnarray*}
1054: where $B = \Op_h^w(b)$ for a Weyl symbol $b = \kappa^*a + \O(h^2)$. Further, $F$ is microlocally invertible
1055: and $F^{-1}A F = B$ in $N \times \kappa(N)$.
1056: \end{theorem}
1057: \begin{proof}
1058: The proof will follow from the proof of Proposition \ref{AF=FB}. Let $\kappa_t$ be the homotopy in the
1059: Proposition, $\kappa_0 = \id$ and $\kappa_1 = \kappa$. We need only verify that $\kappa_t$ is generated
1060: by a Hamiltonian. Set $V_t = \frac{d}{dt} \kappa_t$, and calculate
1061: \begin{eqnarray*}
1062: 0 = \frac{d}{dt} \kappa_t^* \omega = \mathcal{L}_{V_t} \omega = V_t \contraction d \omega + d ( V_t \contraction \omega ).
1063: \end{eqnarray*}
1064: Hence $\lambda_t = V_t \contraction \omega$ is closed and further $\left. \lambda_t \right|_{N} = 0$ so we may
1065: apply Lemma \ref{poincare-lemma} to obtain a $0$-form $I( \lambda_t)$ so that
1066: \begin{eqnarray*}
1067: d I(\lambda_t) = \lambda_t,
1068: \end{eqnarray*}
1069: or
1070: \begin{eqnarray*}
1071: V_t = H_{I(\lambda_t)}.
1072: \end{eqnarray*}
1073: \end{proof}
1074:
1075:
1076: We will make use of the following proposition (see \cite{EvZw} \S
1077: $10.5$ for a proof).
1078: \begin{proposition}
1079: \label{hDx-prop}
1080: Let $P \in \Psi_h^{k,0}(X)$ be a semiclassical operator of real principal type ($p = \sigma_h(P)$ is real
1081: and independent of $h$), and assume $dp \neq 0 $ whenever $p=0$. Then for any $\rho_0 \in \{ p^{-1}(0) \}$,
1082: there exists a symplectomorphism $\kappa : T^*X \to T^* \reals^n$ defined from a neighbourhood of $\rho_0$
1083: to a neighbourhood of $(0,0)$ and an $h$-FIO $T$ associated to its graph such that \\
1084: \indent (i) $\kappa^* \xi_1 = p$, \\
1085: \indent (ii) $TP = h D_{x_1} T$ microlocally near $(\rho_0; (0,0))$, \\
1086: \indent (iii) $T^{-1}$ exists microlocally near $((0,0); \rho_0)$.
1087: \end{proposition}
1088:
1089: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1090:
1091: \section{Symplectic Geometry and Quadratic Forms}
1092: \label{symplectic-geometry}
1093: We now return to the setup of the introduction. Let $P(h)$ satisfy all the assumptions from \S \ref{intro}.
1094: The main tool at our disposal is to use symplectomorphisms to transform the Weyl principal symbol into a
1095: different Weyl principal symbol which is in a more tractible form. Then by Propositions \ref{AF=FB} and
1096: \ref{U-FIO}, any estimates we prove about the quantization of the transformed principal symbol will hold for
1097: the original operator modulo $\O (h^2)$.
1098:
1099: It is classical (see, for example \cite{AbMa}) that using our
1100: assumptions on $p$, the Implicit Function Theorem guarantees that there is an $\epsilon_0 >0$ such that for
1101: $\epsilon \in [-\epsilon_0, \epsilon_0]$, the energy surface $\{p^{-1}(\epsilon)\}$ is regular and contains a
1102: closed loxodromic orbit $\gamma^\epsilon$. Further,
1103: \begin{eqnarray*}
1104: \overline{\gamma} := \bigcup_{-\epsilon_0 \leq \epsilon \leq \epsilon_0} \gamma^\epsilon
1105: \end{eqnarray*}
1106: is a smooth, 2-dimensional symplectic manifold diffeomorphic to $\SS^1
1107: \times [-\epsilon_0, \epsilon_0] \subset T^* \SS^1$. Choose symplectic
1108: coordinates $(t, \tau, x, \xi)$ in a neighbourhood of $\overline{\gamma}$ so that $\gamma$ is the image of the unit circle, $\SS^1 \ni t \mapsto
1109: \gamma(t)$, $t$ parametrizes $\gamma^\epsilon$, and $\gamma = \{t, 0;
1110: 0,0 \}$. In \cite{AbMa} it is shown that $S=\{ t=0\}$ is a contact manifold with the contact form
1111: $\tom_{(x, \xi)} = i^* \omega$, where $i: S \hookrightarrow X$ is the inclusion.
1112: Then the Poincar\'e map preserves $p$ and $\tom$, modulo a term encompassing the period shift for
1113: $\epsilon \in [- \epsilon_0, \epsilon_0]$ different from zero and $(x,
1114: \xi) \neq (0,0)$. This
1115: motivates our next change of variables. Similar to \cite{Sjo3}, we
1116: observe that $\tau$ depends only on the energy surface in which
1117: $\gamma^\epsilon$ lies: $\tau = g( \epsilon )$. $H_p$ is tangent to
1118: the energy surface $\{ p^{-1} ( \epsilon) \}$ for each $\epsilon \in
1119: [- \epsilon_0, \epsilon_0]$, so that
1120: \begin{eqnarray*}
1121: && \partial_t p(t, \tau, x, 0) = \partial_t p(t, \tau, 0, \xi) = 0,
1122: \,\,\, \text{and} \\
1123: && \partial_x p(t, \tau, 0,0) = 0, \,\, \partial_\xi p(t, \tau, 0,0) =
1124: 0,
1125: \end{eqnarray*}
1126: so that
1127: \begin{eqnarray*}
1128: p(t, \tau, 0,0) = f( \tau ) \,\,\, \text{and } p(t, 0, x, \xi ) = f(0)
1129: + \O_t( x^2 + \xi^2).
1130: \end{eqnarray*}
1131: Thus, there exists a smooth nonvanishing function $a( t, \tau, x,
1132: \xi)$ defined in a neighbourhood of $\overline{\gamma}$ such that
1133: \begin{eqnarray*}
1134: a(t, \tau, x, \xi) p(t, \tau, x, \xi) = f(\tau) + \O_t(x^2 + \xi^2).
1135: \end{eqnarray*}
1136:
1137: %Let $g(\epsilon)$ be the action of $\gamma^\epsilon$, that is
1138: %\begin{eqnarray*}
1139: %g(\epsilon) = \int_{\gamma^\epsilon} \lambda, \quad d \lambda = \omega.
1140: %\end{eqnarray*}
1141: %Now $g'(\epsilon) = T(\epsilon)$, where $T(\epsilon)$ is the period of $\gamma^\epsilon$,
1142: %so that $H_\tau = g'(p) H_p = T(p) H_p$, and $\tau = g(p)$. Further,
1143: %since $g(0) = 0$, we can write $p = a(\tau)$ on $\overline{\gamma}$ with $a(\tau) \geq C^{-1}
1144: %>0$ and smooth
1145: %for $\epsilon_0$ sufficiently small. We
1146: %can use Darboux's theorem to complete the coordinates. We choose $\{x_j\}_{j=1}^{n-1}$ and $\{\xi_k\}_{k=1}^{n-1}$ so that
1147: %\begin{eqnarray*} \begin{array}{c}\frac{d}{dt} \kappa_t = (\kappa_t)_* H_{g_t}
1148: %H_{x_j} x_k = H_{\xi_j} \xi_k = 0, \\
1149: %H_{x_j} \tau = H_{x_j} t = H_{\xi_j} \tau = H_{\xi_j} t = 0, \\
1150: %H_{x_j} \xi_k = \delta_{jk}, \,\,\, \text{and} \\
1151: %\left. x_j \right|_{\overline{\gamma}} = \left. \xi_k \right|_{\overline{\gamma}} = 0, \,\,\, \forall j,k.
1152: %\end{array}
1153: %\end{eqnarray*}
1154:
1155:
1156:
1157: Since the Hamiltonian vector field of $p$, $H_p$ is tangent to
1158: $\{p=0\}$, we can choose a Poincar\'{e} section contained in $\{p=0\}$, that is,
1159: a $2n-2$ dimensional submanifold $N$, transverse
1160: to $H_p$ on $\{p=0\}$ centered at $\gamma(0)$. Let $S : N \to N$ be the Poincar\'{e} (first return) map near
1161: $\gamma(0)$. Note that $\omega = dt \wedge d \tau + \tom_{(x, \xi)}$ is the symplectic form on $T^*X$ in our
1162: choice of coordinates, so $S$ preserves the $(2n-2)$ dimensional symplectic form $\tom$ on $N$. Thus $S$ is
1163: a symplectic mapping $N \to N$, with $S(0) = 0$. That $\gamma$ is loxodromic means none of the eigenvalues of
1164: $dS(0)$ lie on the unit circle. In this section for simplicity we
1165: consider only the case where all the eigenvalues are distinct, (the general case is handled in \S
1166: \ref{non-distinct-section}).
1167: We think of $dS(0)$ as the linearization of $S$ near $0 \in N$, with $N$
1168: identified with $T_0N$ near $0$. \\
1169: \indent We want to put
1170: $p$ into a normal form in a neighbourhood of $\gamma$. Inspiration for this construction comes from
1171: \cite{guillemin} and \cite{Sjo3}. Let $q(\rho)$ be defined near $0 \in N$ and quadratic
1172: such that $dS(0) = \exp H_q$. Let $\kappa_t$ be a smooth family of symplectomorphisms such that $\kappa_0 = \id$
1173: while $\kappa_1 = S$. Then from the proof of Lemma \ref{deform-lemma}
1174: we can find $q_t(\rho)$ defined near $0 \in N$ so that
1175: \begin{eqnarray*}
1176: q_t(\rho) = q(\rho) + f_t(\rho)
1177: \end{eqnarray*}
1178: with $f_t(\rho) = \O_t( |\rho|^3)$ and
1179: \begin{eqnarray*}
1180: \frac{d}{dt} \kappa_t = (\kappa_t)_* H_{q_t}.
1181: \end{eqnarray*}
1182: \begin{remark}
1183: Here we see the first obstacle to extending these techniques to
1184: include negative real eigenvalues: We want to write $dS(0) = \exp H_q$
1185: for a real quadratic form $q$. But this is impossible for some linear
1186: symplectic transformations with negative eigenvalues as the example
1187: \begin{eqnarray*}
1188: dS(0) = \left( \begin{array}{cc}
1189: -e^{2} & 0 \\ 0 & -e^{-2}
1190: \end{array} \right)
1191: \end{eqnarray*}
1192: shows. Here $dS(0)$ is symplectic, but cannot be written as $\exp
1193: H_q$ with $q$ real. Roughly, negative eigenvalues may be realized
1194: only by deforming
1195: a family of symplectomorphisms $\kappa_t$ {\it through an elliptic
1196: component}.
1197: \end{remark}
1198:
1199: Set $\tilde{p}(s, \sigma, \rho) = \sigma + q_s(\rho)$. We will show $p$ and $\tilde{p}$
1200: are equivalent under a symplectic change of coordinates on the set
1201: $p^{-1}(0)$. Then since both $p$ and $\tilde{p}$ have nonvanishing
1202: differentials, we can write
1203: \begin{eqnarray}
1204: \label{p=ptilde}
1205: \kappa^* p = b(t, \tau, x, \xi) \tilde{p}
1206: \end{eqnarray}
1207: for a smooth, positive function $b$ and a symplectomorphism $\kappa$.
1208: Indeed, we claim
1209: \begin{eqnarray*}
1210: \exp (tH_{p})(s, \sigma, \rho) = \left( s+t, \sigma_t(\rho,s,\sigma), \kappa_{t+s} \circ \kappa_s^{-1}(\rho)\right)
1211: \end{eqnarray*}
1212: for some $\sigma_t(s, \sigma, \rho)$, giving \eqref{p=ptilde}. To see this, set
1213: \begin{eqnarray*}
1214: \Phi_t(s, \rho) := \left( s+t, \kappa_{t+s} \circ \kappa_s^{-1}(\rho) \right).
1215: \end{eqnarray*}
1216: We need to check that $\left. \Phi_t \right|_{N \times \SS^1}$ is a $1$-parameter group. We compute
1217: \begin{eqnarray*}
1218: \left. \Phi_{t_1 + t_2} \right|_{N \times \SS^1} ( s, \rho ) = \left( s + t_1 + t_2, \kappa_{t_1 + t_2 + s}
1219: \circ \kappa_{s}^{-1}(\rho) \right).
1220: \end{eqnarray*}
1221: But we check
1222: \begin{eqnarray*}
1223: \lefteqn{ \left. \Phi_{t_1} \right|_{N \times \SS^1} \circ \Phi_{N \times \SS^1} ( s, \rho) =} \\
1224: & = & \left. \Phi_{t_1} \right|_{N \times \SS^1} \left( s+ t_2, \kappa_{t_2 + s} \circ \kappa_s^{-1}(\rho) \right) \\
1225: & = & \left( s+t_1+t_2 , \kappa_{t_1 + t_2 + s} \circ \kappa_{t_2 + s}^{-1}( \kappa_{t_2+s} \circ
1226: \left(\kappa_{t_2 + s} \circ \kappa_s^{-1}(\rho) \right) \right),
1227: \end{eqnarray*}
1228: so the group law holds. We need only verify that $p$ and $\tilde{p}$ have the same Poincar\'{e} map, so we check:
1229: \begin{eqnarray*}
1230: \left. \left( \frac{d}{dt} \left. \Phi_t \right|_{N \times \SS^1}( s, \rho ) \right) \right|_{t=0} =
1231: \left(1, H_{q_s}(\rho) \right),
1232: \end{eqnarray*}
1233: which is clear. Note this construction depends only on the Poincar\'{e} map $S$ and is unique up to symplectomorphism. \\
1234: \indent Next we want to examine what form the quadratic part $q(\rho)$ can take. The fact that
1235: $S(0) = 0$ implies we can write
1236: \begin{eqnarray}
1237: \label{q-1}
1238: q ( \rho) = \frac{1}{2} \langle q''(0) \rho, \rho \rangle .
1239: \end{eqnarray}
1240: Now we define the Hamilton matrix $B$ by
1241: \begin{eqnarray}
1242: \label{b-matrix}
1243: q (\rho) = \frac{1}{2} \tom(\rho, B \rho)
1244: \end{eqnarray}
1245: so that the symplectic transpose of $B$, $\,^{\tom}B$, is equal to
1246: $-B$. Note that $B$ is the matrix representation of $H_q$, and so has
1247: eigenvalues which are
1248: the logarithms (with a suitably chosen branch cut) of the eigenvalues
1249: of $dS(0)$. Thus the condition
1250: that $\gamma$ be loxodromic implies none of the eigenvalues of $B$ have nonzero real part. Recall that
1251: since $dS(0)$ is a symplectic transformation, if $\mu$ is an eigenvalue of $dS(0)$, then so are
1252: $\overline{\mu}$, $\mu^{-1}$, and $\overline{\mu}^{-1}$. This implies
1253: for the corresponding Hamilton matrix $B$
1254: in (\ref{b-matrix}), if $\lambda$ is an eigenvalue of $B$, then so are $-\lambda$, $\overline{\lambda}$,
1255: and $-\overline{\lambda}$. Thus the analysis of $B$ in the loxodromic, or complex hyperbolic case amounts
1256: to analyzing the eigenvalues in sets of $2$ or $4$. For this we
1257: follow the appendix in \cite{IaSj}, and recall for this section we are
1258: assuming the eigenvalues are distinct. There are $2$ cases. First, assume $\lambda_j > 0$ is real. Then $-\lambda_j$ is also an eigenvalue.
1259: Let $e_j$ and $f_j$ be the respective eigenvectors such that $\widetilde{\omega}(e_j, f_j)=1$. Then $e_j$
1260: and $f_j$ span a real symplectic vector space of dimension $2$. For a point $\rho$ in this vector space,
1261: write $\rho = x_je_j + \xi_j f_j$. Then $(x_j, \xi_j)$ are symplectic coordinates, in which $q_j(\rho)$,
1262: the projection of $q$ onto the $j$th coordinates becomes $q_j( \rho) = \lambda_j x_j \xi_j$. We call the
1263: \begin{eqnarray*}
1264: \lambda_j x_j \xi_j
1265: \end{eqnarray*}
1266: the {\it action variables}. \\
1267: \indent Now we would like to see what these actions look like when the eigenvalues have nonzero imaginary part.
1268: Suppose $\lambda_j$ is an eigenvalue with $\Re \lambda_j > 0$, $\Im \lambda_j >0$. Then $-\lambda_j$,
1269: $\overline{\lambda}_j$, and $-\overline{\lambda}_j$ are eigenvalues. Let $e_j$, $f_j$, $\overline{e}_j$,
1270: and $\overline{f}_j$ be the respective eigenvectors. Note $\tom (e_j, \overline{e}_j)= \tom(e_j, \overline{f}_j) =
1271: \tom (f_j, \overline{f}_j) = 0$. Scale $f_j$ so that $\tom (e_j, f_j) = 1$. Then $\{e_j, f_j \}$ and $
1272: \{\overline{e}_j, \overline{f}_j \}$ span complex conjugate symplectic vector spaces of complex dimension $2$.
1273: Thus $\{ e_j, \overline{e}_j, f_j, \overline{f}_j \}$ span a symplectic vector space of complex dimension $4$
1274: which is the complexification of a real symplectic vector space. Write a point $\rho$ in this space in this basis,
1275: $\rho = z_j e_j + \zeta_j f_j + w_j \overline{e}_j + \eta_j \overline{f}_j$. Then $(z_j, \zeta_j, w_j, \eta_j)$
1276: become symplectic coordinates, in which the projection $q_j$ becomes $q_j(\rho) = \lambda_j z_j \zeta_j +
1277: \overline{\lambda}_j w_j \eta_j$. Now write
1278: \begin{eqnarray*}
1279: e_j = \frac{1}{\sqrt{2}} \left( e_j^1 + i e_j^2 \right), \quad f_j = \frac{1}{\sqrt{2}} \left( f_j^1 - if_j^2 \right),
1280: \end{eqnarray*}
1281: for real $e_j^k$, $f_j^k$. This is a symplectic change of basis, and writing $\rho$ in this basis:
1282: \begin{eqnarray*}
1283: \rho = z_j e_j + \zeta_j f_j + w_j \overline{e}_j + \eta_j \overline{f}_j = \sum_{k=1}^2
1284: \left( x_j^k e_j^k + \xi_j^k f_j^k \right),
1285: \end{eqnarray*}
1286: we have
1287: \begin{eqnarray*}
1288: q_j(\rho) = \Re \lambda_j \left( x_j^1 \xi_j^1 + x_j^2 \xi_j^2 \right) - \Im \lambda_j
1289: \left( x_j^1 \xi_j^2 - x_j^2 \xi_j^1 \right).
1290: \end{eqnarray*}
1291: This is summarized in the following proposition (using the notation of
1292: \cite{IaSj}). Let $n_{hc}$ be the number of complex hyperbolic eigenvalues
1293: $\mu_j$ of $dS(0)$ with $|\mu_j|>1$, and $n_{hr}$ the number of real
1294: hyperbolic eigenvalues $\mu_j$ of $dS(0)$ such that $\mu_j >1$. Thus
1295: we have $2n-2 = 4n_{hc} + 2 n_{hr}$.
1296: \begin{proposition}
1297: \label{normal-prop-1}
1298: Let $p \in \Ci ( T^*X)$, $\gamma \subset \{p = 0 \}$ as in the introduction, with
1299: the linearized Poincar\'{e} map having distinct eigenvalues $\mu_j$ not on the
1300: unit circle. Assume for $1 \leq j \leq n_{hc}$ we have $|\mu_j|>1$
1301: and $\Im \mu_j >0$, and for $2n_{hc}+1 \leq j \leq 2n_{hc} + n_{hr}$
1302: we have $\mu_j >1$. Then there exists a neighbourhood, $U$, of $\gamma$ in $T^*X$,
1303: a smooth positive function $b \geq C^{-1} >0$ defined in $U$, and a
1304: symplectomorphism $\kappa : U \to \kappa(U) \subset T^*\SS_{(t, \tau)}^1 \times T^* \reals_{(x, \xi)}^{n-1}$ such that
1305: \begin{eqnarray*}
1306: \kappa(\gamma) = \{ (t,0;0,0) : t \in \SS^1 \},
1307: \end{eqnarray*}
1308: and $b(t, \tau, x, \xi) p = \kappa^*(g + r)$, with
1309: \begin{eqnarray}
1310: \lefteqn{ g(t, \tau; x, \xi) =} \nonumber \\
1311: & = & \tau + \sum_{j=1}^{n_{hc}}\left( \Re \lambda_j \left( x_{2j-1} \xi_{2j-1} + x_{2j} \xi_{2j} \right)
1312: - \Im \lambda_j \left( x_{2j-1} \xi_{2j} - x_{2j}\xi_{2j-1} \right) \right) \label{p-pullback-def-1} \\
1313: & & \quad + \sum_{j = 2n_{hc}+1}^{2n_{hc} + n_{hr}} \lambda_j x_j \xi_j , \quad \mathrm{with} \,\,\,
1314: 2n_{hc} + n_{hr} = n-1 \,\,\, \mathrm{and} \label{p-pullback-def-2} \\
1315: r & = & \O (|x|^3 + |\xi|^3). \nonumber
1316: \end{eqnarray}
1317: Here $\lambda_j = \log (\mu_j)$ for $|\mu_j|>1$ and $\Im \lambda_j
1318: \geq 0$.
1319: \end{proposition}
1320: \begin{remark}
1321: The quadratic form (\ref{p-pullback-def-1}-\ref{p-pullback-def-2}) in Proposition \ref{normal-prop-1}
1322: is the leading part of the real Birkhoff normal form for a symplectomorphism near a loxodromic fixed point.
1323: With a non-resonance condition and the addition of some higher order
1324: ``action'' variables (see, for example, \cite{hz} and \cite{IaSj}), the error $r$ could be taken to be
1325: \begin{eqnarray*}
1326: r & = & \O (|x|^4 + |\xi|^4),
1327: \end{eqnarray*}
1328: or even $\O(|x|^\infty + |\xi|^\infty)$.
1329: \end{remark}
1330: \begin{remark}
1331: We think of $p(t, \tau, x, \xi) \in \Ci( \reals^4)$, $p = \tau +
1332: \lambda x \xi$, for $\lambda>0$ as our ``model case''. The feature we
1333: are going to exploit about this model case is that if $G(t, \tau, x,
1334: \xi) = \half ( x^2 - \xi^2)$, then
1335: \begin{eqnarray}
1336: \label{hp-quad}
1337: H_pG = \lambda ( x^2 + \xi^2),
1338: \end{eqnarray}
1339: which is a positive definite quadratic form. However,
1340: the growth of $x^2 - \xi^2$ will force us to use instead $G(x, \xi) =
1341: \log (1 + x^2) - \log(1 + \xi^2)$. Suppose $p = \tau + \lambda x
1342: \xi + x^3 -\xi^3 = \tau + \lambda x\xi + \O(x^3 + \xi^3)$ in a
1343: neighbourhood of $\gamma$ of size $\epsilon >0$ as in
1344: Proposition \ref{normal-prop-1}. Then
1345: \begin{eqnarray*}
1346: H_p G = \lambda \frac{x^2}{1+ x^2} + \lambda \frac{\xi^2}{1 + \xi^2} +
1347: 3 \frac{\xi^2 x}{1 + x^2} + 3 \frac{ x^2 \xi}{1 + \xi^2}.
1348: \end{eqnarray*}
1349: Motivated by \eqref{hp-quad}, we would like to write this as
1350: \begin{eqnarray*}
1351: H_pG = \lambda \frac{x^2}{1+ x^2}(1 + \O( \epsilon)) + \lambda
1352: \frac{\xi^2}{1 + \xi^2}(1 + \O(\epsilon)),
1353: \end{eqnarray*}
1354: which we clearly cannot do in this example.
1355: \end{remark}
1356:
1357: As the last remark indicates, in order to deal with the error terms, we will need a more refined form than that given in
1358: Proposition \ref{normal-prop-1}. Inspiration for this development, and in particular Proposition \ref{normal-prop-2}
1359: comes from \cite{GeSj} and \cite{sjostrand}.
1360:
1361: Let $\{\mu_j\}$ be the eigenvalues of the linearized Poincar\'{e} map at $\gamma(0)$. They come in pairs $\mu_j$,
1362: $\mu_j^{-1}$ for the real $\mu_j$ and in sets of four $\mu_j$, $\overline{\mu_j}$, $\mu_j^{-1}$, and
1363: $\overline{\mu}_j^{-1}$ for the complex $\mu_j$. The Stable/Unstable Manifold Theorem guarantees we will get two
1364: $n$-dimensional, transversal, flow-invariant sub-manifolds $\Lambda_+$ and $\Lambda_-$ such that $\exp tH_p$ is
1365: expanding on $\Lambda_+$ and contracting on $\Lambda_-$. Since the $\Lambda_\pm$ are invariant under the flow
1366: $\Phi_t = \exp tH_p$ which is symplectic, the symplectic form $\omega$ vanishes on the $\Lambda_\pm$, that is,
1367: the $\Lambda_\pm$ are Lagrangian submanifolds.
1368: \begin{lemma}
1369: \label{zero-cov}
1370: Assume $p$ is in the form of Proposition \ref{normal-prop-1}. Then there exists a local symplectic coordinate system $(t,\tau,x, \xi)$ near $\gamma$ such that
1371: $\Lambda_+ = \{ \tau = 0, \xi = 0 \}$ and $\Lambda_- = \{ \tau = 0, x = 0 \}$.
1372: \end{lemma}
1373: \begin{proof}
1374: We claim the $\Lambda_\pm$ are orientable and embedded
1375: in $T^* \SS^1 \times T^* \reals^{n-1}$. Since $dS(0)$ describes how the flow of $H_p$ has acted at time $t = 1$,
1376: we know the evolution of a tangent frame of $\Lambda_\pm$ will be
1377: described by $dS(0)$. Using the action variables in Proposition
1378: \ref{normal-prop-1}, we have
1379: \begin{eqnarray*}
1380: dS(0) = \left( \begin{array}{cc}
1381: A & 0 \\0 & B \end{array} \right)
1382: \end{eqnarray*}
1383: with
1384: \begin{eqnarray*}
1385: A = \diag( \mu_{1},
1386: \bar{\mu}_{1}, \ldots \mu_{n_{hc}}, \bar{\mu}_{n_{hc}} ; \mu_{2n_{hc}+1}, \ldots, \mu_{2n_{hc}+n_{hr}}),
1387: %\left( \begin{array}{cccccccc} \mu_1 & 0 & \multicolumn{5}{c}\dotfill
1388: % & 0 \\
1389: %0 & \ddots & 0 & \multicolumn{4}{c}\dotfill & 0 \\
1390: %0 & \dotfill & \mu_{n_{hr}} & 0 & \multicolumn{3}{c}\dotfill & 0 \\
1391: %0 & \multicolumn{2}{c}\dotfill & \mu_{n_{hr}+1} & 0 &
1392: %\multicolumn{2}{c}\dotfill & 0 \\
1393: %0 & \multicolumn{3}{c}\dotfill & \bar{\mu}_{n_{hr}+1} & 0 & \dotfill &
1394: %0 \\
1395: %0 & \multicolumn{4}{c}\dotfill & \ddots & 0 & \dotfill \\
1396: %0 & \multicolumn{5}{c}\dotfill & \mu_{n_{hc} + n_{hr}} & 0 \\
1397: %0 & \multicolumn{6}{c}\dotfill & \bar{\mu}_{n_{hc} + n_{hr}}
1398: %\end{array} \right)
1399: \end{eqnarray*}
1400: describing the time $1$ evolution of $\Lambda_+$ and $| \mu_j | >1$ for each $1
1401: \leq j \leq n_{hr}+n_{hc}$ by our choice of coordinates. Similarly,
1402: \begin{eqnarray*}
1403: B = \diag( \mu_{1}^{-1},
1404: \bar{\mu}_{1}^{-1}, \ldots \mu_{n_{hc}}^{-1}, \bar{\mu}_{n_{hc}}^{-1} ; \mu_{2n_{hc}+1}^{-1}, \ldots, \mu_{2n_{hc}+n_{hr}}^{-1} )
1405: \end{eqnarray*}
1406: describes the time $1$ evolution of $\Lambda_-$ with $|\mu_j^{-1}| <
1407: 1$ for each $j$. But we've assumed there are no negative real eigenvalues, so $\det A >0$ implies $\Lambda_+$ is
1408: orientable. Similarly, $\det B >0$ and $\Lambda_-$ is orientable. Now our assumptions on $p$ mean the flow has no critical
1409: points in a neighbourhood of $\gamma$ so the $\Lambda_\pm$ can have no self intersections and hence are embedded. \\
1410: \indent Let $\tilde{\Lambda} \subset T^*\SS^1 \times T^* \reals^{n-1}$, $\tilde{\Lambda} = \{ \tau = 0, \xi=0 \}$.
1411: Since $\Lambda_+$ is a closed, $n$-dimensional submanifold of $T^*X$, the tubular neighbourhood theorem guarantees
1412: there is a diffeomorphism $f$ (not necessarily symplectic) taking a neighbourhood $U$ of $\gamma$ into itself so that
1413: $f$ fixes $t$ and
1414: \begin{eqnarray*}
1415: f ( \Lambda_+ \cap U ) = \tilde{\Lambda} \cap U.
1416: \end{eqnarray*}
1417: Further, since $T_{\gamma(t)}\Lambda_+ = T_{\gamma(t)}\tilde{\Lambda}$ for $0 \leq t \leq 1$, we can choose $f$ satisfying
1418: \begin{eqnarray}
1419: \label{f-pb-nondeg}
1420: \left[ (f^{-1})^* \widetilde{\omega} \right]_{\gamma(t)}= \widetilde{\omega}_{\gamma(t)}, \,\,\, 0 \leq t \leq 1.
1421: \end{eqnarray}
1422: The statement in the lemma about $\Lambda_+$ now follows directly from the more general Theorem 4.1 in \cite{Wei2}, but
1423: we include a proof of this concrete case. We have $\tilde{\Lambda} \subset T^*\SS^1 \times T^* \reals^{n-1}$, a
1424: Lagrangian submanifold with two distinct symplectic structures, $\omega_0 = (f^{-1})^* \widetilde{\omega}$ and the
1425: standard symplectic structure $\omega_1$ inherited from $T^*\SS^1 \times T^* \reals^{n-1}$. We want to find a
1426: diffeomorphism $g: U \to U$ such that $g( \tilde{\Lambda}) = \tilde{\Lambda}$ and $g^* \omega_1 = \omega_0$. \\
1427: \indent Set $\omega_s = s \omega_0 + (1 - s) \omega_1$. We have $d \omega_s = 0$ and
1428: $\left. \omega_s \right|_{\tilde{\Lambda}} = 0$. Note \eqref{f-pb-nondeg} implies $\omega_s$ is nondegenerate
1429: in a neighbourhood of $\gamma$ for $0 \leq s \leq 1$. Let $\widehat{\omega}_s: TX \to T^*X$ denote
1430: the isomorphism generated by $\omega_s$, $\widehat{\omega}_s : Z \mapsto Z \contraction \omega_s$. We use the general
1431: Poincar\'{e} Lemma \ref{poincare-lemma} to obtain a $1$-form $\phi = I(\omega_0 - \omega_1)$ so that
1432: $d\phi = \omega_0 - \omega_1$ and set $Y_s = \widehat{\omega}_s^{-1}( \phi )$. Then
1433: $\left. \phi \right|_{\tilde{\Lambda}} = 0$ implies
1434: \begin{eqnarray*}
1435: Y_s \contraction \omega_s & = & \widehat{\omega} ( Y_s ) \\
1436: & = & \phi,
1437: \end{eqnarray*}
1438: so that $Y_s$ is tangent to $\tilde{\Lambda}$. Thus if $g_s = \exp (s Y_s)$ for $0 \leq s \leq 1$ is the
1439: integral of $Y_s$, $g_s(\tilde{\Lambda}) = \tilde{\Lambda}$. We calculate:
1440: \begin{eqnarray*}
1441: \frac{d}{dr} \left. \left( g_r^* \omega_r \right) \right|_{r = s} & = & g_s^* \left.
1442: \left( \frac{d}{dr} \omega_r \right) \right|_{r = s} + g_s^* \left( d ( Y_s \contraction \omega_s)\right) \\
1443: & = & g_s^* \left( \omega_0 - \omega_1 + d( -\phi) \right) \\
1444: & = & 0.
1445: \end{eqnarray*}
1446: Setting $g = g_1$ gives $g^* \omega_1 = \omega_0$ as desired. Now taking $g^{-1} \circ f$ gives a
1447: diffeomorphism of a neighbourhood of $\gamma$ taking $\Lambda_+$ to $\tilde{\Lambda}$ such that
1448: $g^* \circ (f^{-1})^* \widetilde{\omega} = \widetilde{\omega}$. \\
1449: \indent After this change of coordinates, we still need to put $\Lambda_-$ in the desired form.
1450: Since $\Lambda_-$ is transversal to $\Lambda_+$ and all of our transformations so far leave
1451: $\{ \tau = 0\}$ invariant, we can write $\Lambda_-$ as a graph over $\{x=0\}$:
1452: \begin{eqnarray}
1453: \label{lambda-minus-1}
1454: \Lambda_- = \left\{ ( t, 0, x, \xi): x = g(\xi, t) \right\}.
1455: \end{eqnarray}
1456: Further, since for each fixed $t$, \eqref{lambda-minus-1} is Lagrangian and the first de Rham cohomology
1457: group $H^1_{dR}(\{\tau = 0,x=0\}) \simeq H^1_{dR}( \reals^{n-1})$ vanishes, it is classical that we can
1458: write $g(\xi, t) = \partial_\xi h(\xi, t)$ for a smooth $h(\xi, t)$ (see, for example, \cite{Lee}). Then we write
1459: \begin{eqnarray*}
1460: \Lambda_- = \left\{ ( t, 0, x, \xi): x = \partial_\xi h(\xi, t) \right\},
1461: \end{eqnarray*}
1462: and observe $h$ must satisfy $\partial_\xi h(0,t) = 0$. This determines $h$ up to a constant, which we take
1463: to be $0$ so that $h(0,t) = 0$. Now let $b(\xi, t)$ be a smooth function satisfying $b(\xi, t) =
1464: \partial_t h(\xi, t)$, and note $b(0,t) = 0$. Then we perform the following change of variables:
1465: \begin{eqnarray*}
1466: \left\{ \begin{array}{rcl}
1467: t' & = & t \\
1468: \tau' & =& \tau + b(\xi,t) \\
1469: x' & = & x - \partial_\xi h(\xi, t) \\
1470: \xi' & = & \xi . \end{array} \right.
1471: \end{eqnarray*}
1472: We calculate:
1473: \begin{eqnarray*}
1474: d \tau' \wedge dt' + d \xi' \wedge dx' & = & \left(d \tau + \sum_j \partial_{\xi_j} b(\xi, t) d\xi_j +
1475: \partial_t b(\xi, t) dt \right) \wedge dt \\
1476: && + \sum_j d \xi_j \wedge \left(dx_j - \sum_i \partial_{\xi_i} \partial_{\xi_j} h(\xi ,t) d\xi_i -
1477: \partial_t \partial_{\xi_j} h(\xi, t) dt\right) \\
1478: & = & d \tau \wedge d t + d \xi \wedge dx,
1479: \end{eqnarray*}
1480: by the symmetry of the Hessian $\partial_{\xi_i} \partial_{\xi_j} h(\xi, t)$. Thus this change of
1481: variables is symplectic and the Lemma is proved.
1482: \end{proof}
1483:
1484: Using the change of variables in Lemma \ref{zero-cov}, we have the following proposition.
1485: \begin{proposition}
1486: \label{normal-prop-2}
1487: Let $p \in \Ci ( T^*X)$, $\gamma \subset \{p = 0 \}$ as above, with
1488: the Poincar\'{e} map having distinct eigenvalues $\mu_j$ not on the
1489: unit circle.
1490: Then there exists a neighbourhood, $U$, of $\gamma$ in $T^*X$, a smooth
1491: positive function $b\geq C^{-1} >0$ defined in $U$,
1492: a symplectomorphism
1493: $\kappa : U \to \kappa(U) \subset T^*\SS_{(t, \tau)}^1 \times T^* \reals_{(x, \xi)}^{n-1}$,
1494: and a smooth, $n \times n$-matrix valued function $B_t$ such that
1495: \begin{eqnarray}
1496: \kappa(\gamma) & = & \{ (t,0;0,0) : t \in \SS^1 \}, \quad
1497: \mathrm{and} \,\,\, b(t, \tau, x, \xi) p = \kappa^*g, \,\,\,
1498: \mathrm{with} \nonumber \\
1499: g(t, \tau; x, \xi) & = &\tau + \langle B_t(x, \xi) x, \xi \rangle \label{normal-form-2},
1500: \end{eqnarray}
1501: with $B_t$ satisfying
1502: \begin{eqnarray}
1503: \lefteqn{\left\langle B_t(0,0)x, \xi \right\rangle = } \nonumber \\
1504: & = & \sum_{j=1}^{n_{hc}}\left( \Re \lambda_j \left( x_{2j-1} \xi_{2j-1} + x_{2j} \xi_{2j} \right)
1505: - \Im \lambda_j \left( x_{2j-1} \xi_{2j} - x_{2j}\xi_{2j-1} \right) \right) \label{B-condition.a} \\
1506: & & + \sum_{j = 2n_{hc}+1}^{2n_{hc} + n_{hr}} \lambda_j x_j \xi_j \label{B-condition.b}.
1507: \end{eqnarray}
1508: Here $\lambda_j = \log (\mu_j)$ for $|\mu_j|>1$ and $\Im \lambda_j
1509: \geq 0$.
1510: \end{proposition}
1511: \begin{proof}
1512: Recall that the Poincar\'{e} map $S$ is linear in lowest order, and let $dS(0)$ be the linearized map.
1513: Let $q_0$ satisfy $dS(0) = \exp H_{q_0}$. After a linear symplectic change of variables, $q_0$ can be
1514: written in block-diagonal form
1515: \begin{eqnarray*}
1516: q_0(x, \xi) & =& \langle bx ,\xi\rangle \\
1517: & = & \sum_{j=1}^{n_{hc}}\left( \Re \lambda_j \left( x_{2j-1} \xi_{2j-1} + x_{2j} \xi_{2j} \right)
1518: - \Im \lambda_j \left( x_{2j-1} \xi_{2j} - x_{2j}\xi_{2j-1} \right) \right) \\
1519: & & + \sum_{j = 2n_{hc}+1}^{2n_{hc} + n_{hr}} \lambda_j x_j \xi_j , \quad \mathrm{with} \,\,\, 2n_{hc} + n_{hr} = 2n-2.
1520: \end{eqnarray*}
1521: According to Lemma \ref{zero-cov}, we may symplectically change
1522: variables so $\Lambda_+ = \{\tau=0, \xi=0 \}$ and $\Lambda_- = \{ \tau=0, x=0\}$.
1523: The linearization of the Hamiltonian vector field of $p$ is $H_{q_0}$, which implies we have a quadratic
1524: form as in the proposition.
1525: \end{proof}
1526:
1527: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1528:
1529: \section{Proof of Theorem \ref{main-theorem-1} }
1530: \label{main-theorem-1-proof}
1531: \numberwithin{equation}{section}
1532:
1533:
1534:
1535:
1536: \begin{proof}[Proof of Theorem \ref{main-theorem-1} with Distinct Eigenvalues]
1537: First we assume $P(h)$ has principal symbol given by
1538: \begin{eqnarray}
1539: \label{assume-p}
1540: p(t, \tau; x, \xi) = \tau +\langle B_t(x, \xi) x, \xi \rangle,
1541: \end{eqnarray}
1542: with $B_t$ satisfying (\ref{B-condition.a}-\ref{B-condition.b}) as in Proposition \ref{normal-prop-2}. Let $U$ be
1543: a neighbourhood of $\gamma$, $U \subset T^* \SS^1 \times T^* \reals^{n-1}$, and assume
1544: \begin{eqnarray*}
1545: U \subset U_{\epsilon/2} := \left\{ (t, \tau, x, \xi): \left( d(x, x(\gamma(t)))^2 + |\xi - \xi(\gamma(t))|^2
1546: + \tau^2 \right)^\half < \frac{\epsilon}{2} \right\}
1547: \end{eqnarray*}
1548: for $\epsilon >0$. Let $\psi_0$ be a microlocal cutoff function to a neighbourhood of $U$, that is, take
1549: $\psi_0 \in \Ci_c(\reals^{2n})$, $\psi_0 \equiv 1$ on $U_{\epsilon/2}$ with support in $U_{ \epsilon}$.
1550: Then we assume throughout that we are working in $U_\epsilon$. With $\tilde{h}$ small (fixed later
1551: in the proof), we do the following rescaling:
1552: \begin{eqnarray}
1553: \label{rescaling1}
1554: X : = \hsc^\half x, \quad \Xi = \hsc^\half \xi.
1555: \end{eqnarray}
1556: and assume for the remainder of the proof that $|(X, \Xi)| \leq \hsc^\half \epsilon$. We use the unitary operator
1557: $T_{h, \tilde{h}}$ defined in \eqref{U-hsc} to introduce the second parameter into $P(h)$. Following \cite{BZ} we define the
1558: operator $\widetilde{P}(h)$ by
1559: \begin{eqnarray*}
1560: \widetilde{P}(h) = T_{h, \tilde{h}} P(h) T_{h, \tilde{h}}^{-1},
1561: \end{eqnarray*}
1562: \noindent so that the principal symbol of $\widetilde{P}(h)$ is
1563: \begin{eqnarray}
1564: \label{tildep-expression}
1565: \lefteqn{\widetilde{p}(t, \tau; X, \Xi) = } \\
1566: && = \tau + \left\langle B_t\left(
1567: \csh^{\half}( X, \Xi) \right) \csh^\half X, \csh^\half \Xi \right\rangle, \nonumber
1568: \end{eqnarray}
1569: \noindent and $\widetilde{p} \in \s_{-\half}^{-\infty,0,0}$ microlocally. We have
1570: \begin{eqnarray}
1571: \label{tildep-est}
1572: \left|\partial_{X, \Xi}^\alpha \widetilde{p}\right| \leq C_\alpha\csh^{|\alpha|/2}
1573: \end{eqnarray}
1574: for $(X, \Xi) \in U_{\hsc^\half \epsilon}$ by Lemma \ref{U-hsc-lemma}. \\
1575: \indent We will use the following escape function, which we define in the $(X, \Xi)$ coordinates:
1576: \begin{eqnarray*}
1577: G(X, \Xi) : = \frac{1}{2} \left( \log ( 1 + |X|^2) - \log( 1 + |\Xi |^2 ) \right).
1578: \end{eqnarray*}
1579: $G$ satisfies
1580: \begin{eqnarray*}
1581: \left| \partial_X^\alpha \partial_\Xi^\beta G(X, \Xi) \right| \leq C_{\alpha \beta} \langle X \rangle^{-|\alpha|}\langle
1582: \Xi \rangle^{-|\beta|}, \,\,\,\, \,\,\, \mathrm{for} \,\,\, (\alpha, \beta) \neq (0,0),
1583: \end{eqnarray*}
1584: and since $\langle X \rangle^2 \langle \Xi \rangle^{-2} $ is an order function, $G$ satisfies the assumptions of
1585: Lemma \ref{etG-lemma} so we may construct the family $e^{sG^w}$ for sufficiently small $s$. \\
1586: \indent Now for $|(X, \Xi)| \leq \hsc^{\half} \epsilon$ we have
1587: \begin{eqnarray}
1588: \lefteqn{ H_{\widetilde{p}} G(X, \Xi) = } \nonumber \\
1589: & = & \csh\left[ \left\langle B_t X, \frac{\partial}{\partial X} \right\rangle
1590: - \left\langle B_t \frac{\partial}{\partial \Xi}, \Xi \right\rangle \right] G(X, \Xi) \label{hpg-main-term} \\
1591: & & \quad + \csh^{\frac{3}{2}}\left[ \sum_{j = 1}^{n-1} \left\langle \frac{\partial}{\partial \Xi_j}B_t( \cdot, \cdot )
1592: X, \Xi \right\rangle \frac{\partial}{\partial X_j}G(X, \Xi)\right] \label{hpg-err1} \\
1593: & & \quad - \csh^{\frac{3}{2}}\left[ \sum_{j = 1}^{n-1} \left\langle \frac{\partial}{\partial X_j} B_t( \cdot,\cdot ) X,
1594: \Xi \right\rangle \frac{\partial}{\partial \Xi_j} G(X, \Xi)\right]. \label{hpg-err2}
1595: \end{eqnarray}
1596: \indent For $s$ sufficiently small, we define a family of operators
1597: \begin{eqnarray}
1598: \label{Pth1}
1599: \widetilde{P}_s(h) &=& e^{-s G^w} \widetilde{P}(h) \Op_{\tilde{h}}^w \left(\psi_0\left(\csh^{\half} \bullet \right) \right)
1600: e^{sG^w} \nonumber \\
1601: & = & \exp \left(-s \ad_{G^w}\right) \widetilde{P}(h) \Op_{\tilde{h}}^w \left( \psi_0 \left(\csh^{\half} \bullet \right) \right),
1602: \end{eqnarray}
1603: \noindent where $\Op_{\tilde{h}}^w$ and $G^w$ are quantizations in the $\tilde{h}$-Weyl calculus. Now owing to
1604: Lemma \ref{2-param-lemma} and
1605: \eqref{tildep-est} we have microlocally to leading order in $h$:
1606: \begin{eqnarray*}
1607: \ad_{G^w}^k \left( \widetilde{P}(h)
1608: \Op_{\tilde{h}}^w \left(\psi_0\left(\csh^{\half} \bullet
1609: \right)\right) \right) = \O_{L^2 \to L^2}\left( h \tilde{h}^{k-1} \right),
1610: \end{eqnarray*}
1611: and in particular,
1612: \begin{eqnarray}
1613: \label{PG-comm}
1614: \left[ \widetilde{P}(h), G^w \right] = -i \tilde{h} \Op_{\tilde{h}}^w \left( H_{\widetilde{p}} G\right)
1615: + \O (h^{3/2} \tilde{h}^{3/2}).
1616: \end{eqnarray}
1617: \indent Now near $(0,0)$, $B_t$ is positive definite, $\langle B_t X,X \rangle \geq C^{-1} |X|^2$, so
1618: \begin{eqnarray*}
1619: \left\langle B_t X,X \right\rangle^{-1} \leq C|X|^{-2}.
1620: \end{eqnarray*}
1621: Applying this to the errors (\ref{hpg-err1}-\ref{hpg-err2}) we get
1622: \begin{eqnarray*}
1623: \csh^{\frac{3}{2}}\left[ \sum_{j = 1}^{n-1} \left\langle \frac{\partial}{\partial \Xi_j}B_t( \cdot, \cdot ) X,
1624: \Xi \right\rangle \frac{\partial}{\partial X_j}G(X, \Xi)\right] = \csh^{\frac{3}{2}}\frac{|X|^2}{1 + |X|^2}\O(|\Xi|),
1625: \end{eqnarray*}
1626: and similarly for \eqref{hpg-err2}. Adding these to \eqref{hpg-main-term}, we get
1627: \begin{eqnarray}
1628: H_{\widetilde{p}}G & = & \csh \left[ \frac{\langle B_t X, X \rangle}{1 + |X|^2}\right]\left( 1 + \csh^{\frac{1}{2}}
1629: \O ( |\Xi|) \right) \label{hpg-err-2-1}\\
1630: & & \quad +\csh \left[\frac{\langle B_t \Xi, \Xi \rangle}{1 + |\Xi|^2} \right]\left( 1 +\csh^{\frac{1}{2}}
1631: \O ( |X|)\right). \label{hpg-err-2-2}
1632: \end{eqnarray}
1633: Now we expand $B_t$ in a Taylor series about $(0,0)$ to get
1634: \begin{eqnarray*}
1635: \lefteqn{H_{\widetilde{p}}G = } \\
1636: & = & \csh \left[ \frac{\langle B_t(0,0) X, X \rangle}{1 + |X|^2} + \csh^{\half} \frac{|X|^2}{1+|X|^2}
1637: \O(|(X, \Xi)|)\right] \cdot \\
1638: && \quad \quad \quad \cdot \left( 1 + \csh^{\half}\O ( |\Xi|) \right) \\
1639: & & +\csh \left[\frac{\langle B_t(0,0) \Xi, \Xi \rangle}{1 + |\Xi|^2} + \csh^{\half} \frac{|\Xi|^2}{1+|\Xi|^2}
1640: \O(|(X, \Xi)|)\right]\cdot \\
1641: && \quad \quad \quad \cdot \left( 1 +\csh^{\half} \O ( |X|)\right),
1642: \end{eqnarray*}
1643: which can again be written as (\ref{hpg-err-2-1}-\ref{hpg-err-2-2}). Recalling that $B_t(0,0)$ is block
1644: diagonal of the form (\ref{B-condition.a}-\ref{B-condition.b}), we get for $|(X, \Xi)| \leq \hsc^{\half} \epsilon$,
1645: \begin{eqnarray}
1646: \lefteqn{ H_{\widetilde{p}} G(X, \Xi) = } \nonumber \\
1647: & = & \left[ \sum_{j=1}^{n_{hc}} \Re \lambda_j \left( \frac{ X_{2j}^2 + X_{2j-1}^2 }{1 + |X|^2}
1648: + \frac{ \Xi_{2j}^2 + \Xi_{2j-1}^2 }{1 + |\Xi|^2} \right)\right] \left( 1 + \tilde{h}^{-\half} \O ( \epsilon )\right)
1649: \label{HpG-2.a} \\
1650: & & + \left[ \sum_{j = 2n_{hc}+1}^{2n_{hc}+n_{hr}} \lambda_j \left( \frac{X_j^2}{1 + |X|^2}
1651: + \frac{\Xi_j^2}{1+ |\Xi|^2} \right)\right] \left( 1 + \tilde{h}^{-\half} \O ( \epsilon) \right) . \label{HpG-2.b}
1652: \end{eqnarray}
1653: Thus
1654: \begin{eqnarray}
1655: \label{Pth2}
1656: \widetilde{P}_s(h) = \widetilde{P}(h) - ish (A( 1 + E_0))^w + s E_1^w + s^2 E_2^w ,
1657: \end{eqnarray}
1658: \noindent with $E_0 = \O(\tilde{h}^{-\half} \epsilon)$, $E_1=\O
1659: (h^{3/2} \tilde{h}^{3/2})$, $E_2 = \O(h \tilde{h})$, and
1660: $A^w = \Op_{\tilde{h}}^w(A)$ for
1661: \begin{eqnarray}
1662: \lefteqn{ A(X, \Xi) := } \nonumber \\
1663: & = & \sum_{j=1}^{n_{hc}} \Re \lambda_j \left( \frac{ X_{2j}^2 + X_{2j-1}^2 }{1 + |X|^2}
1664: + \frac{ \Xi_{2j}^2 + \Xi_{2j-1}^2 }{1 + |\Xi|^2} \right) \label{A-1-hc} \\
1665: & & + \sum_{j=2n_{hc}+1}^{2n_{hc} + n_{hr}} \lambda_j \left( \frac{X_j^2}{1+|X|^2}
1666: + \frac{\Xi_j^2}{1+ |\Xi|^2} \right). \label{A-1-hr}
1667: \end{eqnarray}
1668:
1669: We claim that for $\tilde{h}$ sufficiently small,
1670: \begin{eqnarray}
1671: \label{A-lower-bound}
1672: \langle A^wU, U \rangle \geq \frac{\tilde{h}}{C} \|U \|^2
1673: \end{eqnarray}
1674: for some constant $C>0$, which is essentially the lower bound for the harmonic oscillator
1675: $\tilde{h}^2 D_X^2 + X^2$. Clearly it suffices to prove this inequality for individual $j$ for
1676: the real hyperbolic terms (\ref{A-1-hr}), and in pairs for the complex
1677: hyperbolic terms (\ref{A-1-hc}), which is the content of Lemma
1678: \ref{harm-osc}.
1679:
1680:
1681:
1682: Now fix $\tilde{h}>0$ and $|s|>0$ sufficiently small so that the
1683: estimate \eqref{A-lower-bound} holds and the errors $E_1$ and $E_2$ satisfy
1684: \begin{eqnarray*}
1685: \|shA^w U \|_{L^2} \gg \| s E_1^w U \|_{L^2} + \|s^2 E_2^w U \|_{L^2},
1686: \end{eqnarray*}
1687: and fix $\epsilon>0$ sufficiently small that the error $|E_0| \ll 1$,
1688: independent of $h>0$.
1689:
1690: We now have for smooth $U$ satisfying $\Op_{\tilde{h}}^w(\psi_0(h^{\half} \bullet)) U = U + \O(h^\infty)$,
1691: \begin{eqnarray}
1692: \label{hsc-est}
1693: -\Im \langle \widetilde{P}_s(h) U, U \rangle \geq \frac{h \tilde{h}}{C} \| U \|^2.
1694: \end{eqnarray}
1695: Now define the operator $K_h^w$ by $e^{sK_h^w} := T_{h, \tilde{h}}^{-1}
1696: e^{sG_{\tilde{h}}^w}T_{h, \tilde{h}}$. Translating back
1697: into original coordinates, and with $z \in [-1,1] + i(-c_0h + \infty)$
1698: for sufficiently small $c_0>0$, (\ref{hsc-est}) gives
1699: \begin{eqnarray*}
1700: -\Im \left\langle e^{sK_h^w} \left( P(h) - z \right) e^{-sK_h^w}u, u \right\rangle \geq \frac{h}{C_1} \| u \|^2.
1701: \end{eqnarray*}
1702: Finally, since $\|\exp ( \pm sK_h^w ) \| = \O (h^{-N})$ for some $N$, the theorem follows in the case where $p$ is of
1703: the form \eqref{assume-p}. \\
1704: \indent For general $p$, by Proposition \ref{normal-prop-2}, there is a symplectomorphism $\kappa$ so that up to an
1705: elliptic factor, $\kappa^*p$
1706: is of the form \eqref{assume-p}. Using Theorem \ref{gamma-egorov} to quantize $\kappa$ as an $h$-FIO
1707: $F$, we get
1708: \begin{eqnarray*}
1709: \Op_h^w \left( \kappa^* p + E_1 \right) = F^{-1} P(h) F ,
1710: \end{eqnarray*}
1711: where $E_1 = \O(h^2)$ is the error arising from Theorem
1712: \ref{gamma-egorov}. We may then
1713: use the previous argument for $\kappa^*p$ getting an additional error of $\O(h^2)$ from Theorem \ref{gamma-egorov} in
1714: \eqref{hsc-est}, which is the same order as $E_1$.
1715: \end{proof}
1716:
1717: \begin{remark}
1718: The error arising at the end of the proof of Theorem
1719: \ref{main-theorem-1} from the use of Theorem \ref{gamma-egorov} is of
1720: order $\O(h^2)$ and hence negligible compared to our lower bound of
1721: $h$ for $A$. However, the estimate of $A$ is used for the imaginary
1722: part of $\widetilde{P}_s$, and the error in Theorem \ref{gamma-egorov}
1723: is real, so $\O(h)$ would
1724: have been sufficient.
1725: \end{remark}
1726:
1727: \begin{lemma}
1728: \label{harm-osc}
1729: Let
1730: \begin{eqnarray*}
1731: a_0(y, \eta):= \frac{y_j^2}{ \langle y \rangle^2} + \frac{\eta_j^2}{ \langle \eta \rangle^2},
1732: \end{eqnarray*}
1733: for $(y, \eta) \in \reals^{2n-2}$, and $\langle y \rangle = ( 1 +
1734: |y|^2)^{1/2}$, and let
1735: \begin{eqnarray*}
1736: a_1(y, \eta) := \frac{y_{2j}^2 + y_{2j-1}^2}{\langle y \rangle^2} +
1737: \frac{\eta_{2j}^2 + \eta_{2j-1}^2}{\langle \eta \rangle^2}.
1738: \end{eqnarray*}
1739: Then $a_i$, $i=0,1$ satisfies
1740: \begin{eqnarray}
1741: \label{ai-ineq}
1742: \langle \Op_{\tilde{h}}^w(a_i)U, U \rangle \geq \frac{\tilde{h}}{C} \|U \|^2
1743: \end{eqnarray}
1744: for $\tilde{h}>0$ sufficiently small and a constant $0<C<\infty$.
1745: \end{lemma}
1746: \begin{proof}
1747: The idea of the proof is that $a_i$ is essentially the harmonic
1748: oscillator which satisfies the inequality \eqref{ai-ineq}. We write each $a_i$ as a $a_i = |b|^2$ for $b$ a complex
1749: symbol. Observe $a_0(y, \eta) = |b(y, \eta)|^2$ with
1750: \begin{eqnarray*}
1751: b(y, \eta): = \frac{y_j}{\langle y \rangle} + i \frac{\eta_j}{\langle \eta \rangle}.
1752: \end{eqnarray*}
1753: Thus, using the $\tilde{h}$-Weyl calculus,
1754: \begin{eqnarray}
1755: \label{A_0}
1756: a_0^w(y, \tilde{h}D_y) = b^w(y, \tilde{h}D_y)^* b^w(y, \tilde{h} D_y) + c^w(y, \tilde{h} D_y),
1757: \end{eqnarray}
1758: where
1759: \begin{eqnarray}
1760: c(y, \eta) & = & \tilde{h} \left\{ \frac{\eta_j}{\langle \eta \rangle},
1761: \frac{y_j}{\langle y \rangle}\right\}
1762: + \O ( \tilde{h}^2) \nonumber \\
1763: & = & \tilde{h} \langle y \rangle^{-3} \langle \eta \rangle^{-3}\left( 1 + \O(|y|^2 + |\eta|^2) \right)
1764: + \O( |y|^2|\eta|^2) + \O ( \tilde{h}^2) \label{C}.
1765: \end{eqnarray}
1766: For $(y, \eta)$ small, $c$ is bounded from below by $\tilde{h}$ as in
1767: (\ref{A-lower-bound}), and for large $(y, \eta)$ we have
1768: \be
1769: C^{-1} \leq a_0 \leq C
1770: \ee
1771: for some constant $C>0$. Hence for large $(y, \eta)$, \eqref{A_0} is
1772: bounded from below independent of $\tilde{h}$. Observe $a_1(y, \eta)
1773: = |b_{2j}(y, \eta)|^2 + |b_{2j-1}(y, \eta)|^2$ for
1774: \begin{eqnarray*}
1775: b_k(y, \eta) = \frac{y_k}{\langle y \rangle } -i \frac{\eta_k}{\langle \eta \rangle },
1776: \end{eqnarray*}
1777: and the same argument applies to $a_1$ as to $a_0$.
1778: \end{proof}
1779:
1780:
1781:
1782:
1783: \begin{remark}
1784: It is interesting to note that the estimate (\ref{main-theorem-1-est}) depends only on the real parts of the eigenvalues
1785: $\lambda_j$ above. Unraveling the definitions, the eigenvalues $\lambda_j$ are logarithms of the eigenvalues of the
1786: linearized Poincar\'{e} map $dS(0)$ from above. Then (\ref{main-theorem-1-est}) depends only on the modulis of the
1787: eigenvalues of $dS(0)$.
1788: \end{remark}
1789:
1790:
1791:
1792:
1793: \numberwithin{equation}{subsection}
1794: \subsection{A Return to Quadratic Forms}
1795: Recall the only place we have used that the eigenvalues are distinct
1796: is in determining the possible form of the quadratic form $q(\rho)$
1797: defined by $dS(0) = \exp H_q$. We then considered the Hamilton, or
1798: Fundamental matrix $B$ defined by
1799: \begin{eqnarray}
1800: \label{qB}
1801: q(\rho) = : \half \tom (\rho, B \rho).
1802: \end{eqnarray}
1803: We follow \cite{hormander5} and return to the setup for Proposition
1804: \ref{normal-prop-1}. All of the following changes of variables will
1805: be linear, so we may assume we are working in $\reals^{2n-2}$ and
1806: choose local symplectic coordinates in
1807: which $\tom$ is the standard symplectic form
1808: \begin{eqnarray*}
1809: \tom = \sum_{j=1}^{n-1} d\xi_j \wedge dx_j.
1810: \end{eqnarray*}
1811: Then we can write \eqref{qB} in a more easily manipulated form:
1812: \begin{eqnarray*}
1813: q(\rho) = : \half \langle \rho , JB \rho \rangle
1814: \end{eqnarray*}
1815: where $J$ is the matrix of symplectic structure on $\reals^{2n}$,
1816: \begin{eqnarray*}
1817: J = \left( \begin{array}{cc} 0 & -I \\ I & 0 \end{array} \right).
1818: \end{eqnarray*}
1819: As mentioned previously, the eigenvalues of $B$ are the logarithms of
1820: the eigenvalues of $dS(0)$ (with a suitably chosen branch cut), hence
1821: have nonzero real part, and come in pairs $\lambda, - \lambda$ for the
1822: positive real
1823: hyperbolic eigenvalues, and $4$-tuples $\lambda, - \lambda,
1824: \bar{\lambda}, - \bar{\lambda}$ for the complex hyperbolic. If we allow $\rho$ to be complex for the moment, and
1825: denote by $V_\lambda$ the generalized eigenspace for $\lambda$ real or
1826: complex, we see
1827: \begin{eqnarray*}
1828: \tom (V_\lambda, V_{\lambda'}) = 0
1829: \end{eqnarray*}
1830: unless $\lambda + \lambda' = 0$. We then consider the spaces
1831: $V_\lambda \oplus V_{-\lambda}$, which is symplectic with the
1832: restricted symplectic form $\left. \tom \right|_{V_\lambda \oplus
1833: V_{-\lambda}}$, since $\lambda \neq 0$. As in \S
1834: \ref{symplectic-geometry} we choose the pairs and $4$-tuples of
1835: eigenvalues so that $\Re \lambda >0$ and $\Im \lambda \geq 0$. We thus have a
1836: decomposition of $\reals^{2n-2}$ into symplectic subspaces
1837: \begin{eqnarray*}
1838: \reals^{2n-2} = \left( \bigoplus_{j = 1}^{n_{hc}} V_{\lambda_j}
1839: \oplus V_{-\lambda_j} \oplus V_{\bar{\lambda}_j}
1840: \oplus V_{-\bar{\lambda}_j} \right) \bigoplus \left( \bigoplus_{j =
1841: n_{hc}+1}^{n_{hc} + n_{hr}} V_{\lambda_j}
1842: \oplus V_{-\lambda_j}\right)
1843: \end{eqnarray*}
1844: where $n_{hr}$ is the number of real eigenvalues with $\lambda>0$ and $n_{hc}$ is the
1845: number of complex eigenvalues with $\Re \lambda >0, \Im \lambda >0$. Our notation here means if $\lambda_j$
1846: has multiplicity $k_j$, then
1847: \begin{eqnarray*}
1848: \sum_{j=1}^{n_{hc}} 4k_j + \sum_{j=n_{hc} + 1}^{n_{hc} + n_{hr}} 4 k_j = 2n-2.
1849: \end{eqnarray*}
1850: Fix $\lambda$ real or complex, $\Re \lambda >0$, $\Im \lambda \geq 0$,
1851: with multiplicity
1852: greater than $1$ and consider the
1853: complex symplectic subspace $V_\lambda
1854: \oplus V_{-\lambda}$. Assume $V_\lambda$ has dimension $m$. Note $B$
1855: restricts to a linear map in $V_\lambda$,
1856: $T:= \left. B \right|_{V_{\lambda}}$, such that $T - \lambda I$ is
1857: nilpotent. Our definitions equip $V_\lambda \oplus V_{-\lambda}$
1858: with a symplectic structure in which $V_{-\lambda}$ is dual and
1859: isomorphic to $V_{\lambda}$. We abuse notation and write a point
1860: $(x, \xi) \in V_\lambda \oplus V_{-\lambda}$. Then if we put $T$ into Jordan form in
1861: $V_\lambda$ so that $Tx = \lambda x + (x_2, x_3, \ldots, x_m, 0)$, we
1862: obtain a symplectic change of coordinates by writing
1863: \begin{eqnarray*}
1864: \left. B \right|_{V_\lambda \oplus V_{-\lambda}}(x, \xi) = (\lambda x
1865: + (x_2, \ldots, x_m, 0), -\lambda \xi - (0, \xi_1, \xi_2, \ldots,
1866: \xi_{m-1})),
1867: \end{eqnarray*}
1868: by the symplectic skew symmetry of $B$. In these coordinates we then have
1869: $q_\lambda$, the
1870: projection of $q$ onto $V_\lambda \oplus V_{-\lambda}$,
1871: \begin{eqnarray}
1872: \label{q-form-cx}
1873: q_\lambda (x,\xi) = \lambda \sum_{l=1}^k x_l \xi_l + \sum_{l=1}^{k-1}
1874: x_{l+1}\xi_l,
1875: \end{eqnarray}
1876: where $k$ is the multiplicity of $\lambda$. This is the normal form
1877: in complex variables, with the ``actions'' $\lambda x_j \xi_j$ as in
1878: \S \ref{symplectic-geometry}, but with the additional terms coming from the Jordan
1879: form. In order to understand the real normal form, there are two cases to
1880: examine.
1881:
1882: {\bf Case 1: $\lambda>0$ is real.} Then the space $V_\lambda \oplus
1883: V_{-\lambda}$ is real, the change of variables above is real, and we
1884: get $q_\lambda$ exactly as in \eqref{q-form-cx}. Let the real matrix
1885: $Q_\lambda$ be defined by the real normal form:
1886: \begin{eqnarray}
1887: \label{Q-matrix}
1888: q_\lambda(x,\xi) = : \half \langle (x, \xi), Q (x, \xi) \rangle.
1889: \end{eqnarray}
1890: Then $Q$ takes the special form
1891: \begin{eqnarray*}
1892: Q = \left( \begin{array}{cc} 0 & A \\ A^T & 0 \end{array} \right)
1893: \end{eqnarray*}
1894: where $A$ is the $k \times k$ matrix
1895: \begin{eqnarray}
1896: \label{A-1}
1897: A = \left( \begin{array}{cccc} \lambda & 0 &
1898: \multicolumn{2}{c}\dotfill \\ 1 & \lambda & 0 & \ldots \\
1899: 0 & \ddots & \ddots & 0 \\
1900: \vdots & \ldots & 1 & \lambda
1901: \end{array}
1902: \right)
1903: \end{eqnarray}
1904: and $A^T$ denotes the transpose of $A$.
1905:
1906: {\bf Case 2: $\lambda$ complex, $\Re \lambda >0$, $\Im \lambda >0$.} We use a similar
1907: change of variables to that in \S \ref{symplectic-geometry}. That is, let
1908: $\{e_l, f_l\}$ be the generalized eigenvectors for $\lambda, -\lambda$
1909: respectively. Here, $1 \leq l \leq k$ where $k$ is the multiplicity
1910: of $\lambda$. Then
1911: $\{e_l,f_l,\bar{e}_l, \bar{f}_l \}$ forms a basis for a complex vector space which is
1912: the complexification of a real symplectic vector space. We then
1913: consider the projection $q_\lambda$ of $q$ onto the space
1914: \begin{eqnarray*}
1915: W = V_\lambda
1916: \oplus V_{-\lambda} \oplus V_{\bar{\lambda}} \oplus
1917: V_{-\bar{\lambda}}.
1918: \end{eqnarray*}
1919: Write a point $\rho$ in $W$ as
1920: \begin{eqnarray*}
1921: \rho = \sum_{l=1}^k z_l e_l + \zeta_l f_l + w_l \bar{e}_l + \eta_l \bar{f}_l,
1922: \end{eqnarray*}
1923: so that
1924: \begin{eqnarray*}
1925: q_\lambda(\rho) = \lambda \sum_1^k z_l \zeta_l + \bar{\lambda}
1926: \sum_1^k w_l \eta_l + \sum_1^{k-1} z_{l+1} \zeta_l + \sum_1^{k-1}
1927: w_{l+1} \eta_l.
1928: \end{eqnarray*}
1929: We define as in \S \ref{symplectic-geometry} a real symplectic basis
1930: $\{e_l^1, e_l^2, f_l^1, f_l^2\}$ for $1 \leq l \leq k$ by
1931: \begin{eqnarray*}
1932: e_l = \frac{1}{\sqrt{2}}(e_l^1 + ie_l^2), \quad f_l= \frac{1}{\sqrt{2}} (f_l^1 -
1933: i f_l^2),
1934: \end{eqnarray*}
1935: and write in these new coordinates
1936: \begin{eqnarray*}
1937: \rho = \sum_{l=1}^k \sum_{r=1}^2 x_l^r e_l^r + \xi_l^r r_l^r.
1938: \end{eqnarray*}
1939: Then we get the real normal form of $q_\lambda$ in these coordinates:
1940: \begin{eqnarray*}
1941: q_\lambda(\rho) & = & \Re \lambda \sum_1^k \left( x_{2l-1} \xi_{2l-1} +
1942: x_{2l} \xi_{2l} \right) - \Im \lambda \sum_1^k \left( x_{2l}\xi_{2l-1}
1943: - x_{2l-1} \xi_{2l} \right) \\
1944: && \quad \quad + \sum_1^{k-1} \left( x_{2l+1}\xi_{2l-1} + x_{2l+2}
1945: \xi_{2l} \right).
1946: \end{eqnarray*}
1947: We again define the real matrix $Q$ in terms of the real quadratic
1948: normal form $q_\lambda$ by \eqref{Q-matrix}, which now takes the form
1949: \begin{eqnarray*}
1950: Q = \left( \begin{array}{cc} 0 & A \\ A^T & 0 \end{array} \right),
1951: \end{eqnarray*}
1952: where $A$ is the $2k \times 2k$ matrix
1953: \begin{eqnarray}
1954: \label{A-2}
1955: \left( \begin{array}{cccc}
1956: \Lambda & 0 &
1957: \multicolumn{2}{c}\dotfill \\ I & \Lambda & 0 & \ldots \\
1958: 0 & \ddots & \ddots & 0 \\
1959: \vdots & \ldots & I & \Lambda
1960: \end{array} \right),
1961: \end{eqnarray}
1962: with $I$ the $2 \times 2$ identity matrix and
1963: \begin{eqnarray*}
1964: \Lambda = \left( \begin{array}{cc}
1965: \Re \lambda & - \Im \lambda \\
1966: \Im \lambda & \Re \lambda
1967: \end{array} \right).
1968: \end{eqnarray*}
1969: Putting this discussion together with the proof of Proposition
1970: \ref{normal-prop-1}, we have proved the following:
1971: \begin{proposition}
1972: \label{normal-prop-3}
1973: Let $p \in \Ci ( T^*X)$, $\gamma \subset \{p = 0 \}$ as above, with
1974: the linearized Poincar\'{e} map $dS(0)$ having eigenvalues $\{\mu_j\}$ not on the unit
1975: circle, and suppose $\mu_j$ has multiplicity $k_j$. Then there exists a neighbourhood, $U$, of $\gamma$ in $T^*X$,
1976: a smooth positive function $b \geq C^{-1} >0$ defined in $U$, and a
1977: symplectomorphism $\kappa : U \to \kappa(U) \subset T^*\SS_{(t, \tau)}^1 \times T^* \reals_{(x, \xi)}^{n-1}$ such that
1978: \begin{eqnarray*}
1979: \kappa(\gamma) = \{ (t,0;0,0) : t \in \SS^1 \},
1980: \end{eqnarray*}
1981: and $b(t, \tau, x, \xi) p = \kappa^*(g + r)$, with
1982: \begin{eqnarray*}
1983: \lefteqn{ g(t, \tau; x, \xi) =} \nonumber \\
1984: & = & \tau + \sum_{j=1}^{n_{hc}} \sum_{l=1}^{k_j}\left( \Re \lambda_j \left( x_{2l-1} \xi_{2l-1} + x_{2l} \xi_{2l} \right)
1985: - \Im \lambda_j \left( x_{2l-1} \xi_{2l} - x_{2l}\xi_{2l-1} \right)
1986: \right) \\
1987: && \quad \quad + \sum_{j=1}^{n_{hc}} \sum_{l=1}^{k_j-1} \left( x_{2l+1}\xi_{2l-1} + x_{2l+2}
1988: \xi_{2l} \right) \\
1989: & & \quad \quad + \sum_{j = 2n_{hc}+1}^{2n_{hc} + n_{hr}} \left(\sum_{l=1}^{k_j}
1990: \lambda_j x_l \xi_l +
1991: \sum_{l=1}^{k_j - 1} x_{l+1}\xi_l\right),
1992: \end{eqnarray*}
1993: where $\lambda_j = \log \mu_j$ for each $j$ (with a suitable branch cut) and $r = \O (|x|^3 + |\xi|^3)$.
1994: \end{proposition}
1995:
1996: The proof of Lemma \ref{zero-cov} depends only on the moduli of the
1997: eigenvalues of $dS(0)$ restricted to the stable and unstable
1998: manifolds, hence does not depend on the multiplicities, or the Jordan
1999: form. Consequently we have the analogue of Proposition \ref{normal-prop-2}.
2000:
2001: \begin{proposition}
2002: \label{normal-prop-4}
2003: Under the assumptions of Proposition \ref{normal-prop-3}, there exists
2004: a neighbourhood, $U$, of $\gamma$ in $T^*X$, a smooth
2005: positive function $b\geq C^{-1} >0$ defined in $U$,
2006: a symplectomorphism
2007: $\kappa : U \to \kappa(U) \subset T^*\SS_{(t, \tau)}^1 \times T^* \reals_{(x, \xi)}^{n-1}$,
2008: and a smooth, $n \times n$-matrix valued function $B_t$ such that
2009: \begin{eqnarray*}
2010: \kappa(\gamma) & = & \{ (t,0;0,0) : t \in \SS^1 \}, \quad
2011: \mathrm{and} \,\,\, b(t, \tau, x, \xi) p = \kappa^*g, \,\,\,
2012: \mathrm{with} \\
2013: g(t, \tau; x, \xi) & = &\tau + \langle B_t(x, \xi) x, \xi \rangle,
2014: \end{eqnarray*}
2015: with $B_t$ satisfying
2016: \begin{eqnarray*}
2017: \lefteqn{\left\langle B_t(0,0)x, \xi \right\rangle = } \\
2018: & = & \sum_{j=1}^{n_{hc}} \sum_{l=1}^{k_j}\left( \Re \lambda_j \left( x_{2l-1} \xi_{2l-1} + x_{2l} \xi_{2l} \right)
2019: - \Im \lambda_j \left( x_{2l-1} \xi_{2l} - x_{2l}\xi_{2l-1} \right)
2020: \right) \\
2021: && \quad \quad + \sum_{j=1}^{n_{hc}} \sum_{l=1}^{k_j-1} \left( x_{2l+1}\xi_{2l-1} + x_{2l+2}
2022: \xi_{2l} \right) \\
2023: & & \quad \quad + \sum_{j = 2n_{hc}+1}^{2n_{hc} + n_{hr}} \left( \sum_{l=1}^{k_j}
2024: \lambda_j x_l \xi_l + \sum_{l=1}^{k_j - 1} x_{l+1}\xi_l \right).
2025: \end{eqnarray*}
2026: \end{proposition}
2027:
2028: \subsection{End of the Proof of Theorem \ref{main-theorem-1}}
2029: \label{non-distinct-section}
2030: Now we turn our attention to the proof of Theorem
2031: \ref{main-theorem-1} in the case of non-distinct eigenvalues
2032: of $dS(0)$. Recall the key feature to the proof of Theorem
2033: \ref{main-theorem-1} in the case of distinct eigenvalues was that the normal form given in Proposition
2034: \ref{normal-prop-2} has quadratic part $q(x, \xi)$ with the
2035: property that there exists another quadratic form
2036: \begin{eqnarray*}
2037: w(x, \xi) = \half \langle W(x, \xi), (x, \xi) \rangle
2038: \end{eqnarray*}
2039: such that $H_q w(x, \xi)$ is a positive definite quadratic form. Then
2040: we would like our escape function to be $G(x, \xi) = w(x, \xi)$,
2041: however for technical reasons we
2042: had to use a logarithmic escape function and form the families $e^{\pm
2043: s G^w}$. With the following theorem, the proof of Theorem
2044: \ref{main-theorem-1} is complete.
2045: \begin{theorem}
2046: \label{non-distinct-theorem}
2047: Suppose $q \in \Ci (\reals^{2m})$ is quadratic of the form
2048: \begin{eqnarray}
2049: \lefteqn{ q(x,\xi) = } \nonumber \\
2050: && =\sum_{j=1}^{n_{hc}} \sum_{l=1}^{k_j}\left( \Re \lambda_j \left( x_{2l-1} \xi_{2l-1} + x_{2l} \xi_{2l} \right)
2051: - \Im \lambda_j \left( x_{2l-1} \xi_{2l} - x_{2l}\xi_{2l-1} \right)
2052: \right) \label{p-1a}\\
2053: && \quad \quad + \sum_{j=1}^{n_{hc}} \sum_{l=1}^{k_j-1} \left( x_{2l+1}\xi_{2l-1} + x_{2l+2}
2054: \xi_{2l} \right) \label{p-1b}\\
2055: & & \quad \quad + \sum_{j = 2n_{hc}+1}^{2n_{hc} + n_{hr}} \left( \sum_{l=1}^{k_j}
2056: \lambda_j x_l \xi_l + \sum_{l=1}^{k_j - 1} x_{l+1}\xi_l \right), \label{p-1c}
2057: \end{eqnarray}
2058: and
2059: \begin{eqnarray*}
2060: G(x, \xi) = \half \left( \log (1 + |x|^2) - \log(1 + |\xi|^2 ) \right).
2061: \end{eqnarray*}
2062: Then there exist $m \times m$ nonsingular matrices $A$
2063: and $A'$, positive real numbers $0 < r_1 \leq r_2, \leq \cdots \leq
2064: r_{m}< \infty$, and symplectic coordinates $(x, \xi)$ such that
2065: \begin{eqnarray}
2066: \label{non-distinct-theorem-statement}
2067: H_q(G) = \frac{\sum_{j=1}^m r_j^{-2} x_j^2 }{1 + |A x|^2} +
2068: \frac{\sum_{j=1}^m r_j^{-2} \xi_j^2 }{1 + |A' \xi|^2}.
2069: \end{eqnarray}
2070: \end{theorem}
2071:
2072: \begin{proof}
2073: First, suppose
2074: \begin{eqnarray*}
2075: g(x, \xi) = \half \left\langle \tilde{g} \left( \begin{array}{c} x \\ \xi
2076: \end{array} \right), \left( \begin{array}{c} x \\ \xi
2077: \end{array} \right) \right\rangle
2078: \end{eqnarray*}
2079: is a real quadratic form with $\tilde{g}$ symmetric of the form
2080: \begin{eqnarray*}
2081: \tilde{g} = \left( \begin{array}{cc} P & 0 \\ 0 & -P \end{array} \right),
2082: \end{eqnarray*}
2083: where $P$ is symmetric and nonsingular. Then
2084: \begin{eqnarray*}
2085: \partial_{x} \half \log( 1 + \langle Px,x \rangle ) = \frac{Px}{1 +
2086: \langle Px,x \rangle},
2087: \end{eqnarray*}
2088: and similarly for $\xi$ so studying
2089: \begin{eqnarray*}
2090: H_q \left( \half \left( \log(1 + \langle Px,x \rangle) - \log(1 + \langle
2091: P\xi,\xi \rangle ) \right) \right)
2092: \end{eqnarray*}
2093: can be reduced to studying $H_q g(x, \xi)$, modulo the positive terms
2094: $1 + \langle P \cdot , \cdot \rangle$ in the denominator. If $q(x, \xi)$ is of the
2095: form (\ref{p-1a}-\ref{p-1c}), then we can write $q$ in terms of the
2096: fundamental matrix $B$:
2097: \begin{eqnarray*}
2098: q(x, \xi) = \left\langle
2099: \left( \begin{array}{c} x \\ \xi
2100: \end{array} \right), JB \left( \begin{array}{c} x \\ \xi
2101: \end{array} \right) \right\rangle,
2102: \end{eqnarray*}
2103: where
2104: \begin{eqnarray*}
2105: J = \left( \begin{array}{cc} 0 & -I \\ I & 0 \end{array} \right)
2106: \end{eqnarray*}
2107: as usual. Then the vector field $H_q$ can be written as
2108: \begin{eqnarray*}
2109: H_q = \left\langle B \left( \begin{array}{c} x \\ \xi
2110: \end{array} \right), \left( \begin{array}{c} \partial_x \\ \partial_\xi
2111: \end{array} \right) \right\rangle,
2112: \end{eqnarray*}
2113: and
2114: \begin{eqnarray*}
2115: H_qg & = & \left\langle B \left( \begin{array}{c} x \\ \xi
2116: \end{array} \right), \left( \begin{array}{c} \partial_x \\ \partial_\xi
2117: \end{array} \right) \right\rangle \left( \half \left\langle \tilde{g} \left( \begin{array}{c} x \\ \xi
2118: \end{array} \right), \left( \begin{array}{c} x \\ \xi
2119: \end{array} \right) \right\rangle \right) \\
2120: & = & \left\langle B \left( \begin{array}{c} x \\ \xi
2121: \end{array} \right), \tilde{g} \left( \begin{array}{c} x \\ \xi
2122: \end{array} \right) \right\rangle \\
2123: & = & \left\langle \tilde{g}B \left( \begin{array}{c} x \\ \xi
2124: \end{array} \right), \left( \begin{array}{c} x \\ \xi
2125: \end{array} \right) \right\rangle,
2126: \end{eqnarray*}
2127: since $\tilde{g}$ is symmetric.
2128:
2129: Now from the discussion preceding the statement of Theorem
2130: \ref{non-distinct-theorem}, we know $B = -JQ$ for $Q$ of the form
2131: \begin{eqnarray}
2132: \label{Q-full}
2133: Q = \left( \begin{array}{cc} 0 & A \\ A^T & 0 \end{array} \right),
2134: \end{eqnarray}
2135: where $A$ is block diagonal with diagonal elements of the form
2136: \eqref{A-1} or \eqref{A-2}. Thus with the same $A$ as \eqref{Q-full},
2137: \begin{eqnarray*}
2138: B = \left( \begin{array}{cc} A^T & 0 \\ 0 & -A \end{array} \right).
2139: \end{eqnarray*}
2140: Now we have reduced the problem to finding nonsingular $P$ such that $PA^T$ and
2141: $PA$ are both positive definite. But we know that if $\lambda$ is an
2142: eigenvalue of $A$, then $\Re \lambda >0$, so $A$ is positive definite
2143: and $P = I$ suffices. \eqref{non-distinct-theorem-statement} then
2144: follows immediately from Lemma \ref{HZ-ell-lemma}.
2145: \end{proof}
2146: We have used the following classical lemma (see, for example,
2147: \cite{hz} for a proof).
2148:
2149: \begin{lemma}
2150: \label{HZ-ell-lemma}
2151: Let
2152: \begin{eqnarray*}
2153: q(x, \xi ) = \half \langle Q (x, \xi), (x, \xi) \rangle
2154: \end{eqnarray*}
2155: be a positive definite quadratic form, where $Q$ is symmetric. Then
2156: there are positive numbers $0 < r_1 \leq r_2 \leq \cdots \leq r_{m}
2157: < \infty$ and a linear symplectic transformation $T$ such that
2158: \begin{eqnarray*}
2159: q(T(x, \xi)) = \sum_{j=1}^{m} \frac{1}{r_j^2} (x_j^2 + \xi_j^2 ).
2160: \end{eqnarray*}
2161: Further, if $T'$ is another linear symplectic transformation such that
2162: \begin{eqnarray*}
2163: q(T'(x, \xi)) = \sum_{j=1}^{m} \frac{1}{r_j'^2} (x_j^2 + \xi_j^2 )
2164: \end{eqnarray*}
2165: for $0 < r_1' \leq \cdots \leq r_{m}' < \infty$, then $r_j = r_j'$ for
2166: all $j$ and $T = T'$.
2167: \end{lemma}
2168:
2169:
2170:
2171:
2172:
2173:
2174:
2175:
2176:
2177:
2178:
2179:
2180:
2181:
2182:
2183:
2184:
2185:
2186:
2187: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2188:
2189:
2190:
2191:
2192:
2193:
2194:
2195:
2196:
2197:
2198:
2199:
2200: \section{Proof of Theorem \ref{main-theorem-2} and the Main Theorem}
2201: \label{main-theorem-2-proof}
2202: \subsection{Proof of Theorem \ref{main-theorem-2}}
2203: In this section we show how to use Theorem \ref{main-theorem-1} with a few other results to deduce
2204: Theorem \ref{main-theorem-2}. This is similar to \cite{BZ}, with the generalization of the
2205: loxodromic assumption. First we need the following standard lemma.
2206: \begin{lemma}
2207: \label{wf-lemma}
2208: Suppose $V_0 \Subset T^*X$, $p$ is a symbol, $T >0$, $A$ an operator, and $V \Subset T^*X$ a neighbourhood of $\gamma$ satisfying
2209: \begin{eqnarray}
2210: \left\{ \begin{array}{l}
2211: \forall \rho \in \{ p^{-1}(0) \} \setminus V, \,\,\, \exists \, 0 <t<T \,\, \text{and} \,\, \epsilon = \pm 1 \,\,\,
2212: \text{such that} \\
2213: \exp(\epsilon s H_p)(\rho) \subset \{ p^{-1}(0) \} \setminus V \,\, \text{for} \,\, 0 < s < t, \,\, \text{and} \\
2214: \exp(\epsilon t H_p)(\rho) \in V_0; \\
2215: \end{array} \right.
2216: \end{eqnarray}
2217: and $A$ is microlocally elliptic in $V_0 \times V_0 $. If $B \in \Psi^{0,0}(X, \Omega_X^\half)$ and $\WF (B) \subset T^*X
2218: \setminus V$, then
2219: \begin{eqnarray*}
2220: \left\| Bu \right\| \leq C \left( h^{-1} \left\| Pu \right\| + \| Au \| \right) + \O (h^\infty) \|u\|.
2221: \end{eqnarray*}
2222: \end{lemma}
2223: Figure \ref{fig-p0} is a picture of the setup of Lemma \ref{wf-lemma}.
2224: \begin{figure}
2225: \include{fig2}
2226: \caption{\label{fig-p0} The energy surface $\{ p^{-1}(0) \}$.}
2227: \end{figure}
2228: \begin{proof}
2229: Since $\{ p^{-1}(0) \}$ is compact, we can replace $V_0$ with a precompact neighbourhood of $V_0 \cap \{p^{-1}(0) \}$.
2230: We will prove a local version which can be pasted together to get the global estimate. We may assume
2231: $\WF(A) \subset U$, where $U$ is a small open neighbourhood of some point $\rho_0 \in V_0$, and
2232: \begin{eqnarray}
2233: \WF (B) \Subset \bigcup_{0 \leq t \leq t_0} \exp (\epsilon t H_p) ( U_1 ) \subset T^*X \setminus V,
2234: \end{eqnarray}
2235: where $U_1 \Subset U$ and $A$ is microlocally elliptic on $U_1\times U_1$. For $|t|\leq t_1$
2236: sufficiently small, by Proposition \ref{hDx-prop} there is a
2237: microlocally invertible $h$-FIO $T$ which
2238: conjugates $P$ to $hD_{x_1}$. Set $\tilde{u} = Tu$, and let
2239: $\tilde{B} \in \Psi^{0,0}$ be microlocally $1$ on $\WF (B) \times \WF (B)$
2240: and $0$ microlocally outside $\left( \cup_{0 \leq t \leq t_1} \exp
2241: (\epsilon t H_p) ( U_1 ) \right)^2 \subset (T^*X \setminus V)^2$.
2242: We calculate
2243: \begin{eqnarray*}
2244: \frac{1}{2} \partial_{x_1} \|\tilde{u} \|^2 & = & \langle \partial_{x_1} \tilde{u}, \tilde{u} \rangle \\
2245: & \leq & \left\| \partial_{x_1} \tilde{u} \right\| \|\tilde{u}\| \\
2246: & \leq & \frac{1}{4} h^{-1} \left\| TPT^{-1}\tilde{u} \right\|^2 + \|\tilde{u} \|^2 + \O(h^\infty) \|u\|_{L^2(X)}^2 \\
2247: \implies \| \tilde{B}T^{-1}\tilde{u} \|^2_{L^2(X)} & \leq & C_{t_1} \Big( h^{-1} \left\| TPT^{-1}\tilde{u} \right\|^2_{L^2(X)}
2248: + \left\|A T^{-1}\tilde{u} \right\|^2_{L^2(X)} + \\
2249: && \quad \quad \quad + \O(h^\infty)\|u\|^2_{L^2(X)}\Big),
2250: \end{eqnarray*}
2251: where the last inequality follows from Gronwall's inequality. But $\|BT^{-1}\tilde{u}\|^2_{L^2(X)} \leq
2252: \|\tilde{B}T^{-1}\tilde{u} \|^2_{L^2(X)}$
2253: gives the result for small $t$. Then we partition $[0, t_0]$ into finitely many subintervals and apply the small
2254: $t$ argument to each one.
2255: \end{proof}
2256: Using this lemma, we can deduce the following proposition.
2257: \begin{proposition}
2258: \label{Q(z)u-prop}
2259: Suppose $\psi_0 \in \s^{0,0}(T^*X) \cap \Ci_c(T^*X)$ is a microlocal cutoff function to a small neighbourhood
2260: of $\gamma \subset \{p^{-1}(0) \}$. For $Q(z) = P(h) - z - iCha^w$ as above with $z \in [-1,1] + i(-c_0h, \infty)$,
2261: $c_0 >0$ and $C>0$ sufficiently large, we have
2262: \begin{eqnarray}
2263: \label{Q(z)u-prop-est}
2264: Q(z)u = f \Longrightarrow \left\| (1 - \psi_0 )^w u \right\| \leq C h^{-1} \|f \| + \O(h^\infty) \|u \|.
2265: \end{eqnarray}
2266: \end{proposition}
2267: For this proposition and the proof, we use the convenient shorthand notation: for a symbol $b$, $b^w := \Op_h^w(b)$.
2268: \begin{remark}
2269: Note that Proposition \ref{Q(z)u-prop} is the best possible situation. It says roughly that away from $\gamma$,
2270: $Q^{-1}$ is bounded by $Ch^{-1}$. Thus the global statement in Theorem \ref{main-theorem-2} represents a loss of
2271: $\sqrt{\log (1/h)}$.
2272: \end{remark}
2273: \begin{proof}
2274: Choose $c_0 >0$ from Theorem \ref{main-theorem-1}, microlocal cutoff
2275: functions $\psi_1$, $\psi_2$ such that $\WF (1-\psi_j) \cap \gamma =
2276: \emptyset$, and $C>0$ sufficiently large so that
2277: \begin{eqnarray*}
2278: (Ca-c_0)^w (1 - \psi_1)^w \geq \left\{ \begin{array}{l}
2279: c_0 (1 - \psi_1)^w /2\\
2280: c_0 ((1 - \psi_2)^w)^*(1-\psi_2)^w/2, \end{array} \right.
2281: \end{eqnarray*}
2282: and $\supp \psi_1 \subset \{ \psi_2 = 1 \}$ (see Figure \ref{fig-cutoffs-1}).
2283: \begin{figure}
2284: \include{fig4}
2285: \caption{\label{fig-cutoffs-1} (a) The cutoff functions $a$, $\psi_1$, and $\psi_2$. (b) $(1-\psi_2)^2 \leq (1-\psi_1)$.}
2286: \end{figure}
2287: Then we calculate
2288: \begin{eqnarray*}
2289: \frac{1}{2} c_0 h \int_X \left| (1 - \psi_2)^w u \right|^2 dx & \leq & h \int_X \left( Ca^w + h^{-1} \Im z \right)
2290: u \overline{ (1- \psi_1)^w u } dx \\
2291: & = & -\Im \int_X Q(z) u \overline{ (1-\psi_1)^w u} dx \\
2292: & = & -\Im \int_X f \overline{ (1-\psi_1)^w u} dx \\
2293: & \leq & \|f\| \left( \left\| (1 - \psi_1)^w u \right\| + \O(h^\infty) \|u \| \right) \\
2294: & \leq & (4\epsilon h)^{-1} \|f \|^2 + \epsilon h \left\| (1 -\psi_1)^w u \right\|^2 + \O(h^\infty)\|u\|^2
2295: \end{eqnarray*}
2296: Now we use Lemma \ref{wf-lemma} with $A = (1-\psi_2)^w$, $B = (1-\psi_1)^w$, and $P = Q(z)$, which we may do since
2297: the perturbation terms in $Q(z)$ are all of lower order. Thus
2298: \begin{eqnarray*}
2299: \left\|(1-\psi_1)^w u \right\| & \leq & Ch^{-1} \left\|Q(z) u \right\| + \left\| (1 - \psi_2)^w u \right\|
2300: + \O(h^\infty) \|u \| \\
2301: \implies \left\|(1-\psi_1)^w u \right\|^2 & \leq & Ch^{-1} \|f\| \left( Ch^{-1} \|f\| + \left\| (1 - \psi_2)^w
2302: u \right\| \right) + \\
2303: && \quad \quad + \left\|( 1 -\psi_2)^w u \right\|^2 + \O(h^\infty) \|u \|^2 \\
2304: & \leq & Ch^{-2} \|f\|^2 + \left\|(1 - \psi_2)^w u \right\|^2 + \O(h^\infty) \|u \|^2 \\
2305: & \leq & Ch^{-2} \|f\|^2 + \epsilon \left\| (1 -\psi_1)^w u \right\|^2 + \O(h^\infty) \|u \|^2,
2306: \end{eqnarray*}
2307: which gives \eqref{Q(z)u-prop-est} with $\psi_0$ replaced by $\psi_1$. Another application of Lemma \ref{wf-lemma}
2308: with $A = (1 - \psi_2)^w$, $B = (\psi_1 - \psi_0)^w$, and $P = Q(z)$ shows the error $\|(\psi_1 - \psi_0)^w u \|$ is
2309: bounded by the same estimate as in \eqref{Q(z)u-prop-est}.
2310: \end{proof}
2311: We will need the next lemma, which is essentially an operator version of the classical Three-Line Theorem from complex
2312: analysis. We include the proof here for the reader's convenience,
2313: collected from \cite{BZ}, \cite{Bur}), and \cite{TaZw}.
2314: \begin{lemma}
2315: \label{BZ-lemma}
2316: Let $\mathcal{H}$ be a Hilbert space, and assume $A,B: \mathcal{H} \to \mathcal{H}$ are bounded, self-adjoint operators
2317: satisfying $A^2 = A$ and $BA = AB = A$. Suppose $F(z)$ is a family of bounded operators satisfying
2318: $F(z)^* = F(\bar{z})$, $\Re F \geq C^{-1} \Im z$ for $\Im z >0$, and further assume
2319: \begin{eqnarray*}
2320: BF^{-1}(z) B \,\, \text{is holomorphic in}\,\, \Omega:= [-\epsilon, \epsilon] + i[-\delta, \delta], \,\,
2321: \text{for} \,\, \frac{\delta}{\epsilon} \ll M^{-\frac{1}{N_1}} < 1
2322: \end{eqnarray*}
2323: for some $N_1>0$, where $\|B F^{-1}(z)B \| \leq M$. Then for $|z| < \epsilon/2$,
2324: $\Im z = 0$,
2325: \begin{eqnarray*}
2326: (a) \quad \left\| B F^{-1}(z) B \right\| & \leq & C \frac{\log M}{\delta}, \\
2327: (b) \quad \left\| B F^{-1}(z) A \right\| & \leq & C \sqrt{ \frac{\log M}{\delta}}.
2328: \end{eqnarray*}
2329: \end{lemma}
2330: \begin{proof}
2331: For the proof of part (a), consider the holomophic operator-valued function $f(z) = B F(z)^{-1} B$. Choose
2332: $\psi \in \Ci_c([-3 \epsilon/4, 3 \epsilon/4])$, $\psi \equiv 1$ on $[-\epsilon/2, \epsilon/2]$, and for
2333: $z \in \Omega$, set
2334: \begin{eqnarray*}
2335: \phi(z) = \delta^{-\half} \int e^{-(x-z)^2/\delta} \psi(x) dx.
2336: \end{eqnarray*}
2337: $\phi(z)$ has the following properties: \\
2338: \indent (a) $\phi(z)$ is holomorphic in $\Omega$, \\
2339: \indent (b) $|\phi(z)| \leq C$ in $\Omega$, \\
2340: \indent (c) $|\phi(z)| \geq C^{-1} >0$ on $[-\epsilon/2, \epsilon/2]$, and \\
2341: \indent (d) $|\phi(z)| \leq Ce^{-C/\delta}$ on $\Omega \cap \{ \Re z = \pm \epsilon \}$. \\
2342: Now for $z \in \tilde{\Omega}:= [-\epsilon, \epsilon] + i[- \delta, \delta/ \log M]$ set
2343: \begin{eqnarray*}
2344: g(z) = e^{-i N z \log M / \delta} \phi(z) f(z),
2345: \end{eqnarray*}
2346: and note that $g(z)$ satisfies \\
2347: \indent (a) $|g(z)| \leq C M^{1-N}$ on $\tilde{\Omega} \cap \{ \Im z = - \delta \}$, \\
2348: \indent (b) $|g(z)| \leq C_N e^{-C/\delta}$ on $\tilde{\Omega} \cap \{ \Re z = \pm \epsilon\}$, and \\
2349: \indent (c) $|g(z)| \leq C_N \log(M) / \delta$ on $\tilde{\Omega} \cap \{ \Im z = \delta/ \log M \}$. \\
2350: Then the classical maximum principle implies for $\delta$ sufficiently small and $N$ sufficiently large,
2351: $|g(z)| \leq C \log(M) / \delta$, which in turn implies
2352: \begin{eqnarray*}
2353: |f(z)| \leq C \frac{ \log M}{\delta} \,\,\, \text{on} \,\, \left[ -\frac{\epsilon}{2}, \frac{\epsilon}{2} \right]
2354: \subset \reals.
2355: \end{eqnarray*}
2356: \indent For part (b), note that our assumptions on $F(z)$ imply
2357: \begin{eqnarray*}
2358: \Im z \| u \|^2 \leq C \Re \langle F(z)u, u \rangle.
2359: \end{eqnarray*}
2360: We have
2361: \begin{eqnarray*} \|B F^{-1} A \|_{L^2 \to L^2} = \sup_{\{\|b\|_{L^2}
2362: = 1\}} \|B F^{-1} A b \|_{L^2} = \sup \| BF^{-1} A^2 b\|_{L^2},
2363: \end{eqnarray*}
2364: since $A^2 = A$. Suppose $F(z) u(x) = A b(x,z)$.
2365: Then $u = F(z)^{-1} A b$ and $Bu = B F^{-1} AA b$, and for $\Im z >0$,
2366: \begin{eqnarray*}
2367: \|Bu\|^2 & \leq & C \| u\|^2 \\
2368: & \leq & \frac{C}{\Im z} \langle \Re F(z) u, u \rangle \\
2369: & \leq & \frac{C}{\Im z} \left| \langle F(z) u, u \rangle \right|\\
2370: & = & \frac{C}{\Im z} \left| \langle Ab, u \rangle \right|\\
2371: & = & \frac{C}{\Im z} \left| \langle Ab, Au \rangle \right| \\
2372: & \leq & \frac{C}{\Im z} \| Ab\|^2
2373: \end{eqnarray*}
2374: where we have used $A^*A = A^2 = A$. Thus we have
2375: \begin{eqnarray*}
2376: \|B F(z)^{-1}A\|_{L^2 \to L^2} & \leq & \frac{C}{ \sqrt{ \Im z}}, \,\,\, \text{for} \,\, \Im z >0 \,\, \text{and} \\
2377: \|B F(z)^{-1}A\|_{L^2 \to L^2} & = & \sup_{\{ \| u \| = 1 \}} \|
2378: BF^{-1}A u \|_{L^2} \\
2379: & = & \sup_{\{ \| u \| = 1 \} } \| BF^{-1} BA u \|_{L^2
2380: \to L^2} \\
2381: & \leq & M \sup_{\{ \| u \| = 1 \} } \| A u \|_{L^2} \\
2382: & \leq & CM,
2383: \end{eqnarray*}
2384: and we can apply the proof of part (a) to $f(z) = B F(z)^{-1} A$ to get (b).
2385: \end{proof}
2386: \begin{proof}[Proof of Theorem \ref{main-theorem-2}]
2387: Let $\psi_0$ satisfy the assumptions of Proposition \ref{Q(z)u-prop}. Then
2388: \begin{eqnarray*}
2389: \| (1 -\psi_0)^w u \| \leq Ch^{-1} \| Q(z) u\| + \O(h^\infty) \|u\|.
2390: \end{eqnarray*}
2391: Further, since
2392: \begin{eqnarray*}
2393: \left\| \left[ Q, \psi_0^w \right] u \right\| \leq \left\| \left[ Q, \psi_0^w \right] (1 - \tilde{\psi}_0^w)u \right\|
2394: + \O(h^\infty) \|u \|,
2395: \end{eqnarray*}
2396: for some $\tilde{\psi}_0$ satisfying the assumptions of Proposition \ref{Q(z)u-prop} and $\WF \tilde{\psi}_0 \subset
2397: \{ \psi_0 = 1\}$, so using Theorem \ref{main-theorem-1} and the fact that $[Q, \psi_0^w]$ is compactly supported and
2398: of order $h$, we have
2399: \begin{eqnarray*}
2400: \| \psi_0^w u \| & \leq & Ch^{-N_0} \left( \| \psi_0^w Q u \| + \left\| \left[ Q, \psi_0^w \right] u \right\| \right) \\
2401: & \leq & C h^{-N_0} \left( \|\psi_0^w Q u\| + h^{-1} \| h Q u\| \right) + \O(h^\infty) \\
2402: & \leq & C h^{-N_0} \| Qu \| + \O(h^\infty)\| u \|.
2403: \end{eqnarray*}
2404: \indent Now let $F(w)$ be the family of operators $F(w) = ih^{-1} Q(z_0 + hw)$, $A = \chi_{\supp \phi}^w$, $B = \id$.
2405: Fix $\delta >0$ independent of $h$, $\epsilon = (Ch)^{-1}$, $M = h^{-N_0}$, and apply Lemma \ref{BZ-lemma} to get
2406: \begin{eqnarray*}
2407: \|B F^{-1} B \| & \leq & C \log (h^{-N_0}); \\
2408: \|B F^{-1} A \| & \leq & C \sqrt{ \log (h^{-N_0}) },
2409: \end{eqnarray*}
2410: and (\ref{main-theorem-2-est-1}-\ref{main-theorem-2-est-2}) follows.
2411: \end{proof}
2412: \subsection{Proof of the Main Theorem}
2413: The Main Theorem is an easy consequence of Theorem \ref{main-theorem-2}.
2414: \begin{proof}[Proof of the Main Theorem]
2415: Recall $A$ is $0$ microlocally away from $\gamma \times \gamma$. Let $\tilde{A} \in \Psi_h^{0,0}$ be a pseudodifferential
2416: operator so that $\tilde{A} = I$ microlocally on a neighbourhood of
2417: $\WF (A) \times \WF (A)$. Let $a^w$ be as in
2418: Theorems \ref{main-theorem-1} and \ref{main-theorem-2}. Choosing $A$ and $\tilde{A}$ so that $\WF (a^w)$ is
2419: disjoint from $\WF(\tilde{A})$, we have for $Q = Q(0)$
2420: \begin{eqnarray}
2421: \label{QAu}
2422: Q\tilde{A} u = P \tilde{A} u.
2423: \end{eqnarray}
2424: The right hand side of \eqref{QAu} is $[ P, \tilde{A}] u + \tilde{A} Pu$. Now $[P, \tilde{A}]$ is
2425: supported away from $\gamma$ since $\tilde{A}$ is constant near $\gamma$, so
2426: \begin{eqnarray}
2427: %\label{QAu-2}
2428: \left\| P\tilde{A} u \right\|_{L^2(X)} & \leq & \left\| \left[ P,
2429: \tilde{A} \right] u \right\|_{L^2(X)} + \|Pu\|_{L^2(X)} \nonumber \\
2430: & \leq & C h \left\|
2431: (I - A) u \right\|_{L^2(X)} + \| Pu \|_{L^2(X)}. \label{QAu-2}
2432: \end{eqnarray}
2433: From Theorem \ref{main-theorem-2}, we have
2434: \begin{eqnarray}
2435: \label{QAu-3}
2436: \left\| Q \tilde{A} u \right\|_{L^2(X)} \geq C^{-1} \frac{h}{\sqrt{\log (1/h)} } \left\| \tilde{A} u \right\|_{L^2(X)}.
2437: \end{eqnarray}
2438: Combining \eqref{QAu-2} and \eqref{QAu-3}, we have
2439: \begin{eqnarray*}
2440: C^{-1} \|u \|_{L^2(X)} & \leq & C^{-1} \left( \left\|\tilde{A} u \right\|_{L^2(X)} + \left\| (I - A) u \right\|_{L^2(X)}
2441: \right) \\
2442: & \leq & C \left( \sqrt{\log (1/h)} + C^{-1} \right) \left\| (I - A ) u
2443: \right\|_{L^2(X)} \\
2444: && \quad \quad \quad \quad + C \frac{ \sqrt{ \log(1/h)}}{h} \| P u \|_{L^2(X)} ,
2445: \end{eqnarray*}
2446: which for $0 < h < h_0$ is \eqref{main-theorem-3-est}.
2447: \end{proof}
2448: \begin{remark}
2449: In the calculation \eqref{QAu-2}, we have only used $\|[P, \tilde{A}]u \| \leq Ch \| (I-A)u\|$. If we could determine
2450: a global geometric condition which would allow us to choose $\tilde{A}$ in a manner which improves this, but doesn't have
2451: too much interaction with $a^w$ in the definition of $Q(z)$, we could eliminate the $\log(h^{-1})$ in
2452: \eqref{main-theorem-3-est}.
2453: \end{remark}
2454:
2455:
2456:
2457:
2458:
2459:
2460: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2461: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2462: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2463: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2464:
2465: \section{An Application: The Damped Wave Equation}
2466: \label{wave-section}
2467: \numberwithin{equation}{section}
2468: In this section we adapt the techniques from \S
2469: \ref{main-theorem-1-proof}-\ref{main-theorem-2-proof}
2470: to study the damped wave equation. Let $X$ be a compact manifold
2471: without boundary, $a(x) \in \Ci(X)$, $a(x) \geq 0$,
2472: and consider the following problem:
2473: \begin{eqnarray}
2474: \label{wave-equation-1}
2475: \left\{ \begin{array}{l}
2476: \left( \partial_t^2 - \Delta + 2a(x) \partial_t \right) u(x,t) = 0, \quad (x,t) \in X \times (0, \infty) \\
2477: u(x,0) = 0, \quad \partial_t u(x,0) = f(x).
2478: \end{array} \right.
2479: \end{eqnarray}
2480: Let $p \in \Ci (T^*X)$,
2481: $p= |\xi|^2$, be the microlocal principal symbol of $-\Delta$ and suppose the classical flow
2482: (geodesic flow) of $H_p$ admits a single closed, loxodromic orbit $\gamma$ in the
2483: level set $\{ p^{-1}(1)\}$. Assume throughout that $a(x)$ is supported away from the
2484: projection $\tilde{\gamma}$ of $\gamma$ onto $X$ (see Figure \ref{figure:x-gamma}).
2485: \begin{figure}
2486: \include{fig5}
2487: \caption{\label{figure:x-gamma} The manifold $X$ and the projection
2488: $\tilde{\gamma}$ of $\gamma$ onto $X$.}
2489: \end{figure}
2490: We recall that the $H^s$ inner product on $X$ is given by the local formula
2491: \be
2492: \lll u, v \rrr_{H^s} = \int_{\reals^n} (1 + |\xi|^2)^s \hat{u}
2493: \bar{\hat{v}} d\xi,
2494: \ee
2495: where $\hat{u}$ is the Fourier transform of $u$.
2496: If $u$ solves \eqref{wave-equation-1}, we define the {\it $s$-energy}
2497: $E^s(t)$ of $u$ at time $t$ to be
2498: \begin{eqnarray*}
2499: E^s(t) = \half \left( \left\| \partial_t u \right\|^2_{H^s(X)} + \left\|
2500: \sqrt{-\Delta} u \right\|^2_{H^s(X)} \right).
2501: \end{eqnarray*}
2502:
2503: \begin{lemma}
2504: If $a(x) \equiv 0$, $E^s(t)$ is constant. If $a(x)$ is not identically zero, then $E^s(t)$ is
2505: decreasing.
2506: \end{lemma}
2507: \begin{proof}
2508: \be
2509: \frac{d}{dt} E^s(t) & = & \lll \partial_t^2 u, \partial_t u \rrr_{H^s} + \lll
2510: \partial_t \sqrt{-\Delta}u, \sqrt{-\Delta} u \rrr_{H^s} \\
2511: & = & \lll \partial_t u , ( \partial_t^2 - \Delta ) u \rrr \\
2512: & = & -\lll \partial_t u, 2 a(x) \partial_t u \rrr.
2513: \ee
2514: %\begin{eqnarray*}
2515: %E'(t) & = & \int_X \left(\partial_t^2 u \partial_tu \right) + \left( \partial_t \partial_x u \cdot \partial_x u \right) dx \\
2516: %& = & \int_X \partial_t u \left( \partial_t^2 u - \Delta u \right) dx \\
2517: %& = & - \int_X a(x)\left( \partial_t u \right)^2 dx.
2518: %\end{eqnarray*}
2519: \end{proof}
2520:
2521:
2522: We make an important dynamical assumption, which amounts to a
2523: geometric control condition similar to that given by
2524: Rauch-Taylor in \cite{RT}. We assume:
2525: \begin{eqnarray}
2526: \label{RT}
2527: \left\{ \begin{array}{l} \text{There exists a time }T>0 \text{ and a
2528: neighbourhood } V \\
2529: \text{of } \gamma \text{ such that for all } |\xi| = 1,\,\,
2530: (x, \xi) \in T^*X \setminus V,\\
2531: \exp(tH_p)(x, \xi) \cap \{ a>0\}
2532: \neq \emptyset
2533: \text{ for some } |t| \leq T.
2534: \end{array} \right.
2535: \end{eqnarray}
2536:
2537: In \cite{EvZw} \S $5.3$, it is shown that with a global Rauch-Taylor condition,
2538: we have exponential decay in zero-energy. Here we have a region
2539: without geometric control, so we expect some loss.
2540: \begin{theorem}
2541: \label{exp-decay}
2542: Assume \eqref{RT} holds and $a(x)$ is not identically zero. Then for
2543: any $\epsilon>0$, there is a constant $C>0$ such that
2544: \begin{eqnarray*}
2545: E^0(t) \leq C e^{ -t/C } \| f\|_{H^\epsilon}^2.
2546: \end{eqnarray*}
2547: \end{theorem}
2548:
2549: The damped wave equation in the context of a global Rauch-Taylor
2550: condition has been studied in
2551: \cite{RT}, \cite {Sjo2}, \cite{Leb}, and \cite{Hit}. The difference
2552: here is the presence of $\gamma$ and a neighbourhood in which the
2553: Rauch-Taylor condition doesn't hold.
2554:
2555: Formally, if $u \equiv 0$ for $t<0$, we apply the Fourier transform to \eqref{wave-equation-1} in
2556: the $t$ variable and integrating by parts motivates us to study the
2557: equation
2558: \begin{eqnarray}
2559: \label{wave-equation-2}
2560: P(\tau)\hat{u}(x, \tau) :=(-\tau^2 - \Delta + 2ia(x) \tau) \hat{u}(x,
2561: \tau) = f.
2562: \end{eqnarray}
2563:
2564: We use the techniques of the previous sections to gain estimates on
2565: the resolvent $P(\tau)^{-1}$. We call the poles of $P(\tau)^{-1}$
2566: {\it eigenfrequencies} for \eqref{wave-equation-1}. Note if $\tau$ is
2567: an eigenfrequency, then $0 \leq \Im \tau \leq 2 \|a\|_{L^\infty}$.
2568: Further, \eqref{wave-equation-2} is invariant under the transformation
2569: $(\hat{u}, \tau) \mapsto (\bar{\hat{u}}, - \bar{\tau})$, so the set of
2570: eigenfrequencies is symmetric about the imaginary axis. We
2571: therefore study only those in the right half plane. For $0 < h \leq
2572: h_0$ and $z \in
2573: \Omega: = [\alpha, \beta] + i[-\gamma, \gamma]$ where $0 <
2574: \alpha<1<\beta<\infty$ and $\gamma>0$, set $\tau = \sqrt{z}/h$.
2575: \eqref{wave-equation-2} becomes
2576: \begin{eqnarray}
2577: \label{wave-equation-3}
2578: \frac{1}{h^2} Q(z, h)\hat{u} = f
2579: \end{eqnarray}
2580: where
2581: \begin{eqnarray}
2582: \label{Q(z,h)}
2583: Q(z,h) = P(h) - z + 2ih \sqrt{z} a(x)
2584: \end{eqnarray}
2585: and the principal symbol
2586: of $P(h)$ is $p(x, \xi) = |\xi|^2$. The next Corollary follows
2587: directly from the proof of Theorem \ref{main-theorem-1}, replacing $s$
2588: in the conjugation \eqref{Pth1} with $-s$.
2589: \begin{corollary}
2590: Suppose $u$ has wavefront set sufficiently close to $\gamma$. Then
2591: there exists $c_0 >0$, $C < \infty$, and $N \geq 0$ such that
2592: for $z \in [\alpha, \beta] + i[-c_0h, c_0h]$,
2593: \begin{eqnarray*}
2594: \Im \left\langle (P(h) - z) u, u \right\rangle \geq C^{-1} h^N
2595: \|u\|^2.
2596: \end{eqnarray*}
2597: In particular, $\|Q(z,h) u\| \geq C^{-1} h^N \|u\|$.
2598: \end{corollary}
2599:
2600: We observe that for $u$ as in the theorem and $-c_0 h <\Im z <0$, $\|Q(z, h) u \| \geq C^{-1} \Im z \|u\|$.
2601:
2602: The proof of Proposition
2603: \ref{Q(z)u-prop} relies on the assumption that the symbol $a(x,
2604: \xi)$ in \eqref{Q(z)} is elliptic away from $\gamma$. The function
2605: $a(x)$ in \eqref{Q(z,h)} is not assumed to be elliptic anywhere, so we
2606: will use a technique from \cite{Leb} to replace
2607: $a(x)$ with its average over trajectories of $\exp(tH_p)$.
2608:
2609: For $T >0$, we define the $T$-trajectory average of a smooth function
2610: $b$:
2611: \begin{eqnarray*}
2612: \langle b \rangle_T (x, \xi) = \frac{1}{T} \int_0^T b \circ
2613: \exp(tH_p)(x, \xi) dt.
2614: \end{eqnarray*}
2615: Set $q(z) = 2 \sqrt{z}a(x)$, and for $z \in \widetilde{\Omega}:= [\alpha, \beta] + i[0, c_1h]$, where
2616: $c_1 >0$ will be chosen later, and $(x, \xi) \in \{p^{-1}([\alpha -
2617: \delta, \beta+ \delta])\}$ for $\delta>0$, let $g_{\Re z} \in \s (1)$
2618: depending on $T$ solve
2619: \begin{eqnarray*}
2620: q( \Re z) - H_pg_{\Re z} = \langle q( \Re z) \rangle_T.
2621: \end{eqnarray*}
2622: (See \cite{Sjo2} for details on the construction of $g_{\Re z}$.) Now we form the elliptic
2623: operator $A: = \Op_h^w (e^g) \in \Psi^{0,0}$, and observe
2624: \begin{eqnarray*}
2625: A^{-1} P A & = & P + A^{-1}[P,A] \\
2626: & = & P - ih \Op_h^w(e^g)^{-1} \Op_h^w (\{p,e^g\}) \\
2627: & = & P - ihB,
2628: \end{eqnarray*}
2629: with $\sigma_h(B) = e^{-g} \{p, e^g \} + \O(h) = H_pg + \O(h)$. Thus
2630: \begin{eqnarray*}
2631: A^{-1} \left( P + ih q(z) \right) A & = & P + ih \Op_h^w \left( q( \Re z)
2632: - H_pg \right) + \O(h^2) \\
2633: & = & P + ih\Op_h^w \left( \langle q( \Re z ) \rangle_T \right),
2634: \end{eqnarray*}
2635: since $\Im z = \O(h) \Re z$. Following \cite{Hit}, we claim there
2636: exists a time $T>0$ such that
2637: \begin{eqnarray}
2638: \label{inf-ave}
2639: \langle a \rangle_T (x, \xi) \geq C^{-1} >0
2640: \end{eqnarray}
2641: for $(x, \xi) \in \{ p^{-1}([\alpha - \delta/2, \beta + \delta/2])
2642: \}\setminus V$, where $V$ is as in the statement of Theorem \ref{exp-decay}. To see this, recall $p = | \xi|^2$ means $H_p = 2 \langle \xi,
2643: \partial_x \rangle$ and $p^{-1}(E) = \{ | \xi| = \sqrt{E} \}$, which
2644: means
2645: \begin{eqnarray*}
2646: \inf_{p^{-1}(E)} \langle a \rangle_T = \inf_{p^{-1}(1)} \langle a
2647: \rangle_{\sqrt{E}T}.
2648: \end{eqnarray*}
2649: By Assumption \eqref{RT},
2650: \begin{eqnarray*}
2651: \inf_{p^{-1}(1)} \langle a
2652: \rangle_{\sqrt{E}T} \geq C^{-1} >0
2653: \end{eqnarray*}
2654: in $\{p^{-1}(1)\} \setminus V$ for $T$ sufficiently large and
2655: $\sqrt{E}$ close to $1$. Choosing
2656: $\alpha$ and $\beta$ sufficiently close to $1$ proves \eqref{inf-ave}.
2657: \begin{corollary}
2658: \label{Q(z)u-prop-2}
2659: Suppose $\psi_0 \in \s^{0,0}(T^*X) \cap \Ci_c(T^*X)$ is a microlocal cutoff function to a small neighbourhood
2660: of $\gamma \subset \{p^{-1}(1) \}$. For $Q(z,h) = P(h) - z
2661: +2ih\sqrt{z}a$ as above with $z \in [\alpha, \beta] + i(-c_1h, c_1h)$,
2662: $c_1 >0$, we have
2663: \begin{eqnarray}
2664: \label{Q(z)u-prop-est-2}
2665: Q(z,h)u = f \Longrightarrow \left\| (1 - \psi_0 )^w u \right\| \leq C h^{-1} \|f \| + \O(h^\infty) \|u \|.
2666: \end{eqnarray}
2667: \end{corollary}
2668: \begin{proof}
2669: Selecting $T >0$ sufficiently large and $c_1 >0$ such that
2670: \begin{eqnarray*}
2671: 0 < c_1 < \inf_{p^{-1}([\alpha - \delta/2, \beta + \delta/2])} \langle a \rangle_T (x, \xi),
2672: \end{eqnarray*}
2673: we apply the proof of Proposition \ref{Q(z)u-prop} to the conjugated
2674: operator $A^{-1} Q(z,h)A$.
2675: \end{proof}
2676: We now have good resolvent estimates for $z$ in an $h$ interval below the
2677: real axis, as well as weaker estimates above.
2678: \begin{corollary}
2679: (i) There exist constants $C>0$ and $N >0$ such that the resolvent $Q(z,h)^{-1}$ satisfies
2680: \begin{eqnarray*}
2681: \|Q(z,h)^{-1}\|_{L^2 \to L^2} \leq C h^{-N} , \quad z \in
2682: [\alpha, \beta] + i(-c_0h, c_0h).
2683: \end{eqnarray*}
2684: (ii) In addition, there is a
2685: constant $C_1$ such that
2686: \begin{eqnarray*}
2687: \|Q(z,h)^{-1}\|_{L^2 \to L^2} \leq C_1 \frac{\log(1/h)}{h}, \quad z
2688: \in [\alpha, \beta] +i[-C_1^{-1}h / \log(1/h), C_1^{-1}h].
2689: \end{eqnarray*}
2690: \end{corollary}
2691: This is an immediate consequence of the proof of Theorem
2692: \ref{main-theorem-2}, together with the slight modification of Lemma
2693: \ref{BZ-lemma} given in Lemma \ref{holo-lemma-2}.
2694:
2695: \begin{lemma}
2696: \label{holo-lemma-2}
2697: Let $f(z)$ be a holomorphic function on $\Omega = [-\epsilon,
2698: \epsilon] + i [-\delta, \delta]$, with
2699: \be
2700: \frac{\delta}{\epsilon} \ll M^{-\frac{1}{N_1}}
2701: \ee
2702: for some $N_1>0$, and suppose
2703: $f$ satisfies $|f(z) | \leq M$ on
2704: $\Omega$ with $|f(z)| \leq C|\Im z|$ for $\Im z <0$. Then there
2705: exists a constant $0<C_1<\infty$ such that if $ - C_1^{-1}\delta/\log M \leq \Im z
2706: \leq C_1^{-1}\delta$ we have
2707: \be
2708: |f(z)| \leq C \frac{\log M}{\delta}.
2709: \ee
2710: \end{lemma}
2711: \begin{proof}
2712: Let $\psi(x)$ be as in the proof of Lemma \ref{BZ-lemma}, and for
2713: $C_1^{-1}\ll c_0$, let
2714: \be
2715: \phi(z) = \delta^{-\half} \int e^{-(x -z + iC_1^{-1} \delta )^2/ \delta} \psi(x) dx.
2716: \ee
2717: We observe if $C_1>0$ is sufficiently large, for $|\Im z| \leq
2718: C_1^{-1} \delta $,
2719: \be
2720: \lefteqn{(x - z + i C_1^{-1} \delta)^2 = } \\ & = & (x - \Re z)^2 - (C_1^{-1}\delta -
2721: \Im z)^2 + 2i(x - \Re z)(C_1^{-1}\delta - \Im z)
2722: \ee
2723: and
2724: \be
2725: \left| (x - \Re z) ( C_1^{-1} \delta - \Im z) \right| \leq 4 C_1^{-1}
2726: \epsilon \delta,
2727: \ee
2728: so if $C_1>0$ is sufficiently large,
2729: \be
2730: \Re e^{-(x -z + iC_1^{-1} \delta )^2/ \delta} & \geq & e^{-(x-\Re
2731: z)^2/\delta + (C_1^{-1} \delta - \Im z )^2/\delta} \cos(4C_1^{-1}
2732: \epsilon) \\
2733: & \geq & C^{-1}
2734: e^{-(x-\Re z)^2/\delta + (C_1^{-1} \delta - \Im z )^2/\delta}.
2735: \ee
2736: Thus$\phi(z)$ satisfies
2737:
2738: (a) $\phi(z)$ is holomorphic in $\Omega$,
2739:
2740: (b) $|\phi(z)| \leq C$ in $\Omega$,
2741:
2742: (c) $|\phi(z)| \geq C^{-1}$ for $z \in [-\epsilon/2, \epsilon/2] +
2743: i[-C_1^{-1}\delta, C_1^{-1} \delta]$,
2744:
2745: (d) $|\phi(z)| \leq C e^{-C/\delta}$ on $\{ \pm \epsilon \} \times
2746: i[-C_1^{-1} \delta, C_1^{-1} \delta]$, if $C_1>0$ is chosen large enough.
2747:
2748:
2749: Now similar to the proof of Lemma \ref{BZ-lemma}, for
2750: \be
2751: z \in
2752: \tilde{\Omega}:= [-\epsilon, \epsilon] + i[- C_1^{-1} \delta/ \log M,
2753: C_1^{-1}\delta]
2754: \ee
2755: set
2756: \begin{eqnarray*}
2757: g(z) = e^{i N z \log M / \delta} \phi(z) f(z).
2758: \end{eqnarray*}
2759:
2760: Then as in the proof of Lemma \ref{BZ-lemma}, the classical maximum principle implies for $\delta$ sufficiently small and $N$ sufficiently large,
2761: $|g(z)| \leq C \log(M) / \delta$, which in turn implies
2762: \begin{eqnarray*}
2763: |f(z)| \leq C \frac{ \log M}{\delta} \,\,\, \text{on} \,\, \left[
2764: -\frac{\epsilon}{2}, \frac{\epsilon}{2} \right] + i [-C_1^{-1}\delta / \log
2765: M , C_1^{-1} \delta ].
2766: \end{eqnarray*}
2767: \end{proof}
2768:
2769:
2770:
2771:
2772:
2773: With these resolvent estimates, we have the following estimates in
2774: terms of $\tau$.
2775: \begin{proposition}
2776: \label{P-tau-bnds}
2777: Fix $\epsilon>0$. There exist constants $0 < C, C_1< \infty $ such that if
2778: \be
2779: -(\log \lll \tau \rrr)^{-1} \leq \Im \tau \leq C_1^{-1},
2780: \ee
2781: \begin{eqnarray}
2782: \label{P-L2} \| P( \tau)^{-1} \|_{L^2 \to L^2} & \leq &
2783: \frac{C\log \lll \tau \rrr}{\langle \tau \rangle }, \\
2784: \label{P-H2} \| P( \tau)^{-1} \|_{L^2 \to H^2} & \leq &
2785: {C}{\langle \tau \rangle \log \lll \tau \rrr}, \,\, \text{and} \\
2786: \label{P-H1} \| P( \tau)^{-1} \|_{H^s \to H^{s+1-\epsilon}} & \leq & C.
2787: \end{eqnarray}
2788: \end{proposition}
2789: \begin{proof}
2790: \eqref{P-L2} follows directly from rescaling. To see \eqref{P-H2},
2791: calculate
2792: \begin{eqnarray*}
2793: \|u \|_{H^2} & \leq & C ( \| \Delta u \|_{L^2} + \| u \|_{L^2} ) \\
2794: & \leq & C \left( \left\| P(\tau) u \right\|_{L^2} + \left\| ( -
2795: \tau^2 + 2ia(x) \tau ) u \right\|_{L^2} + \frac{\log \langle \tau \rangle}{\langle \tau
2796: \rangle } \left\| P(
2797: \tau) u \right\|_{L^2} \right) \\
2798: & \leq & C \left(1 + | \tau |\log \lll \tau \rrr + \langle \tau
2799: \rangle^{-1} \log \lll \tau \rrr \ \right)
2800: \left\| P( \tau) u \right\|_{L^2}.
2801: \end{eqnarray*}
2802: For \eqref{P-H1}, let $\epsilon>0$ be given. From Lemma
2803: \ref{inter-lemma}, we have
2804: \be
2805: \left\| P(\tau)^{-1} u \right\|_{H^{1-\epsilon}}^2 & \leq & C \left \|
2806: P(\tau)^{-1} u
2807: \right\|_{H^2}^{1-\epsilon} \left\| P(\tau)^{-1}u \right\|_{L^2}^{1 +
2808: \epsilon} \\
2809: & \leq & C_\epsilon \|u\|_{L^2(X)}^2.
2810: \ee
2811: To get the estimates for $H^s \to H^{s+1-\epsilon}$, we conjugate
2812: $P(\tau)^{-1}$ by the operators
2813: \be
2814: \Lambda^s = (1 - \Delta)^\frac{s}{2}
2815: \ee
2816: and apply to $v = \Lambda^s u$:
2817: \be
2818: \left\|P(\tau)^{-1} u \right\|_{H^{s+1-\epsilon}} & = &
2819: \left\|\Lambda^{1-\epsilon} \Lambda^sP(\tau)^{-1} \Lambda^{-s} v
2820: \right\|_{L^2} \\
2821: & = & \left\| \Lambda^{1-\epsilon} \left( P(\tau)^{-1}
2822: + \Lambda^s [ P(\tau)^{-1} , \Lambda^{-s}] \right) v \right\|_{L^2} \\
2823: & \leq & \left\| v \right\|_{L^2} \\
2824: & \leq & C \left\|u \right\|_{H^s}.
2825: \ee
2826: \end{proof}
2827: We have used the following interpolation lemma.
2828: \begin{lemma}
2829: \label{inter-lemma}
2830: Let $\epsilon>0$ be given, and suppose $f \in H^2(X) \cap
2831: L^2(X)$. Then
2832: \be
2833: \left\|f\right\|_{H^{1-\epsilon}}^2 \leq C \left \| f
2834: \right\|_{H^2}^{1-\epsilon} \left\| f \right\|_{L^2}^{1 + \epsilon}.
2835: \ee
2836: \end{lemma}
2837: \begin{proof}
2838: We use the local formula for $H^s$ norms and calculate:
2839: \be
2840: \left\|f\right\|_{H^{1-\epsilon}}^2 &=& \int_{\reals^n} (1 +
2841: |\xi|^2)^{1-\epsilon} \hat{f} \bar{\hat{f}} d \xi \\
2842: & = & \int \left( (1 + |\xi|^2) |\hat{f}| \right)^{1-\epsilon}
2843: |\hat{f}|^{1+\epsilon} d \xi \\
2844: & \leq & C \left\|\left( (1 + |\xi|^2) |\hat{f}| \right)^{1-\epsilon}
2845: \right\|_{L^{\frac{2}{1-\epsilon}}} \left\| |\hat{f}|^{1+\epsilon}
2846: \right\|_{L^{\frac{2}{1+\epsilon}}} \\
2847: & = & \leq C \left \| f
2848: \right\|_{H^2}^{1-\epsilon} \left\| f \right\|_{L^2}^{1 + \epsilon} .
2849: \ee
2850: \end{proof}
2851:
2852:
2853:
2854:
2855:
2856:
2857:
2858:
2859:
2860:
2861: We are now in position to prove Theorem \ref{exp-decay}. This proof
2862: comes almost directly from \cite{EvZw} \S $5.3$.
2863: \begin{proof}[Proof of Theorem \ref{exp-decay}]
2864: Assume $u(x,t)$ solves \eqref{wave-equation-1}.
2865: Choose $\chi \in \Ci( \reals )$, $0 \leq \chi \leq 1$, $\chi \equiv 1$
2866: on $[1, \infty)$, and $\chi \equiv 0$ on $(-\infty, 0]$. Set $u_1 (x,t)=
2867: \chi(t) u(x,t)$. We apply the damped wave operator to $u_1$:
2868: \begin{eqnarray}
2869: \lefteqn{ \left( \partial_t^2 - \Delta + 2a \partial_t \right) u_1 =}
2870: \label{wave-equation4.a} \\ & = & \chi''
2871: u + 2 \chi' u_t + 2a \chi' u + \chi \left( \partial_t^2 - \Delta + 2a \partial_t
2872: \right) u \nonumber \\
2873: & = & \chi''
2874: u + 2 \chi' u_t + 2a \chi' u =: g_1. \label{wave-equation4.b}
2875: \end{eqnarray}
2876: With $g_1$ supported in $X \times (0,1)$ and $u_1 \equiv 0$ for $t\leq
2877: 0$, we have
2878: \begin{eqnarray}
2879: \label{g1-bnd}
2880: \|g_1\|_{L^2((0, \infty); H^\epsilon)}^2 \leq C \left( \| u \|_{L^2((0,1);
2881: H^\epsilon)}^2 + \|\partial_t u\|_{L^2((0,1);H^\epsilon)}^2 \right).
2882: \end{eqnarray}
2883: Now
2884: \begin{eqnarray*}
2885: \partial_t \langle u,u \rangle_{H^\epsilon(X)} & = & 2 \langle \partial_t u, u
2886: \rangle_{H^\epsilon(X)} \\
2887: & \leq & \| \partial_t u \|_{H^\epsilon(X)}^2 + \| u \|_{H^\epsilon(X)}^2 \\
2888: & \leq & CE^\epsilon(t) + \|u \|_{H^\epsilon(X)}^2,
2889: \end{eqnarray*}
2890: so by Gronwall's inequality,
2891: \begin{eqnarray*}
2892: \|u(t, \cdot) \|_{H^\epsilon(X)}^2 & \leq& C e^t\left( \|u(0,\cdot)
2893: \|_{H^\epsilon(X)}^2 + \int_0^t E^\epsilon(s)ds \right) \\
2894: & \leq & Ct e^t \| f \|_{H^\epsilon(X)}^2.
2895: \end{eqnarray*}
2896: Thus \eqref{g1-bnd} is bounded by $C \|f \|_{H^\epsilon(X)}^2$.
2897:
2898: We now apply the Fourier transform to
2899: (\ref{wave-equation4.a}-\ref{wave-equation4.b}) to write $\hat{u}_1 =
2900: P(\tau)^{-1} \hat{g}_1$. By Proposition \ref{P-tau-bnds}, we have for $\Im \tau =
2901: C^{-1}>0 $
2902: \begin{eqnarray*}
2903: \left\| e^{t/C} u_1 \right\|_{L^2((0, \infty); H^1)} & = & \left\|
2904: \hat{u}_1( \cdot + iC^{-1}) \right\|_{L^2((-\infty, \infty);H^1)} \\
2905: & = & \left\| P(\cdot + iC^{-1})^{-1} \hat{g}_1( \cdot + i C^{-1})
2906: \right\|_{L^2((-\infty, \infty); H^1)} \\
2907: & \leq & C \|\hat{g}_1 \|_{L^2((-\infty, \infty); H^\epsilon)} \\
2908: & \leq & C\| g_1\|_{L^2((0, \infty); H^\epsilon)} \\
2909: & \leq & C\|f\|_{H^\epsilon(X)}.
2910: \end{eqnarray*}
2911: Thus
2912: \begin{eqnarray*}
2913: \|e^{t/C} u \|_{L^2((1, \infty);H^1)} \leq C \|f\|_{H^\epsilon(X)}.
2914: \end{eqnarray*}
2915: Now for $T>2$, choose $\chi_2 \in \Ci( \reals)$, $0 \leq \chi_2 \leq 1$,
2916: such that $\chi_2 \equiv 0$ for $t \leq T-1$, and $\chi_2 \equiv 1$ for $t
2917: \geq T$. Set $u_2 (x,t) = \chi_2(t) u(x,t)$. We have
2918: \begin{eqnarray*}
2919: \left( \partial_t^2 - \Delta + 2a \partial_t \right) u_2 = g_2
2920: \end{eqnarray*}
2921: for $g_2 = \chi_2'' u + 2 \chi_2' u_t + 2a \chi_2' u$, and $\supp g_2
2922: \subset X \times [T-1,T]$. Define
2923: \begin{eqnarray*}
2924: E_2(t) = \frac{1}{2} \int_X \left( \partial_t u_2 \right)^2 + \left|
2925: \sqrt{-\Delta} u_2 \right|^2 dx,
2926: \end{eqnarray*}
2927: and observe
2928: \begin{eqnarray*}
2929: E_2'(t) & = & \left\langle \partial_t^2 u_2, \partial_t u_2
2930: \right\rangle_X - \left\langle \Delta u_2, \partial_t u_2
2931: \right\rangle_X \\
2932: & = & - \left\langle 2a(x) \partial_t u_2, \partial_t u_2
2933: \right\rangle_X + \left\langle g_2, \partial_t u_2 \right\rangle_X \\
2934: & \leq & C \int_X \left| \partial_t u_2 \right| \left( \left|
2935: \partial_t u \right| + |u| \right) \\
2936: & \leq & C \left( E_2(t) + \int_X \left( \left| \partial_t u \right|^2 + |u|^2
2937: \right) dx\right).
2938: \end{eqnarray*}
2939: Now since $E_2(T-1) = 0$ and $E_2(T) = E(T)$, Gronwall's inequality
2940: gives
2941: \begin{eqnarray}
2942: \label{ET-1}
2943: E(T) \leq C \left( \left\| \partial_t u \right\|_{L^2((T-1,T);L^2)}^2 +
2944: \| u \|_{L^2((T-1,T); L^2)}^2 \right).
2945: \end{eqnarray}
2946: We need to bound the first term on the right hand side of
2947: \eqref{ET-1}. Choose $\chi_3 \in \Ci( \reals)$ such that $\chi_3
2948: \equiv0$ for $t \leq T-2$ and $t \geq T+1$, $\chi_3 \equiv 1$ for $T-1
2949: \leq t \leq T$. Then
2950: \begin{eqnarray*}
2951: 0 & = & \int_{T-2}^{T+1} \int_X \chi_3^2 u \left( \partial_t^2 u -
2952: \Delta u + 2a \partial_t u \right) dxdt \\
2953: & = & \int_{T-2}^{T+1} \int_X -\chi_3^2 (\partial_t u)^2 - 2 \chi_3
2954: \chi_3' u \partial_t u + 2 \chi_3^2a \partial_t u + \chi_3^2 |
2955: \sqrt{-\Delta} u |^2 dxdt,
2956: \end{eqnarray*}
2957: whence
2958: \begin{eqnarray*}
2959: \left\| \partial_t u \right\|_{L^2((T-1, T);L^2)} \leq C \| u
2960: \|_{L^2((T-2, T+1); H^1)},
2961: \end{eqnarray*}
2962: giving
2963: \begin{eqnarray*}
2964: E(T) \leq C \|u \|_{L^2((T-2, T+1);H^1)}^2 \leq C e^{-T/C}
2965: \|f\|_{H^\epsilon(X)}^2
2966: \end{eqnarray*}
2967: as claimed.
2968: \end{proof}
2969:
2970: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2971: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2972:
2973:
2974:
2975:
2976:
2977:
2978:
2979:
2980:
2981:
2982:
2983:
2984:
2985:
2986:
2987:
2988:
2989:
2990:
2991:
2992:
2993:
2994:
2995:
2996:
2997:
2998:
2999:
3000:
3001:
3002:
3003:
3004:
3005:
3006:
3007:
3008:
3009:
3010:
3011:
3012:
3013:
3014:
3015:
3016:
3017:
3018:
3019:
3020:
3021:
3022:
3023:
3024:
3025:
3026:
3027:
3028:
3029:
3030:
3031:
3032:
3033:
3034:
3035:
3036:
3037:
3038:
3039:
3040:
3041:
3042:
3043:
3044:
3045:
3046: \begin{thebibliography}{\hspace{1cm}}
3047:
3048: \bibitem[AbRo]{AbRo}{\sc Abraham, R. and Robbin, J.} {\it Transversal Mappings and Flows}.
3049: W. A. Benjamin, Inc., New York, 1967.
3050:
3051: \bibitem[AbMa]{AbMa}{\sc Abraham, R. and Marsden, J.} {\it Foundations of Mechanics}.
3052: W. A. Benjamin, Inc., New York, 1967.
3053:
3054: \bibitem[AuMa]{AuMa}{\sc Aurich, R. and Marklof, J.} Trace Formulae
3055: for $3$-dimensional Hyperbolic Lattices and Application to a
3056: Strongly Chaotic Tetrahedral Billiard. {\it Phys. D.} {\bf 92},
3057: No. 1-2, 1996. p. 101-129.
3058:
3059: \bibitem[BoCh]{BoCh}{\sc Bony, J.-M. and Chemin, J.-Y.} Espaces fonctionnels associ\'{e}s au calcul de Weyl-H\"{o}rmander.
3060: {\it Bull. Soc. math. France.} {\bf 122}, 1994, p. 77-118.
3061:
3062: \bibitem[BFRZ]{BFRZ} {\sc Bony, J.-F., Ramond, T., Fujiie, S., and Zerzeri, M.} Quantum Monodromy for a Homoclinic Orbit. \\ {\tt http://www.math.u-psud.fr/$\sim$ramond/articles/qmho.pdf}
3063:
3064: \bibitem[Bur]{Bur}{\sc Burq, N.} Smoothing Effect for Schr\"{o}dinger Boundary Value Problems. {\it Duke Math. Journal.}
3065: {\bf 123}, No. 2, 2004, p. 403-427.
3066:
3067: \bibitem[BuZw]{BZ}{\sc Burq, N. and Zworski, M.} Geometric Control in the Presence of a Black Box.
3068: {\it J. Amer. Math. Soc.} {\bf 17}, 2004, p. 443-471.
3069:
3070: \bibitem[CVP]{CVP}{\sc Colin de Verdi\`ere, Y. and Parisse, B.} Equilibre Instable en R\'egime Semi-classique: I -
3071: Concentration Microlocale. {\it Commun. PDE.} {\bf 19}, 1994, p. 1535-1563;
3072: Equilibre Instable en R\'egime Semi-classique: II - conditions de Bohr-Sommerfeld.
3073:
3074: {\it Ann. Inst. Henri Poincar\'e (Phyique th\'eorique).} {\bf 61}, 1994, 347-367.
3075:
3076: %\bibitem[CVP3]{CVP3}{\sc Colin de Verdi\`ere, Y. and Parisse, B.} Singular Bohr-Sommerfeld Rules. {\it Commun. Math Phys.} {\bf 205}, 1999, 459-500.
3077:
3078: \bibitem[DiSj]{DiSj}{\sc Dimassi, M. and Sj\"{o}strand, J.} {\it Spectral Asymptotics in the Semi-classical Limit}.
3079: Cambridge University Press, Cambridge, 1999.
3080:
3081: \bibitem[Dui]{duistermaat}{\sc Duistermaat, J. J.} {\it Fourier Integral Operators}. Birkh\"{a}user, Boston, 1996.
3082:
3083: \bibitem[EvZw]{EvZw}{\sc Evans, L.C. and Zworski, M.} {\it Lectures on Semiclassical Analysis}. \\
3084: {\tt http://math.berkeley.edu/$\sim$evans/semiclassical.pdf}.
3085:
3086: \bibitem[GeSj]{GeSj}{\sc G\'{e}rard, C. and Sj\"{o}strand, J.} Semiclassical Resonances Generated by a Closed
3087: Trajectory of Hyperbolic Type. {\it Communications in Mathematical Physics.} {\bf 108}, 1987, p. 391-421.
3088:
3089: \bibitem[Gui]{guillemin}{\sc Guillemin, V.} Wave-Trace Invariants. {\it Duke Mathematical Journal}.
3090: {\bf 83}, No. 2, 1996, p. 287-352.
3091:
3092: \bibitem[Hit]{Hit}{\sc Hitrik, M.} Eigenfrequencies and Expansions for Damped Wave Equations.
3093: {\it Methods and Applications of Analysis.} {\bf 10}, No. 4, 2003, p. 543-564.
3094:
3095: \bibitem[HoZe]{hz}{\sc Hofer, H. and Zehnder, E.} {\it Symplectic Invariants and Hamiltonian Dynamics}.
3096: Birkh\"{a}user Verlag, Basel, 1994.
3097:
3098: \bibitem[Hor1]{hormander}{\sc H\"{o}rmander, L.} {\it The Analysis of Linear Partial Differential Operators I}.
3099: Springer-Verlag, Berlin, 1983.
3100:
3101: \bibitem[Hor3]{hormander3}{\sc H\"{o}rmander, L.} {\it The Analysis of Linear Partial Differential Operators III}.
3102: Springer-Verlag, Berlin, 1985.
3103:
3104: \bibitem[Hor5]{hormander5}{\sc H\"{o}rmander, L.} Symplectic
3105: Classification of Quadratic Forms, and General Mehler Formulas.
3106: {\it Mathematische Zeitschrift.} {\bf 219}, 1995. p. 413-449.
3107:
3108: \bibitem[IaSj]{IaSj}{\sc Iantchenko, A. and Sj\"{o}strand, J.} Birkhoff Normal Forms for Fourier Integral Operators II.
3109: {\it American Journal of Mathematics}. {\bf 124}, 2002. p. 817-850.
3110:
3111: \bibitem[ISZ]{ISZ}{\sc Iantchenko, A., Sj\"{o}strand, J., and Zworski,
3112: M.} Birkhoff Normal Forms in Semi-classical Inverse Problems. {\it
3113: Math. Res. Lett.} {\bf 9}, No. 2-3, 2002. p. 337-362.
3114:
3115: \bibitem[KaHa]{kaha}{\sc Katok, A. and Hasselblatt, B.} {\it Introduction to the Modern Theory of Dynamical Systems}.
3116: University of Cambridge Press, Cambridge, 1995.
3117:
3118: \bibitem[Kl]{Kl}{\sc Klingenberg, W.} {\it Riemannian Geometry}.
3119: Walter de Gruyter, Berlin, 1995.
3120:
3121: \bibitem[Leb]{Leb}{\sc Lebeau, G.} Equation des Ondes Amorties.
3122: {\em Algebraic and Geometric Methods in Mathematical Physics}, A Boutet de Monvel and V. Marchenko (eds.)
3123: Kluwer Academic Publishers, Netherlands, 1996. 73-109.
3124:
3125: \bibitem[Lee]{Lee}{\sc Lee, J. M.} {\it Introduction to Smooth Manifolds}. Springer-Verlag, New York, 2003.
3126:
3127: \bibitem[RT]{RT}{\sc Rauch, J. and Taylor, M.} Decay of Solutions to
3128: Nondissipative Hyperbolic Systems on Compact Manifolds.
3129: {\it Comm. Pure Appl. Math.} {\bf 28}, No. 4, 1975, p. 501-523.
3130:
3131: \bibitem[Sj\"{o}]{sjostrand}{\sc Sj\"{o}strand, J.} Semiclassical Resonances Generated by Non-degenerate Critical Points. {\it Pseudodifferential operators,
3132: Oberwolfach, 1986).} Lecture Notes in Math., 1256. Springer, Berlin, 1987. 402-429.
3133:
3134: \bibitem[Sj\"{o}2]{Sjo2a}{\sc Sj\"{o}strand, J.} Geometric Bounds on
3135: the Density of Resonances for Semiclassical Problems. {\it Duke
3136: Mathematical Journal.} {\bf 60}, No. 1, 1990. p. 1-57.
3137:
3138: \bibitem[Sj\"{o}3]{Sjo2}{\sc Sj\"{o}strand, J.} Asymptotic Distribution of Eigenfrequencies for Damped
3139: Wave Equations. {\it Publ. RIMS, Kyoto Univ.} {\bf 36}, 2000.
3140: 573-611.
3141:
3142: \bibitem[Sj\"{o}4]{Sjo3}{\sc Sj\"{o}strand, J.} Resonances Associated
3143: to a Closed Hyperbolic Trajectory in Dimension $2$. {\it
3144: Asymptot. Anal.} {\bf 36}, No. 2, 2003. p. 93-113.
3145:
3146:
3147: \bibitem[SjZw]{SjZw}{\sc Sj\"{o}strand, J. and Zworski, M.} Quantum Monodromy and Semi-classical Trace Formul{\ae}.
3148: {\it Journal de Math\'{e}matiques Pures et Appliques}. {\bf 81}, 2002. 1-33.
3149:
3150: \bibitem[SjZw2]{SjZw2}{\sc Sj\"{o}strand, J. and Zworski, M.} Fractal Upper Bounds on the Density of Semiclassical
3151: Resonances. \\
3152: {\tt http://xxx.lanl.gov/pdf/math.SP/0506307}.
3153:
3154: \bibitem[TaZw]{TaZw}{\sc Tang, S.H. and Zworski, M.} From Quasimodes to Resonances. {\it Math. Res. Lett.}
3155: {\bf 5}, 1998. 261-272.
3156:
3157: \bibitem[Tay]{Tay}{\sc Taylor, M.} {\it Pseudodifferential
3158: Operators}. Princeton University Press, Princeton, 1981.
3159:
3160: \bibitem[Wei]{Wei}{\sc Weinstein, A.} Neighborhood Classification of Isotropic Embeddings.
3161: {\it Journal of Differential Geometry.} {\bf 16}, 1981. 125-128.
3162:
3163: \bibitem[Wei2]{Wei2}{\sc Weinstein, A.} Symplectic Manifolds and Their Lagrangian Submanifolds.
3164: {\it Advances in Mathematics.} {\bf 6}, 1971. 329-346.
3165:
3166: \bibitem[Ze]{Ze}{\sc Zelditch, S.} Wave Invariants for Non-degenerate
3167: Closed Geodesics. {\it Geom. Func. Anal.} {\bf 8}, No. 1, 1998.
3168: p. 179-217.
3169:
3170: \end{thebibliography}
3171:
3172:
3173: \end{document}
3174: