1: \documentclass{amsart}
2:
3:
4: \parskip 0.2cm
5: \usepackage{amssymb}
6: \usepackage{amscd}
7: \usepackage{amsmath}
8: \usepackage{amsfonts}
9: \usepackage{a4}
10: \usepackage{amsthm}
11: \usepackage{color}
12:
13: %\usepackage[hypertex]{hyperref}
14: %\usepackage[active]{srcltx}
15:
16:
17: %\usepackage[inner]{showlabels}
18: %\usepackage[notref, notcite]{showkeys} beaucoup mieux que le pr{\'e}c{\'e}dent
19:
20:
21: \theoremstyle{plain}
22: \newtheorem{Hyp}{H}
23: \newtheorem{theo}{Theorem}[section]
24: \newtheorem{prop}[theo]{Proposition}
25: \newtheorem{lemme}[theo]{Lemme}
26: \newtheorem{cor}[theo]{Corollaire}
27: \newtheorem{defin}[theo]{Definition}
28:
29:
30: \theoremstyle{definition}
31:
32: \newtheorem{rem}[theo]{Remarque}
33: \newtheorem{exe}[theo]{Exemple}
34: \newtheorem{exo}[theo]{Exercice}
35: \newtheorem{commentaire}[theo]{Commentaire}
36:
37: \newcommand{\eval}[1]{\lvert_{#1}}
38:
39: \newcommand{\Hi}{{\mathcal{H}}}
40: \newcommand{\Ci}{C^{\infty}}
41:
42:
43: \newcommand{\R}{{\mathbb{R}}}
44: \newcommand{\N}{{\mathbb{N}}}
45: \newcommand{\C}{{\mathbb{C}}}
46: \newcommand{\Z}{{\mathbb{Z}}}
47:
48: \newcommand{\de}{{\delta}}
49: \newcommand{\be}{{\beta}}
50: \newcommand{\al}{{\alpha}}
51: \newcommand{\la}{{\lambda}}
52: \newcommand{\De}{{\Delta}}
53: \newcommand{\si}{{\sigma}}
54: \newcommand{\hb}{{\hbar}}
55: \newcommand{\ka}{{\kappa}}
56: \newcommand{\ga}{{\gamma}}
57: \newcommand{\te}{{\theta}}
58: \newcommand{\om}{{\omega}}
59: \newcommand{\Om}{\Omega}
60: \newcommand{\ph}{{\varphi}}
61: \newcommand{\Ph}{{\Phi}}
62: \newcommand{\Ga}{{\Gamma}}
63: \newcommand{\eps}{{\epsilon}}
64: \newcommand{\La}{{\Lambda}}
65: \newcommand{\Si}{{\Sigma}}
66:
67: \newcommand{\operateur}{{\mathcal{Q}}}
68: \newcommand{\identite}{{\operatorname{Id}}}
69: \newcommand{\Trace}{{\operatorname{Tr}}}
70: \newcommand{\trace}{{\operatorname{tr}}}
71: \newcommand{\volume}{{\operatorname{Vol}}}
72: \newcommand{\squar}{{\operatorname{sq}}}
73: \newcommand{\argument}{{\operatorname{Arg }}}
74: \newcommand{\diagonale}{{\operatorname{Diag}}}
75:
76: \newcommand{\Hilbert}{{\mathcal{H}}}
77: \newcommand{\Top}{\text{top}}
78: \newcommand{\Toeplitz}{{\mathcal{T}}}
79: \newcommand{\Lie}{{\mathcal{L}}}
80: \newcommand{\equivalence}{{\mathcal{E}}}
81: \newcommand{\lagrangien}{{\mathcal{L}}}
82: \newcommand{\Hessien}{{\operatorname{Hess}}}
83:
84: \newcommand{\contravariant}{{\operatorname{cont}}}
85: \newcommand{\normalised}{{\operatorname{norm}}}
86:
87: \newcommand{\End}{{\operatorname{End}}}
88:
89: \newcommand{\Lag}{{\operatorname{Lag}}}
90:
91: \newcommand{\Mas}{{\mathcal{M}}}
92:
93: \newcommand{\ideal}{{\mathcal{I}}}
94: \newcommand{\Sym}{\operatorname{Sym}}
95: \newcommand{\no}{{\mathcal{N}}}
96:
97: %\newcommand{\scratch}[1]{\textsl{[\textcolor{blue}{#1}]}}
98: \newcommand{\scratch}[1]{}
99:
100:
101:
102: \author{L. Charles}
103:
104: \address{Universit{\'e} Pierre et Marie Curie-Paris6, UMR 7586 Institut de
105: Math{\'e}matiques de Jussieu, Paris, F-75005 France.}
106:
107: \email{charles@math.jussieu.fr}
108:
109: \keywords{Geometric Quantization, Toeplitz operators, Bohr-Sommerfeld conditions, Half-form bundle}
110:
111: \subjclass{53D12, 53D50, 53D55, 81S30, 47L80, 35P20}
112:
113: \title{Symbolic calculus for Toeplitz operators with half-forms}
114:
115: \begin{document}
116:
117: \begin{abstract}
118: This paper is devoted to the use of half-form bundles in the symbolic
119: calculus of Berezin-Toeplitz operators on K{\"a}hler manifolds. We state
120: the Bohr-Sommerfeld conditions and relate them to the
121: functional calculus of Toeplitz operators, a trace formula and the
122: characteristic classes in deformation quantization. We also develop
123: the symbolic calculus of Lagrangian sections, with the crucial
124: estimates of the subprincipal terms.
125: \end{abstract}
126:
127: \maketitle
128:
129:
130: \bibliographystyle{plain}
131:
132: In semi-classical analysis we usually consider (pseudo) differential
133: operators depending on a small parameter and acting on a $L^2$ space,
134: the underlying classical limit being a cotangent space with its
135: canonical symplectic structure. In this paper we are interested in a
136: similar theory where the classical phase space is a compact K{\"a}hler
137: manifold endowed with a prequantum line bundle $L$. Here the quantum
138: Hilbert space consists of the holomorphic sections of $L^k$. The
139: small parameter is the inverse of the power $k$. The operators of
140: interest are the
141: Berezin-Toeplitz operators. This setting was mainly introduced by
142: Kostant \cite{Ko},
143: Souriau \cite{So} and Berezin \cite{Be} and the suitable microlocal techniques were
144: developed by Boutet de Monvel and Guillemin \cite{BoGu}. Since then many standard
145: results for pseudo-differential operators have been adapted to this
146: context, like for instance the Schnirelman theorem \cite{Ze}, the Gutzwiller trace formula
147: \cite{BP2} or
148: the Bohr-Sommerfeld conditions \cite{oim2}. The statements of these results are
149: easily predictable as far as only the symplectic structure of the
150: phase space is concerned, because they are the same
151: for the cotangent and K{\"a}hler spaces. But the semi-classical results
152: for the pseudo-differential operators may involve also some invariants,
153: like the subprincipal symbol or the Maslov index, which
154: do not only depend on the symplectic structure and consequently
155: are difficult to identify in the K{\"a}hler setting. Furthermore these
156: quantities generally appear as quantum corrections and are
157: difficult to compute. Nevertheless in the papers \cite{oim1} and \cite{oim2}, we
158: carried out successfully some techniques to handle this. To formulate our
159: result we used the Riemannian metric of the K{\"a}hler structure instead
160: of the vertical polarization of the cotangent bundle. Typically we
161: proved some Bohr-Sommerfeld conditions where the Maslov index is
162: replaced with a curvature integral. Actually we missed the right
163: formulation which uses the half-form bundles.
164: The main purpose of this paper and the sequel \cite{new} is to develop this point
165: of view. In this part we focus on the Bohr-Sommerfeld conditions
166: whereas \cite{new} is devoted to the dependence of the quantization on the complex structure.
167:
168: Concretely we alter the usual setting by defining the quantum space as the
169: space of holomorphic sections of $L^k \otimes L_1 \otimes \delta
170: \rightarrow M$. Here
171: $L$ is the prequantum bundle as previously, $L_1$ is an auxiliary
172: line bundle and $\delta$ is a half-form bundle, {\em i.e.} a square
173: root of the canonical bundle of $M$. A priori artificial, this
174: decomposition enlightens the semi-classical results, even in the usual
175: case where $L_1 \otimes
176: \delta$ is the trivial bundle. Roughly speaking the contribution of
177: $L_1$ in the semi-classical limit is the same as the one
178: of $L^k$ (one can view $L^k \otimes L_1$ as a first order
179: deformation of $L^k$) whereas the half-form bundle contributes in a
180: specific way. This principle will be confirmed in all our
181: results. Another important point is that there
182: is a topological obstruction to the existence of half-form bundles. To
183: avoid this problem we consider globally the bundle $L^k \otimes K$ and
184: write locally $K = L_1 \otimes \delta$.
185: The situation here is analogous to that in Riemannian geometry where we
186: think any Clifford module, at least locally, as the spinor bundle
187: twisted with an auxiliary bundle.
188:
189: The first section is devoted to basic properties of Toeplitz
190: operators and their symbolic calculus. In particular an important
191: subprincipal symbol is defined. We state the Bohr-Sommerfeld
192: conditions in section 2 and relate them to the symbolic calculus and
193: trace formula by adapting an argument of Colin de Verdi{\`e}re
194: \cite{Co}. Here the formulation with half-forms permits to check easily the
195: consistency of the results. The next sections contain the proof of the
196: Bohr-Sommerfeld conditions. In section 4, we introduce the Lagrangian
197: sections, which are similar to the Lagrangian distributions, and
198: develop their symbolic calculus. Bohr-Sommerfeld conditions follows immediately.
199: A comparison with the usual setting
200: is included, where a $\Z_4$-bundle plays a role analogous to the
201: Maslov bundle. In section 5, the technical part of the paper, we
202: provide the proof for the symbolic calculus of the Lagrangian
203: sections. We follow essentially the method of \cite{oim2} but avoid the
204: complicated computations involving the derivatives of the K{\"a}hler
205: metric. These simplifications rely on a version of the stationary
206: phase lemma stated in an appendix of this paper. In view of this
207: proof, we think that our result should generalize mutatis mutandis to the case where the
208: symplectic manifold doesn't admit any integrable complex structure.
209:
210:
211:
212: {\bf Acknowledgments} We thank Y. Colin de Verdi{\`e}re who provided us
213: his preprint \cite{Co} and suggested us to adapt his argument to the Toeplitz
214: operators. This was actually one of our original motivations to develop
215: the half-form formalism. We also thank F. Faure for his kind interest.
216:
217:
218:
219:
220:
221:
222:
223:
224:
225: \section{The setting}
226: \subsection{Square root of line bundle}
227:
228: Let $M$ be a manifold and $F \rightarrow M$ be a complex line bundle. A {\em square root} $(\delta,
229: \ph)$ of $F$ is a line bundle $\delta \rightarrow M$ together with an
230: isomorphism of line bundle $\ph : \delta^{\otimes 2} \rightarrow
231: F$. If $M$ is a complex manifold, a square root of its canonical bundle
232: $\La ^{n,0} T^*M$ is called a {\em half-form bundle}.
233: Let us state basic properties of square roots.
234:
235: If $F$ has a Hermitian structure and $(\delta,
236: \ph)$ is a square root of $F$, then $\delta$ has a unique Hermitian structure
237: such that $\ph$ is a isomorphism of Hermitian line bundle. In the same
238: way, if $F$ is holomorphic or flat, $\delta$ inherits the same
239: structure. If $D^F$ is a first order differential operator acting on
240: sections of $F \rightarrow M$, then there exists a unique first
241: order differential operator $D^\delta$ acting on section of $\delta$ such that
242: $$ D^F \ph (s \otimes s ) = 2 \ph ( s \otimes D^\delta s), \quad \forall \; s \in \Ci
243: (M,\delta).$$
244: A line bundle admits a square root if and only if its Chern class is divisible by 2 in $H^2(M, \Z)$.
245: Two square
246: roots $(\delta, \ph)$ and $(\delta' , \ph')$ of $F$ are {\em equivalent}
247: if there exists an isomorphism $\Psi : \delta \rightarrow
248: \delta'$ such that $\ph' \circ \Psi^2 = \ph$.
249:
250: \begin{prop} Assume that $F$ admits a square root. Then the set
251: of
252: equivalence classes of square roots of $F$ is a principal
253: homogeneous space for the first group of cohomology of $M$ with coefficient
254: in $\Z_2$.
255: \end{prop}
256:
257: \begin{proof}
258: First, if $(\delta, \ph)$ is a square root of the trivial line bundle
259: $1_M = M \times \C$, then $\delta$ inherits a flat structure from $1_M$
260: with structure group $\Z_2$. Furthermore this flat structure
261: determines $\ph$. It is easily proved that this induces an
262: isomorphism between the set of equivalence classes of square roots of
263: $1_M$ and the set of equivalences of flat line bundles with structure
264: group $\Z_2$. The latter is isomorphic to $H^1(M, \Z_2)$. Now, observe
265: that the tensor product of a square root of $L$ with a square root of
266: $1_M$ is a square root of $L$. This defines an action of $H^1(M,
267: \Z_2)$ on $\equivalence _L$, which is easily shown to be free and transitive.
268: \end{proof}
269:
270:
271: \subsection{Quantum spaces}
272: Let $M$ be a connected compact K{\"a}hler manifold of complex dimension
273: $n$. Denote by $\om \in \Om ^2 (M, \R)$ the fundamental form of $M$. Assume $M$ is endowed with a prequantization bundle $$L
274: \rightarrow M,$$ that is a
275: Hermitian line bundle with a connection $\nabla^L$ of curvature $\frac{1}{i}
276: \om$. Since $\om$ is a $(1,1)$-form, $L$ has a natural holomorphic
277: structure defined in such a way that the (local) holomorphic sections
278: satisfy the Cauchy-Riemann equations: $\nabla _{\bar{Z}} s
279: = 0$ for every holomorphic vector field $Z$ of $M$.
280:
281: Let $K \rightarrow M$ be a Hermitian holomorphic line bundle. For
282: every positive integer $k$ define the quantum space
283: $\Hilbert_k$ :
284: $$ \Hilbert_k = \bigl\{ \text{holomorphic section of } L^k \otimes K
285: \bigr\}.$$
286: Assume that $M$ carries a half-form bundle $(\delta, \varphi)$.
287: $\delta \rightarrow M$ inherits a Hermitian scalar product and a holomorphic structure from
288: $\La^{ n, 0 }T^* M$. Introduce the Hermitian holomorphic line bundle $L_1$
289: such that $$K = L_1 \otimes \delta$$
290: and let $\frac{1}{i} \om_1 $ be the curvature of the Chern connection of
291: $L_1$.
292:
293: Since $M$ is
294: compact, $\Hilbert_k$ is finite dimensional and it follows
295: from the Riemann-Roch-Hirzebruch theorem and
296: Kodaira vanishing theorem that
297: \begin{gather} \label{Riemann-Roch}
298: \operatorname{dim} \Hilbert_k = \Bigl( \frac{k}{2 \pi} \Bigr)^{n}
299: \int_M ( \om + k^{-1} \om_1 )^{\wedge n} /n! + O(k^{n-2}) \end{gather}
300: To interpret this formula, we consider $L^k \otimes L_1$ and $\om +
301: \hbar \om_1$ as
302: deformations of $L^k$ and $\om$ which give the first quantum
303: corrections in the semi-classical limit. Indeed the leading term
304: $$\bigl( \tfrac{k}{2 \pi} \bigr)^{n} \int \om^{\wedge n} /n!$$ gives
305: the second-order correction when we replace $\om$ with $\om + k^{-1}
306: \om_1$. Furthermore in the case $M$ doesn't carry any half-form
307: bundle, equation \eqref{Riemann-Roch} is still
308: valid if we define $\om_1$ by
309: $$ \om_1 := \om_K - \om_c/2,$$
310: where $\frac{1}{i} \om_K$
311: and $\frac{1}{i} \om_c $ are the curvatures of the Chern connections of
312: $K$ and $\La^{n,0} T^*M$.
313:
314:
315:
316: \subsection{Toeplitz operators}
317: Denote by $\Pi_k$ the orthogonal projector of $L^2(M,
318: L^k \otimes K)$ onto ${\mathcal{H}}_k$, where the scalar product of two sections of $L^k \otimes K$ is
319: defined from the Hermitian structures of $L$ and $K$ and the
320: Liouville form $\mu_M$.
321:
322: A Toeplitz operator is any sequence $(T_k: \Hilbert_k \rightarrow
323: \Hilbert_k )$ of operators of the form
324: $$ T_k = \Pi_k f(.,k) + R_k , $$
325: where $f(.,k)$ is a sequence of $\Ci( M)$ with an asymptotic expansion
326: $f_0 + k^{-1} f_1 +...$ for the $\Ci$ topology and the norm of $R_k$
327: is $O(k^{-\infty})$.
328:
329: The set $\Toeplitz$ of Toeplitz operators is a semi-classical algebra associated
330: to $(M, \om)$ in the following sense.
331:
332: \begin{theo} \label{sc_algebra}
333: $\Toeplitz$ is closed under the
334: formation of product. So it is a star algebra, the identity is
335: $(\Pi_k)$. The symbol
336: map $$\si_{ \contravariant} : \Toeplitz \rightarrow \Ci(M)[[\hb]],$$ sending $T_k$ into the
337: formal series $f_0 + \hb f_1 + ...$ where the functions $f_i$ are the
338: coefficients of the asymptotic expansion of the multiplicator
339: $f(.,k)$, is
340: well defined. It is onto and its kernel is the ideal consisting of
341: $O(k^{-\infty})$
342: Toeplitz operators. More precisely for any integer $\ell$,
343: $$ \| T_k \| = O(k^{-\ell} ) \text{ if and only if } \si_{\contravariant} (T_k)
344: = O( \hb^{\ell}). $$
345: Furthermore, the induced product $*_{\contravariant}$ on $\Ci (M) [[\hbar]]$ is a
346: star-product.
347: \end{theo}
348: Following the terminology of Berezin in \cite{Be}, we call $\si_{
349: \contravariant}$ the
350: contravariant symbol map. This result is essentially a consequence of the works of Boutet de Monvel and Guillemin
351: \cite{BoGu} (cf. also \cite{Gu}, \cite{BoMeSc} and \cite{oim1}).
352: Let us recall that equivalence classes of star-products on $(M,\om)$
353: are parametrized by elements in
354: $$ \frac{1}{i \hb} [\om] + H^2 (M, \C )[[\hb]] $$
355: called Fedosov characteristic classes. The following theorem was
356: proved by Karabegov and Schlichenmaier in \cite{ka} and \cite{KaSc},
357: in the case $K$ is the trivial line bundle.
358: \begin{theo} \label{Fedosov}
359: The Fedosov class of the star-product $*_{\contravariant}$ is $\frac{1}{i \hb} ( [\om] + \hb [\om_1])$.
360: \end{theo}
361: Again it is interesting to note the appearance of $\om + \hb
362: \om_1$. We do not need this result but some related facts.
363: Let us define the {\em normalized symbol} of a Toeplitz operator by
364: $$ \si_{\normalised} (T_k ) := (\identite + \tfrac{\hbar}{2} \Delta ) \si_{\contravariant} (T_k) $$
365: where $\Delta$ is the holomorphic Laplacian acting on $\Ci(M)$.
366: Actually we are only interested in the leading and second order terms
367: of $\si_{\normalised} (T_k)$ and modifying the definition of
368: $\si_{\normalised} (T_k)$ by a $O(\hb^2)$ term wouldn't change the statements
369: of our results. To compare with our previous article \cite{oim2}, the Weyl symbol that
370: we introduced when $K$ is
371: the trivial line bundle is equal to the normalized symbol modulo $O(\hb^2)$.
372:
373: The map $\si_{\normalised}: \Toeplitz \rightarrow \Ci(M)[[\hb]]$ satisfies
374: the same properties as $\si_{\contravariant}$ stated in theorem \ref{sc_algebra}. Denote
375: by $*_{\normalised}$ the associated star product.
376:
377: \begin{theo} \label{subprincipal_calculus}
378: Let $f$ and $g$ belong to $\Ci(M)[[\hb]]$, then
379: $$ f *_{\normalised} g = f.g + \tfrac{\hb}{2i} \langle \pi, df \wedge
380: dg \rangle + O(\hb^2) $$
381: and
382: $$ i \hb^{-1} \bigl( f *_{\normalised} g - g *_{\normalised} f
383: \bigr) = \langle \pi + \hb \pi_1 , df \wedge
384: dg \rangle + O(\hb^2).$$
385: where $\pi$ is the Poisson bivector and $\pi_1$ is the bivector such that
386: $\langle \pi_1, df \wedge dg
387: \rangle + \langle X_f \wedge X_g , \om_1 \rangle =0$
388: for every $f, g \in \Ci(M)$.
389: \end{theo}
390:
391: So $*_{\normalised}$ is a normalized star-product, in the sense that
392: the second order term in the first formula is antisymmetric, which explains
393: our terminology.
394: Observe that $\pi + \hb \pi_1$ is the Poisson bivector associated to $\om
395: + \hb \om_1$ in the sense that
396: $$ \langle \pi + \hb \pi_1, df \wedge dg
397: \rangle = \langle (X_f + \hb X_f^1) \wedge (X_g + \hb X_g^1) , \om +
398: \hb \om_1
399: \rangle +O(\hb^2),$$
400: where $X_f + \hb X_f^1+O(\hb^2)$ is
401: the Hamiltonian vector field of $f$ with respect to $\om +
402: \hb \om_1$, that is
403: $$df + \langle \om + \hb \om_1 , X_f + \hb X_f^1
404: \rangle = O(\hb^2),$$ and the same holds for $g$ and $X_g + \hb
405: X_g^1$. So it follows from theorem \ref{Fedosov} that there exists a star
406: product equivalent to $*$ satisfying the formulas of theorem
407: \ref{subprincipal_calculus}. This last result is more precise because the
408: equivalence is specified.
409: We can prove it using the methods of \cite{oim1} or
410: \cite{KaSc}. But this leads to complicated computations. We will present in
411: \cite{new} a more conceptual proof. We stated this result because we
412: can deduce from it a part of the Bohr-Sommerfeld conditions
413: (cf. sections \ref{trace}, \ref{func}, \ref{var}).
414:
415: \scratch{Last remark is that $\si_{\normalised}$ isn't the only symbol map which
416: defines a normalized star-product. Indeed consider a symbol
417: map $\si$ of the form
418: $$ \si (T_k) = (\identite + \hb P ) \si_{\normalised} (T_k) +O(\hb^2)$$
419: where $P$ is a differential operator.
420: Then the associated star-product $*$ is normalized if and only if $P$ acts
421: as a vector field $X$.}
422:
423:
424:
425: \subsection{Relation with geometric Quantization}
426:
427: Our definition of the normalized symbol agrees in some sense with the
428: usual procedure to quantize observables in geometric
429: quantization. Assume that the Hamiltonian flow of $f \in \Ci (M)$
430: preserves the complex structure of $M$. Assume also that $K = L_1
431: \otimes \delta$, where $(\delta, \varphi)$ is a half-form bundle.
432: Then the following operator is well-defined
433: \begin{gather}
434: \label{geometric_quantization}
435: \operateur (f) := f + \frac{1}{ik} \bigl( \nabla_{X_f}^{L^k \otimes L_1}\otimes
436: \identite + \identite \otimes \Lie_{X_f} \bigr) : \Hilbert_k \rightarrow
437: \Hilbert_k. \end{gather}
438: Here $\Lie_{X}$ acts on sections of
439: $\delta$ by $ \varphi ( (\Lie _X s) \otimes s) = \frac{1}{2} \Lie _X
440: \varphi ( s^{\otimes 2} )$, or
441: equivalently as a Lie derivative where the pull-back of sections of
442: $\delta$ by a
443: complex diffeomorphism $\zeta $ is defined in such a way that $\varphi
444: ((\zeta^* s)^{\otimes 2} ) =
445: \zeta^* \varphi (s^{\otimes 2} )$.
446:
447: The definition \eqref{geometric_quantization}
448: is natural for the following reason. Denote by $\Phi_t$ the
449: Hamiltonian flow of $X_f$. Let $\tilde{\Phi}_t$ be the lift
450: of $\Phi_t$ to $L^k \otimes L_1 \otimes \delta$
451: defined by the tensor product of the parallel transport along the trajectories of $X_f$ in $L^k
452: \otimes L_1$ and by the pull-back in $\delta$. Then the solution of the
453: Schr{\"o}dinger equation
454: $$ \frac{1}{ik} \frac{d}{dt} \Psi(.,t) + \operateur (f) \Psi (.,t) = 0$$
455: with initial condition $\Psi \in \Hilbert _k$ is given by
456: $$\Psi (x,t) = e^{\frac{k}{i} \int_0^t f ( \Phi_{s-t} (x) ) \sl{ds}} \tilde{\Phi}_{t} \bigr( \Psi
457: (\Phi_{-t}(x)) \bigl) .$$
458:
459: \scratch{La formule est v{\'e}rifi{\'e}e. Regarder dans le Woodhouse
460: s'il introduit qqch de ressemblant.}
461:
462: The important point for us is that $\operateur (f)$ is a Toeplitz
463: operators whose normalized symbol is $f$ modulo $O(\hb^2)$. We will prove a
464: more general result for every smooth function $f$, which simplifies
465: some further proofs.
466:
467: If $f$ is an arbitrary smooth function, formula \eqref{geometric_quantization} doesn't
468: make sense, because the Lie derivative with respect to $X_f$ doesn't
469: necessarily preserve $\Om^{n,0} (M)$. So we define for any vector field $X$ the
470: operator $D_X$,
471: $$ D_X \al = p (\Lie _X \al), \qquad \al \in \Om^{n,0} (M) $$
472: where $p$ is the projection from $\La^{n} T^*M \otimes \C$ onto
473: $\La^{n,0} T^*M$ with kernel the sum
474: $$ \La^{n-1,1}T^* M \oplus \La^{n-2,2}T^* M \oplus ... \oplus
475: \La^{0,n} T^* M .$$
476: Next we let $D_X$ act on the sections of $\delta$, as the first-order
477: differential operator such that $ 2 \varphi( s \otimes D_X s) = D_X
478: \varphi( s^2)$ .
479:
480: \begin{theo} \label{normalised_symbol_quantisation}
481: For any $f\in \Ci(M)$, the operator
482: $$ \operateur (f) := \Pi_k \Bigl( f + \frac{1}{ik} \bigl( \nabla_{X_f}^{L^k \otimes L_1}\otimes
483: \identite + \identite \otimes D_{X_f} \bigr) \Bigr) : \Hilbert_k \rightarrow
484: \Hilbert_k $$
485: is a Toeplitz operator with principal symbol $f$ and vanishing subprincipal symbol.
486: \end{theo}
487:
488: This theorem is a consequence of the following lemma and an argument
489: of Tuynman \cite{Tu}.
490:
491: \begin{lemme} \label{l1}
492: Let $s$ be a half-form, then
493: $$ D_{X_f} s = \nabla^{\delta}_{X_f} s + \tfrac{i}{2} (\Delta f) s $$
494: where $\nabla^\delta$ is the Chern connection of $\delta$ and $\Delta$
495: is the holomorphic Laplacian of $M$.
496: \end{lemme}
497:
498: It follows that
499: \begin{gather} \label{defe}
500: \nabla_{X_f}^{ L^k \otimes L_1}\otimes
501: \identite + \identite \otimes D_{X_f} = \nabla_{X_f}^{ L^k \otimes
502: L_1 \otimes \delta } + \tfrac{i}{2} (\Delta f) = \nabla_{X_f}^{ L^k \otimes
503: K} + \tfrac{i}{2} (\Delta f) \end{gather}
504: Now we have for every $\Psi \in \Hilbert_k$,
505: $$ \Pi_k ( \nabla_{X_f}^{ L^k \otimes
506: K} \Psi) = \tfrac{1}{i} \Pi_k ( \Delta f . \Psi ), $$
507: cf. \cite{Tu} or \cite{BoMeSc} for a proof. Hence
508: $$ \operateur (f) \Psi = \Pi_k \bigl(
509: ( f - \tfrac{1}{2k} \Delta f ) \Psi \bigr)
510: $$
511: which proves theorem \ref{normalised_symbol_quantisation}.
512:
513: The last expression in \eqref{defe} shows that the definition of
514: $\operateur (f)$ is
515: independent of the choice of the half-form bundle and generalizes in
516: the cases where no such bundle exists.
517:
518: \begin{proof}[Proof of lemma \ref{l1}]
519: It suffices to prove that for every $\alpha \in \Om^{n,0} (M)$, we have
520: $$ D_{X_f} \al = \nabla_{X_f} \al + i (\Delta f) \al $$
521: Introduce normal complex coordinates $z^1,...,z^n$ centered at
522: $x_0$. So if $\alpha = dz^1 \wedge ... \wedge dz^n$, then $\nabla \al =
523: 0$ at $x_0$. Let us write $\om = i G_{j,k} dz^j \wedge d
524: \bar{z}^k$. Then
525: $$X_f = - i G^{j,k} (\partial_{z^j} f) \partial_{\bar{z}^k} + i
526: G^{j,k}(\partial_{\bar{z}^k} f) \partial_{z^j}
527: $$
528: Using that the first derivatives of $G^{j,k}$ vanish at $x_0$, we
529: obtain easily
530: $$ D_{X_f} \al = i G^{j,k} (\partial_{z^j}\partial_{\bar{z}^k} f) = i
531: \Delta f.$$
532: The result follows.
533: \end{proof}
534:
535: %\newpage
536: \section{Bohr-Sommerfeld Conditions} \label{BS_section}
537:
538: \subsection{The result}
539:
540: Assume that $M$ is $2$-dimensional. Let $(\delta,\varphi)$ be a
541: half-form bundle and let us write $K = L_1 \otimes \delta$ as
542: previously. Consider a self-adjoint Toeplitz
543: operator $(T_k)$. Its normalized symbol $$f_0 + \hb f_1 +...$$ is
544: real-valued. Bohr-Sommerfeld conditions give the spectrum of $T_k$ on
545: every open interval $I$ of regular values of $f_0$ in the
546: semi-classical limit. To simplify the statements, assume that $f_1$ vanishes.
547:
548: Let $\Ga^1$, ...,
549: $\Ga^m$ be the components of $f_0^{-1}(I)$. For every $i \in
550: \{ 1,...,m\}$, the map $$f_0 : \Ga^i \rightarrow I$$ is a trivial
551: fibration with fiber diffeomorphic to $
552: S^1$. For every $\la \in I$, fix an orientation on the fiber
553: $\Ga^{i}_\la = f_0^{-1} (\la) \cap \Ga^i$ depending continuously on
554: $\la$.
555:
556: Let $a^i \in \Ci(I)$ be the {\em principal action}, defined in such a
557: way that the parallel transport in $L$ along $\Ga^i_\la$ is the
558: multiplication by $\exp \bigl( {i a^i(\la)}\bigr)$. Using $L_1$
559: instead of $L$, define in the same way the
560: {\em subprincipal action} $a^i_1 \in \Ci(I)$.
561:
562: Let us define an index $\epsilon^i$ from the half-form bundle
563: $(\delta, \varphi)$. Observe that the restriction $\delta_{i, \la}$ of
564: $\delta$ to $ \Ga^{i}_\la$ is a square root of $T^*
565: \Ga^{i}_\la \otimes \C$. Indeed, let us denote by $\iota$ the embedding $\Ga^{i}_\la
566: \rightarrow M$, then the map
567: $$ \varphi_{i,\la} :\delta_{i, \la}^2 \rightarrow T^*
568: \Ga^{i}_\la \otimes \C, \qquad u \rightarrow \iota^* \varphi (u)$$
569: is an isomorphism of line bundle. The set
570: $$ \{ u \in \delta_{i,\la}; \; \varphi_{i,\la} (u ^{\otimes 2}) >0 \} $$
571: has one or two connected components. In the first case,
572: we set $\epsilon^{i}_\la =1$ and in the second case $\epsilon^{i}_\la
573: =0$. Observe that $\epsilon^{i}_\la$ doesn't depend on $\la$.
574:
575: The Bohr-Sommerfeld conditions are
576: \begin{gather} \label{BS_conditions}
577: a^i(\la) + k^{-1}( a^i_1 (\la) + \epsilon^i \pi ) \in \frac{2 \pi}{k} \Z \end{gather}
578: Denote by $\Si^i(k)$ the set of $\la \in I$ satisfying
579: \eqref{BS_conditions}. When $k$ is sufficiently large, $\Si^i(k)$ is a
580: finite set containing
581: $$\frac{k}{2\pi} \volume (\Ga^i ) + O( k^{-1})$$ points.
582: Let $\Si(k)$ be the union of the $\Si^i(k)$. Define the
583: {\em multiplicity} of $\la \in \Si(k)$ as the number of $\Si^i(k)$ which contains
584: $\la$. The points of $\Si (k)$ approximate the eigenvalues of $T_k$
585: in the following sense.
586:
587: \begin{theo} \label{BS_theo}
588: Let $\la_-(k)$ and $\la_+(k)$ be two sequences of $I$ such that
589: \begin{gather} \label{E1}
590: d ( \la_-(k) , \Si (k)) \geqslant C k^{-1}, \quad d ( \la_+(k) ,
591: \Si (k))
592: \geqslant C k^{-1} \end{gather}
593: for some positive $C$. Assume furthermore that there exists $\la_-,
594: \la_+ \in I$ such that
595: \begin{gather} \label{E2}
596: \la_- \leqslant
597: \la_-(k) \leqslant \la_+(k) \leqslant \la_+ .\end{gather}
598: Denote by $\la_1(k) \leqslant \la_2(k) \leqslant ....\leqslant \la_{N(k)}(k)$
599: (resp. $\la'_1(k) \leqslant \la'_2(k) \leqslant ....\leqslant \la'_{N'(k)}(k)$)
600: the eigenvalues of $(T_k)$ (resp. points of $\Si (k)$) contained in
601: $(\la_-(k), \la_+(k))$ and counted with multiplicities. Then, when $k$ is
602: sufficiently large, $N(k) = N'(k)$. Furthermore
603: \begin{gather} \label{Approximation}
604: \la_j(k) = \la_j' (k) + O(k^{-2}) \end{gather}
605: uniformly with respect to $j$.
606: \end{theo}
607:
608: The interest of condition \eqref{E1} is to avoid any ambiguity in the counting of eigenvalues near the
609: endpoints of $(\la_- (k), \la_+ (k))$. It is not restrictive. Indeed if
610: $\la_-(k)$, $\la_+ (k)$ are arbitrary sequences satisfying \eqref{E2},
611: then by modifying them by suitably chosen $O(k^{-1})$ sequences, we
612: obtain sequences satisfying both estimates \eqref{E1} and \eqref{E2}.
613:
614: Since the definition of the Toeplitz operators and of their normalized
615: symbol only depend on $K$ and not on the choice of the half form bundle, it is likely that the same holds for the Bohr-Sommerfeld
616: conditions. This is easily checked using that any other half-form bundle is of the form $\delta' =\delta \otimes
617: F$, where $F$ is a flat Hermitian line bundle with holonomy in $\Z_2$. So
618: $L_1' = L_1 \otimes F^{-1}$, and straightforward computations show
619: that the functions $a^i_1 +
620: \epsilon^i$ do not depend on the choice of $\delta$.
621:
622: To compare with our previous results in \cite{oim2}, when $K$ is
623: the trivial bundle, we defined the function $a^i_1 +
624: \epsilon^i$ as the integral of the geodesic curvature of
625: $\Ga_{\la}^i$.
626:
627: We can also approximate the eigenvalues up to a $O(k^{-\infty})$ error. More precisely, there exist
628: sequences $(S^i(.,k))_k$ of $\Ci (I)$ such that the Bohr-Sommerfeld
629: conditions $$S^i(.,k) \in \tfrac{2 \pi}{ k} \Z$$
630: instead of \eqref{BS_conditions} lead to the same
631: result with a $O(k^{-\infty})$ error in \eqref{Approximation}. Furthermore, the sequences $S^i(.,k)$ admit asymptotic
632: expansions of the form
633: $$S_0^i + k^{-1} S^i_1 + k^{-2} S^i_2 +...$$
634: The leading and second order terms are the same like in \eqref{BS_conditions}. Applying an argument of Colin de
635: Verdi{\`e}re \cite{Co}, we can prove that the derivatives of the $S^i_j$ only depend on the star-product $*_\normalised$ and the
636: normalized symbol of $T_k$. In particular, assuming that $S_0^i$ is
637: increasing, we will deduce from
638: proposition \ref{subprincipal_calculus} that necessarily
639: \begin{gather} \label{Colin}
640: S^i_0 (\la) - S^i_0 (\la') = \int_{D^i} \om \quad \text{and} \quad S^i_1 (\la ) - S^i_1 (\la') = \int_{D^i} \om_1 \end{gather}
641: if $\la , \la' \in I$ are such that $\la' \leqslant \la$ and $D^i =
642: \Ga^i \cap f_0^{-1} ((\la , \la'))$.
643: This determines the Bohr-Sommerfeld conditions \eqref{BS_conditions} modulo a
644: constant term.
645: Note that only
646: the derivatives of $S^i_0$ and $S^i_1$ are determined by the
647: star-product. Indeed, if we twist $L$ or $L_1$ by a flat Hermitian
648: line bundle, the actions of the non-contractible loops may change although the star-product remains the same.
649:
650: As a first step to deduce \eqref{Colin} from proposition
651: \ref{subprincipal_calculus}, let us state some results on traces
652: and functional calculus of Toeplitz operators.
653:
654: \subsection{Traces} \label{trace}
655:
656: In this section and the next one, we do not necessarily assume that $M$
657: is $2$-dimensional.
658: It is a known result that the trace of a Toeplitz operator $T_k$ with normalized symbol $f_0 +
659: \hb f_1 +...$ admits an asymptotic
660: expansion of the following form:
661: $$
662: \Trace ( T_k) = \Bigl( \frac{k}{2 \pi} \Bigr)^n \int_M (f_0 +
663: k^{-1} f_1 +...) (1+ k^{-1} d_1 + k^{-2} d_2 +...) \mu_M
664: +O(k^{-\infty}) $$
665: where $d_1$, $d_2$, $d_3$, ... are functions of $\Ci(M)$ which do not depend of
666: $T_k$ (cf. for instance \cite{oim1}).
667:
668: These functions may be computed in terms of the K{\"a}hler metric
669: by using the methods of \cite{oim1}, but it is more convenient to relate them to the star-product $*_\normalised$.
670: To do this, observe that the $\C[[\hb]]$-linear map
671: \begin{gather} \label{Trace} \trace: \Ci(M) [[\hb ]] \rightarrow \C[\hb^{-1}, \hb]], \quad f(\hb)
672: \rightarrow (2 \pi \hb )^{-n} \int_M f(\hb) d(\hb)\; \mu_M \end{gather}
673: where $d( \hb) =1 + \hb d_1 + \hb^2 d_2 +...$ is a trace for the
674: star-product $*_{\normalised}$, in the sense that it satisfies
675: $$\trace (f
676: *_{\normalised} g) = \trace( g *_{\normalised} f).$$
677: Following Fedosov
678: \cite{Fe} or Nest-Tsygan \cite{NeTs}, such a trace is
679: unique up to multiplication by an element of $\C [\hb^{-1}, \hb]]$ and
680: there exists a canonical one determined by the
681: following normalization condition: for every local equivalence $\Phi$
682: between $*_{\normalised}$ and the
683: Weyl star-product, we have
684: \begin{gather} \label{trace_normalised}
685: \trace (f) = (2 \pi \hb)^{-n} \int_M \Ph (f)\; \mu_M, \quad \text{
686: where } \mu_M = \om^n/n!. \end{gather}
687: We claim that the trace defined in
688: \eqref{Trace} is the canonical trace. This follows from the fact that
689: the quantization by Toeplitz operators is microlocally equivalent to
690: the usual Weyl quantization.
691:
692: So the functions $d_i$ are determined by $*_\normalised$. In
693: particular, it follows from
694: proposition \ref{subprincipal_calculus} that
695: \begin{gather} \label{sub_trace}
696: \trace (f) = (2 \pi \hb)^{-n} \int_M f \; ( \om + \hb \om_1 )^{\wedge
697: n}/ n! + O( \hb ^{-n +2}) \end{gather}
698:
699: \begin{proof}[Proof of formula \eqref{sub_trace}] Consider an equivalence of star-product of the form
700: $$\Phi = \identite + \hb X + O(\hb^2)$$ where $X$ is a vector field. It is easily
701: checked that the star-product
702: $$ f *' g = \Phi \bigl( \Phi^{-1} (f) *_{\normalised} \Phi^{-1} (g) \bigr)$$
703: is normalized and satisfies
704: $$ i \hb^{-1} \bigl( f *' g - g *' f
705: \bigr) = \langle \pi + \hb( \pi_1+ X.\pi) , df \wedge
706: dg \rangle + O(\hb^2).$$
707: Locally on can choose $X$ in such a way that $ \pi_1 + X.\pi = 0$. Indeed this
708: equation is equivalent to
709: $\om_1 + d\al = 0,$ where $\al$ is the
710: $1$-form such that $ \om(X,.) = \al$.
711:
712: Next step is to introduce local Darboux coordinates, which define a Weyl star-product $*_{\operatorname{Weyl}}$. And modifying $\Phi$
713: by a $O(\hb^2)$ term, one has $*_{\operatorname{Weyl}} = *'$
714: (cf. \cite{BeCaGu}). Then it follows from
715: \eqref{trace_normalised} that
716: $$ \trace (f) = (2 \pi \hb)^{-n} \int_M f\; \bigl(\mu_M - \hb (X\mu_M)
717: +O(\hb^2)\bigr) .$$
718: By definition of $\pi_1$ in theorem
719: \ref{subprincipal_calculus}, one has $\langle \pi_1 , \om \rangle +
720: \langle \pi, \om_1 \rangle = 0 $. Since $X. \langle \pi, \om \rangle
721: =0$, we obtain
722: $$\langle \pi , X \om \rangle = - \langle X \pi, \om
723: \rangle = \langle \pi_1, \om \rangle = - \langle \om_1, \pi \rangle.$$
724: Then a straightforward computation leads to
725: $$\langle \pi^ {\wedge n}, \mu_M - \hb X
726: \mu_M \rangle = \langle \pi^{\wedge n}, ( \om + \hb \om_1 )^{\wedge
727: n}/ n! \rangle + O(\hb^2),$$
728: which proves the result.
729: \end{proof}
730:
731: As a consequence of \eqref{sub_trace}, we obtain the estimate
732: \eqref{Riemann-Roch} of the dimension of $\Hilbert_k$, since this
733: dimension is the trace of the projector $\Pi_k$, whose normalized symbol is
734: $1$.
735: Actually the index theorems of deformation quantization proved
736: in \cite{Fe} or \cite{NeTs} yield the asymptotic expansion of $\Trace (
737: \Pi_k)$ modulo $O(k^{-\infty})$ in terms of the Fedosov class
738: of $*_\normalised$. Since this trace is an integer, the
739: $O(k^{-\infty})$ error vanishes when $k$ is sufficiently
740: large. In this way we can deduce Riemann-Roch-Hirzebruch theorem from
741: theorem \ref{Fedosov} .
742:
743:
744: \subsection{Functional calculus} \label{func}
745:
746: Let $T_k$ be a self-adjoint Toeplitz operator with normalized symbol
747: $$f =f_0 + \hb f_1 +....$$ and let $g$ be a function of $\Ci
748: (\R, \C)$. Then it is known that $g(T_k)$ is a Toeplitz
749: operator (cf. for instance \cite{oim1}). Furthermore, the normalized symbol of $g(T_k)$ is given by
750: the following non-commutative Taylor formula:
751: \begin{gather} \label{NC_Taylor}
752: g^{*_{\normalised}}(f) (x)= \sum \frac{1}{\ell!}g^{(\ell)} (f_0 (x)) \; ( f(y) - f_0 (x)
753: )^{*_{\normalised} \ell} \eval{y = x}, \end{gather}
754: where $g^{(\ell)}$ is the $\ell$th derivative of $g$ and $h^{*_{\normalised}
755: \ell} = h*_{\normalised}...*_{\normalised} h$ repeated $\ell$ times.
756: In particular, an easy computation
757: from proposition \ref{subprincipal_calculus} leads to
758: \begin{gather} \label{sub_functional_calculus}
759: g^{*_{\normalised}}(f)= g(f_0) + \hb g'(f_0)f_1 +
760: O(\hb^2). \end{gather}
761: \scratch{
762: Notice that formula \eqref{NC_Taylor} makes sense for an arbitrary
763: star product $*$. This defines a formal functional calculus
764: satisfying the expected properties. For instance, one can check that $g_1^* (f) * g_2^*(f) = (g_1
765: g_2)^{*}(f)$. }
766:
767: \subsection{On the variation of $S_0$ and $S_1$} \label{var}
768:
769: Let us deduce formulas \eqref{Colin} on the variations of $S_0$ and
770: $S_1$ from the trace formula \eqref{sub_trace} and the functional symbolic calculus.
771: Assume that
772: $f_0^{-1}(I)$ is connected.
773:
774: If $g \in \Ci_o (I, \C)$, then we deduce from
775: \eqref{sub_functional_calculus}, \eqref{sub_trace} and the fact that
776: $f_1 =0$ that
777: \begin{xalignat}{1} \notag
778: \Trace (g(T_k)) = & \frac{k}{2 \pi} \int_M (g \circ f_0) ( \om + k^{-1}
779: \om_1 ) \; + O( k^{-1})\\\label{eq1}
780: =& \frac{k}{2 \pi} \int_I g \; \bigl( f_{0*} [ \om + k^{-1}
781: \om_1 ] \bigr) \; + O( k^{-1})
782: .\end{xalignat}
783: where $f_{0*}$ is the push-forward $\Om^2 (M) \rightarrow \Om^1 (I)$
784: defined by
785: $$ \int_M (f_0^* h ) \ \al = \int_I g. f_{0*} \al, \qquad \forall h
786: \in \Ci_o(I) .$$
787:
788:
789: On the other hand, assume the Bohr-Sommerfeld condition is $$ S(\la,k) \in 2 \pi k^{-1} \Z $$ with
790: $S(\la,k) =S_0 + k^{-1}S_1 + O(k^{-2})$. If the derivative of $S_0$
791: doesn't vanish, then one can invert the functions $S(.,k)$ when $k$ is
792: sufficiently large and
793: $$ \Trace (g(T_k)) = \sum_{x \in 2\pi k^{-1} \Z} g ( S^{-1}(x,k))
794: + O(k^{-1}) $$
795: Interpreting this as a Riemann sum, it follows that
796: \begin{xalignat}{1} \notag
797: \Trace (g(T_k)) =& \frac{k}{2 \pi} \int_\R g (S^{-1}(x,k )) dx
798: +O(k^{-1}) \\ \label{eq2}
799: = & \frac{k}{2 \pi} \int_I g (\la) S' (\la, k ) d\la
800: +O(k^{-1})
801: \end{xalignat}
802: with the same orientation of $I$ as before if $S'_0$ is positive.
803: Since \eqref{eq1} and \eqref{eq2} hold for any function $g$ of $\Ci_o( I,
804: \C)$, we have
805: $$ d S_0 + \hb d S_1 = f_{0*} ( \om + \hb \om_1 )$$
806: and \eqref{Colin} follows.
807:
808:
809:
810:
811:
812:
813:
814:
815: %\newpage
816: \section{Lagrangian states}
817:
818:
819: \subsection{Definitions and symbolic calculus} \label{sec:resultat}
820:
821: First we recall the definition of a local Lagrangian section
822: associated to a closed Lagrangian embedding $\iota: \Ga \rightarrow M$.
823:
824: Let $U$ be an
825: open set of $M$ such that $U_\Gamma := \iota^{-1} (U )$ is contractible. Since the
826: curvature of $\iota^* L$ vanishes, there exists a
827: flat unitary section $t_\Gamma$ of $ \iota^* L \rightarrow U_\Gamma$.
828: Introduce a formal
829: series
830: $$\sum_{\ell =0 }^{\infty} \hb^\ell g_\ell \in \Ci ( U_\Gamma, \iota^*K)[[\hb]].$$
831: Let $V$ be an open set of $M$ such that $\overline{V} \subset U$.
832: Then a
833: sequence $\Psi_k \in {\mathcal{H}}_k$ is a {\em Lagrangian
834: section}
835: over $V$ associated to $(\Gamma, t_\Gamma)$ with symbol $\sum \hb^\ell
836: g_\ell$ if
837: \begin{gather*} % \label{Ansatz_Lag}
838: \Psi_k(x) = \Bigl( \frac{k}{2\pi} \Bigr)^{\frac{n}{4}} F^k (x)
839: \tilde{g} (x,k) + O(k^{-\infty}) \text{ over $V$, } \end{gather*}
840: where
841: \begin{itemize}
842: \item $F$ is a section of $L \rightarrow U$ such that
843: $$\iota^* F =
844: t_\Gamma \quad \text{ and } \quad \bar{\partial} F \equiv 0$$ modulo a section which
845: vanishes to every order along $\iota (\Gamma)$. Furthermore $|F(x)|< 1$ if $x \notin \iota(\Gamma)$.
846: \item
847: $\tilde{g}(.,k )$ is a sequence of $\Ci ( U,K)$
848: with an asymptotic expansion $\textstyle{\sum} k^{-\ell} \tilde{g}_{\ell}$
849: in the $\Ci$ topology such that
850: $$\iota^* \tilde{g}_\ell = g_\ell \text{
851: and } \quad \bar{\partial} \tilde{g}_\ell \equiv 0$$ modulo a section which vanishes at
852: every order along $\iota(\Gamma)$.
853: \end{itemize}
854: We assume furthermore that $\Psi_k$ is admissible in the sense that $\Psi_k(x)$ is uniformly
855: $O(k^{N})$ for some $N$ and the same holds for its successive
856: covariant derivatives.
857:
858:
859: It is not obvious that such a sequence exists.
860: \begin{theo} \label{existence}
861: For every series $\sum \hb^\ell g_\ell$ of $\Ci ( U_\Gamma,
862: \iota^* K)[[\hb]]$, there exists a Lagrangian
863: section over $V$ associated to $(\Gamma, t_\Gamma)$ with symbol $\sum \hb^\ell
864: g_\ell$. It is unique modulo a section which is $O(k^{-\infty})$
865: over $V$. \end{theo}
866:
867: In the statement of the following theorems, we consider that $K = L_1
868: \otimes \delta $ over $U$, where $(\delta, \varphi)$ is a half-form bundle. Recall that
869: $\iota^* \delta = \delta_{\Ga}$ is a square root of $\La ^n T^*\Ga \otimes
870: \C$ through the isomorphism
871: $$ \varphi_{\Ga} : \delta^2_ \Ga
872: \rightarrow \La ^n T^*\Ga \otimes \C, \quad u \rightarrow \iota^*
873: \varphi( u).$$
874: Let us associate
875: to the {\em principal symbol} $g_0$ of a Lagrangian section a density
876: $m(g_0)$ where $m$ is the map:
877: $$ m : \iota^* L_1 \otimes \delta_{\Ga} \rightarrow | \La | (\Ga) , \quad
878: u \otimes v \rightarrow \| u \|^2_{L_1} \ | \varphi_{\Ga} ( v^{\otimes
879: 2})|$$
880: We then have the following estimate of the norm of $\Psi_k$.
881: \begin{theo} \label{norme}
882: Let $\xi \in \Ci_o ( V)$, then we have
883: $$ \int_M \xi \ \| \Psi_k \|^2 _{L^k \otimes L_1 \otimes \delta } \; \mu_M
884: = \int_{\Gamma} (\iota^* \xi) \ m(g_0) +
885: O(k^{-1}).$$
886: \end{theo}
887: Next results describe how a Toeplitz operator acts on a Lagrangian
888: section.
889:
890: \begin{theo} \label{Toep_Lag}
891: Let $T_k$ be a Toeplitz operator with principal symbol
892: $f_0$. Then $T_k \Psi_k$ is a Lagrangian section over $V$ associated
893: to $(\Gamma, t_\Gamma)$ with symbol $(\iota^* f_0) g_0 +O(\hb)$.
894: \end{theo}
895:
896:
897: To prove the Bohr-Sommerfeld conditions, we need to compute
898: the subsequent coefficient of the symbol of $T_k \Psi_k$, in the case
899: where $f_0$ is constant over $\Ga$.
900:
901: \begin{theo} \label{resultat_clef}
902: Let $T_k$ be a Toeplitz operator with normalized symbol $f_0 + \hb
903: f_1 + O(\hb^2)$. Assume that $f_0$ is constant along $\Gamma$. Then
904: the symbol of $ T_k \Psi_k$ is
905: $$ (\iota^* ( f_0 + \hb f_1)).( g_0 + \hb g_1) + \hb \tfrac{1}{i} (
906: \nabla_{X}^{\iota^* L_1}
907: \otimes \identite + \identite \otimes \Lie^{\delta_{\Ga}}_X ).g_0 + O(\hb^2) $$
908: where
909: \begin{itemize}
910:
911: \item $X$ is the Hamiltonian vector field of $f_0$,
912: \item $\nabla^{\iota^* L_1}$ is the pull-back of the Chern connection of $L_1$,
913: \item $ \Lie^{\delta_{\Ga}}_X$ is the first order differential operator
914: acting on sections of $\delta_{\Ga}$
915: such that $$ \varphi_{\Ga} ( \Lie^{\delta_{\Ga}}_X g \otimes g) = \tfrac{1}{2}
916: \Lie_X \varphi_{\Ga} ( g ^{\otimes 2}) $$
917: for every section $g$.
918: \end{itemize}
919: \end{theo}
920:
921: It is easily checked that the operator $\nabla_{X}^{\iota^* L_1}
922: \otimes \identite + \identite \otimes \Lie^{\delta_{\Ga}}_X$ doesn't
923: depend of the choice of the half-form bundle if we consider that it
924: acts on sections of $\iota^* K = \iota^* L_1 \otimes
925: \delta_\Ga$. The
926: same holds with the map $m$.
927:
928: \subsection{Proof of Bohr-Sommerfeld conditions}
929: Let us deduce from the previous theorems the Bohr-Sommerfeld conditions for $n$
930: self-adjoint commuting Toeplitz operators $T^1$, $T^2$,..., $T^n$, which is a slight
931: generalization of \eqref{BS_conditions}.
932:
933: Denote by $f^i_0$ and $f^i_1$
934: the principal and subprincipal symbols of $T^i$. Let $E$ be a
935: regular value of $f=(f^1_0,...,f^n_0)$ and $\iota : \Ga \rightarrow M$ be
936: an embedding with image a connected component
937: of $f^{-1} (\{ E\})$. Since $M$ is compact, $\iota (\Ga)$ is a Lagrangian
938: torus. So there exists a half-form bundle $(\delta, \varphi)$ defined over
939: a neighborhood of $\iota (\Ga)$. It is not unique but as usual the
940: final result doesn't depend on the choice of $(\delta, \varphi)$.
941: Introduce like in the previous section two open sets $U,V$ and a flat section
942: $t_\Ga$. Let us try to solve the eigenvalues equation
943: \begin{gather} \label{vap_equation}
944: T^i \Psi = E^i \Psi + O(k^{-\infty}) \text{ over $V$} \end{gather}
945: where $\Psi$ is a Lagrangian section associated to $(\Ga, t_{\Ga})$. By
946: theorem \ref{resultat_clef}, the symbol of $(T^i
947: -E^i) \Psi$ is $O(\hb)$ because $\iota ^* f^i = E^i$. Furthermore it
948: is $O(\hb^2)$ if and only if it satisfies the following transport equation
949: \begin{gather} \label{transport_equation}
950: \bigr[ f^i_1 + ( \nabla_{X^i}^{\iota^*L_1 }
951: \otimes \identite + \identite \otimes \Lie^{\delta_{\Ga}}_{X^i} ) \bigl] g_0 = 0
952: \text{ over $V \cap \Ga$} \end{gather}
953: where $g_0$ is the principal symbol of $\Psi$. This equation can be
954: interpreted as $g_0$ being flat for a connection on $ \iota^* L_1 \otimes \delta_{\Ga}$ that we
955: describe now.
956:
957: First a section $g$ of $\delta_{\Ga}$ is flat if $\mathcal{L}_{X^i}^{\delta_{\Ga}} g
958: = 0$ for every $i$. With this definition, $\varphi_{\Ga} : \delta_{\Ga}^2
959: \rightarrow \La ^n T^*\Ga \otimes \C $ is a morphism of flat bundles, if
960: we endow $\La ^n T^* \Ga \otimes \C $ with the Weinstein connection. Since
961: the form $\be$ such that
962: $$ \langle \be, X^1 \wedge ... \wedge X^n \rangle =1 $$
963: is a global non vanishing flat section of $\La ^n T^* \Ga \otimes \C $, $\delta_\Ga$ has holonomy in $\Z_2 \subset U(1)$.
964:
965: Let $\al \in \Om^1(\Gamma)$ be such that $ \frac{1}{i} \langle \al ,
966: X^j \rangle = f^j_1$. Consider the connection $\nabla ^{ \iota^* L_1} + \al$
967: on $\iota^* L_1$. Its flat sections satisfy
968: $$(f_1^j + \tfrac{1}{i} \nabla^{\iota^* L_1}_{X^j}) s = 0.$$
969: Furthermore its curvature vanishes. Indeed, since $[X^i, X^j] = 0$, we have
970: \begin{xalignat*}{2}
971: \frac{1}{i} \langle d \al , X^i \wedge X^j \rangle = & ( X^i. f_1^j - X^j. f_1^i
972: ) \\
973: = & \om_1 ( X^i, X^j) \end{xalignat*}
974: which follows from proposition \ref{subprincipal_calculus} and the fact that
975: $[T^i, T^j] = 0$.
976:
977: This defines a structure of flat line bundle for $\iota^*L_1 \otimes
978: \delta_{\Ga} $, whose flat sections are the solutions of \eqref{transport_equation}.
979: Recall that
980: $$ \Psi(x,k) = t^k_\Ga (x) (g_0(x) + O(k^{-1}))
981: \text{ over $V \cap \iota(\Gamma)$},$$
982: where $t_\Ga$ is a flat section of $\iota^* L $. The condition
983: to patch together these sections along $\Ga$ is the Bohr-Sommerfeld
984: condition:
985: $$ \text{$\iota^*( L^k \otimes L_1) \otimes \delta_{\Ga} \rightarrow \Ga$ is trivial as a flat bundle.} $$
986: When $M$ is two-dimensional, this is equivalent to
987: \eqref{BS_conditions}. To prove theorem \ref{BS_theo} or a similar result in the
988: $2n$-dimensional case for the joint spectrum, we should consider
989: Lagrangian sections depending continuously on $\Ga$. Furthermore, we
990: can show by using a local normal form that the solutions of
991: \eqref{vap_equation} are necessarily Lagrangian sections associated to
992: $\Ga$. A complete proof is in \cite{oim2}. The only novelty here is
993: the formulation of theorems \ref{resultat_clef} and \ref{norme}, and
994: consequently of the Bohr-Sommerfeld conditions.
995:
996: \subsection{Comparison with the cotangent case}
997: To compare theorems \ref{norme} and \ref{resultat_clef} with the similar statements in
998: the case of pseudo-differential operators, we can introduce some
999: kind of Maslov
1000: bundle in the following way. Recall that we denote by $\varphi_{\Ga}$ the
1001: isomorphism $\delta_{\Ga}^2 \rightarrow \La ^n T^*\Ga \otimes
1002: \C$. Introduce
1003: $$ P := \bigl\{ u \in \delta_{\Ga} ; \; \varphi_{\Ga} ( u ^{\otimes 2} )
1004: \in \La^n T^* \Ga - \{0 \} \bigr\} .$$
1005: Let $\Z_4$ be the subgroup $\{ 1, -1, i, -i \}$ of $\C^*$. Then $P$ is
1006: a principal bundle with structure group $ \Z_4 \times \R_+$. Introduce
1007: the complex line bundles $| \delta_{\Ga} |$ and $\arg ( \delta_{\Ga}
1008: )$ associated to $P$ via the homomorphism $\Z_4 \times \R_+ \rightarrow
1009: \Z_4$ and $\Z_4 \times \R_+ \rightarrow \R_+$ respectively. Following Weinstein in \cite{We}, we call $\arg (\delta_{\Ga})$ the unitarization
1010: of $\delta_{\Ga}$. We have a
1011: canonical isomorphism
1012: $$ \delta_{\Ga} \rightarrow | \delta_{\Ga} | \otimes \arg
1013: (\delta_{\Ga} ).$$
1014: Furthermore, the map
1015: $$ |\delta_{\Ga} | = P
1016: \times_{\R_+} \C \ni [u,z] \rightarrow z.| \varphi_{\Ga} (u^{\otimes
1017: 2})|^{\frac{1}{2}}\in |\La |^{\frac{1}{2}} ( \Ga)$$
1018: is an isomorphism between $| \delta_{\Ga} |$ and
1019: the bundle of half-densities of $\Ga$. So we obtain an isomorphism
1020: $$ \zeta: \delta_{\Ga} \rightarrow |\La |^{\frac{1}{2}} ( \Ga) \otimes \arg
1021: (\delta_{\Ga} ) .$$
1022: The bundle $\arg (\delta_{\Ga})$ is a line bundle with structure group $\Z_4$ like the
1023: Maslov bundle.
1024: The isomorphism $\zeta$ intertwines the operator
1025: $\Lie_X^{\delta_{\Ga}}$ of theorem \ref{resultat_clef} with the Lie
1026: derivative of half-densities. So in the case $L_1$ is trivial, the
1027: theorem \ref{resultat_clef} is similar to the formula 1.3.13 in
1028: \cite{Du3} p223, computing the symbol of an oscillatory integral acted
1029: on under a differential operator.
1030: Furthermore the map $ \delta_{\Ga}
1031: \rightarrow |\La |(\Ga)$ used in theorem \ref{norme} is the composition of
1032: $\zeta$ with the squaring map from half-densities into
1033: densities. Again theorem \ref{norme} is similar to formula 1.3.15 in
1034: \cite{Du3} p224, computing the norm of an oscillatory integral.
1035:
1036: To end this comparison, we apply the previous construction to a
1037: symplectic vector space $E$ and prove that we obtain the usual Maslov
1038: bundle. Consider a one-dimensional vector space $\delta$ with an
1039: isomorphism $\varphi: \delta^2 \rightarrow \La^{n, 0} E^*$.
1040: Let $\Lag (E)$ be the Lagrangian Grassmannian and $\eta
1041: \rightarrow \Lag (E) $ be the tautological vector bundle, that is the
1042: bundle whose fiber over $x$ is $x$ itself. In the same way we defined
1043: $\varphi_{\Ga}$, one has an isomorphism
1044: $$ \varphi_{\Lag} : \Lag (E) \times \delta^2 \rightarrow \La ^n \eta
1045: \otimes \C, $$
1046: sending $(x,u)$ into $\iota_x^* \varphi(u)$ where $\iota_x$ is the
1047: embedding $ \eta_x \rightarrow E$. Introduce the
1048: $\Z_4 \times \R_+$ bundle
1049: $$ \bigl\{ (x,u) \in \Lag (E)
1050: \times \delta; \; \varphi_{\Lag} (x,u^2) \in \La_x^n \eta \text{ and }
1051: u \neq 0 \bigr\} $$
1052: Dividing by $\R_+$ we get a $\Z_4$-principal bundle $\Mas$. We claim
1053: that the holonomy in
1054: $\Mas$ of a loop of $\Lag (E)$ is the mod 4 reduction of its Maslov index.
1055:
1056: \begin{proof} Introduce linear
1057: Darboux coordinates $(p^i,q^i)$ and identify $E$ with $\R^{2n}$. Set $z^j= p^j + i q^j$ and let
1058: $$ \delta := (dz^1\wedge...\wedge dz^n)^{\frac{1}{2}} \C$$
1059: be the square root of $\La^{n,0} E = (dz^1\wedge...\wedge dz^n)\
1060: \C$. Recall that $\Lag (E)$ is isomorphic to $U(n)/O(n)$ (cf. lemma 2.31 of
1061: \cite{McSa}), through the map sending the unitary matrix $U =P +i Q$
1062: into the range of
1063: \begin{gather*}
1064: A_U = \left(
1065: \begin{array}{c} P \\ Q
1066: \end{array} \right) \end{gather*}
1067: Let us denote by $\al^1_U$,...,$\al^n_U$ the base of
1068: $\eta_x^*$ dual to the column vectors of $A_U$. A
1069: straightforward computation shows that $\varphi_{\Lag}$ sends the
1070: square of
1071: $$([U], (dz^1\wedge...\wedge dz^n)^{\frac{1}{2}})
1072: \in \Lag (E) \times \delta $$ into
1073: $\det (U) \ \al^1_U \wedge... \wedge \al^n_U$. Consequently,
1074: $$ \Mas \simeq \bigl\{ ( [U], v ) \in \Lag (E)
1075: \times \C^* ;
1076: \; U \in U(n) \text{ and } v^2 \det ( U) = \pm 1 \bigr\} $$
1077: Recall now that the
1078: Maslov index of a loop $x: \R/ \Z \rightarrow \Lag (E)$ is the degree of $\rho \circ x : S^1 \rightarrow
1079: S^1$ where $\rho$ is the map
1080: $$ \Lag (E) \rightarrow S^1, \quad [U] \rightarrow \operatorname{det}^2 (U)$$
1081: (cf. \cite{McSa} page 53). Its mod 4 reduction is the holonomy of $x$
1082: in $ \Mas$.
1083: \end{proof}
1084:
1085:
1086: This result is related to the paper
1087: \cite{We} of Weinstein, where it is observed that the Maslov bundle of
1088: $\Lag (E)$ is a unitarization of a square root of $\La ^n \eta
1089: \otimes \C$.
1090:
1091: Last remark is that in general the
1092: Maslov bundle of a Lagrangian submanifold of a cotangent space can be
1093: different of the bundle we construct. Indeed notice that the structure
1094: group of $\arg (\delta_{\Ga})$ reduces to $\Z_2$ if and only if $\Ga$
1095: is orientable. Consider a non-orientable manifold $Q$. Then as the
1096: null section of $T^*Q$, $Q$ is a
1097: Lagrangian submanifold and its Maslov bundle is the flat trivial
1098: bundle. So it can not be a unitarization of a square root of $\La^n T^*Q
1099: \otimes \C$.
1100:
1101:
1102: %\newpage
1103: \section{Proof}
1104:
1105: We assume in the whole section
1106: that there exists a globally defined half-form bundle $(\delta, \varphi)$ and
1107: $K = \delta$. There is no difficulty to generalize to the case where
1108: $K = \delta \otimes L_1$.
1109:
1110: \subsection{A preliminary result}
1111:
1112: Consider a sequence $\Psi_k \in \Ci(M,L^k \otimes \delta)$ of the
1113: form
1114: $$\Psi_k(x) = \Bigl( \frac{k}{2\pi} \Bigr)^{\frac{n}{4}} F^k (x)
1115: \tilde{g} (x,k) + O(k^{-\infty}) \text{ over $V$, } $$
1116: where $F$ and $\tilde{g}(.,k)$
1117: satisfies the same assumptions as in section \ref{sec:resultat} except that the coefficients
1118: $\tilde{g}_k$ do not necessarily satisfy $\bar{\partial} \tilde{g}_k \equiv
1119: 0$. Assume furthermore that $\Psi_k$ is admissible.
1120:
1121: \begin{theo} \label{theo:main}
1122: Let $T_k$ be a Toeplitz operator with principal symbol
1123: $f_0$. Then $T_k \Psi_k$ is a Lagrangian section over $V$ with
1124: symbol $\iota^* (f_0 \tilde{g}_0) + O(\hb)$.
1125:
1126: Furthermore, if $\tilde{g}_0$
1127: and its first derivatives vanish along $\iota(\Ga)$, then the
1128: symbol of $T_k \Psi_k$ is
1129: $$ \hb (\iota^* f_0) \bigl( \square \tilde{g}_0 + \iota^* \tilde{g}_1) ) + O(\hb^2)
1130: $$
1131: where $\square \tilde{g}_0 \in \Ci(\Ga, \delta_{\Ga})$ and at every $x \in \Ga$
1132: $$ \square \tilde{g}_0 (x) = - \frac{1}{2} \sum \bar{\partial}_i \bar{\partial}_i
1133: \tilde{g}_0 (\iota(x))$$
1134: if $\partial_1,...,\partial_n$ is a base of vectors of
1135: $T^{1,0}_{\iota(x)} M$ such that
1136: $\frac{1}{i} \om (\partial_{i}, \bar{\partial}_j)
1137: = \delta_{ij}$
1138: and the vectors $\partial_{i} + \bar{\partial}_i$ are
1139: tangent to $\Ga$.
1140: \end{theo}
1141:
1142: The proof starts from the following representation of the Schwartz
1143: kernel of the Toeplitz operator $T_k$:
1144: \begin{gather} \label{eq:noyau_Toeplitz}
1145: T_k (x,y) = \Bigl( \frac{k}{2 \pi} \Bigr)^n E^{k} (x,y) \tilde{f}
1146: (x,y,k) + O (k^{-\infty}) \end{gather}
1147: where, if we consider $M^2$ as a complex manifold with complex
1148: structure $(j , -j)$,
1149: \begin{itemize}
1150: \item
1151: $E$ is a section of $L \boxtimes \bar{L}
1152: \rightarrow M^2$ satisfying
1153: $$ E (x,x) = u \otimes \bar{u}, \quad \forall u \in L_x \text{ such
1154: that }
1155: \| u \| = 1,
1156: $$ $ \bar{\partial} E \equiv 0 $
1157: modulo a section vanishing to every order along the diagonal $\De$ of
1158: $M^2$ and $\| E(x,y)
1159: \| <1 $ if $x \neq y$.
1160: \item
1161: $\tilde{f}(.,k)$ is a sequence of sections of $ \delta \boxtimes
1162: \bar{\delta} \rightarrow M^2$ with an asymptotic expansion of the form
1163: $$ \tilde{f}(.,k) = \tilde{f}_0 + k^{-1} \tilde{f}_1 + ...$$
1164: whose coefficient satisfy $\bar{\partial} f_l \equiv 0 $
1165: modulo a section vanishing to every order along $\De$. Furthermore, $$\tilde{f}_0 (x,x) = f_0 (x),$$
1166: where $f_0$ is the principal symbol of $T_k$.
1167: \end{itemize}
1168: In other words, $T_k (.,.)$ is a
1169: Lagrangian section associated to the diagonal $\De$ of $M^2$. This
1170: result was proved in \cite{oim1}, without the additional bundle
1171: $\delta$. The generalization is straightforward.
1172:
1173: Since the norm of $E$ is $<1$ outside the diagonal and $\Psi_k$ is
1174: admissible, one has for every $x$ in $V$
1175: $$ (T_k \Psi _k ) (x)= \Bigl( \frac{k}{2 \pi} \Bigr)^{n+ \frac{n}{4}} \int E^k
1176: (x,y).F^k(y) \ \tilde{f}(x,y,k). \tilde{g} (y,k)\ \mu_M(y) +
1177: O(k^{-\infty})$$
1178: where we integrate on a neighborhood of $x$. Introduce a unitary
1179: section $t$ of $L$ over $U$ such that $\iota^* t = t_\Gamma$ over
1180: $U_{\Gamma}$ and let us write
1181: \begin{gather} \label{eq:def_phi}
1182: E(x,y).F(y) = e^{i \phi(x,y) } t(x) .\end{gather}
1183: Then the imaginary part of $\phi$ is non positive and vanishes only if
1184: $(x,y)$ belongs to
1185: $$ C:= \{ (x,x) \in M^2; \; x \in \iota(\Ga) \} .$$
1186: To compute the derivatives of $\phi$ along $C$, recall the following
1187: lemma proved in \cite{oim2} (cf. proposition 2.2 p.1535).
1188: \begin{lemme} \label{lem:derivee_F}
1189: If $\nabla^{L} F = \tfrac{1}{i} \al_F \otimes F$, then
1190: $\al_F$ vanishes along $\iota(\Ga)$ and for every vector field $X,Y$
1191: $$ \Lie_X \langle \al_F , Y \rangle = \om ( q X, Y)$$
1192: at $x \in \iota(\Ga)$, where $q$ is the projection onto $T_x^{0,1}M$ with kernel
1193: $T_x \iota(\Ga) \otimes \C$.
1194: \end{lemme}
1195: As a corollary, we have the following
1196: \begin{lemme} \label{lem:derivee_E}
1197: If $\nabla^{L \boxtimes \bar{L}} E =
1198: \tfrac{1}{i} \al_E \otimes E$, then $\al_E$ vanishes along the
1199: diagonal, and for every vector fields $X_1,Y_1,X_2,Y_2$ of
1200: $M$,
1201: $$ \Lie_{(X_1,Y_1)}. \langle \al_E, (X_2,Y_2) \rangle = \om (
1202: X_1^{0,1} - Y_1^{0,1}, X_2) + \om (
1203: X_1^{1,0} - Y_1^{1,0}, Y_2)$$
1204: along the diagonal, where we denoted by $X^{1,0}$ and $X^{0,1}$ the
1205: holomorphic and anti-holomorphic parts of $X$ respectively.
1206: \end{lemme}
1207: We deduce from both lemmas that $d_y \phi$ vanishes along $C$. Furthermore
1208: the kernel of the tangent map to $d_y \phi $ at $(x,x) \in C$ is
1209: $$ \bigl( T_x ^{0,1}M \times (0) \bigr) \oplus \bigl( T_{(x,x)} C \otimes \C \bigr) $$
1210: Finally, we have along $C$,
1211: $$ d^2_y \phi (Y_1,Y_2) = \om ( Y_1^{1,0}, Y_2) - \om ( q Y_1, Y_2) $$
1212: and $d^2_y \phi $ is non-degenerate. So we can apply the stationary
1213: phase lemma (cf. \cite{Ho} section 7.7 or theorem \ref{theo:PS1}).
1214: One gets
1215: $$ (T_k \Psi_k ) (x) = \Bigl( \frac{k}{2 \pi} \Bigr)^{\frac{n}{4}} e^{ik
1216: \phi_r (x) } t^k(x) \ \tilde{h} (x,k) + O( k^{-\infty})$$
1217: where
1218: \begin{gather} \label{eq:def_phir}
1219: \phi_r (x) \equiv \phi (x,y) \end{gather}
1220: modulo a linear combination with
1221: $\Ci$ coefficients of the
1222: $\partial_{y^i} \phi (x,y)$.
1223: \begin{lemme} $e^{ik
1224: \phi_r (x) } t(x)$ satisfies the same assumption as the section
1225: $F$.
1226: \end{lemme}
1227: \begin{proof}
1228: Since $\phi$ and $d_y \phi$ vanishes along $C$, $\phi_r$ vanishes
1229: along $\iota(\Ga)$
1230: and consequently $$ e^{i
1231: \phi_r (x) } t(x) = t_{\Ga} (x)$$ for every $x$ in
1232: $\iota(\Ga)$. Introduce complex coordinates $x^1,...,x^n$ and write
1233: $$ \nabla t = \tfrac{1}{i} t \otimes \sum a_j (x) dx^j + \bar{a}_j
1234: (x) d\bar{x}^j .$$
1235: Derivating \eqref{eq:def_phi}, it follows from $\bar{\partial} E
1236: \equiv 0$ that
1237: $$ \partial_{\bar{x}^i} \phi (x,y) \equiv \bar{a}_j (x) \mod
1238: \ideal_{\De} (\infty),$$ i.e. modulo a function vanishing to infinite
1239: order along the diagonal. Derivating again, one has
1240: $$\partial_{\bar{x}^i} \partial_{y^j} \phi (x,y) \equiv 0 \mod
1241: \ideal_{\De} (\infty) $$
1242: Then we deduce from the two previous equations and
1243: \eqref{eq:def_phir} that for every multi-index $\al$,
1244: $$ \partial_{\bar{x}^1}^{\al(1)}...\partial_{\bar{x}^n}^{\al(n)} \bigl(
1245: \partial_{\bar{x}^i} \phi_r - \bar{a}_i \bigr) (x) = 0 $$
1246: along $\iota(\Ga)$. And consequently
1247: $$ \partial_{\bar{x}^i} \phi_r \equiv \bar{a}_i \mod \ideal_{\iota(\Ga)}
1248: (\infty)$$
1249: which proves the result.
1250: \end{proof}
1251:
1252: \begin{lemme}
1253: The sequence $\tilde{h}(.,k)$ admits an asymptotic
1254: expansion $\tilde{h}_0 + k^{-1} \tilde{h}_1 +...$ whose
1255: coefficients satisfy $\bar{\partial} \tilde{h}_\ell \equiv 0$ modulo a
1256: section vanishing to every order along $\Ga$.
1257: \end{lemme}
1258:
1259: \begin{proof} First one deduces from lemma \ref{lem:derivee_F} that
1260: the imaginary part of $\phi_r$ and its first derivatives vanishes
1261: along $\iota(\Ga)$. Furthermore the Hessian of $\Im \phi_r$ along $\iota(
1262: \Ga)$ is non-degenerate in the
1263: transverse direction to $\iota(\Ga)$. As a consequence, if $\tilde{e}
1264: (x,k)$ is a sequence with an asymptotic expansion $\tilde{e}_0 +
1265: k^{-1} \tilde{e}_1 +...$ such that
1266: $$ e^{ik \phi_r (x)} \tilde{e} (x,k) = O ( k^{-\infty}) $$
1267: then the coefficients $\tilde{h}_\ell$ vanish to every order along
1268: $\iota(\Ga)$. This was proved in \cite{oim1} (cf. lemma 1, p.6). We apply this to the
1269: sequence $\bar{\partial} T_k \Psi_k$ which vanishes, since $ \Pi_k T_k
1270: = T_k$ implies that $T_k \Psi_k$ belongs to
1271: $\Hilbert_k$.
1272: \end{proof}
1273:
1274: The two previous lemmas imply that $T_k \Psi_k$ is a Lagrangian section.
1275: Then applying theorems \ref{theo:PS1} and \ref{theo:PS2}, we obtain
1276: the symbol of $T_k\Psi_k$ by computations of linear algebra, which are easily done using
1277: the tangent vectors $\partial_i$ introduced in the statement of
1278: theorem \ref{theo:main}.
1279:
1280: \scratch{Quelques d{\'e}tails suppl{\'e}mentaires: Pour d{\'e}duire des
1281: lemmes \ref{lem:derivee_E} et \ref{lem:derivee_F} l'application tangente {\`a} $d_y \phi$, on
1282: d{\'e}rive \eqref{eq:def_phi}, ce qui nous donne
1283: $$ \al_E + \pi_y^* \al_F = - d \phi + \pi_x^* \al_t $$
1284: o{\`u} $\nabla t = \frac{1}{i} t \otimes \al_t$, et $\pi_x$ et $\pi_y$
1285: sont les projections de $M^2$ sur le premier et second facteur
1286: respectivement. On en d{\'e}duit que
1287: $$ \Lie_{(X_1,Y_1)}. \langle d_y \phi , Y_2 \rangle = - \om (
1288: X_1^{0,1} - Y_1^{0,1}, X_2) - \om ( q Y_1,Y_2 ) $$
1289: On v{\'e}rifie imm{\'e}diatement que $ \bigl( T_x ^{0,1}M \times (0)
1290: \bigr) \oplus \bigl( T_{(x,x)} C \otimes \C \bigr) $ est inclus dans le
1291: noyau. Pour l'inclusion r{\'e}ciproque, il suffit de raisonner sur les dimension
1292: une fois que l'on a montr{\'e} que $d_y^2 \Phi$ est
1293: non-d{\'e}g{\'e}n{\'e}r{\'e}e. Pour montrer ceci, on calcule la matrice de
1294: $\frac{1}{i} d_y^2 \Phi$ dans la base $\frac{1}{\sqrt{2}} ( \partial_i
1295: + \bar{\partial}_i), \frac{i}{\sqrt{2}} ( \partial_i
1296: - \bar{\partial}_i)$, on obtient
1297: $$ \frac{1}{2} \left(
1298: \begin{array}{cc} 1 & -i \\ -i & 3
1299: \end{array} \right)
1300: $$
1301: dont le d{\'e}terminant est $1$. Et ceci nous calcule de plus
1302: $\tilde{h}_0$ sur la diagonale (il ne faut pas oublier la mesure de
1303: Liouville qui ne contribue pas car la base pr{\'e}c{\'e}dente est orthonorm{\'e}e pour
1304: la m{\'e}trique $\om (X, JY)$). Ensuite pour calculer le terme suivant
1305: lorsque $\tilde{g}_0$ s'annule au second ordre le long de
1306: $\iota(\Ga)$, on applique le th{\'e}or{\`e}me \ref{theo:PS2}. Comme base
1307: de $ T_x ^{0,1}M \times (0)
1308: $ qui est dans le noyau de l'application tangente {\`a} $d_y
1309: \phi$ et en somme directe avec $ T_{(x,x)} C \otimes \C$, on choisit
1310: les $(\bar{\partial}_i,0)$. Come base de $(0) \times (T_x M \otimes
1311: \C)$, on choisit les $\partial_i, \bar{\partial_i}$. La matrice de
1312: $\frac{1}{i} d^2_y \phi$ dans cette base est
1313: $$ \left(
1314: \begin{array}{cc} -1 & 1 \\ 1 & 0
1315: \end{array} \right)
1316: $$
1317: qui admet pour inverse
1318: $$ \left(
1319: \begin{array}{cc} 0 & 1 \\ 1 & 1
1320: \end{array} \right)
1321: $$
1322: On obtient le r{\'e}sultat du th{\'e}or{\`e}me avec $$\square \tilde{g}_0 = \Delta
1323: \tilde{g}_0 = \sum \partial_i \bar{\partial}_i \tilde{g}_0 +
1324: \frac{1}{2}\bar{\partial}_i\bar{\partial}_i \tilde{g}_0$$
1325: On conclut en utilisant que $ \partial_i \bar{\partial}_i \tilde{g}_0
1326: = - \bar{\partial}_i\bar{\partial}_i \tilde{g}_0$ car $\partial_i +
1327: \bar{\partial}_i$ tangent {\`a} $\iota( \Ga)$. }
1328:
1329: \subsection{Proofs of the theorems of part \ref{sec:resultat}}
1330:
1331: A first corollary of theorem \ref{theo:main} is the existence of a
1332: Lagrangian section with an arbitrary symbol: applying theorem \ref{theo:main} with the Toeplitz operator $\Pi_k$,
1333: we construct a Lagrangian section with a prescribed principal symbol,
1334: then theorem \ref{existence} follows from Borel resummation. Theorem
1335: \ref{Toep_Lag} is a particular case of theorem \ref{theo:main}.
1336:
1337: To prove theorem \ref{resultat_clef}, we can assume that $f_0$
1338: vanishes along $\Ga$. Since we compute the symbol of
1339: $T_k \Psi_k$ modulo $O(\hb^2)$, we can replace $T_k$ with every
1340: Toeplitz operators of symbol $f_0 + \hb f_1 + O(\hb^2)$. So by theorem
1341: \ref{normalised_symbol_quantisation}, we can choose
1342: $$ T_k = \Pi_k \Bigr( f_0 + k^{-1} f_1 + \frac{1}{ik} ( \nabla^{L^k}_X
1343: \otimes \identite + \identite \otimes D_X ) \Bigl) $$
1344: where $X$ is the Hamiltonian vector field of $f_0$. So $T_k \Psi_k$ is
1345: equal to
1346: \begin{gather} \label{La}
1347: \Pi_k \Bigl[ \bigl( \tfrac{k}{2 \pi} \bigr)^{\frac{n}{4}} \bigl( [f_0 + \tfrac{1}{ik} \nabla_{X}^{L^k} ] F^k
1348: \bigr) \tilde{g}(.,k) \Bigr] + k^{-1} \Pi_k \Bigl[ \bigl( \tfrac{k}{2
1349: \pi} \bigr)^{\frac{n}{4}} F^k \bigl( [f_1 +
1350: \tfrac{1}{i} D_X ] \tilde{g}(.,k) \bigr) \Bigr] \end{gather}
1351: By theorem \ref{theo:main}, each term of the sum is a Lagrangian
1352: section. Furthermore by the first part of this theorem, the symbol of the second one is
1353: $$ \hb\ \bigl( (\iota^* f_1).g_0 + \iota^*(\tfrac{1}{i} D_X
1354: \tilde{g}_0) \bigr) +O(\hb^2).$$
1355: Since $X$ is tangent to $\iota(\Ga)$, $\iota ^* (D_X
1356: \tilde{g}_0)$ only depends on the restriction of $\tilde{g}_0$
1357: to $\iota(\Ga)$. So we can define the operator $\iota^* D_X$ acting on $\Ci
1358: (\Ga, \delta_\Ga)$ which sends $g_0$ to $ \iota ^* (D_X
1359: \tilde{g}_0)$. And the previous symbol is
1360: \begin{gather} \label{2}
1361: \hb\ \bigl( (\iota^* f_1) + \tfrac{1}{i} (\iota^* D_X) \bigr).g_0
1362: +O(\hb^2).
1363: \end{gather}
1364:
1365:
1366: To compute the symbol of the first term of \eqref{La}, let us write
1367: $$ [f_0 + \tfrac{1}{ik} \nabla_{X}^{L^k} ] F^k
1368: = F^k a $$
1369: with $a$ defined on a neighborhood of $\iota(\Ga)$.
1370: \begin{lemme} \label{li2}
1371: The function $a$ and its first derivatives vanish along $\iota(\Ga)$. If $Z$
1372: and $W$ are holomorphic vector fields of $M$, then
1373: $$ \Lie_{\bar{W}} \Lie_{\bar{Z}} a = \om( \bar{W} , \Lie_X \bar{Z} ) $$
1374: on $\iota( \Ga)$.
1375: \end{lemme}
1376:
1377: \begin{proof}
1378: Denote by $\alpha _F$ be the one-form such that $ \nabla^{L} F =
1379: \frac{1}{i}\al_F \otimes F$. Then
1380: $$ a = f_0 - \langle \al_{F} , X \rangle .$$
1381: This vanishes along $\iota ( \Ga)$ because $\al_F$ vanishes along $\iota(\Ga)$
1382: (cf. lemma \ref{lem:derivee_F}).
1383:
1384: Since $X$ is the Hamiltonian vector field of $f_0$, one has $\Lie_Y
1385: f_0 + \om (X,Y) =0$. Since the curvature of $L$ is $\frac{1}{i}\om$,
1386: one has $d \al_F = \om$ and consequently
1387: $$\Lie_Y \langle \al_F, X
1388: \rangle = \Lie_X \langle \al_F, Y
1389: \rangle + \om(Y,X) + \langle \al_F, [Y,X] \rangle.$$ It follows that
1390: $$ \Lie_Y a = - \Lie_X \langle \al_F, Y \rangle - \langle \al_F, [Y,X]
1391: \rangle .$$
1392: This vanishes along $\iota(\Ga)$, because $\al_F$ vanishes along $\iota(\Ga)$
1393: and $X$ is tangent to $\iota(\Ga)$.
1394:
1395: Since $\bar{\partial} F \equiv 0$ modulo a flat section and $Z$ is holomorphic, $ \langle
1396: \al_F, \bar{Z} \rangle$ vanishes to every order along $\iota(\Ga)$. So choosing
1397: $Y =\bar{Z}$ in the previous equation, we obtain
1398: $$ \Lie_{\bar{W}}.\Lie_{\bar{Z}} a = - \Lie_{\bar{W}} \langle \al_F,
1399: [\bar{Z}, X] \rangle \qquad \text{ along } \iota(\Ga) .$$
1400: Using again that $\om = d\al_F$ and $\al_F $ vanishes along $\iota (\Ga)$, it
1401: follows that
1402: $$ \Lie_{\bar{W}}.\Lie_{\bar{Z}} a = - \om ( \bar{W}, [\bar{Z}, X ] )
1403: - \Lie_{[\bar{Z},X ]} \langle \al_F, \bar{W} \rangle $$
1404: along $\iota ( \Ga)$. The second term of the right side vanishes along
1405: $\iota ( \Ga)$ because $\bar{W}$ is an anti-holomorphic vector
1406: field. This gives the result.
1407: \end{proof}
1408:
1409: Since $\iota^* D_X$ and $\Lie_X^{\delta_{\Ga}}$ are first order
1410: differential operators which have the same symbol,
1411: $$ \iota^* D_X - \Lie_X^{\delta_{\Ga}} = b $$
1412: where $b$ is a function of $\Ga$.
1413:
1414: \begin{lemme} \label{li3}
1415: The symbol of $ \Pi_k \bigl[ F^k a \tilde{g}(.,k) \bigr]$ is
1416: $ i \hb b. g_0 + O(\hb^2)$
1417: \end{lemme}
1418:
1419:
1420: So the symbol of $T_k \Psi_k$ is the sum of \eqref{2} and $ i
1421: \hb b g_0 + O(\hb^2)$, which is equal to
1422: $$ \hb(\iota^* f_1 + \tfrac{1}{i} \Lie_X^{\delta_{\Gamma}}) . g_0 + O(\hb^2). $$
1423: Theorem \ref{resultat_clef} follows.
1424:
1425: \begin{proof}
1426: Let us start with a local computation of the function $b$.
1427: Let $u$ be a non-vanishing section of $\delta_{\Ga} \rightarrow
1428: \Ga$. Then one has
1429: $$ b = \frac{ \bigl( \iota^* D_X - \Lie_X^{\delta_{\Ga}} \bigr) .u
1430: }{u}$$
1431: Let $\beta$ be a non-vanishing $(n,0)$-form of $M$ such that $\varphi
1432: (u^{\otimes 2}(x)) = \beta (\iota(x))$ if $x$ be\-longs to $\Ga$. Then
1433: $\iota^* D_X $ is defined in such a way that
1434: $$ \frac{ ( \iota^* D_X).u}{u} = \frac{1}{2} \iota^* \Bigl(\frac{ p
1435: \Lie_{X} \be}{\be}\Bigr) = \frac{1}{2} \frac{ \iota^* (p
1436: \Lie_{X} \be)}{\iota^* \be} .$$
1437: where $p$ is the projection from $\La^{n} M \otimes \C$ onto
1438: $\La^{n,0} M$ with kernel
1439: $ \La^{0,n} M \oplus ... \oplus \La^{n-1,1} M.$
1440: On the other hand, since
1441: $\varphi_{\Ga}(u^{\otimes 2} (x)) = \iota^* \varphi
1442: (u^{\otimes 2} (x) ) = \iota^* \be (x) $,
1443: one has
1444: $$
1445: \frac{ \Lie_X^{\delta_{\Ga}}.u
1446: }{u} = \frac{1}{2} \frac{ \Lie_X \iota^* \be}{\iota^* \be} = \frac{1}{2} \frac{ \iota^* \Lie_X \be}{\iota^* \be} .$$
1447: Consequently
1448: $$ b = \frac{1}{2} \frac{ \iota^* \bigl(p \Lie_X \be- \Lie_X
1449: \be\bigr) }{\iota ^* \be}. $$
1450:
1451: Now let us choose a frame $(\partial_1,...,\partial_n)$ of holomorphic
1452: vector fields of $M$ such that the vectors $\partial_i + \bar{\partial_i}$ are tangent
1453: to $\iota(\Ga)$. Denote by $(\te^{1},..., \te^n)$ the dual frame and set
1454: $$ \be = \te^1 \wedge... \wedge \te^n .$$
1455: Then using that $\Lie_X \te^i \equiv - \sum \langle \te^i, \Lie_X
1456: \bar{\partial}_j \rangle \bar{\te}^j $ modulo a linear combination of the
1457: $\te^i$, we obtain that $\iota^* (p \Lie_X \be- \Lie_X
1458: \be)$ is equal to
1459: \begin{gather*}
1460: \sum \langle \te^1, \Lie_X \bar{\partial}_j \rangle
1461: \bar{\te}^j \wedge \te^2 \wedge ...\wedge \te^n + \langle \te^2,
1462: \Lie_X \bar{\partial}_j \rangle \te^1 \wedge \bar{\te}^j \wedge
1463: \te^3 \wedge ...\wedge \te^n + ...\\
1464: + \langle \te^n,
1465: \Lie_X \bar{\partial}_j \rangle \te^1 \wedge ...\wedge \te^{n-1}
1466: \wedge \bar{\te}^j \end{gather*}
1467: It follows then from $\iota^* \te^i = \iota^* \bar{\te}^i$ that
1468: $$ b = \frac{1}{2} \sum \langle \te^i, \Lie_X
1469: \bar{\partial}_i \rangle .$$
1470:
1471: To end the proof, assume furthermore that $\frac{1}{i} \om ( \partial_i
1472: ,\bar{\partial}_j) = \de_{ij}$. Then it follows from the second part of theorem \ref{theo:main}
1473: that the symbol of $ \Pi_k \bigl[ F^k a \tilde{g}(.,k) \bigr]$ is
1474: $$ - \hb g_0 \frac{1}{2} \ \iota^* \sum \Lie_{\bar{\partial}_i}
1475: \Lie_{\bar{\partial}_i} a $$
1476: which by lemma \ref{li2} is equal to
1477: $$ - \hb g_0 \frac{1}{2} \ \iota^* \sum \om ( \bar{\partial}_i, \Lie_X
1478: \bar{\partial}_i)$$
1479: Using again that $\frac{1}{i} \om ( \partial_i
1480: ,\bar{\partial}_j) = \de_{ij}$, we obtain
1481: $$ \hb g_0 \frac{i}{2} \ \iota^* \sum \langle \te^i, \Lie_X
1482: \bar{\partial}_i \rangle .$$
1483: The final result follows.
1484: \end{proof}
1485:
1486: Finally, let us prove theorem \ref{norme}.
1487: One has
1488: $$ \int_M \xi \ \bigl\| \Psi_k \bigr\|_{L^k \otimes \delta} ^2 \; \mu_M =
1489: \Bigl( \frac{k}{2\pi} \Bigr)^{\frac{n}{2}} \int_M e^{-k c} \ \xi \
1490: \bigl\| \tilde{g}(.,k) \bigr\|_{\delta}^2 \;
1491: \mu_M +O(k^{-\infty}) $$
1492: where $c(x) = -2 \ln \| F(x) \| _{L}$. The following is a
1493: consequence of lemma
1494: \ref{lem:derivee_F}.
1495: \begin{lemme} \label{lem:norme_F}
1496: The function $c$ and its first derivatives vanish along
1497: $\iota(\Ga)$. Furthermore, its Hessian at $x \in \iota(\Ga)$ is definite positive on
1498: $J T_x \iota(\Ga)$ and is given by
1499: $$ X.Y.c = 2 \om (X, JY), \quad X,Y \in J T_x \iota(\Ga).$$
1500: \end{lemme}
1501: So integrating along transversal directions to $\Ga$, it follows from the
1502: stationary phase lemma that
1503: $$ \int_M \; \xi \ \bigl\| \Psi_k \bigr\|_{L^k \otimes \delta}
1504: ^2 \; \mu_M = \int_{\Gamma} (\iota^* \xi)\ d +O(k^{-1})$$
1505: where $d$ is the density of $\Ga$ such that
1506: $$ d\eval{x} ( X) = \bigl\| g_0(x) \bigr\|^2_{ \delta}
1507: \ \mu_{M} \eval{x} (X \wedge Y) \ 2^{-\frac{n}{2}} \Bigl( \det \bigl[ \om (Y_i,J Y_j)
1508: \bigr]\Bigr)^{-\frac{1}{2}}.$$
1509: Here $(X_i)$ and $(Y_i)$ are bases of $T_x \Ga$ and $J T_x \Ga$
1510: respectively, and $X = X_1 \wedge...\wedge X_n$, $Y = Y_1
1511: \wedge ... \wedge Y_n$. To deduce
1512: theorem \ref{norme}, we have to check that
1513: \begin{gather*}
1514: d = m (g_0)
1515: \end{gather*}
1516: This is easily done, by introducing the same vector fields
1517: $\partial_i$ and forms $\te^i$ like in the
1518: proof of lemma \ref{li3}, setting $X_i = \frac{1}{\sqrt{2}}(
1519: \partial_i + \bar{\partial}_i)$, $Y_j = J X_j$ and choosing $g_0$ such that $ \varphi (g_0^{\otimes
1520: 2} ) = \theta_1 \wedge ...\wedge \theta^n$.
1521:
1522: \scratch{Quelques d{\'e}tails: Pour d{\'e}duire le lemme \ref{lem:norme_F}
1523: du lemme \ref{lem:derivee_F}, on utilise que si $X \in J
1524: T\iota(\Ga)$, alors $q(JX) = 0$ et $q(X+ i JX) = X+iJX$, donc $q(X)
1525: = X + iJX$. Apr{\`e}s, on calcule
1526: $$ X.Y .c = - \frac{X .Y. \| F \|^2}{\| F \|^2} = - \frac{1}{i} X. \langle \al_F -
1527: \bar{\al}_F, Y \rangle .$$
1528: Ce qui donne avec le lemme \ref{lem:derivee_F}
1529: $$ -\frac{1}{i} \om (X + iJX - (X - iJX), Y)
1530: = -2 \om ( JX,Y) =
1531: 2 \om (X, JY).$$
1532: Voil{\`a}. Ensuite, pour la v{\'e}rification de $d = m(g_0)$, avec les
1533: choix que j'indique $\om (Y_i, J Y_j) =\de_{ij}$, $ \mu_M ( X \wedge
1534: Y) = 1$, et $\te^1 \wedge ...\wedge \te^n$ {\'e}tant de norme 1, il
1535: vient $d(X) = 2^{-\frac{n}{2}}$. D'autre part $\langle \iota^*
1536: \te^i, X_j \rangle = \frac{1}{\sqrt{2}} \de_{ij} $, donc $\langle \iota^*
1537: \varphi (g_0^2) , X \rangle = 2^{-\frac{n}{2}}$, autrement dit $m(
1538: g_0) (X) = 2^{- \frac{n}{2}}$.
1539: }
1540:
1541: %\newpage
1542: \section{Appendix}
1543:
1544: Let $W$ be an open set of $\R^{n} \times \R^{k} \ni (x,y)$. Denote by
1545: $p$ the projection from $W$ onto $\R^n$. Let $\ph (x,y)$ be a $\Ci$
1546: function on $W$ whose imaginary part is $ \geqslant 0$. Let $a(x,y)$
1547: be a $\Ci$ function with compact support in $W$. Stationary phase
1548: lemma gives the asymptotic expansion of
1549: $$ I(a, \ph)(x, \tau ) = \int_{\R^k} e^{i \tau \ph (x,y) } a(x,y) |dy| $$
1550: when $\tau \rightarrow \infty$.
1551: First, if the support of $a$ doesn't meet the critical locus
1552: $$ C := \{ (x,y)\in W ; \; d_y \ph (x,y) = 0 \text{ and } \Im \ph (x,y) = 0 \}, $$
1553: then $I(a, \ph)$ is $O(\tau^{-\infty})$. Introduce the functions
1554: $$\ph_{i,j}(x,y) = \partial_{y^i} \partial_{y^j} \ph
1555: (x,y), \quad i,j =1,...,k.$$
1556: The following theorem is proved in \cite{Ho} section 7.7.
1557:
1558: \begin{theo} \label{theo:PS1}
1559: Assume that at $(x_0,y_0) \in C$, the matrix $
1560: \bigl(\ph_{i,j}(x_0,y_0)\bigr)$ is invertible. Then there exists a
1561: neighborhood $U$ of $(x_0, y_0)$ such that if the support of $a$ is a
1562: subset of $ U$, one has
1563: $$ I(\ph, a) (x, \tau) =( \tfrac{2 \pi }{ \tau} )^{\frac{k}{2}} \ d (x) \
1564: e^{i \tau \ph_r (x) }\ b(x,\tau) + O(\tau^{-\infty}) \qquad \text{ over
1565: } p(U)
1566: $$
1567: where $d$, $\ph_r$ et $b(.,\tau)$ are $\Ci$ functions such that
1568: \begin{itemize}
1569: \item $d$ only depends on $\ph$. In particular,
1570: $$ d(x) = \operatorname{det}^{-\frac{1}{2}} [ \tfrac{1}{i}
1571: \varphi_{j,k} (x,y)]_{j,k}, \quad \text {if } (x,y) \in C \cap U.$$
1572: \item $\ph_r$ is such that $ \ph (x,y) \equiv \ph_r (x) $ on $U$
1573: modulo a linear combination with $\Ci$ coefficient of the functions $\partial_{y^j} \ph$.
1574: \item $b(.,\tau)$ has an asymptotic expansion for the $\Ci$ topology of the form
1575: $$ b_0 (x) + \tau b_1(x) + \tau^2 b_2 (x) +...$$
1576: Furthermore, $b_0(x) = a(x,y)$
1577: if $(x,y) \in C \cap U$.
1578: \end{itemize}
1579: \end{theo}
1580: In \cite{Ho} the various terms of the asymptotic expansion are
1581: completely determined and not only their restriction at $p(C)$. But in the applications this leads to
1582: complicated computations that we prefer to avoid. Let us introduce an
1583: additional assumption.
1584:
1585: Denote by $E_{(x,y)}$ the complexification of the tangent space to the
1586: fiber of $p$ at $(x,y)$.
1587: At $(x,y) \in C$ the tangent map to the section $d_y
1588: \ph$ of $E^*$
1589: $$ T_{(x,y)} d_{y} \ph : T_{(x,y)} W \otimes \C \rightarrow E^*_{(x,y)} $$
1590: is well-defined. Assume that $(\varphi_{i,j})$ is invertible along
1591: $C$, that is the kernel $F_{(x,y)}$ of $ T_{(x,y)} d_{y} \ph$
1592: satisfies
1593: \begin{gather} \label{H1}
1594: \forall \ (x,y) \in C, \quad F_{(x,y)} \oplus E_{(x,y)} =
1595: T_{(x,y)} W \otimes \C. \end{gather}
1596: Assume furthermore that
1597: \begin{gather} \label{H2}
1598: \text{$C$ is a submanifold of $W$ and }
1599: T C \otimes \C = F \cap
1600: \bar{F} . \end{gather}
1601: Finally, these two assumptions imply that the restriction $p : C \rightarrow
1602: \R^n$ is an immersion. We assume it is an embedding.
1603:
1604: Observe that when the phase takes real values, the assumption
1605: \eqref{H2} is a consequence of
1606: \eqref{H1}. We are interested in the opposite case, typically when the
1607: Hessian of the imaginary part of the phase is non-degenerate in the
1608: transverse directions to $C$, for instance with $$\ph (x,y) = xy + \tfrac{i}{2}(x^2 +
1609: y^2).$$
1610: We can also consider intermediary cases, for example $\ph (x,y) = xy + \frac{i}{2} y^2$.
1611:
1612: Under the previous assumptions, when the amplitude $a$ vanishes to
1613: order $m$ along $C$, i.e. when the partial derivatives of $a$ of order
1614: $\leqslant m-1$ vanish along $C$, it follows from the result of
1615: \cite{Ho} that the functions $b_i$ vanish to order $m-2i$ along
1616: $C$. Furthermore one can easily compute
1617: $b_i$ modulo a function vanishing to order $m-2i+1$ along
1618: $C$.
1619:
1620: To state the result, consider a free family $\partial_1,...,
1621: \partial_l$ of complex tangent vectors to $W$ at $(x,y) \in
1622: C$ such that
1623: $$ \operatorname{Vect}_{\C} (\partial_1,..., \partial_l) \oplus (TC
1624: \otimes \C) = F_{(x,y)}.$$
1625: If $a$ vanishes to order $m$ along $C$, we define the polynomial
1626: $$ [a] (Z,Y) = \sum_{|\al| + |\be|=m} \frac{1}{\al ! \be !} \Bigl(
1627: \partial_1^{\al(1)}...\partial_l^{\al(l)}\partial_{y^1}^{\be(1)}...\partial_{y^k}^{\be(l)}
1628: a(x,y) \Bigr) \ Z^{\al} Y^{\be}$$
1629: at $(x,y) \in C$.
1630: Similarly, if $b(x)$ vanishes to order $l$ along $p(C)$, we set
1631: $$ [b] (Z) = \sum_{|\al|=l} \frac{1}{\al ! \be !} \Bigl(
1632: (p_*\partial_1)^{\al(1)}...(p_* \partial_l)^{\al(l)} b(x) \Bigr) \ Z^{\al}.$$
1633: at $x \in p(C)$.
1634: \begin{theo} \label{theo:PS2}
1635: Under the assumptions \eqref{H1} and \eqref{H2}, if $a$ vanishes to
1636: order $m$ along $C$, then for every $i \leqslant \frac{m}{2}$ the
1637: function $b_i$ vanishes to order $m-2i$ along $p(C)$. Furthermore
1638: $$ [b_i](Z) = \frac{1}{i!} \De^i A_{2i}(Z,Y) $$
1639: at $(x,y) \in C$, where
1640: \begin{itemize}
1641: \item $[a](Z,Y) = \sum_{l=0}^{m} A_l(Z,Y)$ and $A_l$ is homogeneous of
1642: degree $l$ in $Y$ and of degree $m-l$ in $Z$.
1643: \item $\De = \tfrac{i}{2} \textstyle{\sum}_{j,k}
1644: \ph^{j,k}(x,y) \partial_{Y^j} \partial_{Y^k}$, with $(\ph^{j,k}(x,y))$
1645: the inverse of $ (\ph_{j,k}(x,y))$.
1646: \end{itemize}
1647: \end{theo}
1648:
1649: More intrinsically, denote by $\ideal ^m(C) \subset \Ci( W)$ the ideal
1650: of functions vanishing to order $l$ along $C$. Then $\ideal ^m (C)/
1651: \ideal ^{m-1} (C)$ is isomorphic to the space of sections of the $m$-th symmetric power of the
1652: complex conormal bundle $\no^* (C)$. By \eqref{H1}, we have an
1653: isomorphism of vector bundle over $C$,
1654: $$ \no^* (C) \simeq \no ^* ( p (C)) \oplus E^* ,$$
1655: which associates $\ga$ and $\al \oplus \be$ if
1656: $$ \langle \ga, U +
1657: V \rangle = \langle \al, p_* U \rangle + \langle \beta, V \rangle,
1658: \quad \text{$U \in F$ and $V \in E$} $$
1659: Consequently,
1660: \begin{gather*} %\label{dec_sym}
1661: \Sym_m ( \no^* (C) ) = \bigoplus_{l=0}^{m} \Sym_{m-l} \bigl( \no^* (p(C) ) \bigr) \otimes \Sym_{l} ( E^* ) \end{gather*}
1662: $ \Delta = \tfrac{i}{2} \textstyle{\sum}_{j,k} \ph^{j,k}
1663: \partial_{y^j} \partial_{y^k}$ defines a section of $\Sym_{2}(E)$, so
1664: $\Delta ^i$ acts as an operator
1665: $$\Sym_{2i} (
1666: E^* ) \rightarrow \C .$$ In theorem \ref{theo:PS2}, we consider
1667: $[a]$ as a section of $\Sym_m ( \no^*
1668: (C) )$ and $[b_i]$ as a section of $\Sym_{m-2i} ( \no^* (p(C)))$ . Then we have
1669: $$ [b_i] = \frac{1}{i!} (\identite \otimes \De^i) A_{2i} $$
1670: where
1671: $$ [a] = \sum_{l=0}^{m} A_l, \quad A_l \in \Ci (C,\Sym_{m-l} \bigl( \no^*
1672: (p(C) ) \bigr) \otimes \Sym_{l} ( E^* )) .$$
1673:
1674:
1675:
1676: \bibliography{biblio}
1677:
1678:
1679: \end{document}
1680:
1681:
1682:
1683:
1684: %%% Local Variables:
1685: %%% mode: latex
1686: %%% TeX-master: t
1687: %%% End:
1688: