This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

Resolvents of Bochner Laplacians in the semiclassical limit

L. Charles
Abstract

We introduce a new class of pseudodifferential operators, called Heisenberg semiclassical pseudodifferential operators, to study the space of sections of a power of a line bundle on a compact manifold, in the limit where the power is large. This class contains the Bochner Laplacian associated to a connection of the line bundle, and when the curvature is nondegenerate, its resolvent and some associated spectral projections, including generalised Bergman kernels.

1 Introduction

The spectral analysis of the Bochner Laplacians acting on sections of a line bundle with a large curvature, has many applications ranging from complex geometry to mathematical physics: holomorphic Morse inequalities and Bergman kernels [20], dynamical systems [13], geometric quantization [2] or large magnetic field limit of Schrödinger operators [23], [22], [17] to quote just a few references. In this paper we introduce an algebra of pseudodifferential operators, shaped to study the bottom of the spectra of these Laplacian at small scale.

To understand how the scales matter, let us state two Weyl laws corresponding to two different regimes. Let (Mn,g)(M^{n},g) be a closed Riemannian manifold and LML\rightarrow M a Hermitian line bundle with a connection \nabla. For any integer kk, the Bochner Laplacian Δk=12\Delta_{k}=\tfrac{1}{2}\nabla^{*}\nabla acting on sections of LkL^{k} is an elliptic differential operator. Hence Δk\Delta_{k} with domain 𝒞(M,Lk){\mathcal{C}}^{\infty}(M,L^{k}), is essentially self-adjoint. Its spectrum is a subset of 0{\mathbb{R}}_{\geqslant 0} and consists only of eigenvalues with finite multiplicity.

When kk is large, the structure of this spectrum depends essentially on the curvature 1iω\frac{1}{i}{\omega} of \nabla. Since \nabla is assumed to preserve the Hermitian structure, ω{\omega} is a real closed 22-form of MM. For the first regime we will need the twisted symplectic form Ω\Omega of TMT^{*}M defined by

Ω=dξidxi+pω\displaystyle\Omega=\sum d\xi_{i}\wedge dx_{i}+p^{*}{\omega} (1)

where pp is the projection TMMT^{*}M\rightarrow M. For the second one, we will assume that ω{\omega} and gg are compatible in the sense that ω(x,y)=g(jx,y){\omega}(x,y)=g(jx,y) for an (almost) complex structure jj. The Weyl laws at the energy scales k2k^{2} and kk are respectively:

  1. 1.

    For any λ>0\lambda>0, the number Nk(λ)N_{k}(\lambda) of eigenvalues of k2Δkk^{-2}\Delta_{k} smaller than λ\lambda and counted with multiplicities satisfies

    Nk(λ)(k2π)nvol{ξTM,12|ξ|x2λ} as k\displaystyle N_{k}(\lambda)\sim\Bigl{(}\frac{k}{2\pi}\Bigr{)}^{n}\operatorname{vol}\{\xi\in T^{*}M,\;\tfrac{1}{2}|\xi|_{x}^{2}\leqslant\lambda\}\quad\text{ as $k\rightarrow\infty$} (2)

    with the volume computed with respect to the Liouville form 1n!Ωn\frac{1}{n!}\Omega^{n}.

  2. 2.

    if ω{\omega} and gg are compatible, then for any M>0M>0 there exists C>0C>0 such that for any kk,

    spec(k1Δk)],M](12+)+Ck12]1,1[,\displaystyle\operatorname{spec}(k^{-1}\Delta_{k})\cap]-\infty,M]\subset(\tfrac{1}{2}+{\mathbb{N}})+Ck^{-\frac{1}{2}}]-1,1[, (3)

    and for any mm\in{\mathbb{N}},

    (spec(k1Δk)(12+m+]14,14[))(k2π)n/2(m+n1m)\displaystyle\sharp\Bigl{(}\operatorname{spec}(k^{-1}\Delta_{k})\cap\bigl{(}\tfrac{1}{2}+m+]-\tfrac{1}{4},\tfrac{1}{4}[\bigr{)}\Bigr{)}\sim\Bigl{(}\frac{k}{2\pi}\Bigr{)}^{n/2}{m+n-1\choose m} (4)

The estimate (2) does not appear in the literature, but it is actually a small variation of the Weyl law of semiclassical pseudodifferential operators on a compact manifold with semiclassical parameter h=k1h=k^{-1}, as we will explain below. The cluster structure (3) and the estimate (4) of the number of eigenvalues in each cluster have been proved in [13]. This eigenvalue number is actually given by a Riemann-Roch formula when kk is sufficiently large [6], [7].

These results rely on a good description of the resolvents (kϵΔkz)1(k^{-\epsilon}\Delta_{k}-z)^{-1}, which allows to study the f(kϵΔk)f(k^{-\epsilon}\Delta_{k}) for ff a smooth compactly supported function, where ϵ=2\epsilon=2 or 11 according to the regime. In the first case, f(k2Δk)f(k^{-2}\Delta_{k}) and (k2Δkz)1(k^{-2}\Delta_{k}-z)^{-1} are semiclassical twisted pseudodifferential operators, a class of operators which was introduced in [4, Chapitre 4] and used recently in [14]. The local theory of these operators is exactly the same as the one of the standard semiclassical pseudodifferential operators with h=k1h=k^{-1}, whereas the global theory involves the twisted symplectic form (1). Similar operators have been studied as well on n{\mathbb{R}}^{n} under the name of magnetic pseudodifferential operators, cf. [19] for an introduction and many references.

For the second regime, the theory is much less developed. The proof of (3) is based on an approximation of the resolvent of k1Δkk^{-1}\Delta_{k}, which is obtained by gluing local resolvents of some model operators deduced from Δk\Delta_{k} by freezing the coordinates in an appropriate way. The main difficulty in this construction is that the number of terms we have to glue increases with kk. Still these approximations have been used successfully to prove (3) and later to describe the spectral projections onto the various clusters as generalizations of Bergman kernels [6], [7], [18].

In this paper, we introduce a new class of pseudodifferential operators, and we prove that it contains the resolvent of k1Δkk^{-1}\Delta_{k} when ω{\omega} is nondegenerate, as well as the spectral projector associated to the cluster (3). This operator class is in a sense a semiclassical version of the class of Heisenberg pseudodifferential operators, which has been developed for the study of the ¯b\overline{\partial}_{b}-complex [1], [12]. For this reason, we call our operators semiclassical Heisenberg pseudodifferential operators.

Just as the usual Heisenberg operators have a symbol of type (12,12)(\frac{1}{2},\frac{1}{2}), the semiclassical ones have an exotic symbol of type 12\frac{1}{2}: at each derivative, a power of h12h^{\frac{1}{2}} is lost. Recall that the semiclassical operators with a symbol of type 12\frac{1}{2} form a limit class: they are closed under product, the usual operator norm estimates hold, but the standard expansions in the symbolic calculus do not hold because all the terms in these expansions have the same magnitude, cf. for instance [10, Proposition 7.7, Theorem 7.9 and Theorem 7.11]. As we will see in the sequel [3] of this paper, the semiclassical Heisenberg operators are closed under product. However the composition of their principal symbols, which are functions on TMT^{*}M, is not the usual product. Instead it is a fiberwise product, whose restriction to each fiber TxMT_{x}^{*}M depends on the curvature ωx{\omega}_{x}: when ωx{\omega}_{x} is nondegenerate, it is essentially the Weyl product whereas when ωx=0{\omega}_{x}=0, this is the usual function product. So in general the principal symbol composition is not commutative.

Our definition of the Heisenberg operators is based on the usual semiclassical pseudodifferential operators, from which we deduce easily many of their properties, except what regards their composition. In this paper, we only address the Heisenberg composition of differential operators with pseudodifferential ones, because this suffices to show that the resolvent and cluster spectral projectors of k1Δkk^{-1}\Delta_{k} are Heisenberg operators. We have also included an exposition of the theory of twisted pseudodifferential operators, because this is not really standard material, and this helps to understand the specificity of the Heisenberg pseudodifferential operators. In the remainder of the introduction, we state our main result.

Twisted pseudodifferential operators

As the usual Laplace-Beltrami operator, Δk\Delta_{k} has the following local expression in a coordinate chart (U,xi)(U,x_{i})

k2Δk=12gj,=1nπjgjgπk^{-2}\Delta_{k}=-\tfrac{1}{2\sqrt{g}}\sum_{j,\ell=1}^{n}\pi_{j}g^{j\ell}\sqrt{g}\,\pi_{\ell}

where for any j=1j=1, πj\pi_{j} is the dynamical moment πj:=(ik)1jLk\pi_{j}:=(ik)^{-1}\nabla^{L^{k}}_{j} with jLk\nabla^{L^{k}}_{j} the covariant derivative of LkL^{k} with respect to xj\partial_{x_{j}}.

Let s𝒞(U,L)s\in{\mathcal{C}}^{\infty}(U,L) be such that |s|=1|s|=1 and let βΩ1(U,)\beta\in\Omega^{1}(U,{\mathbb{R}}) be the corresponding connection 1-form, s=iβs\nabla s=-i\beta\otimes s. Identifying 𝒞(U,Lk){\mathcal{C}}^{\infty}(U,L^{k}) with 𝒞(U){\mathcal{C}}^{\infty}(U) through the frame sks^{k},

πj=(ik)1xjβj\pi_{j}=(ik)^{-1}\partial_{x_{j}}-\beta_{j}

where βj(x)=β(x)(xj)\beta_{j}(x)=\beta(x)(\partial_{x_{j}}). Under these local identifications, the πj\pi_{j}’s and consequently k2Δkk^{-2}\Delta_{k} are semiclassical differential operators with h=k1h=k^{-1}, their symbols being respectively ξjβj\xi_{j}-\beta_{j} and 12|ξβ(x)|2\frac{1}{2}|\xi-\beta(x)|^{2}.

These symbols become independent of ss if we pull-back them by the momentum shift Tβ:TUTUT_{\beta}:T^{*}U\rightarrow T^{*}U, (x,ξ)(x,ξ+β(x))(x,\xi)\rightarrow(x,\xi+{\beta}(x)). The same method will be used to define the symbol of a twisted pseudodifferential operator, cf. (8). Similarly, the shift TβT^{*}_{\beta} is used in classical mechanic to write the motion equations of a particle in a magnetic field in an invariant way: TβT_{\beta} sends the Hamiltonian 12|ξβ|2\frac{1}{2}|\xi-\beta|^{2} with the standard symplectic form dξidxi\sum d\xi_{i}\wedge dx_{i} to the usual kinetic energy 12|ξ|2\frac{1}{2}|\xi|^{2} with the twisted symplectic form Ω\Omega.

Let us briefly introduce the twisted pseudodifferential operators of LL, the complete definition will be given in Section 2. Let us start with the residual class. An operator family

P=(Pk:𝒞(M,Lk)𝒞(M,Lk),k)\displaystyle P=(P_{k}:{\mathcal{C}}^{\infty}(M,L^{k})\rightarrow{\mathcal{C}}^{\infty}(M,L^{k}),\;k\in{\mathbb{N}}) (5)

belongs to kΨ(L)k^{-\infty}\Psi^{-\infty}(L) if for each kk, the Schwartz kernel of PkP_{k} is smooth, its pointwise norm is in 𝒪(k)\mathcal{O}(k^{-\infty}) uniformly on MM and the same holds for its successive derivatives. For any local data δ=(U,s,ρ)\delta=(U,s,\rho) consisting of an open set UU of MM, a frame s𝒞(U,L)s\in{\mathcal{C}}^{\infty}(U,L) such that |s|=1|s|=1 and a function ρ𝒞0(U)\rho\in{\mathcal{C}}^{\infty}_{0}(U), we define the local form of a family PP as in (5) as:

Pkδ:𝒞(U)𝒞(U),(Pkδf)sk=ρPk(ρfsk),k\displaystyle P^{\delta}_{k}:{\mathcal{C}}^{\infty}(U)\rightarrow{\mathcal{C}}^{\infty}(U),\qquad(P^{\delta}_{k}f)s^{k}=\rho P_{k}(\rho fs^{k}),\qquad k\in{\mathbb{N}} (6)

A twisted pseudodifferential operator of order mm is a family PP as in (5) such that for any ρ1\rho_{1}, ρ2𝒞(M)\rho_{2}\in{\mathcal{C}}^{\infty}(M) with disjoint supports, (ρ1Pkρ2)(\rho_{1}P_{k}\rho_{2}) belongs to kΨ(L)k^{-\infty}\Psi^{-\infty}(L) and for any local data δ=(U,s,ρ)\delta=(U,s,\rho), Pkδ=Qk1δP_{k}^{\delta}=Q^{\delta}_{k^{-1}} where (Qhδ:𝒞(U)𝒞(U),h(0,1])(Q_{h}^{\delta}:{\mathcal{C}}^{\infty}(U)\rightarrow{\mathcal{C}}^{\infty}(U),\;h\in(0,1]) is a semiclassical pseudodifferential operator of order mm. So in terms of coordinates (xi)(x_{i}) on UU the Schwartz kernel of PkδP_{k}^{\delta} has the form

Pkδ(x,y)=(k2π)neikξ(xy)a(k1,12(x+y),ξ)𝑑ξ\displaystyle P_{k}^{\delta}(x,y)=\Bigl{(}\frac{k}{2\pi}\Bigr{)}^{n}\int e^{ik\xi\cdot(x-y)}a(k^{-1},\tfrac{1}{2}(x+y),\xi)\;d\xi (7)

for some semiclassical polyhomogeneous symbol (a(h,)𝒞(U×n),h(0,1])(a(h,\cdot)\in{\mathcal{C}}^{\infty}(U\times{\mathbb{R}}^{n}),\;h\in(0,1]) of order mm. The (principal) symbol of PP is the function σ𝒞(TM){\sigma}\in{\mathcal{C}}^{\infty}(T^{*}M) such that for any local data δ\delta as above

a(h,x,ξ+β(x))=ρ(x)σ(x,ξ)+𝒪(h),\displaystyle a(h,x,\xi+\beta(x))=\rho(x){\sigma}(x,\xi)+\mathcal{O}(h), (8)

where βΩ1(U,)\beta\in\Omega^{1}(U,{\mathbb{R}}) is the connection one-form of ss.

Examples of twisted (pseudo)differential operators of respective order 0, 11 and 22 are the multiplications by the functions of 𝒞(M){\mathcal{C}}^{\infty}(M), the covariant derivatives (ik)1XLk(ik)^{-1}\nabla^{L^{k}}_{X} where X𝒞(M,TM)X\in{\mathcal{C}}^{\infty}(M,TM), the symbol being (x,ξ)ξ,X(x)(x,\xi)\rightarrow\langle\xi,X(x)\rangle, and the normalised Laplacian k2Δkk^{-2}\Delta_{k}, its symbol is (x,ξ)12|ξ|x2(x,\xi)\rightarrow\frac{1}{2}|\xi|_{x}^{2}.

It is not difficult to adapt the standard results [8], [27] on the resolvents of elliptic operators and their functional calculus in this setting. Denoting by Ψtscm(L)\Psi_{\operatorname{tsc}}^{m}(L) the space of twisted pseudodifferential operators of order mm, we have

  • for any z0z\in{\mathbb{C}}\setminus{\mathbb{R}}_{\geqslant 0}, the resolvent (zk2Δk)1(z-k^{-2}\Delta_{k})^{-1} belongs to Ψtsc2(L)\Psi_{\operatorname{tsc}}^{-2}(L) and its symbol is (z12|ξ|2)1(z-\frac{1}{2}|\xi|^{2})^{-1}.

  • for any f𝒞0()f\in{\mathcal{C}}^{\infty}_{0}({\mathbb{R}}), f(k2Δk)f(k^{-2}\Delta_{k}) belongs to Ψtsc(L)\Psi_{\operatorname{tsc}}^{-\infty}(L) and its symbol is f(12|ξ|2)f(\frac{1}{2}|\xi|^{2}).

In particular trf(k2Δk)=(k2π)nTMf(12|ξ|2)1n!Ωn+𝒪(k1+n)\operatorname{tr}f(k^{-2}\Delta_{k})=\bigl{(}\frac{k}{2\pi}\bigr{)}^{n}\int_{T^{*}M}f(\frac{1}{2}|\xi|^{2})\frac{1}{n!}\Omega^{n}+\mathcal{O}(k^{-1+n}). The Weyl law (2) follows.

Heisenberg pseudodifferential operators

A (semiclassical) Heisenberg pseudodifferential operator of (L,)(L,\nabla) of order mm is by definition a family PP of operators of the form (5) such that for any ρ1\rho_{1}, ρ2𝒞(M)\rho_{2}\in{\mathcal{C}}^{\infty}(M) with disjoint supports, (ρ1Pkρ2)(\rho_{1}P_{k}\rho_{2}) belongs to kΨ(L)k^{-\infty}\Psi^{-\infty}(L) and for any local data δ=(U,s,ρ)\delta=(U,s,\rho) as above with a coordinate set (xi)(x_{i}) on UU, the Schwartz kernel of PkδP_{k}^{\delta} has the form

eikβ(x+y2)(xy)(k2π)nneikξ(xy)a(k12,12(x+y),ξ)𝑑ξ\displaystyle e^{ik{\beta}\bigl{(}\tfrac{x+y}{2}\bigr{)}\cdot(x-y)}\Bigl{(}\frac{\sqrt{k}}{2\pi}\Bigr{)}^{n}\int_{{\mathbb{R}}^{n}}e^{i\sqrt{k}\;\xi\cdot(x-y)}a(k^{-\frac{1}{2}},\tfrac{1}{2}(x+y),\xi)\;d\xi (9)

where

  1. -

    β=βi(x)dxi{\beta}=\sum\beta_{i}(x)dx_{i} is the connection one-form of ss defined as above and viewed as the n{\mathbb{R}}^{n}-valued function x(β1(x),,βn(x))x\rightarrow({\beta}_{1}(x),\ldots,{\beta}_{n}(x))

  2. -

    (a(h,)𝒞(U×n),h(0,1])(a(h,\cdot)\in{\mathcal{C}}^{\infty}(U\times{\mathbb{R}}^{n}),\;h\in(0,1]) is a semiclassical polyhomogeneous symbols of order mm, so in particular xαξβa=𝒪α,β(ξm|β|)\partial_{x}^{{\alpha}}\partial_{\xi}^{\beta}a=\mathcal{O}_{{\alpha},{\beta}}(\langle\xi\rangle^{m-|\beta|}) and a=0a(x,ξ)a\sim\sum_{\ell=0}^{\infty}\hbar^{\ell}a_{\ell}(x,\xi) with polyhomogeneous coefficients aa_{\ell} of order mm-\ell.

As we see, the Schwartz kernel (9) is the product of an oscillatory factor depending on the frame ss with the Schwartz kernel of a semiclassical operator where the semiclassical parameter is k12k^{-\frac{1}{2}}. We will prove that this formula is consistent with change of frame and that we can define a (principal) symbol σ𝒞(TM){\sigma}\in{\mathcal{C}}^{\infty}(T^{*}M) such that for any local data as above,

a(h,x,ξ)=ρ(x)σ(x,ξ)+𝒪(h).a(h,x,\xi)=\rho(x){\sigma}(x,\xi)+\mathcal{O}(h).

Let us denote by ΨHeism(L,)\Psi_{\operatorname{Heis}}^{m}(L,\nabla) the space of Heisenberg pseudodifferential operators of order mm, and by ΨHeis(L,)\Psi_{\operatorname{Heis}}^{-\infty}(L,\nabla) the intersection mΨHeism(L,)\cap_{m}\Psi_{\operatorname{Heis}}^{m}(L,\nabla).

To state our main result, we need to introduce some symbols Rd,zR_{d,z} and πd,E\pi_{d,E}. Recall that for any tempered distribution a𝒮(sd×ςd)a\in\mathcal{S}^{\prime}({\mathbb{R}}_{s}^{d}\times{\mathbb{R}}_{\varsigma}^{d}), the Weyl quantization of aa is the operator aw:𝒮(d)𝒮(d)a^{w}:\mathcal{S}({\mathbb{R}}^{d})\rightarrow\mathcal{S}^{\prime}({\mathbb{R}}^{d}) with Schwartz kernel at (s,t)(s,t):

(2π)deiς(st)a(12(s+t),ς)𝑑ς.(2\pi)^{-d}\int e^{i\varsigma\cdot(s-t)}a(\tfrac{1}{2}(s+t),\varsigma)\;d\varsigma.

The (quantum) harmonic oscillator is HwH^{w} with H(s,ς)=12i=1d(si2+ςi2)H(s,\varsigma)=\tfrac{1}{2}\sum_{i=1}^{d}(s_{i}^{2}+\varsigma_{i}^{2}). As an operator of L2(d)L^{2}({\mathbb{R}}^{d}) with domain the Schwartz space, HwH^{w} is essentially self-adjoint with spectrum spHw=d2+\operatorname{sp}H^{w}=\frac{d}{2}+{\mathbb{N}}. Then for any zspHwz\in{\mathbb{C}}\setminus\operatorname{sp}H^{w} and EspHwE\in\operatorname{sp}H^{w}, Rd,zR_{d,z} and πd,E\pi_{d,E} are the tempered distributions such that

(Hwz)1=Rd,zw,1{E}(Hw)=πd,Ew\displaystyle(H^{w}-z)^{-1}=R_{d,z}^{w},\qquad 1_{\{E\}}(H^{w})=\pi^{w}_{d,E} (10)

By Weyl calculus, Rd,zR_{d,z} belongs to the symbol class S2(2d)S^{-2}({\mathbb{R}}^{2d}) and Rd,z=(Hz)1R_{d,z}=(H-z)^{-1} modulo S3(2d)S^{-3}({\mathbb{R}}^{2d}). Moreover πd,E\pi_{d,E} belongs to the Schwartz space, being the Weyl symbol of an orthogonal projector onto a finite dimensional subspace of 𝒮(d)\mathcal{S}({\mathbb{R}}^{d}). The analytic Fredholm theory can be developed in this setting and it says that the function zRd,zz\rightarrow R_{d,z} with values in 𝒞(2d){\mathcal{C}}^{\infty}({\mathbb{R}}^{2d}), or better the symbol space S2(2d)S^{-2}({\mathbb{R}}^{2d}), is meromorphic on {\mathbb{C}} with simple poles at d2+\frac{d}{2}+{\mathbb{N}} whose residues are the πd,E\pi_{d,E}.

Theorem 1.1.

Assume ω{\omega} and gg are compatible so that n=2dn=2d with dd\in{\mathbb{N}}. Then

  1. 1.

    For any z(d2+)z\in{\mathbb{C}}\setminus(\tfrac{d}{2}+{\mathbb{N}}), there exists Q(z)ΨHeis2(L,)Q(z)\in\Psi_{\operatorname{Heis}}^{-2}(L,\nabla) such that

    1. -

      (k1Δkz)Qk(z)id(k^{-1}\Delta_{k}-z)Q_{k}(z)\equiv\operatorname{id} and Qk(z)(k1Δkz)idQ_{k}(z)(k^{-1}\Delta_{k}-z)\equiv\operatorname{id} mod kΨ(L)k^{-\infty}\Psi^{-\infty}(L).

    2. -

      (k1Δkz)Qk(z)=Qk(z)(k1Δkz)=id(k^{-1}\Delta_{k}-z)Q_{k}(z)=Q_{k}(z)(k^{-1}\Delta_{k}-z)=\operatorname{id} when kk is large.

    3. -

      the symbol of Q(z)Q(z) restricted to TxMT_{x}^{*}M is the Weyl symbol Rd,zR_{d,z} of the resolvent of the harmonic oscillator with symbol 12|ξ|2\tfrac{1}{2}|\xi|^{2}, cf. (10).

  2. 2.

    For any Ed2+E\in\frac{d}{2}+{\mathbb{N}}, the spectral projector family

    (1[E1/2,E+1/2](k1Δk),k)(1_{[E-1/2,E+1/2]}(k^{-1}\Delta_{k}),k\in{\mathbb{N}})

    belongs to ΨHeis(L,)\Psi_{\operatorname{Heis}}^{-\infty}(L,\nabla). The restriction of its symbol to TxMT_{x}^{*}M is the Weyl symbol πd,E\pi_{d,E} of the spectral projector on the EE-eigenspace of the harmonic oscillator with symbol 12|ξ|2\tfrac{1}{2}|\xi|^{2}, cf. (10)

In this statement, we view HH, Rd,zR_{d,z} and πd,E\pi_{d,E} as functions on TxMT^{*}_{x}M as follows: choose an orthosymplectic basis (ei,fi)(e_{i},f_{i}) of TxMT_{x}M, that is (ei,fi)(e_{i},f_{i}) is an orthonormal basis and for any i,ji,j, ω(x)(ei,ej)=0=ω(x)(fi,fj){\omega}(x)(e_{i},e_{j})=0={\omega}(x)(f_{i},f_{j}), ω(x)(ei,fj)=δij{\omega}(x)(e_{i},f_{j})=\delta_{ij}. Let (si,ςi)(s_{i},{\varsigma}_{i}) be the associated coordinates of TxMT^{*}_{x}M, so si(ξ):=ξ(ei)s_{i}(\xi):=\xi(e_{i}) and ςi(ξ):=ξ(fi){\varsigma}_{i}(\xi):=\xi(f_{i}) for any ξTxM\xi\in T_{x}^{*}M. Then any function f:2df:{\mathbb{R}}^{2d}\rightarrow{\mathbb{C}} identifies with the function of TxMT_{x}^{*}M

ξTxMf(s(ξ),ς(ξ)).\displaystyle\xi\in T^{*}_{x}M\rightarrow f(s(\xi),\varsigma(\xi)). (11)

In particular H(s(ξ),ς(ξ))=12|ξ|2H(s(\xi),\varsigma(\xi))=\frac{1}{2}|\xi|^{2} because (ei,fi)(e_{i},f_{i}) is orthonormal. The fact that Rd,z(s(ξ),ς(ξ))R_{d,z}(s(\xi),\varsigma(\xi)) and πd,E(s(ξ),ς(ξ))\pi_{d,E}(s(\xi),\varsigma(\xi)) are independent of the choice of the basis (ei,fi)(e_{i},f_{i}) follows from the symplectic invariance of the Weyl quantization.

It follows as well from the symplectic invariance of Weyl quantization and the O(n)O(n) invariance of HH that Rd,zR_{d,z} and πd,E\pi_{d,E} are radial functions. A computation from Mehler formula leads to [9]

Rd,z=01(112s)d2z1(1+12s)d2+z1esH𝑑s, if Rez<d\displaystyle R_{d,z}=\int_{0}^{1}(1-\tfrac{1}{2}s)^{\frac{d}{2}-z-1}(1+\tfrac{1}{2}s)^{\frac{d}{2}+z-1}e^{-sH}ds,\quad\text{ if }\operatorname{Re}z<d (12)

We can also compute πd,E\pi_{d,E} in terms of Laguerre polynomial [26] :

πd,E(ξ)=2d(1)me|ξ|2Lmd1(2|ξ|2), where m=Ed2\displaystyle\pi_{d,E}(\xi)=2^{d}(-1)^{m}e^{-|\xi|^{2}}L_{m}^{d-1}(2|\xi|^{2}),\qquad\text{ where }m=E-\tfrac{d}{2} (13)

and Lmα(x)=1m!exxαxm(exxm+α).L_{m}^{\alpha}(x)=\frac{1}{m!}e^{x}x^{-{\alpha}}\partial_{x}^{m}(e^{-x}x^{m+{\alpha}}).

Remark 1.2.

1. The proof of the first part of Theorem 1.1 is an adaptation of the standard parametrix construction of an elliptic pseudodifferential operator, the main change being in the symbolic calculus: if (Pk)(P_{k}) belongs to ΨHeism(L,)\Psi_{\operatorname{Heis}}^{m}(L,\nabla) and has symbol σ{\sigma}, then (k1ΔkPk)(k^{-1}\Delta_{k}P_{k}) belongs to ΨHeism2(L,)\Psi_{\operatorname{Heis}}^{m-2}(L,\nabla) and its symbol restricted to TxMT_{x}^{*}M is the Weyl product of 12|ξ|x2\tfrac{1}{2}|\xi|_{x}^{2} and σ(x,){\sigma}(x,\cdot). By Weyl product, we mean the product of symbols in Weyl quantization, and the identification of functions of TxMT_{x}^{*}M with symbols is done through (11). This explains how the symbols Rd,zR_{d,z} and πd,E\pi_{d,E} appear.

2. A remarkable fact is that the proof of the second assertion of Theorem 1.1 is a direct application of Cauchy formula for the spectral projector of an operator with compact resolvent. This part is much simpler than the proof that f(k2Δk)Ψtsc(L)f(k^{-2}\Delta_{k})\in\Psi_{\operatorname{tsc}}^{-\infty}(L), even with the modern approach through Helffer-Sjöstrand formula.

3. The Schwartz kernel of the cluster spectral projectors was described in [18] and [7] as a generalisation of the Bergman kernel. The advantage of considering these projectors as Heisenberg pseudodifferential operators is merely that it connects them directly to the Laplacian and its resolvent. Moreover, in [3], we will explain how can use the Heisenberg calculus instead of the algebra (A)\mathcal{L}(A) of [6] to compute the dimension of each cluster and develop the theory of Toeplitz operators as it was done in [7]. ∎

Outline of the paper

In Section 2, we introduce notations and basic analytical tools to address the large kk limit of the space of sections of the kk-th power of LL, including the theory of semiclassical twisted pseudodifferential operators with their Sobolev spaces. The study of Heisenberg pseudodifferential operators starts in Section 3, from their Schwartz kernel asymptotic to their mapping properties. In Section 4, we introduce the symbol product, which is then used in Section 5 for the composition of differential operators with pseudodifferential operators. This is applied to resolvents and spectral projections in Section 6. In Section 7, we explain how we can add auxiliary bundles to the theory, which provides some important examples.

Acknowledgment

I would like to thank Clotilde Fermanian Kammerer, Colin Guillarmou and Thibault Lefeuvre for useful discussions.

2 Twisted pseudodifferential operators

Symbols

We will use the class of semiclassical polyhomogeneous symbols introduced in [11, Section E.1.2], cf. also [8, Section 6.1]. Let VV be an open set of p{\mathbb{R}}^{p} and mm\in{\mathbb{R}}. For any ξn\xi\in{\mathbb{R}}^{n}, let |ξ||\xi| and ξ\langle\xi\rangle be the Euclidean norm and Japanese bracket, so |ξ|2=ξi2|\xi|^{2}=\sum\xi_{i}^{2}, ξ2=1+|ξ|2\langle\xi\rangle^{2}=1+|\xi|^{2}. Let Sm(V,n)S^{m}(V,{\mathbb{R}}^{n}), Sphm(V,n)S^{m}_{\operatorname{ph}}(V,{\mathbb{R}}^{n}) and Sscm(V,n)S^{m}_{\operatorname{sc}}(V,{\mathbb{R}}^{n}) be the spaces of symbols (resp. polyhomogeneous symbols, semiclassical polyhomogeneous symbols) of order mm. By definition

  • Sm(V,n)S^{m}(V,{\mathbb{R}}^{n}) consists of the families (a(h,),h(0,1])(a(h,\cdot),\;h\in(0,1]) of 𝒞(V×n){\mathcal{C}}^{\infty}(V\times{\mathbb{R}}^{n}) such that for any compact set KK of VV, αp,βn{\alpha}\in{\mathbb{N}}^{p},{\beta}\in{\mathbb{N}}^{n}, there exists C>0C>0 such that

    |xαξβa(h,x,ξ)|Cξm|β|,xK,ξn,h(0,1]|\partial_{x}^{\alpha}\partial_{\xi}^{\beta}a(h,x,\xi)|\leqslant C\langle\xi\rangle^{m-|{\beta}|},\qquad\forall x\in K,\;\xi\in{\mathbb{R}}^{n},\;h\in(0,1]
  • bSphm(V,n)b\in S_{\operatorname{ph}}^{m}(V,{\mathbb{R}}^{n}) if bSm(V,n)b\in S^{m}(V,{\mathbb{R}}^{n}), bb is independent on hh and for every NN, b=j=0N1bjmodSmN(V,n)b=\sum_{j=0}^{N-1}b_{j}\mod S^{m-N}(V,{\mathbb{R}}^{n}) with coefficients bj𝒞(V×n)b_{j}\in{\mathcal{C}}^{\infty}(V\times{\mathbb{R}}^{n}) such that bj(x,tξ)=tmjbj(x,ξ)b_{j}(x,t\xi)=t^{m-j}b_{j}(x,\xi) when |ξ|1|\xi|\geqslant 1 and t>0t>0.

  • aSscm(V,n)a\in S^{m}_{\operatorname{sc}}(V,{\mathbb{R}}^{n}) if aSm(V,n)a\in S^{m}(V,{\mathbb{R}}^{n}) and a=0N1hahNSmN(V,n)a-\sum_{\ell=0}^{N-1}h^{\ell}a_{\ell}\in h^{N}S^{m-N}(V,{\mathbb{R}}^{n}) for some coefficients aSphm(V,n)a_{\ell}\in S_{\operatorname{ph}}^{m-\ell}(V,{\mathbb{R}}^{n}).

More generally these definitions make sense for a real vector bundle ENE\rightarrow N instead of the product V×nV\times{\mathbb{R}}^{n}. We denote by Sm(N,E)S_{*}^{m}(N,E) the corresponding spaces and set S(N,E)=mSm(N,E)S^{\infty}_{*}(N,E)=\bigcup_{m}S^{m}_{*}(N,E) for =,ph,sc*=\emptyset,\operatorname{ph},\operatorname{sc}. An easy remark is that for any section uu of EE, the translation TuT_{u} of 𝒞(E,){\mathcal{C}}^{\infty}(E,{\mathbb{C}}) given by Tuf(x,v)=f(x,vu(x))T_{u}f(x,v)=f(x,v-u(x)) preserves Sm(N,E)S_{*}^{m}(N,E). When VV is reduced to a point, we set Sm(n):=Sm({},n)S^{m}_{*}({\mathbb{R}}^{n}):=S^{m}_{*}(\{\cdot\},{\mathbb{R}}^{n}).

Negligible families

We say that a family (fh,h(0,1])(f_{h},h\in(0,1]) of 𝒞(N){\mathcal{C}}^{\infty}(N) is negligible, and we write fh=𝒪(h)f_{h}=\mathcal{O}_{\infty}(h^{\infty}), if all its 𝒞{\mathcal{C}}^{\infty}-seminorms are in 𝒪(h)\mathcal{O}(h^{\infty}). This definition is meaningful if (fh)(f_{h}) is only defined for hDh\in D where DD is any subset of (0,1](0,1] whose closure contains 0.

Let LML\rightarrow M be a Hermitian line bundle and AMA\rightarrow M a complex vector bundle with rank rr. A family (tk𝒞(M,LkA),k)(t_{k}\in{\mathcal{C}}^{\infty}(M,L^{k}\otimes A),\;k\in{\mathbb{N}}) is said to be negligible if for any open set UU of MM, any s𝒞(U,L)s\in{\mathcal{C}}^{\infty}(U,L) with pointwise norm |s|=1|s|=1 and any frame (ai𝒞(U,A),i=1,,r)(a_{i}\in{\mathcal{C}}^{\infty}(U,A),i=1,\ldots,r), tk=fi,k1skait_{k}=\sum f_{i,k^{-1}}s^{k}\otimes a_{i} on UU where each coefficient fi,h𝒪(h)f_{i,h}\in\mathcal{O}_{\infty}(h^{\infty}). We denote by 𝒪(k)\mathcal{O}_{\infty}(k^{-\infty}) the space of negligible families.

Let PP be a family of operators

P=(Pk:𝒞(M,Lk)𝒞(M,Lk),k),\displaystyle P=(P_{k}:{\mathcal{C}}^{\infty}(M,L^{k})\rightarrow{\mathcal{C}}^{\infty}(M,L^{k}),\;k\in{\mathbb{N}}), (14)

The Schwartz kernel of each PkP_{k} is a section of (LkL¯k)(M|Λ|(M))(L^{k}\boxtimes\overline{L}^{k})\otimes({\mathbb{C}}_{M}\boxtimes|\Lambda|(M)), where we denote by \boxtimes the external tensor product of vector bundles, by M{\mathbb{C}}_{M} the trivial line bundle over MM and by |Λ|(M)|\Lambda|(M) the density bundle. Since LkL¯k=(LL¯)kL^{k}\boxtimes\overline{L}^{k}=(L\boxtimes\overline{L})^{k}, the previous definition of a negligible family applies to the family (Pk)(P_{k}) of Schwartz kernels.

We denote by kΨ(L)k^{-\infty}\Psi^{-\infty}(L) the space consisting of operator families of the form (14) such that each PkP_{k} is smoothing with a Schwartz kernel family in 𝒪(k)\mathcal{O}_{\infty}(k^{-\infty}). As we will see, kΨ(L)k^{-\infty}\Psi^{-\infty}(L) is both the residual space of twisted pseudodifferential operators and of Heisenberg pseudodifferential operators.

Semiclassical pseudodifferential operators

Let Ψscm(M)\Psi^{m}_{\operatorname{sc}}(M) be the space of semiclassical pseudodifferential operators of order mm acting on smooth functions of MM. By definition PΨscm(M)P\in\Psi^{m}_{\operatorname{sc}}(M) is a family of operators (Ph:𝒞(M)𝒞(M),h(0,1])(P_{h}:{\mathcal{C}}^{\infty}(M)\rightarrow{\mathcal{C}}^{\infty}(M),h\in(0,1]) with a Schwartz kernel Kh(x,y)K_{h}(x,y) satisfying for any ρ𝒞(M2)\rho\in{\mathcal{C}}^{\infty}(M^{2}),

  1. 1.

    if suppρdiagM=\operatorname{supp}\rho\cap\operatorname{diag}M=\emptyset, then ρKh\rho K_{h} is smooth and negligible.

  2. 2.

    if suppρU2\operatorname{supp}\rho\subset U^{2} where (U,xi)(U,x_{i}) is a coordinate chart of MM, then on U2U^{2}

    (ρKh)(x,y)=(2πh)neih1ξ(xy)a(h,x,y,ξ)𝑑ξ\displaystyle(\rho K_{h})(x,y)=(2\pi h)^{-n}\int e^{ih^{-1}\xi\cdot(x-y)}a(h,x,y,\xi)\;d\xi (15)

    with aSscm(U2,n)a\in S_{\operatorname{sc}}^{m}(U^{2},{\mathbb{R}}^{n}).

Here and in the sequel, when the Schwartz kernel is written in a coordinate chart, we implicitly use the density |dx1dxn||dx_{1}\ldots dx_{n}|. The principal symbol of PP is the function σSphm(M,TM){\sigma}\in S_{\operatorname{ph}}^{m}(M,T^{*}M) such that a(h,x,x,ξ)=ρ(x)σ(x,ξ)+𝒪(h)a(h,x,x,\xi)=\rho(x){\sigma}(x,\xi)+\mathcal{O}(h) on UU.

Twisted pseudodifferential operators

Let LML\rightarrow M be a Hermitian line bundle.

Definition 2.1.

A semiclassical twisted pseudodifferential operator PP of LL is a family having the form (14) such that for any ρ𝒞(M2)\rho\in{\mathcal{C}}^{\infty}(M^{2}),

  1. 1.

    if suppρdiagM=\operatorname{supp}\rho\cap\operatorname{diag}M=\emptyset, then ρPk\rho P_{k} is smooth and negligible.

  2. 2.

    if suppρU2\operatorname{supp}\rho\subset U^{2} where (U,xi)(U,x_{i}) is a coordinate chart of MM and s𝒞(U,L)s\in{\mathcal{C}}^{\infty}(U,L) is such that |s|=1|s|=1, then on U2U^{2}

    (ρPk)(x,y)=(k2π)neikξ(xy)a(k1,x,y,ξ)𝑑ξsk(x)s¯k(y)(\rho P_{k})(x,y)=\Bigl{(}\frac{k}{2\pi}\Bigr{)}^{n}\int e^{ik\xi\cdot(x-y)}a(k^{-1},x,y,\xi)\;d\xi\;s^{k}(x)\otimes\overline{s}^{k}(y)

    with aSscm(U2,n)a\in S_{\operatorname{sc}}^{m}(U^{2},{\mathbb{R}}^{n}).

To understand the dependence of the oscillatory integral with respect to the choice of the frame ss, consider a new frame t=eiφst=e^{i\varphi}s where φ𝒞(U,)\varphi\in{\mathcal{C}}^{\infty}(U,{\mathbb{R}}). Then φ(x)φ(y)=jψj(x,y)(xjyj)\varphi(x)-\varphi(y)=\sum_{j}\psi_{j}(x,y)(x_{j}-y_{j}) where ψj𝒞(U2,)\psi_{j}\in{\mathcal{C}}^{\infty}(U^{2},{\mathbb{R}}) is such that ψj(x,x)=xjφ(x)\psi_{j}(x,x)=\partial_{x_{j}}\varphi(x). Using that s(x)s¯(y)=exp(iψ(x,y)(xy))t(x)t¯(y))s(x)\otimes\overline{s}(y)=\exp(-i\psi(x,y)\cdot(x-y))t(x)\otimes\overline{t}(y)) and changing the variable ξ\xi into ξ+ψ(x,y)\xi+\psi(x,y), we obtain

eikξ(xy)a(k1,x,y,ξ)𝑑ξsk(x)s¯k(y)=eikξ(xy)a(k1,x,y,ξ+ψ(x,y))𝑑ξtk(x)t¯k(y)\displaystyle\begin{split}&\int e^{ik\xi\cdot(x-y)}a(k^{-1},x,y,\xi)\;d\xi\;s^{k}(x)\otimes\overline{s}^{k}(y)\\ &=\int e^{ik\xi\cdot(x-y)}a(k^{-1},x,y,\xi+\psi(x,y))\;d\xi\;t^{k}(x)\otimes\overline{t}^{k}(y)\end{split} (16)

So multiplying ss by eiφe^{i\varphi} amounts to change the amplitude aa to bb such that b(h,x,y,ξ)=a(h,x,y,ξ+ψ(x,y))b(h,x,y,\xi)=a(h,x,y,\xi+\psi(x,y)). This relation writes on the diagonal

b(h,x,x,ξ)=a(h,x,x,ξ+dφ(x)).b(h,x,x,\xi)=a(h,x,x,\xi+d\varphi(x)).

Let \nabla be a connection on LL preserving the metric. Then s=1iβss\nabla s=\frac{1}{i}\beta_{s}\otimes s, where βs\beta_{s} is a real one-form of UU. Observe that t=1iβtt\nabla t=\frac{1}{i}\beta_{t}\otimes t where βt=βsdφ\beta_{t}=\beta_{s}-d\varphi. So we can define the principal symbol as follows.

Definition 2.2.

The principal symbol σ(P){\sigma}_{\nabla}(P) of PΨtscm(L)P\in\Psi_{\operatorname{tsc}}^{m}(L) is the element of Sphm(M,TM)S^{m}_{\operatorname{ph}}(M,T^{*}M) such that for any local data (ρ,U,s,a)(\rho,U,s,a) as in Definition 2.1, we have

a(h,x,x,ξ+βs(x))=ρ(x)σ(P)(x,ξ)+𝒪(h)a(h,x,x,\xi+\beta_{s}(x))=\rho(x){\sigma}_{\nabla}(P)(x,\xi)+\mathcal{O}(h)

If \nabla^{\prime} is another connection of LL preserving the metric, then =+1iα\nabla^{\prime}=\nabla+\frac{1}{i}{\alpha} with αΩ1(M,){\alpha}\in\Omega^{1}(M,{\mathbb{R}}) and σ(P)(x,ξ)=σ(P)(x,ξ+α(x)){\sigma}_{\nabla^{\prime}}(P)(x,\xi)={\sigma}_{\nabla}(P)(x,\xi+{\alpha}(x)). It is easy to extend the basic properties of pseudodifferential operators to our setting:

  • -

    If PΨtscm(L)P\in\Psi_{\operatorname{tsc}}^{m}(L), then σ(P)=0{\sigma}_{\nabla}(P)=0 if and only if Pk1Ψtscm1(L)P\in k^{-1}\Psi_{\operatorname{tsc}}^{m-1}(L).

  • -

    mkmΨtscm(L)=kΨ(L)\bigcap_{m}k^{-m}\Psi_{\operatorname{tsc}}^{-m}(L)=k^{-\infty}\Psi^{-\infty}(L).

  • -

    if PΨtscm(L)P\in\Psi_{\operatorname{tsc}}^{m}(L) and QΨtscp(L)Q\in\Psi_{\operatorname{tsc}}^{p}(L), then

    1. i)

      (PkQk)(P_{k}Q_{k}) belongs to Ψtscm+p(L)\Psi_{\operatorname{tsc}}^{m+p}(L) and its principal symbol is the product of the principal symbols of PP and QQ.

    2. ii)

      ik[Pk,Qk]ik[P_{k},Q_{k}] belongs to Ψtscm+p1(L)\Psi_{\operatorname{tsc}}^{m+p-1}(L) and its symbol is the Poisson bracket for the twisted symplectic form (1) where 1iω\frac{1}{i}{\omega} is the curvature of \nabla.

It is possible to define the twisted pseudodifferential operators without using local frames. Recall that the Schwartz kernel of an operator acting on 𝒞(M,Lk){\mathcal{C}}^{\infty}(M,L^{k}) is a section of (LL¯)k(M|Λ|1(M))(L\boxtimes\overline{L})^{k}\otimes({\mathbb{C}}_{M}\boxtimes|\Lambda|^{1}(M)). Introduce an open neighborhood VV of the diagonal of M2M^{2} and a section F𝒞(V,LL¯)F\in{\mathcal{C}}^{\infty}(V,L\boxtimes\overline{L}) such that |F|=1|F|=1 on VV and F(x,x)=1F(x,x)=1 for any xMx\in M, in the sense that F(x,x)=uu¯F(x,x)=u\otimes\overline{u} for any uLxu\in L_{x} with norm 11. We claim that PΨtscm(L)P\in\Psi_{\operatorname{tsc}}^{m}(L) if and only if its Schwartz kernel has the form

Fk(x,y)ϕ(x,y)Kk1(x,y)+𝒪(k)\displaystyle F^{k}(x,y)\phi(x,y)K_{k^{-1}}(x,y)+\mathcal{O}_{\infty}(k^{-\infty}) (17)

where ϕ𝒞0(V)\phi\in{\mathcal{C}}^{\infty}_{0}(V) is equal to 11 on a neighborhood of the diagonal and (Kh,h(0,1])(K_{h},\;h\in(0,1]) is the Schwartz kernel family of a semiclassical pseudodifferential operator QΨscm(M)Q\in\Psi_{\operatorname{sc}}^{m}(M). If furthermore \nabla is a connection of LL such that the corresponding covariant derivative of FF is zero on the diagonal, then σ(P)=σ(Q)\sigma_{\nabla}(P)={\sigma}(Q).

These facts follow from a computation similar to (16), by writing F(x,y)=exp(iφ(x,y))s(x)s¯(y)F(x,y)=\exp(i\varphi(x,y))s(x)\otimes\overline{s}(y) where s𝒞(U,L)s\in{\mathcal{C}}^{\infty}(U,L) is such that |s|=1|s|=1 on UU and φ𝒞(U2,)\varphi\in{\mathcal{C}}^{\infty}(U^{2},{\mathbb{R}}) satisfies φ(x,x)=0\varphi(x,x)=0.

Semiclassical Sobolev norms

Let mm\in{\mathbb{R}}. Denote by Hm(M,Lk)H^{m}(M,L^{k}) the Sobolev space of sections of LkL^{k} of order mm. Let us give three equivalent definitions of the semiclassical Sobolev norms of a section uu of LkL^{k}. First the norm of H0(M,Lk)=L2(M,Lk)H^{0}(M,L^{k})=L^{2}(M,L^{k}) is defined by

uL2(M,Lk)2=M|u(x)|2𝑑μ(x)\displaystyle\|u\|_{L^{2}(M,L^{k})}^{2}=\int_{M}|u(x)|^{2}d\mu(x) (18)

where μ\mu is a volume element of MM independent of kk.

  1. 1.

    only for integral exponent mm\in{\mathbb{N}}: choose a connection \nabla of LL, vector fields (Xi)i=1N(X_{i})_{i=1}^{N} of MM which generates TxMT_{x}M at each xx, and set

    um:=|α|mk|α|XαuL2(M,Lk)\|u\|_{m}:=\sum_{|{\alpha}|\leqslant m}k^{-|{\alpha}|}\|\nabla_{X}^{\alpha}u\|_{L^{2}(M,L^{k})}

    where for any αN{\alpha}\in{\mathbb{N}}^{N}, Xα=X1α(1)XNα(N)\nabla_{X}^{\alpha}=\nabla_{X_{1}}^{{\alpha}(1)}\ldots\nabla_{X_{N}}^{{\alpha}(N)}.

  2. 2.

    based on local semi-norms: for any chart (U,χ)(U,\chi) of MM, frame s𝒞(U,L)s\in{\mathcal{C}}^{\infty}(U,L) such that |s|=1|s|=1 and ρ𝒞o(U)\rho\in{\mathcal{C}}^{\infty}_{o}(U) we set

    um,U,χ,s,ρ=k1ξmv^(ξ)L2(n)where ρu=(χv)sk\|u\|_{m,U,\chi,s,\rho}=\|\langle k^{-1}\xi\rangle^{m}\hat{v}(\xi)\|_{L^{2}({\mathbb{R}}^{n})}\qquad\text{where }\rho u=(\chi^{*}v)s^{k}

    and v^\hat{v} is the Fourier transform of vv. Choose a finite family (Ui,χi,si,ρi)(U_{i},\chi_{i},s_{i},\rho_{i}) of local data such that MM is covered by the {ρi=1}\{\rho_{i}=1\} and set um:=ium,Ui,χi,si,ρi\|u\|_{m}:=\sum_{i}\|u\|_{m,U_{i},\chi_{i},s_{i},\rho_{i}}.

  3. 3.

    based on twisted pseudodifferential operators: choose EΨtscm(L)E\in\Psi_{\operatorname{tsc}}^{m}(L) which is elliptic and invertible for any kk, and set um=EuL2(M,Lk)\|u\|_{m}=\|Eu\|_{L^{2}(M,L^{k})}.

The ellipticity condition is as usual that the principal symbol satisfies for some C>0C>0, |σ(L)(x,ξ)|C1|ξ|m|{\sigma}_{\nabla}(L)(x,\xi)|\geqslant C^{-1}|\xi|^{m} when |ξ|C|\xi|\geqslant C. It does not depend on the choice of \nabla.

We claim that all these norms are equivalent with constants uniform in kk. Furthermore for any twisted pseudodifferential operator PΨtscp(L)P\in\Psi_{\operatorname{tsc}}^{p}(L) and any mm\in{\mathbb{R}}, there exists CC such that for any kk,

PkumCum+p,u𝒞(M,Lk).\displaystyle\|P_{k}u\|_{m}\leqslant C\|u\|_{m+p},\qquad\forall\;u\in{\mathcal{C}}^{\infty}(M,L^{k}). (19)

3 Heisenberg semiclassical operator

In the introduction, we defined the Heisenberg pseudodifferential operators by expressing locally their Schwartz kernels as oscillatory integrals. Here we will start with a global definition which has the advantage that we can deduce some basic properties of these operators directly from the ones of the semiclassical pseudodifferential operators.

Let LML\rightarrow M be a Hermitian line bundle with a connection \nabla preserving the metric. The line bundle LL¯L\boxtimes\overline{L} inherits from LL a Hermitian metric and a connection. Its restriction to the diagonal is the flat trivial bundle with a natural trivialisation obtained by sending uv¯LxL¯xu\otimes\overline{v}\in L_{x}\otimes\overline{L}_{x} to the scalar product of uu and vv. In the sequel we will use a particular extension of this trivialisation.

Lemma 3.1.

There exist a tubular neighborhood VV of the diagonal of M2M^{2} and F𝒞(V,LL¯)F\in{\mathcal{C}}^{\infty}(V,L\boxtimes\overline{L}) such that |F|=1|F|=1 on VV and

F(x,x)=1,F(x,x)=0,YYF(x,x)=0xMF(x,x)=1,\quad\nabla F(x,x)=0,\quad\nabla_{Y}\nabla_{Y}F(x,x)=0\qquad\forall x\in M

for any vector field YY of M2M^{2} having the form Y(x,y)=(X(x),X(y))Y(x,y)=(X(x),-X(y)) with X𝒞(M,TM)X\in{\mathcal{C}}^{\infty}(M,TM). If (V,F)(V^{\prime},F^{\prime}) satisfies the same conditions, then F=Fexp(iψ)F=F^{\prime}\exp(i\psi) where ψ𝒞(VV,)\psi\in{\mathcal{C}}^{\infty}(V\cap V^{\prime},{\mathbb{R}}) vanishes to third order along the diagonal

Proof.

Consider more generally a closed submanifold NN of MM, a flat section EE of L|NL|_{N}, and a subbundle 𝒟\mathcal{D} of TM|NTM|_{N} such that 𝒟TN=TM|N\mathcal{D}\oplus TN=TM|_{N}. Then we can extend EE to a neighborhood of NN in such a way that it satisfies on NN: E=0\nabla E=0 and XXE=0\nabla_{X}\nabla_{X}E=0 for any vector field XX of MM such that X|NX|_{N} is a section of 𝒟\mathcal{D}. To see this, introduce a coordinate chart (U,xi,yj)(U,x_{i},y_{j}) of MM and a unitary frame s:ULs:U\rightarrow L such that NU={x1==xk=0}N\cap U=\{x_{1}=\ldots=x_{k}=0\}, (x1,,xk)(\partial_{x_{1}},\ldots,\partial_{x_{k}}) is a frame of 𝒟\mathcal{D} and ss extends EE. Then the section we are looking for is eiφse^{i\varphi}s with

φ=i=1kβi(0,y)xi+12i,j=1k(xjβi)(0,y)xixj+𝒪(|x|3)\varphi=\sum_{i=1}^{k}\beta_{i}(0,y)x_{i}+\tfrac{1}{2}\sum_{i,j=1}^{k}(\partial_{x_{j}}\beta_{i})(0,y)x_{i}x_{j}+\mathcal{O}(|x|^{3})

where the βi\beta_{i}’s are the functions in 𝒞(U){\mathcal{C}}^{\infty}(U) such that xis=1iβis\nabla_{\partial_{x_{i}}}s=\frac{1}{i}\beta_{i}\,s. Applying this to M2M^{2}, LL¯L\boxtimes\overline{L} and diagM\operatorname{diag}M instead of MM, LL, NN concludes the proof. ∎

Definition 3.2.

A semiclassical Heisenberg pseudodifferential operator of order mm\in{\mathbb{R}} is a family of operators (Pk:𝒞(M,Lk)𝒞(M,Lk),k)(P_{k}:{\mathcal{C}}^{\infty}(M,L^{k})\rightarrow{\mathcal{C}}^{\infty}(M,L^{k}),\;k\in{\mathbb{N}}) whose Schwartz kernels have the form

Fk(x,y)ϕ(x,y)Kk12(x,y)+𝒪(k)\displaystyle F^{k}(x,y)\phi(x,y)K_{k^{-\frac{1}{2}}}(x,y)+\mathcal{O}_{\infty}(k^{-\infty}) (20)

where (V,F)(V,F) satisfies the conditions of Lemma 3.1, ϕ𝒞0(V)\phi\in{\mathcal{C}}^{\infty}_{0}(V) is equal to 11 on a neighborhood of the diagonal and (Kh,h(0,1])(K_{h},\;h\in(0,1]) is the Schwartz kernel family of a semiclassical pseudodifferential operator (Qh)Ψscm(M)(Q_{h})\in\Psi_{\operatorname{sc}}^{m}(M).

The principal symbol σ(P){\sigma}(P) of (Pk)(P_{k}) is defined as the principal symbol of (Qh)(Q_{h}).

We denote by ΨHeism(L,)\Psi^{m}_{\operatorname{Heis}}(L,\nabla) the space of semiclassical Heisenberg pseudodifferential operators of order mm. For any PΨHeism(L,)P\in\Psi^{m}_{\operatorname{Heis}}(L,\nabla), for any fixed kk, PkP_{k} is a pseudo-differential operator of order mm, so PkP_{k} act on 𝒞(M,Lk){\mathcal{C}}^{\infty}(M,L^{k}) and on 𝒞(M,Lk)\mathcal{C}^{-\infty}(M,L^{k}). The definition clearly does not depend on the choice of the cutoff function ϕ\phi. It neither doesn’t depend on the choice of FF as will be explained below. To compare with the twisted pseudodifferential operators, observe first that the section FF in (17) satisfies a weaker condition than in Definition 3.2 and second in (17), the Schwartz kernel of QQ is evaluated at h=k1h=k^{-1}, whereas in (20) we have h=k1/2h=k^{-1/2}.

By defining globally the Heisenberg pseudodifferential operators in terms of scalar pseudodifferential operators as in Definition 3.2 instead of the local oscillatory integrals (9), we avoid the usual discussions on the coordinate changes and the principal symbol and we deduce easily the following three facts:

  • -

    If PΨHeism(L,)P\in\Psi_{\operatorname{Heis}}^{m}(L,\nabla), then σ(P)=0{\sigma}(P)=0 if and only if Pk12ΨHeism1(L,)P\in k^{-\frac{1}{2}}\Psi_{\operatorname{Heis}}^{m-1}(L,\nabla).

  • -

    mkm2ΨHeism(L,)=kΨ(L)\bigcap_{m}k^{-\frac{m}{2}}\Psi_{\operatorname{Heis}}^{-m}(L,\nabla)=k^{-\infty}\Psi^{-\infty}(L).

  • -

    If PΨHeism(L)P\in\Psi_{\operatorname{Heis}}^{m}(L) and ρ𝒞(M2)\rho\in{\mathcal{C}}^{\infty}(M^{2}) is such that suppρdiagM=\operatorname{supp}\rho\cap\operatorname{diag}M=\emptyset, then the kernel (x,y)ρ(x,y)Pk(x,y)(x,y)\rightarrow\rho(x,y)P_{k}(x,y) is smooth and negligible.

Unfortunately, the definition 3.2 does not allow to deduce the composition properties of the Heisenberg operators from the one of the semiclassical pseudodifferential operators.

By Lemma 3.1, FF is uniquely defined modulo a factor eiψe^{i\psi} with ψ𝒞(U2)\psi\in{\mathcal{C}}^{\infty}(U^{2}) vanishing to third order along the diagonal. Write

ψ=|α|=3ψα(x,y)(xy)α\psi=\sum_{|{\alpha}|=3}\psi_{{\alpha}}(x,y)(x-y)^{{\alpha}}

with smooth coefficients ψα\psi_{\alpha}. For any symbol aS(U2,n)a\in S^{\infty}(U^{2},{\mathbb{R}}^{n}), let I(a)I(a) be the oscillatory integral

I(a)(h,x,y)=eih1ξ(xy)a(h,x,y,ξ)𝑑ξ.I(a)(h,x,y)=\int e^{ih^{-1}\xi\cdot(x-y)}a(h,x,y,\xi)\;d\xi.
Lemma 3.3.

For all aSm(U2,n)a\in S^{m}(U^{2},{\mathbb{R}}^{n}), eih2ψ(x,y)I(a)(h,x,y)=I(b)(h,x,y)e^{ih^{-2}\psi(x,y)}I(a)(h,x,y)=I(b)(h,x,y) with bSm(U2,n)b\in S^{m}(U^{2},{\mathbb{R}}^{n}) having the asymptotic expansion

b==0h!L(a), with L=|α|=3ψα(x,y)ξα.\displaystyle b=\sum_{\ell=0}^{\infty}\frac{h^{\ell}}{\ell!}L^{\ell}(a),\qquad\text{ with }\quad L=\sum_{|{\alpha}|=3}\psi_{{\alpha}}(x,y)\partial_{\xi}^{\alpha}. (21)

In particular, if aSscm(U2,n)a\in S^{m}_{\operatorname{sc}}(U^{2},{\mathbb{R}}^{n}), then bSscm(U2,n)b\in S^{m}_{\operatorname{sc}}(U^{2},{\mathbb{R}}^{n}).

This proves that Definition 3.2 does not depend on the choice of FF. Moreover, since b=a+𝒪(h)b=a+\mathcal{O}(h), the principal symbol of (Qk)(Q_{k}) is also independent of FF.

Proof.

By integration by part, (xiyi)I(a)=ihI(ξia)(x_{i}-y_{i})I(a)=ihI(\partial_{\xi_{i}}a), so ih2ψI(a)=hI(L(a))ih^{-2}\psi I(a)=hI(L(a)) with LL given by (21). By Taylor formula, it comes that

eih2ψI(a)==0Nh!I(L(a))+hN+1rNI(LN+1(a))e^{ih^{-2}\psi}I(a)=\sum_{\ell=0}^{N}\frac{h^{\ell}}{\ell!}I(L^{\ell}(a))+h^{N+1}r_{N}I(L^{N+1}(a))

with

rN(h,x,y)=1N!01eith2ψ(x,y)(1t)N𝑑t.r_{N}(h,x,y)=\frac{1}{N!}\int_{0}^{1}e^{ith^{-2}\psi(x,y)}(1-t)^{N}\;dt.

Observe that rNr_{N} is smooth and h2|α|x,yαrN=𝒪(1)h^{2|{\alpha}|}\partial^{\alpha}_{x,y}r_{N}=\mathcal{O}(1). Furthermore, since I(a)I(a) is a genuine integral for m<nm<-n, by derivating under the integral sign, when kk\in{\mathbb{N}} satisfies k+m<nk+m<-n, I(a)𝒞kI(a)\in\mathcal{C}^{k} and for any |α|=k|{\alpha}|=k, h|α|x,yαI(a)=O(1)h^{|{\alpha}|}\partial^{\alpha}_{x,y}I(a)=O(1). Since LN+1(a)L^{N+1}(a) is a symbol of order m3(N+1)m-3(N+1), it comes that for any kk, when NN is sufficiently large, rNI(LN+1(a))r_{N}I(L^{N+1}(a)) is of class 𝒞k\mathcal{C}^{k} and for any |α|=k|{\alpha}|=k, h2|α|x,yα(rNI(LN+1(a)))=𝒪(1)h^{2|{\alpha}|}\partial^{\alpha}_{x,y}(r_{N}I(L^{N+1}(a)))=\mathcal{O}(1).

So for any bSm(U2,n)b\in S^{m}(U^{2},{\mathbb{R}}^{n}) having the asymptotic expansion (21), we have that eih2ψI(a)=I(b)+ρe^{ih^{-2}\psi}I(a)=I(b)+\rho with ρh𝒞(U2)\rho\in h^{\infty}{\mathcal{C}}^{\infty}(U^{2}), and we can absorb ρ\rho in I(b)I(b) by modifying bb by a summand in hS(U2,n)h^{\infty}S^{-\infty}(U^{2},{\mathbb{R}}^{n}). ∎

Let us explain how we recover the local expression (9) of the introduction. Let (U,xi)(U,x_{i}) be a local chart of MM and s𝒞(U,L)s\in{\mathcal{C}}^{\infty}(U,L) such that |s|=1|s|=1 on UU. Let βΩ1(U,)\beta\in\Omega^{1}(U,{\mathbb{R}}) be the connection form, s=1iβs\nabla s=\frac{1}{i}\beta\otimes s. Then we easily check that the section F𝒞(U2,LL¯)F\in{\mathcal{C}}^{\infty}(U^{2},L\boxtimes\overline{L}) given by

F(x,y)=eiβ(x+y2)(xy)s(x)s¯(y)\displaystyle F(x,y)=e^{i\beta\bigl{(}\tfrac{x+y}{2}\bigr{)}\cdot(x-y)}s(x)\otimes\overline{s}(y) (22)

satisfies the condition of Lemma 3.1. Consequently, the Schwartz kernel of an operator in ΨHeism(L)\Psi^{m}_{\operatorname{Heis}}(L) has the form Kksks¯kK_{k}s^{k}\boxtimes\overline{s}^{k} on U2U^{2} with

Kk(x,y)=eikβ(x+y2)(xy)(k2π)neikξ(xy)a(k12,x,y,ξ)𝑑ξ\displaystyle K_{k}(x,y)=e^{ik{\beta}\bigl{(}\tfrac{x+y}{2}\bigr{)}\cdot(x-y)}\Bigl{(}\frac{\sqrt{k}}{2\pi}\Bigr{)}^{n}\int e^{i\sqrt{k}\;\xi\cdot(x-y)}a(k^{-\frac{1}{2}},x,y,\xi)\;d\xi (23)

with aSscm(U2,n)a\in S^{m}_{\operatorname{sc}}(U^{2},{\mathbb{R}}^{n}). Of course, we can assume that aa does not depend on yy (resp. xx) or that it is on the Weyl form a(h,x,y,ξ)=b(h,12(x+y),ξ)a(h,x,y,\xi)=b(h,\frac{1}{2}(x+y),\xi) with bSscm(U,n)b\in S^{m}_{\operatorname{sc}}(U,{\mathbb{R}}^{n}). In this last case, we recover exactly the expression (9).

Another interesting expression is obtained by rescaling the variable ξ\xi by a square root of kk in (23) and absorbing the β\beta factor into the amplitude:

Kk(x,y)=eikβ(x+y2)(xy)(k2π)neikξ(xy)a(k12,x,y,kξ)𝑑ξ=(k2π)neikξ(xy)a(k12,x,y,k(ξβ(x+y2)))𝑑ξ\displaystyle\begin{split}K_{k}(x,y)&=e^{ik{\beta}\bigl{(}\tfrac{x+y}{2}\bigr{)}\cdot(x-y)}\Bigl{(}\frac{k}{2\pi}\Bigr{)}^{n}\int e^{ik\;\xi\cdot(x-y)}a(k^{-\frac{1}{2}},x,y,\sqrt{k}\;\xi)\;d\xi\\ &=\Bigl{(}\frac{k}{2\pi}\Bigr{)}^{n}\int e^{ik\;\xi\cdot(x-y)}a\bigl{(}k^{-\frac{1}{2}},x,y,\sqrt{k}\bigl{(}\xi-{\beta}\bigl{(}\tfrac{x+y}{2}\bigr{)}\bigr{)}\bigr{)}\;d\xi\end{split} (24)

Assume that aa is on the Weyl form, a(h,x,y,ξ)=b(h,12(x+y),ξ)a(h,x,y,\xi)=b(h,\frac{1}{2}(x+y),\xi), then we have that

Kk(x,y)=(k2π)neikξ(xy)b~(k1,12(x+y),ξ)𝑑ξ\displaystyle K_{k}(x,y)=\Bigl{(}\frac{k}{2\pi}\Bigr{)}^{n}\int e^{ik\;\xi\cdot(x-y)}\tilde{b}\bigl{(}k^{-1},\tfrac{1}{2}(x+y),\xi\bigr{)}\;d\xi (25)

where b~(h,x,ξ)=b(h,x,h12(ξβ(x)))\tilde{b}(h,x,\xi)=b(\sqrt{h},x,h^{-\frac{1}{2}}(\xi-{\beta}(x))). So we recognise a semiclassical pseudodifferential operator at k=h1k=h^{-1} with a Weyl symbol b~\tilde{b}.

We call b~\tilde{b} the effective symbol. As we will see, it satisfies some exotic estimates. Let us introduce the symbol semi-norms of Sm(U,n)S^{m}(U,{\mathbb{R}}^{n}),

am,,K=max|α|+|β|supxK,ξn|xαξβa(x,ξ)|ξm+|β|\|a\|_{m,\ell,K}=\max_{|{\alpha}|+|{\beta}|\leqslant\ell}\sup_{x\in K,\xi\in{\mathbb{R}}^{n}}|\partial_{x}^{{\alpha}}\partial_{\xi}^{{\beta}}a(x,\xi)|\langle\xi\rangle^{-m+|{\beta}|}

where KK is a compact subset of UU.

Lemma 3.4.

For any mm\in{\mathbb{R}}, α,βn{\alpha},{\beta}\in{\mathbb{N}}^{n} and compact subset KK of UU, there exists C>0C>0 such that for any a𝒞(U×n)a\in{\mathcal{C}}^{\infty}(U\times{\mathbb{R}}^{n}), the function a~(h,x,ξ)=a(x,h12(ξβ(x))\tilde{a}(h,x,\xi)=a(x,h^{-\frac{1}{2}}(\xi-{\beta}(x)) satisfies

|xαξβa~(h,x,ξ)|Cam,,Kh12(m++)ξm|β||\partial_{x}^{{\alpha}}\partial_{\xi}^{{\beta}}\tilde{a}(h,x,\xi)|\leqslant C\|a\|_{m,\ell,K}h^{-\frac{1}{2}(m_{+}+\ell)}\langle\xi\rangle^{m-|{\beta}|}

for all 0<h10<h\leqslant 1, xKx\in K, ξn\xi\in{\mathbb{R}}^{n} with =|α|+|β|\ell=|{\alpha}|+|{\beta}|, m+=max(m,0)m_{+}=\max(m,0).

Proof.

For any 0<ϵ10<\epsilon\leqslant 1, we have ηϵ1ηϵ1η\langle\eta\rangle\leqslant\langle\epsilon^{-1}\eta\rangle\leqslant\epsilon^{-1}\langle\eta\rangle. Furthermore, if xKx\in K, C1ξξβ(x)CξC^{-1}\langle\xi\rangle\leqslant\langle\xi-{\beta}(x)\rangle\leqslant C\langle\xi\rangle. So for any mm\in{\mathbb{R}},

ϵ1(ξβ(x))mCmϵm+ξm.\displaystyle\langle\epsilon^{-1}(\xi-{\beta}(x))\rangle^{m}\leqslant C_{m}\epsilon^{-m_{+}}\langle\xi\rangle^{m}. (26)

The derivatives of a~ϵ(x,ξ)=a(x,ϵ1(ξβ(x)))\tilde{a}_{\epsilon}(x,\xi)=a(x,\epsilon^{-1}(\xi-{\beta}(x))) have the form xαξβa~ϵ=b~ϵ\partial_{x}^{\alpha}\partial_{\xi}^{\beta}\tilde{a}_{\epsilon}=\tilde{b}_{\epsilon} with

b=α,βϵ|β|fα,βxαξβa\displaystyle b=\sum_{{\alpha}^{\prime},{\beta}^{\prime}}\epsilon^{-|{\beta}^{\prime}|}f_{{\alpha}^{\prime},{\beta}^{\prime}}\partial_{x}^{{\alpha}^{\prime}}\partial_{\xi}^{{\beta}^{\prime}}a (27)

where the coefficients fα,βf_{{\alpha}^{\prime},{\beta}^{\prime}} are in 𝒞(U){\mathcal{C}}^{\infty}(U) and don’t depend on aa, and we sum over the multi-indices satisfying ββ{\beta}\leqslant{\beta}^{\prime} and |α|+|β||α|+|β||{\alpha}^{\prime}|+|{\beta}^{\prime}|\leqslant|{\alpha}|+|{\beta}|. So for xKx\in K,

|xαξβa~ϵ(x,ξ)|\displaystyle|\partial_{x}^{\alpha}\partial_{\xi}^{\beta}\tilde{a}_{\epsilon}(x,\xi)| Cα,βϵ|β|am,,Kϵ1(ξβ(x))m|β| by (27)\displaystyle\leqslant C\sum_{{\alpha}^{\prime},{\beta}^{\prime}}\epsilon^{-|{\beta}^{\prime}|}\|a\|_{m,\ell,K}\langle\epsilon^{-1}(\xi-{\beta}(x))\rangle^{m-|{\beta}^{\prime}|}\quad\text{ by }\eqref{eq:e2}
Cα,βϵ|β|am,,Kξm|β|ϵm+ by (26)\displaystyle\leqslant C\sum_{{\alpha}^{\prime},{\beta}^{\prime}}\epsilon^{-|{\beta}^{\prime}|}\|a\|_{m,\ell,K}\langle\xi\rangle^{m-|{\beta}^{\prime}|}\epsilon^{-m_{+}}\quad\text{ by }\eqref{eq:e1}
Cϵ(|α|+|β|)am,,Kξm|β|ϵm+\displaystyle\leqslant C\epsilon^{-(|{\alpha}|+|{\beta}|)}\|a\|_{m,\ell,K}\langle\xi\rangle^{m-|{\beta}|}\epsilon^{-m_{+}}

because |β||β||α|+|β||{\beta}|\leqslant|{\beta}^{\prime}|\leqslant|{\alpha}|+|{\beta}| and we conclude by setting ϵ=h12\epsilon=h^{\frac{1}{2}}. ∎

So b~\tilde{b} belongs to the class SδS_{\delta} with exponent δ=1/2\delta=1/2, that is at each derivative we loose a factor hδh^{-\delta}. Recall that δ=1/2\delta=1/2 is the critical exponent: the space of pseudodifferential operators with symbol in SδS_{\delta} is an algebra for δ[0,1/2]\delta\in[0,1/2], but the standard asymptotic expansions of the symbolic calculus only hold for δ[0,1/2[\delta\in[0,1/2[, cf. for instance [10, Proposition 7.7]. As we will see in Section 5 and in [5], the Heisenberg pseudodifferential operators form an algebra and have an associated symbol calculus, but this can not be deduced form the usual composition rules of pseudodifferential operators. Nevertheless, Lemma 3.4 has some useful consequences, the first of them being the L2L^{2} mapping property. Recall the definition (18) of the L2L^{2}-norm with a volume element independent of kk.

Theorem 3.5.

For any QΨHeis0(L)Q\in\Psi^{0}_{\operatorname{Heis}}(L), there exists C>0C>0 such that for any kk, Qk(L2(M,Lk))C\|Q_{k}\|_{\mathcal{L}(L^{2}(M,L^{k}))}\leqslant C.

Proof.

Introduce a finite atlas (Ui,ϕi)(U_{i},\phi_{i}) of MM with functions φi,ψi𝒞0(Ui)\varphi_{i},\psi_{i}\in{\mathcal{C}}^{\infty}_{0}(U_{i}) such that φi=1\sum\varphi_{i}=1 and suppφiint{ψi=1}\operatorname{supp}\varphi_{i}\subset\operatorname{int}\{\psi_{i}=1\}. Write

P=ψiPφi+Q.\displaystyle P=\sum\psi_{i}P\varphi_{i}+Q. (28)

Since ψi(x)φi(y)=1\sum\psi_{i}(x)\varphi_{i}(y)=1 when xx is close to yy, QQ is in kΨk^{-\infty}\Psi^{-\infty}. Identifying UiU_{i} with ϕi(Ui)\phi_{i}(U_{i}), the Schwartz kernel of ψiPφi\psi_{i}P\varphi_{i} has the form (25) with a symbol b~i\tilde{b}_{i} satisfying

|xαξβb~i(h,x,ξ)|h12(|α|+|β|)Cα,β.|\partial^{\alpha}_{x}\partial^{\beta}_{\xi}\tilde{b}_{i}(h,x,\xi)|\leqslant h^{-\frac{1}{2}(|{\alpha}|+|{\beta}|)}C_{{\alpha},{\beta}}.

By [10, Theorem 7.11], ψiPφi=𝒪(1):L2(n)L2(n)\psi_{i}P\varphi_{i}=\mathcal{O}(1):L^{2}({\mathbb{R}}^{n})\rightarrow L^{2}({\mathbb{R}}^{n}). ∎

Another consequence of Lemma 3.4 is the following important fact.

Lemma 3.6.

kΨ(L)k^{-\infty}\Psi^{-\infty}(L) is a bilateral ideal of ΨHeis(L,)\Psi_{\operatorname{Heis}}(L,\nabla).

Proof.

Consider a pseudodifferential operator A(h)A(h) of n{\mathbb{R}}^{n} with the Schwartz kernel (2πh)neih1ξ(xy)a(h,x,y,ξ)𝑑ξ(2\pi h)^{-n}\int e^{ih^{-1}\xi(x-y)}a(h,x,y,\xi)d\xi where the amplitude a(h,x,y,ξ)a(h,x,y,\xi) is zero if |x|+|y|C|x|+|y|\geqslant C and satisfies |x,yαa(h,x,y,ξ)|h|α|Cαξm|\partial^{\alpha}_{x,y}a(h,x,y,\xi)|\leqslant h^{-|{\alpha}|}C_{{\alpha}}\langle\xi\rangle^{m} for any α{\alpha}. Then, with the usual regularisation of oscillatory integrals by integration by part, one proves that

h|α|sup|αA(h)u|Cαmax|β|m+n+1+|α|suph|β||βu|,h(0,1]h^{|{\alpha}|}\sup|\partial^{{\alpha}}A(h)u|\leqslant C^{\prime}_{{\alpha}}\max_{|{\beta}|\leqslant m+n+1+|{\alpha}|}\sup h^{|{\beta}|}|\partial^{\beta}u|,\qquad\forall h\in(0,1]

So for any family of operators B(h):𝒞(n)𝒞(n)B(h):{\mathcal{C}}^{\infty}({\mathbb{R}}^{n})\rightarrow{\mathcal{C}}^{\infty}({\mathbb{R}}^{n}), h(0,1]h\in(0,1], if B(h)B(h) has a compactly supported smooth kernel in 𝒪(h)\mathcal{O}_{\infty}(h^{\infty}) then the same holds for A(h)B(h)A(h)\circ B(h). Applying this to the ψiPφi\psi_{i}P\varphi_{i} of (28) proves the result. ∎

To end this section, let us extend the mapping property to the Sobolev space. We denote by m\|\cdot\|_{m} the mm-th semiclassical Sobolev norm of sections of LkL^{k}, defined as in Section 2.

Theorem 3.7.

For any m,pm,p\in{\mathbb{R}} and any QΨHeism(L)Q\in\Psi_{\operatorname{Heis}}^{m}(L), there exists C>0C>0 such that for any kk,

QkupCk12m+up+m,u𝒞(M,Lk).\|Q_{k}u\|_{p}\leqslant Ck^{\frac{1}{2}m_{+}}\|u\|_{p+m},\qquad\forall u\in{\mathcal{C}}^{\infty}(M,L^{k}).

Since for any kk, QkQ_{k} is a pseudodifferential operator of order mm of LkL^{k}, we already know that PkP_{k} is continuous Hp+m(M,Lk)Hm(M,Lk)H^{p+m}(M,L^{k})\rightarrow H^{m}(M,L^{k}). Theorem 3.7 gives a uniform estimate with respect to kk.

Proof.

It suffices to prove that for any EΨtscpm(L)E\in\Psi_{\operatorname{tsc}}^{p-m}(L) and EΨtscp(L)E^{\prime}\in\Psi_{\operatorname{tsc}}^{-p}(L) one has

EkPkEk=𝒪(k12m+):L2(M,Lk)L2(M,Lk).\displaystyle E^{\prime}_{k}P_{k}E_{k}=\mathcal{O}(k^{\frac{1}{2}m_{+}}):L^{2}(M,L^{k})\rightarrow L^{2}(M,L^{k}). (29)

For this it suffices to prove that for any chart domain UU of MM and functions ρj𝒞0(U)\rho_{j}\in{\mathcal{C}}^{\infty}_{0}(U), j=1,,4j=1,...,4, one has

ρ1Ekρ2Pkρ3Ekρ4=𝒪(k12m+):L2(M,Lk)L2(M,Lk).\displaystyle\rho_{1}E^{\prime}_{k}\,\rho_{2}P_{k}\,\rho_{3}E_{k}\,\rho_{4}=\mathcal{O}(k^{\frac{1}{2}m_{+}}):L^{2}(M,L^{k})\rightarrow L^{2}(M,L^{k}). (30)

To show that (30) implies (29), write PP on the form (28), Eψi=ψ~iEψi+(1ψ~i)EψiE^{\prime}\psi_{i}=\tilde{\psi}_{i}E^{\prime}\psi_{i}+(1-\tilde{\psi}_{i})E^{\prime}\psi_{i} with ψ~i𝒞0(Ui)\tilde{\psi}_{i}\in{\mathcal{C}}^{\infty}_{0}(U_{i}) such that suppψiint{ψ~i=1}\operatorname{supp}\psi_{i}\subset\operatorname{int}\{\tilde{\psi}_{i}=1\} and similarly for EE, and use that kΨ(L)k^{-\infty}\Psi^{-\infty}(L) is an ideal of both ΨHeis(L,)\Psi_{\operatorname{Heis}}^{\infty}(L,\nabla) and Ψtsc(L)\Psi_{\operatorname{tsc}}^{\infty}(L).

As in [10, Definition 7.5], for δ[0,1]\delta\in[0,1] and m:n[0,)m:{\mathbb{R}}^{n}\rightarrow[0,\infty) an order function, let Sδ(m)S_{\delta}(m) be the space of families (a(h),h(0,1])(a(h),\;h\in(0,1]) of 𝒞(n){\mathcal{C}}^{\infty}({\mathbb{R}}^{n}) such that |αa(h,x)|Cαhδ|α|m(x)|\partial^{\alpha}a(h,x)|\leqslant C_{{\alpha}}h^{-\delta|{\alpha}|}m(x). Identify UU with ϕ(U)\phi(U) and denote by Opk(b~)\operatorname{Op}_{k}(\tilde{b}) the operator with kernel (25).

Then for any ρ,ρ𝒞0(U)\rho,\rho^{\prime}\in{\mathcal{C}}^{\infty}_{0}(U), ρEkρ\rho E^{\prime}_{k}\,\rho^{\prime}, k12m+ρPkρk^{-\frac{1}{2}m_{+}}\rho P_{k}\,\rho^{\prime} and ρEkρ\rho E_{k}\,\rho^{\prime} are equal to Opk(b~)\operatorname{Op}_{k}(\tilde{b}) with b~\tilde{b} in S0(ξp)S_{0}(\langle\xi\rangle^{-p}), S1/2(ξm)S_{1/2}(\langle\xi\rangle^{m}) and S0(ξpm)S_{0}(\langle\xi\rangle^{p-m}) respectively. By [10, Proposition 7.7, Theorem 7.9], their product is equal to Opk(c)\operatorname{Op}_{k}(c) with cS1/2(1)c\in S_{1/2}(1), which proves (30) by [10, Theorem 7.11]. ∎

Actually, Theorem 3.7 can be improved if we use Sobolev norms associated to the covariant derivative \nabla instead of the semiclassical Sobolev norms. For instance, for any mm\in{\mathbb{N}}, any QΨHeism(L,)Q\in\Psi_{\operatorname{Heis}}^{-m}(L,\nabla) and any vector fields X1X_{1}, …, XmX_{m} of MM, we will see in Proposition 5.1, that P=(km/2X1XmQk)P=(k^{-m/2}\nabla_{X_{1}}\ldots\nabla_{X_{m}}Q_{k}) belongs to ΨHeis0(L,)\Psi_{\operatorname{Heis}}^{0}(L,\nabla), so by Theorem 3.5,

km/2X1XmQk=𝒪(1):L2(M,Lk)L2(M,Lk)\displaystyle k^{-m/2}\nabla_{X_{1}}\ldots\nabla_{X_{m}}Q_{k}=\mathcal{O}(1):L^{2}(M,L^{k})\rightarrow L^{2}(M,L^{k}) (31)

To compare, Theorem 3.7 only implies that the norm of PkP_{k} in (L2(M,Lk))\mathcal{L}(L^{2}(M,L^{k})) is in 𝒪(km/2)\mathcal{O}(k^{m/2}). The generalisation of (31) to fractional exponents mm not necessarily nonnegative will be given in [5].

4 A product associated to an antisymmetric bilinear form

Let EE be a nn-dimensional real vector space and A2EA\in\wedge^{2}E^{*}. Later, we will choose E=TxME=T_{x}M with A=ω(x)A={\omega}(x). Introduce the covariant derivative of EE

A=d+1iβ, where βΩ1(E,),β(x)(Y)=12A(x,Y).\displaystyle\nabla^{A}=d+\tfrac{1}{i}\beta,\qquad\text{ where }{\beta}\in\Omega^{1}(E,{\mathbb{R}}),\quad\beta(x)(Y)=\tfrac{1}{2}A(x,Y). (32)

The curvature of A\nabla^{A} is 1iA\frac{1}{i}A, that is [XA,YA]=1iA(X,Y)[\nabla_{X}^{A},\nabla_{Y}^{A}]=\frac{1}{i}A(X,Y) for any X,YEX,Y\in E. We will define for any tempered distribution g𝒮(E)g\in\mathcal{S}^{\prime}(E^{*}) an operator g(1iA)g(\frac{1}{i}\nabla^{A}).

We assume first that E=nE={\mathbb{R}}^{n}. For any g𝒮(n)g\in\mathcal{S}^{\prime}({\mathbb{R}}^{n}), we denote by g^\widehat{g} and gg^{\vee} its Fourier transform and inverse Fourier transform, with the normalisation

g^(ξ)=neixξg(x)𝑑x,g(x)=(2π)ng^(x).\widehat{g}(\xi)=\int_{{\mathbb{R}}^{n}}e^{-ix\cdot\xi}g(x)\;dx,\qquad g^{\vee}(x)=(2\pi)^{-n}\,\widehat{g}(-x).

Let g(1i)g(\tfrac{1}{i}\partial) be the operator from 𝒮(n)\mathcal{S}({\mathbb{R}}^{n}) to 𝒮(n)\mathcal{S}^{\prime}({\mathbb{R}}^{n}) such that g(1i)u=vg(\tfrac{1}{i}\partial)u=v if and only if g(ξ)u^(ξ)=v^(ξ)g(\xi)\widehat{u}(\xi)=\widehat{v}(\xi). The Schwartz kernel of g(1i)g(\tfrac{1}{i}\partial) is g(xy)g^{\vee}(x-y).

Then for any antisymmetric bilinear form AA of n{\mathbb{R}}^{n}, define g(1iA)g(\frac{1}{i}\nabla^{A}) as the operator with Schwartz kernel

Kg(x,y)=ei2A(x,y)g(xy).\displaystyle K_{g}(x,y)=e^{-\frac{i}{2}A(x,y)}g^{\vee}(x-y). (33)

Since g(xy)g^{\vee}(x-y) is a tempered distribution of xn×yn{\mathbb{R}}^{n}_{x}\times{\mathbb{R}}^{n}_{y}, the same holds for KgK_{g}, so g(1iA)g(\tfrac{1}{i}\nabla^{A}) is continuous from 𝒮(n)\mathcal{S}({\mathbb{R}}^{n}) to 𝒮(n)\mathcal{S}^{\prime}({\mathbb{R}}^{n}). We claim that this definition has an intrinsic meaning for A2EA\in\wedge^{2}E^{*} if we consider that g𝒮(E)g\in\mathcal{S}^{\prime}(E^{*}) and g(1iA)g(\frac{1}{i}\nabla^{A}) is an operator 𝒮(E)𝒮(E)\mathcal{S}(E)\rightarrow\mathcal{S}^{\prime}(E). One way to see this is to write for gg and uu in 𝒮(n)\mathcal{S}({\mathbb{R}}^{n})

(g(1iA)u)(x)=(2π)nn×nei2A(x,y)+iξ(xy)g(ξ)u(y)𝑑y𝑑ξ\displaystyle(g(\tfrac{1}{i}\nabla^{A})u)(x)=(2\pi)^{-n}\int_{{\mathbb{R}}^{n}\times{\mathbb{R}}^{n}}e^{-\frac{i}{2}A(x,y)+i\xi\cdot(x-y)}g(\xi)u(y)\;dy\,d\xi (34)

and to notice that the product ξ(xy)\xi\cdot(x-y) is well-defined for ξE\xi\in E^{*}, x,yEx,y\in E, and the measure dydξdy\,d\xi can be interpreted as the canonical volume form of E×Eyn×ξnE\times E^{*}\simeq{\mathbb{R}}^{n}_{y}\times{\mathbb{R}}^{n}_{\xi}.

Assume again that ExnE\simeq{\mathbb{R}}_{x}^{n}, EξnE^{*}\simeq{\mathbb{R}}_{\xi}^{n} and let

jA:=xjA=xj+12ikxkAkj\displaystyle\nabla_{j}^{A}:=\nabla_{\partial_{x_{j}}}^{A}=\partial_{x_{j}}+\tfrac{1}{2i}\textstyle{\sum_{k}}x_{k}A_{kj} (35)

where (Aij)(A_{ij}) is the matrix of AA, so Aij=A(ei,ej)A_{ij}=A(e_{i},e_{j}) with (ei)(e_{i}) the canonical basis of n{\mathbb{R}}^{n}.

Lemma 4.1.

For any f𝒮(E)f\in\mathcal{S}^{\prime}(E^{*}), we have 1ijAf(1iA)=(ξjAf)(1iA)\frac{1}{i}\nabla_{j}^{A}\circ f(\frac{1}{i}\nabla^{A})=(\xi_{j}\sharp_{A}f)(\frac{1}{i}\nabla^{A}) where

ξjAf=(ξj+i2kAjkξk)f.\displaystyle\xi_{j}\sharp_{A}f=(\xi_{j}+\tfrac{i}{2}\textstyle{\sum_{k}}A_{jk}\partial_{\xi_{k}})f. (36)
Proof.

Simply use the identity

1i(xj+12ikxkAkj)ei2A(x,y)+iξ(xy)=(ξj+12ikAjkξk)ei2A(x,y)+iξ(xy)\tfrac{1}{i}(\partial_{x_{j}}+\tfrac{1}{2i}\textstyle{\sum_{k}}x_{k}A_{kj})e^{-\frac{i}{2}A(x,y)+i\xi\cdot(x-y)}=(\xi_{j}+\frac{1}{2i}\sum_{k}A_{jk}\partial_{\xi_{k}})e^{-\frac{i}{2}A(x,y)+i\xi\cdot(x-y)}

in (34) and integrate by part with respect to the variables ξk\xi_{k}. ∎

The reason for the notation g(1iA)g(\tfrac{1}{i}\nabla^{A}) is that when gg is a monomial, g(1iA)g(\tfrac{1}{i}\nabla^{A}) is merely a symmetrization of covariant derivatives. The precise result is the following proposition which is not really needed in the sequel. Notice first that for g1g\equiv 1, g(1iA)=idg(\frac{1}{i}\nabla^{A})=\operatorname{id} as a direct consequence of the definition.

Proposition 4.2.

For any N1N\geqslant 1 and X1,XNEX_{1},\ldots X_{N}\in E, if g=i=1Nfig=\prod_{i=1}^{N}f_{i} with fi(ξ)=ξXif_{i}(\xi)=\xi\cdot X_{i}, then

g(1iA)=(i)NN!σ𝔖NXσ(1)AXσ(N)A,g(\tfrac{1}{i}\nabla^{A})=\frac{(-i)^{N}}{N!}\sum_{{\sigma}\in\mathfrak{S}_{N}}\nabla_{X_{{\sigma}(1)}}^{A}\ldots\nabla_{X_{{\sigma}(N)}}^{A},

where 𝔖N\mathfrak{S}_{N} is the group of permutations of 1,,N1,\ldots,N.

Proof.

For any XEX\in E we have by (36) with f(ξ)=ξXf(\xi)=\xi\cdot X that

1iXAg(1iA)=(fAg)(1iA)\displaystyle\tfrac{1}{i}\nabla^{A}_{X}\circ g(\tfrac{1}{i}\nabla^{A})=(f\sharp_{A}\,g)(\tfrac{1}{i}\nabla^{A}) (37)

where fAg=(f+i2A(X,ξ))gf\sharp_{A}\,g=(f+\tfrac{i}{2}A(X,\partial_{\xi}))g. Choosing g=1g=1, we obtain the result for N=1N=1. We now proceed by induction over NN and assume the result holds for N1N-1 with N2N\geqslant 2. Thus

(i)NN!σ𝔖NXσ(1)AXσ(N)A=1Nj=1N1iXjAgj(1iXA)\frac{(-i)^{N}}{N!}\sum_{{\sigma}\in\mathfrak{S}_{N}}\nabla_{X_{{\sigma}(1)}}^{A}\ldots\nabla_{X_{{\sigma}(N)}}^{A}=\frac{1}{N}\sum_{j=1}^{N}\tfrac{1}{i}\nabla^{A}_{X_{j}}\circ g_{j}(\tfrac{1}{i}\nabla^{A}_{X})

with gj=g/fjg_{j}=g/f_{j}. By (37), fAgj=fgj+i2jA(X,X)gjf\sharp_{A}g_{j}=fg_{j}+\frac{i}{2}\sum_{\ell\neq j}A(X,X_{\ell})g_{j\ell} where gj=g/(fjf)g_{j\ell}=g/(f_{j}f_{\ell}). So we have

(i)NN!σSNXσ(1)AXσ(N)A=g(1iA)+1NjA(Xj,X)gj(1iA)\frac{(-i)^{N}}{N!}\sum_{{\sigma}\in S_{N}}\nabla_{X_{{\sigma}(1)}}^{A}\ldots\nabla_{X_{{\sigma}(N)}}^{A}=g(\tfrac{1}{i}\nabla^{A})+\frac{1}{N}\sum_{j\neq\ell}A(X_{j},X_{\ell})g_{j\ell}(\tfrac{1}{i}\nabla^{A})

and the sum in the right-hand side is zero because AA is antisymmetric whereas gj=gjg_{j\ell}=g_{\ell j}. ∎

Let 𝒟is(A)\mathcal{D}_{\operatorname{is}}^{\infty}(A) be the filtered algebra generated by the covariant derivatives XA\nabla_{X}^{A} where XEX\in E. More explicitly, 𝒟is(A)=m𝒟ism(A)\mathcal{D}_{\operatorname{is}}^{\infty}(A)=\cup_{m\in{\mathbb{N}}}\mathcal{D}_{\operatorname{is}}^{m}(A) with

𝒟ism(A)=Span(X1AXA/ 0m,X1,XE).\mathcal{D}_{\operatorname{is}}^{m}(A)=\operatorname{Span}(\nabla^{A}_{X_{1}}\ldots\nabla^{A}_{X_{\ell}}/\;0\leqslant\ell\leqslant m,X_{1},\ldots X_{\ell}\in E).

Let m[E]{\mathbb{C}}_{\leqslant m}[E^{*}] be the space of complex polynomial functions of EE^{*} with degree less than or equal to mm. By Lemma 4.1 and Proposition 4.2,

𝒟ism(A)={f(1iA),fm[E]}.\displaystyle\mathcal{D}_{\operatorname{is}}^{m}(A)=\bigl{\{}f(\tfrac{1}{i}\nabla^{A}),\;f\in{\mathbb{C}}_{\leqslant m}[E^{*}]\bigr{\}}. (38)

By Lemma 4.1 again, the left composition by any element of 𝒟is(A)\mathcal{D}_{\operatorname{is}}^{\infty}(A) preserves {g(1i),g𝒮(E)}\{g(\frac{1}{i}\nabla),\;g\in\mathcal{S}^{\prime}(E^{*})\}. This defines the product

A:[E]×𝒮(E)𝒮(E),(fAg)(1i)=f(1i)g(1i)\displaystyle\sharp_{A}:{\mathbb{C}}[E^{*}]\times\mathcal{S}^{\prime}(E^{*})\rightarrow\mathcal{S}^{\prime}(E^{*}),\qquad(f\sharp_{A}g)(\tfrac{1}{i}\nabla)=f(\tfrac{1}{i}\nabla)\circ g(\tfrac{1}{i}\nabla) (39)

In the sequel we will use the basis (α,|α|m)(\nabla^{{\alpha}},|{\alpha}|\leqslant m) of 𝒟ism(A)\mathcal{D}_{\operatorname{is}}^{m}(A), defined by α:=(1A)α(1)(nA)α(n)\nabla^{{\alpha}}:=(\nabla_{1}^{A})^{{\alpha}(1)}\ldots(\nabla_{n}^{A})^{{\alpha}(n)}, αn{\alpha}\in{\mathbb{N}}^{n}. Clearly

i|α|α=f(1i) with f=ξα:=ξ1α(1)ξnα(n),\displaystyle i^{-|{\alpha}|}\nabla^{\alpha}=f(\tfrac{1}{i}\nabla)\qquad\text{ with }f=\xi^{\sharp{\alpha}}:=\xi_{1}^{\sharp{\alpha}(1)}\sharp\ldots\sharp\xi_{n}^{\sharp{\alpha}(n)}, (40)

where we have not written the AA dependence to lighten the notations. Furthermore, if |γ|=m|{\gamma}|=m,

ξγAf=ξγf+|α|+|β|m,|α|m1aα,β,γξαξβf\displaystyle\xi^{\sharp{\gamma}}\sharp_{A}f=\xi^{\gamma}f+\sum_{|{\alpha}|+|{\beta}|\leqslant m,\;|{\alpha}|\leqslant m-1}a_{{\alpha},{\beta},{\gamma}}\xi^{{\alpha}}\partial_{\xi}^{\beta}f (41)

where the coefficients aα,β,γa_{{\alpha},{\beta},{\gamma}}\in{\mathbb{C}} depends smoothly (even polynomially) on AA, which follows from Lemma 4.1 again. Actually there is a closed formula for A\sharp_{A}, cf. (42), but (41) is enough for our purpose.

Introduce the space Ψism(A):={f(1i),fSphm(E)}\Psi_{\operatorname{is}}^{m}(A):=\{f(\frac{1}{i}\nabla),\;f\in S^{m}_{\operatorname{ph}}(E^{*})\}. We have

𝒟ism(A)Ψism(A),𝒟ism(A)Ψisp(A)Ψism+p(A),\mathcal{D}_{\operatorname{is}}^{m}(A)\subset\Psi_{\operatorname{is}}^{m}(A),\qquad\mathcal{D}_{\operatorname{is}}^{m}(A)\circ\Psi_{\operatorname{is}}^{p}(A)\subset\Psi_{\operatorname{is}}^{m+p}(A),

the second assertion being a consequence of (41). This is all what we need to define in the next section the symbolic calculus corresponding to the composition of differential Heisenberg operators with Heisenberg pseudodifferential operators. In the case where A=0A=0, A\sharp_{A} is the usual pointwise product of functions. In Lemma 6.1, we will see that when AA is nondegenerate so that n=2dn=2d, Ψis(A)\Psi_{\operatorname{is}}^{\infty}(A) is an algebra isomorphic to the Weyl algebra of 2d{\mathbb{R}}^{2d}.

In the companion paper [5], we will prove that for any AA, Ψis(A)\Psi_{\operatorname{is}}^{\infty}(A) is a filtered algebra, that is Ψism(A)Ψisp(A)Ψism+p(A)\Psi_{\operatorname{is}}^{m}(A)\circ\Psi_{\operatorname{is}}^{p}(A)\subset\Psi_{\operatorname{is}}^{m+p}(A). Moreover

(fAg)(ξ)=[ei2A(ξ,η)f(ξ)g(η)]ξ=η\displaystyle(f\sharp_{A}\,g)(\xi)=\Bigl{[}e^{\frac{i}{2}A(\partial_{\xi},\partial_{\eta})}f(\xi)g(\eta)\Bigl{]}_{\xi=\eta} (42)

So Ψis(A)\Psi_{\operatorname{is}}^{\infty}(A) is isomorphic with the algebra called the AA-isotropic algebra in [12, Chapter 4, section 2].

Recall the standard and Weyl quantization maps which associate to any f𝒮(2n)f\in\mathcal{S}^{\prime}({\mathbb{R}}^{2n}) the operators f(x,1i)f(x,\frac{1}{i}\partial) and fw(x,1i)f^{w}(x,\frac{1}{i}\partial) with Schwartz kernels

(2π)neiξ(xy)f(x,ξ)𝑑ξ and (2π)neiξ(xy)f(12(x+y),ξ)𝑑ξ(2\pi)^{-n}\int e^{i\xi\cdot(x-y)}f(x,\xi)\;d\xi\quad\text{ and }\quad(2\pi)^{-n}\int e^{i\xi\cdot(x-y)}f(\tfrac{1}{2}(x+y),\xi)\;d\xi

respectively.

Lemma 4.3.

For any g𝒮(n)g\in\mathcal{S}^{\prime}({\mathbb{R}}^{n}), we have

g(1iA)=f(x,1i)=fw(x,1i)g(\tfrac{1}{i}\nabla^{A})=f(x,\tfrac{1}{i}\partial)=f^{w}(x,\tfrac{1}{i}\partial)

where f(x,ξ)=g(ξβ(x))f(x,\xi)=g(\xi-\beta(x)) and β(x){\beta}(x) is defined in (34), equivalently f(x,ξ)=g(ξ112A(x,e1),,ξn12A(x,en))f(x,\xi)=g(\xi_{1}-\frac{1}{2}A(x,e_{1}),\ldots,\xi_{n}-\frac{1}{2}A(x,e_{n})).

Proof.

By the same computation as in (24),

eiξ(xy)g(ξβ(x))𝑑ξ=eiβ(x)(xy)eiξ(xy)g(ξ)𝑑ξ.\int e^{i\xi\cdot(x-y)}g(\xi-\beta(x))\;d\xi=e^{i{\beta}(x)(x-y)}\int e^{i\xi\cdot(x-y)}g(\xi)\;d\xi.

AA being antisymmetric, β(x)(xy)=12A(x,y){\beta}(x)(x-y)=-\frac{1}{2}A(x,y), which proves that g(1iA)=f(x,1i)g(\tfrac{1}{i}\nabla^{A})=f(x,\tfrac{1}{i}\partial). The same proof by using this time that β(12(x+y))(xy)=12A(x,y){\beta}(\frac{1}{2}(x+y))(x-y)=-\frac{1}{2}A(x,y) shows that g(1iA)=fw(x,1i)g(\tfrac{1}{i}\nabla^{A})=f^{w}(x,\tfrac{1}{i}\partial). ∎

5 Heisenberg differential operators

The algebra 𝒟Heis(L,)\mathcal{D}_{\operatorname{Heis}}^{\infty}(L,\nabla) of Heisenberg differential operators consists of families of differential operators

P=(Pk:𝒞(M,Lk)𝒞(M,Lk),k),\displaystyle P=(P_{k}:{\mathcal{C}}^{\infty}(M,L^{k})\rightarrow{\mathcal{C}}^{\infty}(M,L^{k}),\;k\in{\mathbb{N}}), (43)

satisfying some conditions given below. It includes the multiplications by any ff in 𝒞(M){\mathcal{C}}^{\infty}(M), the normalised covariant derivatives k1/2Xk^{-1/2}\nabla_{X} where XX is any vector field of MM and the multiplication by k1/2k^{-1/2}. It is actually generated by these operators but it will be easier to use the following definition.

For any mm\in{\mathbb{N}}, 𝒟Heism(L,)\mathcal{D}_{\operatorname{Heis}}^{m}(L,\nabla) consists of the families PP of differential operators of the form (43) such that for any coordinate chart (U,xi)(U,x_{i}) and frame s𝒞(U,L)s\in{\mathcal{C}}^{\infty}(U,L) with |s|=1|s|=1, we have on UU,

Pk=,αn+|α|mk2f,απ~α\displaystyle P_{k}=\sum_{\begin{subarray}{c}\ell\in{\mathbb{N}},\;{\alpha}\in{\mathbb{N}}^{n}\\ \ell+|{\alpha}|\leqslant m\end{subarray}}k^{-\frac{\ell}{2}}f_{\ell,{\alpha}}\tilde{\pi}^{\alpha} (44)

where f,α𝒞(U)f_{\ell,{\alpha}}\in{\mathcal{C}}^{\infty}(U), π~α=π~1α(1)π~nα(n)\tilde{\pi}^{{\alpha}}=\tilde{\pi}_{1}^{{\alpha}(1)}\ldots\tilde{\pi}_{n}^{{\alpha}(n)} and

π~i=1iki=1ikikβi with s=1iβidxis\displaystyle\tilde{\pi}_{i}=\tfrac{1}{i\sqrt{k}}\nabla_{i}=\tfrac{1}{i\sqrt{k}}\partial_{i}-\sqrt{k}\beta_{i}\qquad\text{ with }\quad\nabla s=\tfrac{1}{i}\sum\beta_{i}dx_{i}\otimes s (45)

Set

𝒟Heis(L,)=m𝒟Heism(L,).\mathcal{D}_{\operatorname{Heis}}^{\infty}(L,\nabla)=\bigcup_{m\in{\mathbb{N}}}\mathcal{D}_{\operatorname{Heis}}^{m}(L,\nabla).

In the sequel to lighten the notations, we omit (L,)(L,\nabla) and write 𝒟Heism\mathcal{D}_{\operatorname{Heis}}^{m}, 𝒟Heis\mathcal{D}_{\operatorname{Heis}}^{\infty}. Since [π~i,π~j]=1i(iβjjβi)[\tilde{\pi}_{i},\tilde{\pi}_{j}]=\frac{1}{i}(\partial_{i}{\beta}_{j}-\partial_{j}\beta_{i}) and [π~i,f]=1ikif[\tilde{\pi}_{i},f]=\frac{1}{i\sqrt{k}}\partial_{i}f, we see that

𝒟Heism𝒟Heisp𝒟Heism+p.\mathcal{D}_{\operatorname{Heis}}^{m}\circ\mathcal{D}_{\operatorname{Heis}}^{p}\subset\mathcal{D}_{\operatorname{Heis}}^{m+p}.

Notice that 𝒟Heis\mathcal{D}_{\operatorname{Heis}}^{\infty} has two filtrations: one ascending 𝒟Heism𝒟Heism+1\mathcal{D}_{\operatorname{Heis}}^{m}\subset\mathcal{D}_{\operatorname{Heis}}^{m+1} and the other descending k/2𝒟Heisk^{-\ell/2}\mathcal{D}_{\operatorname{Heis}}, \ell\in{\mathbb{N}}. The generators ff, k1/2Xk^{-1/2}\nabla_{X} and k1/2k^{-1/2} have orders 0, 11, 11 for the former and 0, 0, 11 for the latter.

By the next proposition, 𝒟Heis\mathcal{D}_{\operatorname{Heis}}^{\infty} is contained in ΨHeis(L,)\Psi_{\operatorname{Heis}}^{\infty}(L,\nabla) and acts on it. Being Heisenberg pseudodifferential operators, the elements of 𝒟Heis\mathcal{D}_{\operatorname{Heis}}^{\infty} have a principal symbol, cf Definition 3.2. As we will see, the product of symbols is the fiberwise product \sharp of TMT^{*}M defined from ω{\omega}. Precisely, we denote by x\sharp_{x} the product

m[TxM]×Sp(TxM)Sm+p(TxM)\displaystyle{\mathbb{C}}_{\leqslant m}[T^{*}_{x}M]\times S^{p}(T^{*}_{x}M)\rightarrow S^{m+p}(T^{*}_{x}M) (46)

associated to ω(x)2TxM{\omega}(x)\in\wedge^{2}T^{*}_{x}M defined in (39). We will need as well the polynomials ξxα\xi^{\sharp_{x}{\alpha}} defined in (40).

Proposition 5.1.

  • -

    for any mm\in{\mathbb{N}}, 𝒟HeismΨHeism\mathcal{D}_{\operatorname{Heis}}^{m}\subset\Psi_{\operatorname{Heis}}^{m}

  • -

    the principal symbols of the operators of 𝒟Heism\mathcal{D}_{\operatorname{Heis}}^{m} are the functions f𝒞(TM)f\in{\mathcal{C}}^{\infty}(T^{*}M) such that f(x,)m[TxM]f(x,\cdot)\in{\mathbb{C}}_{\leqslant m}[T^{*}_{x}M] for any xx. If (44) holds on UU, then σ(P)(x,ξ)=|α|mf0,α(x)ξxα{\sigma}(P)(x,\xi)=\sum_{|{\alpha}|\leqslant m}f_{0,{\alpha}}(x)\xi^{\sharp_{x}{\alpha}}.

  • -

    for any P𝒟HeismP\in\mathcal{D}_{\operatorname{Heis}}^{m}, σ(P)=0{\sigma}(P)=0 if and only if Pk12𝒟Heism1P\in k^{-\frac{1}{2}}\mathcal{D}_{\operatorname{Heis}}^{m-1}.

  • -

    for any mm\in{\mathbb{N}} and pp\in{\mathbb{R}}, 𝒟HeismΨHeispΨHeism+p\mathcal{D}_{\operatorname{Heis}}^{m}\circ\Psi_{\operatorname{Heis}}^{p}\subset\Psi_{\operatorname{Heis}}^{m+p}. Furthermore

    σ(PQ)(x,)=σ(P)(x,)xσ(Q)(x,){\sigma}(P\circ Q)(x,\cdot)={\sigma}(P)(x,\cdot)\,\sharp_{x}\,{\sigma}(Q)(x,\cdot)

    for any P𝒟HeismP\in\mathcal{D}_{\operatorname{Heis}}^{m}, QΨHeispQ\in\Psi_{\operatorname{Heis}}^{p}.

Proof.

We start with the computation of π~iOpHeis(a)\tilde{\pi}_{i}\circ\operatorname{Op}_{\operatorname{Heis}}(a) where OpHeis(a)\operatorname{Op}_{\operatorname{Heis}}(a) is the operator with Schwartz kernel (24). Using (45), we get first that π~iOpHeis(a)=OpHeis(b)\tilde{\pi}_{i}\circ\operatorname{Op}_{\operatorname{Heis}}(a)=\operatorname{Op}_{\operatorname{Heis}}(b) with

b(h,x,y,ξ)=(h1ψi(x,y)+ξi+hixi)a(h,x,y,ξ)b(h,x,y,\xi)=\bigl{(}h^{-1}\psi_{i}(x,y)+\xi_{i}+\tfrac{h}{i}\partial_{x_{i}}\bigr{)}\,a(h,x,y,\xi)

where ψi(x,y)=(iβ)(12(x+y))(xy)+βi(12(x+y))βi(x)\psi_{i}(x,y)=(\partial_{i}{\beta})(\tfrac{1}{2}(x+y))\cdot(x-y)+\beta_{i}(\tfrac{1}{2}(x+y))-\beta_{i}(x). Taylor expanding along x=yx=y, we get

ψi(x,y)=12jωij(x)(xjyj)+i,jrij(x,y)(xiyi)(xjyj)\displaystyle\psi_{i}(x,y)=\tfrac{1}{2}\sum_{j}{\omega}_{ij}(x)(x_{j}-y_{j})+\sum_{i,j}r_{ij}(x,y)(x_{i}-y_{i})(x_{j}-y_{j})

with ωij=xiβjxjβi{\omega}_{ij}=\partial_{x_{i}}\beta_{j}-\partial_{x_{j}}\beta_{i}. Integrating by part, π~iOpHeis(a)=OpHeis(c)\tilde{\pi}_{i}\circ\operatorname{Op}_{\operatorname{Heis}}(a)=\operatorname{Op}_{\operatorname{Heis}}(c) with

c(h,x,y,ξ)=(ξi+i2jωij(x)ξi+hixihijrij(x,y)ξiξj)a(h,x,y,ξ).c(h,x,y,\xi)=(\xi_{i}+\tfrac{i}{2}\sum_{j}{\omega}_{ij}(x)\partial_{\xi_{i}}+\tfrac{h}{i}\partial_{x_{i}}-h\sum_{ij}r_{ij}(x,y)\partial_{\xi_{i}}\partial_{\xi_{j}})a(h,x,y,\xi).

Notice that if aSscm(U2,n)a\in S^{m}_{\operatorname{sc}}(U^{2},{\mathbb{R}}^{n}), then the same holds for cc. Furthermore, if a(h,x,x,ξ)=σ(x,ξ)+𝒪(h)a(h,x,x,\xi)={\sigma}(x,\xi)+\mathcal{O}(h), then c(h,x,x,ξ)=(ξiσ)(x,ξ)+𝒪(h)c(h,x,x,\xi)=(\xi_{i}\sharp{\sigma})(x,\xi)+\mathcal{O}(h).

We claim that everything can be deduced easily from these preliminary observations. Starting from the fact that OpHeis(1)\operatorname{Op}_{\operatorname{Heis}}(1) is the identity, we deduce by induction on |α||{\alpha}| that π~α=OpHeis(aα)\tilde{\pi}^{{\alpha}}=\operatorname{Op}_{\operatorname{Heis}}(a_{\alpha}) with aα(h,x,x,ξ)=ξα+𝒪(h)a_{{\alpha}}(h,x,x,\xi)=\xi^{\sharp{\alpha}}+\mathcal{O}(h). The first two assertions follow. The third assertion is a consequence of the fact that the ξα|x\xi^{\sharp{\alpha}}|_{x}, αn{\alpha}\in{\mathbb{N}}^{n} are linearly independent so that f0,αξα=0\sum f_{0,{\alpha}}\xi^{\sharp{\alpha}}=0 implies that f0,α=0f_{0,{\alpha}}=0. Last assertion follows again from the preliminaries by induction on mm. ∎

6 Resolvent

Let (F,λ)(F,{\lambda}) be a real symplectic vector space with dimension 2d2d. The Weyl product of the Schwartz space 𝒮(F)\mathcal{S}(F) is defined by

(aλb)(ξ)=(π)2de2iλ(η,ζ)a(ξ+η)b(ξ+ζ)𝑑μF(η)𝑑μF(ζ)(a\circ_{\lambda}b)(\xi)=(\pi)^{-2d}\int e^{-2i{\lambda}(\eta,\zeta)}a(\xi+\eta)b(\xi+\zeta)\;d\mu_{F}(\eta)\;d\mu_{F}(\zeta)

where μF=λd/d!\mu_{F}={\lambda}^{\wedge d}/d! is the Liouville measure of FF. For F=td×τdF={\mathbb{R}}^{d}_{t}\times{\mathbb{R}}^{d}_{\tau} with λ(t,τ;s,ς)=τsςt{\lambda}(t,\tau;s,{\varsigma})=\tau\cdot s-{\varsigma}\cdot t, it is the composition law of the Weyl symbols of pseudodifferential operators of d{\mathbb{R}}^{d}, cf. for instance [16, page 152].

This product extends continuously from Sm(F)×Sp(F)S^{m}(F)\times S^{p}(F) to Sm+p(F)S^{m+p}(F) by preserving the subspaces of polyhomogeneous symbols. So the corresponding pseudodifferential operators, fw(x,1i)f^{w}(x,\frac{1}{i}\partial), with fS(2d)f\in S^{\infty}({\mathbb{R}}^{2d}), form an algebra, called sometimes the Shubin class or isotropic class. This algebra is one of the most studied in microlocal analysis, cf. [24, Chapter IV], [15], [21, Chapter 4], [12, Chapter 4], [25, Appendix A] for lecture note references.

The Weyl product appears naturally in our context as the product of the operators f(1iA)f(\frac{1}{i}\nabla^{A}) defined in Section 4 when AA is nondegenerate.

Lemma 6.1.

If A2EA\in\wedge^{2}E^{*} is nondegenerate, then for any ff, gg in S(E)S^{\infty}(E^{*}),

f(1iA)g(1iA)=(fλg)(1iA)f(\tfrac{1}{i}\nabla^{A})\circ g(\tfrac{1}{i}\nabla^{A})=(f\circ_{{\lambda}}g)(\tfrac{1}{i}\nabla^{A})

where λ{\lambda} is the symplectic form of EE^{*} dual to AA.

Proof.

Introduce a symplectic basis (ei,fi)(e_{i},f_{i}) of (E,A)(E,A) and denote by (xi,yi)(x_{i},y_{i}) the associated linear coordinates, so that E=xd×ydE={\mathbb{R}}^{d}_{x}\times{\mathbb{R}}^{d}_{y}. Then the operators

1iei=1ixi+12yi,1ifi=1iyi12xi,1iyi+12xi,1ixi12yi\tfrac{1}{i}\nabla_{e_{i}}=\tfrac{1}{i}\partial_{x_{i}}+\tfrac{1}{2}y_{i},\quad\tfrac{1}{i}\nabla_{f_{i}}=\tfrac{1}{i}\partial_{y_{i}}-\tfrac{1}{2}x_{i},\quad\ \tfrac{1}{i}\partial_{y_{i}}+\tfrac{1}{2}x_{i},\quad\tfrac{1}{i}\partial_{x_{i}}-\tfrac{1}{2}y_{i}

satisfy the same commutation relations as the operators sis_{i}, 1isi\frac{1}{i}\partial_{s_{i}}, tit_{i}, 1iti\frac{1}{i}\partial_{t_{i}} of sd×td{\mathbb{R}}^{d}_{s}\times{\mathbb{R}}^{d}_{t}. So the linear isomorphism Φ:4d4d\Phi:{\mathbb{R}}^{4d}\rightarrow{\mathbb{R}}^{4d},

Φ(x,ξ,y,η)=(ξ+12y,η12x,η+12x,ξ+12y).\Phi(x,\xi,y,\eta)=(\xi+\tfrac{1}{2}y,\eta-\tfrac{1}{2}x,\eta+\tfrac{1}{2}x,\xi+\tfrac{1}{2}y).

is a symplectomorphism. Its metaplectic representation U:L2(E)L2(2d)U:L^{2}(E)\rightarrow L^{2}({\mathbb{R}}^{2d}) satisfies fw=U(fΦ)wUf^{w}=U(f\circ\Phi)^{w}U^{*} for any f𝒮(4d)f\in\mathcal{S}^{\prime}({\mathbb{R}}^{4d}), cf. [16, Theorem 18.5.9]. Applying this to f(x,ξ,y,η)=g(ξ+12y,η12x)=((g1)Φ)(x,ξ,y,η)f(x,\xi,y,\eta)=g(\xi+\frac{1}{2}y,\eta-\frac{1}{2}x)=((g\boxtimes 1)\circ\Phi)(x,\xi,y,\eta), we obtain

fw(x,1ix,y,1iy)=U(gw(s,1is)idtd)Uf^{w}(x,\tfrac{1}{i}\partial_{x},y,\tfrac{1}{i}\partial_{y})=U^{*}(g^{w}(s,\tfrac{1}{i}\partial_{s})\otimes\operatorname{id}_{{\mathbb{R}}^{d}_{t}})U

and by Lemma 4.3, fw(x,1ix,y,1iy)=g(1iA)f^{w}(x,\tfrac{1}{i}\partial_{x},y,\tfrac{1}{i}\partial_{y})=g(\tfrac{1}{i}\nabla^{A}). The result follows. ∎

From now on, we assume that ω{\omega} is nondegenerate. By Lemma 6.1, at any xMx\in M, the product x\sharp_{x} defined in (46) extends continuously

Sm(TxM)×Sp(TxM)Sm+p(TxM).S^{m}(T_{x}^{*}M)\times S^{p}(T_{x}^{*}M)\rightarrow S^{m+p}(T_{x}^{*}M).

Recall that fSm(M,TM)f\in S^{m}(M,T^{*}M) is elliptic if |f(x,ξ)|C1|ξ|m|f(x,\xi)|\geqslant C^{-1}|\xi|^{m} when |ξ|C|\xi|\geqslant C for some positive CC. We say that ff is invertible if at any xMx\in M, f(x,)f(x,\cdot) is invertible in (S(TxM),x)(S^{\infty}(T_{x}^{*}M),\sharp_{x}).

Lemma 6.2.
  1. 1.

    Sph(M,TM)S_{\operatorname{ph}}^{\infty}(M,T^{*}M) endowed with the fibered product (fg)(x)=f(x)xg(x)(f\sharp g)(x)=f(x)\sharp_{x}g(x) is a filtered algebra.

  2. 2.

    For any fSphm(M,TM)f\in S^{m}_{\operatorname{ph}}(M,T^{*}M) which is both elliptic and invertible, the pointwise inverse of ff belongs to Sphm(M,TM)S^{-m}_{\operatorname{ph}}(M,T^{*}M).

Proof.

This holds more generally for Sphm(N,E)S^{m}_{\operatorname{ph}}(N,E) where EE is any symplectic vector bundle with base NN. When NN is a point, Sph(N,E)S^{\infty}_{\operatorname{ph}}(N,E) is isomorphic with the Weyl algebra Sph(2d)S^{\infty}_{\operatorname{ph}}({\mathbb{R}}^{2d}), and the result is well-known as we already mentioned it. In general, we can assume that EE is the trivial symplectic bundle 2d{\mathbb{R}}^{2d} over an open subset UU of d{\mathbb{R}}^{d}. Viewing symbols in Sm(U,2d)S^{m}(U,{\mathbb{R}}^{2d}) as smooth maps from UU to the Fréchet space Sm(2d)S^{m}({\mathbb{R}}^{2d}), and using that the Weyl product is continuous Sm(2d)×Sp(2d)Sm+p(2d)S^{m}({\mathbb{R}}^{2d})\times S^{p}({\mathbb{R}}^{2d})\rightarrow S^{m+p}({\mathbb{R}}^{2d}), we deduce with a little work that the fibered Weyl product \sharp is continuous

Sm(U,2d)×Sp(U,2d)Sm+p(U,2d).\displaystyle S^{m}(U,{\mathbb{R}}^{2d})\times S^{p}(U,{\mathbb{R}}^{2d})\rightarrow S^{m+p}(U,{\mathbb{R}}^{2d}). (47)

which proves the first assertion.

Let fSphm(U,2d)f\in S_{\operatorname{ph}}^{m}(U,{\mathbb{R}}^{2d}) be elliptic and invertible. Let us prove that its pointwise inverse gg is in Sphm(M,TM)S^{-m}_{\operatorname{ph}}(M,T^{*}M). Multiplying ff by f(x0)1f(x_{0})^{-1}, we may assume that m=0m=0. Since Sph(U,2d)S^{\infty}_{\operatorname{ph}}(U,{\mathbb{R}}^{2d}) is a filtered algebra, cf. (47), and by Borel lemma, ff has a parametrix hSph0(U,2d)h\in S^{0}_{\operatorname{ph}}(U,{\mathbb{R}}^{2d}). Let us prove that g=h+S(U,2d)g=h+S^{-\infty}(U,{\mathbb{R}}^{2d}). We have hf=1+rh\sharp f=1+r, fh=1+sf\sharp h=1+s with rr and ss in S(U,2d)S^{-\infty}(U,{\mathbb{R}}^{2d}). So g=hrh+rgsg=h-r\sharp h+r\sharp g\sharp s. By (47) again, rhS(U,2d)r\sharp h\in S^{-\infty}(U,{\mathbb{R}}^{2d}). It remains to prove that rgsS(U,2d)r\sharp g\sharp s\in S^{-\infty}(U,{\mathbb{R}}^{2d}).

By Calderon-Vaillancourt theorem, the Weyl quantization Op:S0(2d)(L2(d))\operatorname{Op}:S^{0}({\mathbb{R}}^{2d})\rightarrow\mathcal{L}(L^{2}({\mathbb{R}}^{d})) is continuous. Op(g(x))\operatorname{Op}(g(x)) being the inverse of Op(f(x))\operatorname{Op}(f(x)) for any xx, Op(g)𝒞(U,(L2(d))\operatorname{Op}(g)\in{\mathcal{C}}^{\infty}(U,\mathcal{L}(L^{2}({\mathbb{R}}^{d})). Moreover, the multilinear map

M:S(2d)×(L2(d))×S(2d)S(2d),M:S^{-\infty}({\mathbb{R}}^{2d})\times\mathcal{L}(L^{2}({\mathbb{R}}^{d}))\times S^{-\infty}({\mathbb{R}}^{2d})\rightarrow S^{-\infty}({\mathbb{R}}^{2d}),

defined by Op(M(σ,A,τ))=Op(σ)AOp(τ)\operatorname{Op}(M({\sigma},A,\tau))=\operatorname{Op}({\sigma})\circ A\circ\operatorname{Op}(\tau), being continuous, we obtain with a little work that rgs=M(r,Op(g),s)r\sharp g\sharp s=M(r,\operatorname{Op}(g),s) belongs to S(U,2d)S^{-\infty}(U,{\mathbb{R}}^{2d}). ∎

Consider now P𝒟Heism(L,)P\in\mathcal{D}_{\operatorname{Heis}}^{m}(L,\nabla) having an elliptic symbol. Then for any fixed kk, PkP_{k} is an elliptic differential operator of 𝒞(M,Lk){\mathcal{C}}^{\infty}(M,L^{k}), so for any ss\in{\mathbb{R}}, PkP_{k} extends to a Fredholm operator of (Hs(M,Lk),Hsm(M,Lk))\mathcal{L}(H^{s}(M,L^{k}),H^{s-m}(M,L^{k})). If we assume that the symbol of PP is invertible, then by the following Theorem, PkP_{k} is invertible when kk is large, and its inverse is a Heisenberg pseudodifferential operator.

Theorem 6.3.

Assume that ω{\omega} is nondegenerate. Let P𝒟Heism(L,)P\in\mathcal{D}_{\operatorname{Heis}}^{m}(L,\nabla) having an elliptic and invertible symbol σ𝒞(TM){\sigma}\in{\mathcal{C}}^{\infty}(T^{*}M). Then there exists QΨHeism(L,)Q\in\Psi_{\operatorname{Heis}}^{-m}(L,\nabla) such that

  1. -

    PQidPQ-\operatorname{id} and QPidQP-\operatorname{id} are in kΨ(L)k^{-\infty}\Psi^{-\infty}(L)

  2. -

    when kk is sufficiently large, QkPk=PkQk=idQ_{k}P_{k}=P_{k}Q_{k}=\operatorname{id}

  3. -

    the symbol of QQ is the inverse of σ{\sigma} for the product \sharp.

Proof.

This follows merely from the previous results, by the standard techniques for elliptic operators. First, using Lemma 6.2, we construct a parametrix QΨHeism(L,)Q\in\Psi_{\operatorname{Heis}}^{-m}(L,\nabla) of PP, so PQ=id+RPQ=\operatorname{id}+R and QP=id+SQP=\operatorname{id}+S with R,SR,S in the residual algebra kΨ(L)k^{-\infty}\Psi^{-\infty}(L). Then, by the Sobolev continuity (19), RkR_{k} and SkS_{k} belongs to (L2(M,Lk))\mathcal{L}(L^{2}(M,L^{k})) and their operator norms are in 𝒪(k)\mathcal{O}(k^{-\infty}). So when kk0k\geqslant k_{0}, PkP_{k} is invertible from Hm(M,Lk)H^{m}(M,L^{k}) to H0(M,Lk)H^{0}(M,L^{k}), which implies by the Fredholm properties of elliptic operators [24, Theorem 8.1], that PkP_{k} is an invertible operator of the distribution space 𝒟(M,Lk)\mathcal{D}^{\prime}(M,L^{k}).

Its inverse satisfies

Pk1=QkRkQk+Rk(QkPk)1QkSk\displaystyle P^{-1}_{k}=Q_{k}-R_{k}Q_{k}+R_{k}(Q_{k}P_{k})^{-1}Q_{k}S_{k} (48)

By Lemma 3.6, (RkQk)(R_{k}Q_{k}) and (QkSk)(Q_{k}S_{k}) are in kΨ(L)k^{-\infty}\Psi^{-\infty}(L). It is a classical fact that if (Ak)(A_{k}), (Bk)(B_{k}) are in kΨ(L)k^{-\infty}\Psi^{-\infty}(L) and Ck=𝒪(1):L2(M,Lk)L2(M,Lk)C_{k}=\mathcal{O}(1):L^{2}(M,L^{k})\rightarrow L^{2}(M,L^{k}), then (AkCkBk)(A_{k}C_{k}B_{k}) is in kΨ(L)k^{-\infty}\Psi^{-\infty}(L). So the last term in (48) belongs to kΨ(L)k^{-\infty}\Psi^{-\infty}(L). So by adding to QkQ_{k} an element of the residual algebra, we have that Qk=Pk1Q_{k}=P_{k}^{-1} when kk is large. ∎

Assume m>0m>0 and consider P𝒟Heism(L,)P\in\mathcal{D}_{\operatorname{Heis}}^{m}(L,\nabla) having an elliptic symbol σ{\sigma} such that for some z0z_{0}\in{\mathbb{C}}, σz0{\sigma}-z_{0} is invertible. Then by Theorem 6.3, when kk is sufficiently large, Pkz0P_{k}-z_{0} has an inverse, which is continuous L2(M,Lk)Hm(M,Lk)L^{2}(M,L^{k})\rightarrow H^{m}(M,L^{k}). So the restriction of PkP_{k} to Hm(M,Lk)H^{m}(M,L^{k}) is a closed unbounded operator of L2(M,Lk)L^{2}(M,L^{k}) having a compact resolvent. So its spectrum is a discrete subset of {\mathbb{C}} and it consists only of eigenvalues with finite multiplicity, the generalised eigenvectors being smooth [24, Theorem 8.4].

To state the next theorem, we need some spectral properties of the symbols themselves. Later we will explain these properties in terms of Weyl quantization, but since this quantization is only auxiliary in what we do, we prefer first to discuss everything intrinsically in terms of the algebra (S(F),λ)(S^{\infty}(F),\circ_{{\lambda}}) where (F,λ)(F,{\lambda}) is a symplectic vector space as above.

The spectrum of aS(F)a\in S^{\infty}(F) is defined by: zsp(a)z\notin\operatorname{sp}(a) if and only if zaz-a is invertible in (S(F),σ)(S^{\infty}(F),\circ_{\sigma}). A family (b(z),zΩ)(b(z),z\in\Omega) of Sm(F)S^{m}(F) is holomorphic if Ω\Omega is an open set of {\mathbb{C}}, bSm(Ω,F)b\in S^{m}(\Omega,F) and z¯b=0\partial_{\overline{z}}b=0. By the analytic Fredholm theory for the isotropic algebra exposed in [21, Chapter 3], for any elliptic aSphm(F)a\in S^{m}_{\operatorname{ph}}(F) with m>0m>0, the spectrum of aa is {\mathbb{C}} or a discrete subset of {\mathbb{C}}. In the latter case, the resolvent ((az)(1)λ,zsp(a))((a-z)^{(-1)_{\circ_{\lambda}}},\;z\in{\mathbb{C}}\setminus\operatorname{sp}(a)) is a holomorphic family of Sm(F)S^{-m}(F) and for any z0sp(a)z_{0}\in\operatorname{sp}(a), we have on a neighborhood of z0z_{0} for some NN\in{\mathbb{N}}

(az)(1)λ=h(z)+r1zz0++rN(zz0)N\displaystyle(a-z)^{(-1)_{\circ_{\lambda}}}=h(z)+\frac{r_{1}}{z-z_{0}}+\ldots+\frac{r_{N}}{(z-z_{0})^{N}} (49)

where (h(z))(h(z)) is a holomorphic family of Sphm(F)S_{\operatorname{ph}}^{-m}(F) and r1r_{1}, …, rNr_{N} are in S(F)S^{-\infty}(F).

Theorem 6.4.

Assume that ω{\omega} is nondegenerate. Let P𝒟Heism(L,)P\in\mathcal{D}_{\operatorname{Heis}}^{m}(L,\nabla) be elliptic with m1m\geqslant 1 and symbol σ{\sigma}. Let Σ\Sigma be the closed set xMsp(σ(x))\bigcup_{x\in M}\operatorname{sp}({\sigma}(x)). Then

  1. 1.

    if KK is a compact subset of {\mathbb{C}} disjoint from Σ\Sigma, then the spectrum of PkP_{k} does not intersect KK when kk is large enough.

  2. 2.

    if Ω\Omega is an open bounded subset of {\mathbb{C}} with a smooth boundary disjoint from Σ{\Sigma}, then there exists ΠΨHeism(L,)\Pi\in\Psi_{\operatorname{Heis}}^{-m}(L,\nabla) such that Πk=(1Ω(Pk))\Pi_{k}=(1_{\Omega}(P_{k})) when kk is large. Furthermore the principal symbol of Π\Pi is at xx

    π(x)=(2iπ)1Ω(σ(x)z)(1)x𝑑z.\displaystyle\pi(x)=(2i\pi)^{-1}\int_{\partial\Omega}({\sigma}(x)-z)^{(-1)_{\sharp_{x}}}dz. (50)
  3. 3.

    if for any kk, PkP_{k} is formally self-adjoint for some volume element of MM, then for any E,E+ΣE_{-},E_{+}\in{\mathbb{R}}\setminus{\Sigma} with E<E+E_{-}<E_{+}, (1[E,E+](Pk))(1_{[E_{-},E_{+}]}(P_{k})) belongs to ΨHeis(L,)\Psi_{\operatorname{Heis}}^{-\infty}(L,\nabla).

Observe that the symbol π(x)\pi(x) is the sum of the residues of the poles in Ω\Omega of the resolvent of σ(x){\sigma}(x). As we will see in the proof, the third assertion is a particular case of the second one, the symbol being the sum of the residues of the poles in [E,E+][E_{-},E_{+}].

In [5], we will prove that ΨHeismΨHeispΨHeism+p\Psi_{\operatorname{Heis}}^{m}\circ\Psi_{\operatorname{Heis}}^{p}\subset\Psi_{\operatorname{Heis}}^{m+p}. So in the second assertion, Π\Pi being idempotent, it belongs to ΨHeis(L,)\Psi_{\operatorname{Heis}}^{-\infty}(L,\nabla).

Proof.

First, Σ{\Sigma} is closed because the Weyl quantization is continuous from Sm(d)S^{m}({\mathbb{R}}^{d}) to (Hisom(d),Hiso0(d))\mathcal{L}(H^{m}_{\operatorname{iso}}({\mathbb{R}}^{d}),H^{0}_{\operatorname{iso}}({\mathbb{R}}^{d})), so that the characterization of the spectrum given below implies that if z0sp(σ(x0))z_{0}\notin\operatorname{sp}({\sigma}(x_{0})) then zsp(σ(x))z\notin\operatorname{sp}({\sigma}(x)) when (z,x)(z,x) is sufficiently close to (z0,x0)(z_{0},x_{0}).

Assume that KK is a compact subset of {\mathbb{C}} disjoint from Σ{\Sigma}. When zKz\in K, (Pkz)(P_{k}-z) satisfies the assumptions of Theorem 6.3, so there exists Q(z)ΨHeism(L,)Q(z)\in\Psi_{\operatorname{Heis}}^{-m}(L,\nabla) such that Qk(z)=(zPk)1Q_{k}(z)=(z-P_{k})^{-1} when kk0(z)k\geqslant k_{0}(z). This proves at least that PkP_{k} has a compact resolvent as explained above when kk is large. Moreover we claim that everything in the proof of Theorem 6.3 can be done continuously with respect to zKz\in K (even holomorphically with respect to zz in a neighborhood of KK). More precisely, the Schwartz kernel of Qk(z)Q_{k}(z) is locally of the form (25) where the dependence in zz in only in the symbol bb, which is continuous in zz. This proves first that we can choose k0(z)k_{0}(z) independent of zz, which shows the first assertion. Second, if Ω\Omega satisfies the assumptions of the second assertion of the Theorem, we can apply the previous consideration to K=ΩK=\partial\Omega and it follows that Πk:=(2iπ)1ΩQk(z)𝑑z\Pi_{k}:=(2i\pi)^{-1}\int_{\partial\Omega}Q_{k}(z)\;dz belongs to ΨHeism(L,)\Psi_{\operatorname{Heis}}^{-m}(L,\nabla) with a symbol given by (50). When kk is large enough, Qk(z)Q_{k}(z) is the resolvent, so by Cauchy formula, Πk=1Ω(Pk)\Pi_{k}=1_{\Omega}(P_{k}). This concludes the proof of the second assertion.

For the last assertion, by assumption, for any fixed kk, PkP_{k} is a formally self-adjoint elliptic differential operator on a compact manifold, so its spectrum is a discrete subset of {\mathbb{R}} and 1[E,E+](Pk)1_{[E_{-},E_{+}]}(P_{k}) is a finite rank projector onto a subspace of 𝒞(M,Lk){\mathcal{C}}^{\infty}(M,L^{k}), [24, Theorem 8.3]. Moreover σ{\sigma} is real valued so Σ{\Sigma}\subset{\mathbb{R}}. So there exists Ω\Omega satisfying the previous assumptions and such that Ω=[E,E+]\Omega\cap{\mathbb{R}}=[E_{-},E_{+}]. So Π=(1[E,E+](Pk),k)\Pi=(1_{[E_{-},E_{+}]}(P_{k}),\;k\in{\mathbb{N}}) belongs to ΨHeism(L,)\Psi_{\operatorname{Heis}}^{-m}(L,\nabla).

For any odd NN\in{\mathbb{N}}, Π=1[EN,E+N](PN)\Pi=1_{[E^{N}_{-},E^{N}_{+}]}(P^{N}) and PN𝒟HeismN(L,)P^{N}\in\mathcal{D}_{\operatorname{Heis}}^{mN}(L,\nabla), which implies by the previous argument that Π\Pi belongs to ΨHeisNm(L,)\Psi_{\operatorname{Heis}}^{-Nm}(L,\nabla), so ΠΨHeis(L,)\Pi\in\Psi_{\operatorname{Heis}}^{-\infty}(L,\nabla). ∎

Let us discuss briefly the convertibility and resolvent of elliptic elements of (S(F),λ)(S^{\infty}(F),\circ_{\lambda}) from the point of view of Weyl quantization. Let Ψiso(d)\Psi_{\operatorname{iso}}^{\infty}({\mathbb{R}}^{d}) be the space of pseudodifferential operators of d{\mathbb{R}}^{d} with a symbol in Sph(2d)S^{\infty}_{\operatorname{ph}}({\mathbb{R}}^{2d}). Any AΨisom(d)A\in\Psi_{\operatorname{iso}}^{m}({\mathbb{R}}^{d}) acts continuously 𝒮(d)𝒮(d)\mathcal{S}({\mathbb{R}}^{d})\rightarrow\mathcal{S}({\mathbb{R}}^{d}), 𝒮(d)𝒮(d)\mathcal{S}^{\prime}({\mathbb{R}}^{d})\rightarrow\mathcal{S}^{\prime}({\mathbb{R}}^{d}) and Hisos(d)Hisosm(d)H^{s}_{\operatorname{iso}}({\mathbb{R}}^{d})\rightarrow H^{s-m}_{\operatorname{iso}}({\mathbb{R}}^{d}), where Hisos(d)H^{s}_{\operatorname{iso}}({\mathbb{R}}^{d}) is the isotropic Sobolev spaces

Hisos(d)={u𝒮(d),AuL2(d),AΨisos(d)},s.H^{s}_{\operatorname{iso}}({\mathbb{R}}^{d})=\{u\in\mathcal{S}^{\prime}({\mathbb{R}}^{d}),\;Au\in L^{2}({\mathbb{R}}^{d}),\;\forall A\in\Psi_{\operatorname{iso}}^{s}({\mathbb{R}}^{d})\},\quad s\in{\mathbb{R}}.

When AA is elliptic, the following Fredholm property holds: kerA\ker A and kerA\ker A^{*} are finite dimensional subspaces of 𝒮(d)\mathcal{S}({\mathbb{R}}^{d}),

𝒮(d)=A(𝒮(d))kerA=A(𝒮(d))kerA\mathcal{S}^{\prime}({\mathbb{R}}^{d})=A(\mathcal{S}^{\prime}({\mathbb{R}}^{d}))\oplus\ker A^{*}=A^{*}(\mathcal{S}^{\prime}({\mathbb{R}}^{d}))\oplus\ker A

and the generalised inverse B:𝒮(d)𝒮(d)B:\mathcal{S}^{\prime}({\mathbb{R}}^{d})\rightarrow\mathcal{S}^{\prime}({\mathbb{R}}^{d}) such that BAidBA-\operatorname{id} and ABidAB-\operatorname{id} are the orthogonal projectors onto kerA\ker A and kerA\ker A^{*} respectively, belongs to Ψisom(d)\Psi_{\operatorname{iso}}^{-m}({\mathbb{R}}^{d}). So AA is invertible in the algebra Ψiso(d)\Psi_{\operatorname{iso}}^{\infty}({\mathbb{R}}^{d}) if and only if kerA=kerA=0\ker A=\ker A^{*}=0 if and only if AA is invertible as an operator in 𝒮\mathcal{S}^{\prime} if and only if AA is invertible in (Hisos(d),Hisosm(d))\mathcal{L}(H^{s}_{\operatorname{iso}}({\mathbb{R}}^{d}),H^{s-m}_{\operatorname{iso}}({\mathbb{R}}^{d})).

When m>0m>0, any elliptic AΨisom(d)A\in\Psi_{\operatorname{iso}}^{m}({\mathbb{R}}^{d}) defines by restriction a closed unbounded operator of L2(d)L^{2}({\mathbb{R}}^{d}) with domain Hisom(d)H^{m}_{\operatorname{iso}}({\mathbb{R}}^{d}). By the previous characterization of invertibility, the spectrum of AA is the same as the spectrum of its symbol aa defined above. Assume it is not empty, then AA has a compact resolvent, and as it was already explained, sp(A)\operatorname{sp}(A) is a discrete subset of {\mathbb{C}} and the resolvent (Az)1(A-z)^{-1} is a holomorphic family of Ψisom(d)\Psi_{\operatorname{iso}}^{-m}({\mathbb{R}}^{d}). Furthermore, for A=aw(x,1ix)A=a^{w}(x,\frac{1}{i}\partial_{x}), the residues rw(x,1ix)r_{\ell}^{w}(x,\frac{1}{i}\partial_{x}) defined in (49) have finite rank and r1w(x,1ix)r_{1}^{w}(x,\frac{1}{i}\partial_{x}) is a projector onto the space of generalised eigenvectors of AA for the eigenvalue z0z_{0}, which is a subspace of 𝒮(d)\mathcal{S}({\mathbb{R}}^{d}).

7 Auxiliary bundles

Let us first define symbols taking values in an auxiliary bundle. Recall the spaces Sm(N,E)S^{m}_{*}(N,E) introduced in Section 2 for a real vector bundle p:ENp:E\rightarrow N and =*=\emptyset, ph\operatorname{ph}, sc\operatorname{sc}. Let BB be a complex vector bundle over NN. By definition Sm(N,E;B)S^{m}(N,E;B) is the space of sections s𝒞(E,pB)s\in{\mathcal{C}}^{\infty}(E,p^{*}B) such that for any frame (uα)(u_{{\alpha}}) of BB over an open set UU of NN, we have over p1(U)p^{-1}(U),

s(x,ξ)=fα(x,ξ)uα(x),xN,ξExs(x,\xi)=\sum f_{{\alpha}}(x,\xi)u_{{\alpha}}(x),\qquad x\in N,\;\xi\in E_{x}

with coefficients fαf_{{\alpha}} in Sm(U,E)S^{m}(U,E). Since Sm(U,E)S^{m}(U,E) is a 𝒞(U){\mathcal{C}}^{\infty}(U)-submodule of 𝒞(U,E){\mathcal{C}}^{\infty}(U,E), this definition is compatible with the frame changes. Similarly, we define Sm(M,E;B)S^{m}_{*}(M,E;B) for =ph*=\operatorname{ph} or sc\operatorname{sc} by requiring that the coefficients fαf_{{\alpha}} belong to Sm(U,E)S^{m}_{*}(U,E). More precisely, in the case of semiclassical symbols where the section ss and its local coefficients depend on hh, we only choose frames (uα)(u_{{\alpha}}) independent of hh.

Let A1A_{1} and A2A_{2} be two complex vector bundles over MM and let us define the pseudodifferential operator spaces Ψscm(M;A1,A2)\Psi^{m}_{\operatorname{sc}}(M;A_{1},A_{2}), Ψtscm(L;A1,A2)\Psi_{\operatorname{tsc}}^{m}(L;A_{1},A_{2}) and ΨHeism(L,;A1,A2)\Psi_{\operatorname{Heis}}^{m}(L,\nabla;A_{1},A_{2}). For A1A_{1}, A2A_{2} being both the trivial line bundle, these are the spaces we introduced previously. In general, set B=A2A1B=A_{2}\boxtimes A_{1}^{*}. Then

  • Ψscm(M;A1,A2)\Psi^{m}_{\operatorname{sc}}(M;A_{1},A_{2}) consists of the families (Ph:𝒞(M,A1)𝒞(M,A2)(P_{h}:{\mathcal{C}}^{\infty}(M,A_{1})\rightarrow{\mathcal{C}}^{\infty}(M,A_{2}), h(0,1])h\in(0,1]) satisfying the same conditions as before except that the amplitude aa appearing in (15) belongs to Sscm(U2,n;B)S_{\operatorname{sc}}^{m}(U^{2},{\mathbb{R}}^{n};B).

  • Ψtscm(L;A1,A2)\Psi_{\operatorname{tsc}}^{m}(L;A_{1},A_{2}) consists of the families

    P=(Pk:𝒞(M,LkA1)𝒞(M,LkA2),k)\displaystyle P=(P_{k}:{\mathcal{C}}^{\infty}(M,L^{k}\otimes A_{1})\rightarrow{\mathcal{C}}^{\infty}(M,L^{k}\otimes A_{2}),\;k\in{\mathbb{N}}) (51)

    satisfying the conditions of Definition 2.1 with aSscm(U2,n;B)a\in S_{\operatorname{sc}}^{m}(U^{2},{\mathbb{R}}^{n};B)

  • ΨHeism(L;A1,A2)\Psi_{\operatorname{Heis}}^{m}(L;A_{1},A_{2}) consists of the families PP of the form (51) satisfying the conditions of Definition 3.2 with QhQ_{h} an operator of Sscm(M;A1,A2)S^{m}_{\operatorname{sc}}(M;A_{1},A_{2})

The symbol of PP is defined as before. Since the restriction of BB to the diagonal is isomorphic with Hom(A1,A2)\operatorname{Hom}(A_{1},A_{2}), in the three cases, the symbol identifies with an element of Sphm(M,TM;Hom(A1,A2))S^{m}_{\operatorname{ph}}(M,T^{*}M;\operatorname{Hom}(A_{1},A_{2})).

The space 𝒟Heism(L,;A1,A2)\mathcal{D}_{\operatorname{Heis}}^{m}(L,\nabla;A_{1},A_{2}) of Heisenberg differential operators consists of the families (51) of differential operators such that for any coordinate chart (U,xi)(U,x_{i}) of MM, we have on UU

Pk=,αn,+|α|mk2f,απ~α\displaystyle P_{k}=\sum_{\ell\in{\mathbb{N}},\;{\alpha}\in{\mathbb{N}}^{n},\;\ell+|{\alpha}|\leqslant m}k^{-\frac{\ell}{2}}f_{\ell,{\alpha}}\tilde{\pi}^{\alpha} (52)

where f,α𝒞(U,Hom(A1,A2))f_{\ell,{\alpha}}\in{\mathcal{C}}^{\infty}(U,\operatorname{Hom}(A_{1},A_{2})), π~α=π~1α(1)π~nα(n)\tilde{\pi}^{{\alpha}}=\tilde{\pi}_{1}^{{\alpha}(1)}\ldots\tilde{\pi}_{n}^{{\alpha}(n)} with π~i=1ikxiLkA2\tilde{\pi}_{i}=\tfrac{1}{i\sqrt{k}}\nabla_{\partial_{x_{i}}}^{L^{k}\otimes A_{2}}. Here we use a connection of A2A_{2}, which induces with the connection of LL a covariant derivative of A2LkA_{2}\otimes L^{k}. Proposition 5.1 still holds: Heisenberg differential operators are Heisenberg pseudodifferential operators, the symbol of (52) is |α|mf0,α(x)ξxα\sum_{|{\alpha}|\leqslant m}f_{0,{\alpha}}(x)\xi^{\sharp_{x}{\alpha}},

𝒟Heism(L,;A2,A3)ΨHeisp(L,;A1,A2)ΨHeism+p(L,;A1,A3),\mathcal{D}_{\operatorname{Heis}}^{m}(L,\nabla;A_{2},A_{3})\circ\Psi_{\operatorname{Heis}}^{p}(L,\nabla;A_{1},A_{2})\subset\Psi_{\operatorname{Heis}}^{m+p}(L,\nabla;A_{1},A_{3}),

and the product of symbols is the fiberwise product x\sharp_{x} tensored by the composition Hom(A2,x,A3,x)×Hom(A1,x,A2,x)Hom(A1,x,A3,x).\operatorname{Hom}(A_{2,x},A_{3,x})\times\operatorname{Hom}(A_{1,x},A_{2,x})\rightarrow\operatorname{Hom}(A_{1,x},A_{3,x}). It is easy to see that the definition of the Heisenberg differential operators and of their symbols do not depend on the choice of the connection of A2A_{2}.

In the sequel we assume that A1=A2=AA_{1}=A_{2}=A and is equipped with a Hermitian metric. We use the notation 𝒟Heism(L,;A)\mathcal{D}_{\operatorname{Heis}}^{m}(L,\nabla;A) instead of 𝒟Heism(L,;A,A)\mathcal{D}_{\operatorname{Heis}}^{m}(L,\nabla;A,A) and similarly for the other operator spaces. Our goal is to generalize Theorem 6.4 for P𝒟Heis2(L,;A)P\in\mathcal{D}_{\operatorname{Heis}}^{2}(L,\nabla;A) having a symbol σ{\sigma} of the form

σ(x,ξ)=12|ξ|x2+V(x)\displaystyle{\sigma}(x,\xi)=\tfrac{1}{2}|\xi|_{x}^{2}+V(x) (53)

where |||\cdot| is the norm of TMT^{*}M for a Riemannian metric of MM not necessarily compatible with ω{\omega} and V𝒞(M,EndA)V\in{\mathcal{C}}^{\infty}(M,\operatorname{End}A) is Hermitian at each point. Example of such operators include Schrödinger operators with magnetic field and electric potential, holomorphic Laplacians or semiclassical Dirac operators, cf. [7, Section 3]. Besides of the numerous examples, the interest of these operators is that we can compute explicitly the spectrum of the symbols σ(x,){\sigma}(x,\cdot)

sp(σ(x,))={i=1nBi(x)(α(i)+12)+Vj(x)/αd,j=1,,r}\operatorname{sp}({\sigma}(x,\cdot))=\Bigl{\{}\sum_{i=1}^{n}B_{i}(x)({\alpha}(i)+\tfrac{1}{2})+V_{j}(x)/\;{\alpha}\in{\mathbb{N}}^{d},j=1,\ldots,r\Bigr{\}}

where 0<B1(x)Bd(x)0<B_{1}(x)\leqslant\ldots\leqslant B_{d}(x) are the eigenvalues of ω(x){\omega}(x) with respect to gxg_{x} and V1(x)Vr(x)V_{1}(x)\leqslant\ldots\leqslant V_{r}(x) are the eigenvalues of V(x)V(x). Moreover, we have

12|ξ|x2=i=1dBi(x)h(si,σi),h(y,η)=12(y2+η2)\tfrac{1}{2}|\xi|^{2}_{x}=\sum_{i=1}^{d}B_{i}(x)h(s_{i},{\sigma}_{i}),\qquad h(y,\eta)=\tfrac{1}{2}(y^{2}+\eta^{2})

where sis_{i} and σi{\sigma}_{i} are the linear coordinates of TxMT_{x}^{*}M associated to a symplectic basis. So the analysis of σ(x,){\sigma}(x,\cdot) boils down to the standard quantum harmonic oscillator hwh^{w} or the Landau Hamiltonian h(1i)h(\frac{1}{i}\nabla).

Theorem 7.1.

Let P𝒟Heis2(L,;A)P\in\mathcal{D}_{\operatorname{Heis}}^{2}(L,\nabla;A) having a symbol σ{\sigma} of the form (53) and such that for each kk, PkP_{k} is formally selfadjoint for a volume element of MM. Assume ω{\omega} is nondegenerate and let Σ=xMsp(σ(x,)){\Sigma}=\bigcup_{x\in M}\operatorname{sp}({\sigma}(x,\cdot)). Then

  • -

    For any zΣz\in{\mathbb{C}}\setminus{\Sigma}, there exists Q(z)ΨHeis2(L,;A)Q(z)\in\Psi_{\operatorname{Heis}}^{-2}(L,\nabla;A) such that (Pkz)Qk(z)=id(P_{k}-z)Q_{k}(z)=\operatorname{id} and Qk(z)(Pkz)=idQ_{k}(z)(P_{k}-z)=\operatorname{id} when kk is large.

  • -

    For any EΣE\in{\mathbb{R}}\setminus{\Sigma}, (1(,E](Pk))(1_{(-\infty,E]}(P_{k})) belongs to ΨHeis(L,;A)\Psi_{\operatorname{Heis}}^{-\infty}(L,\nabla;A).

The proof is the same as the one of Theorem 6.4. The symbols τ(z)\tau(z) and pEp_{E} of Q(z)Q(z) and 1(,E]1_{(-\infty,E]} respectively are such that for any xMx\in M,

τ(z)(x,)w=(σ(x,)wz)1,pE(x,)w=1(,E]((σ(x,)w).\tau(z)(x,\cdot)^{w}=({\sigma}(x,\cdot)^{w}-z)^{-1},\qquad p_{E}(x,\cdot)^{w}=1_{(-\infty,E]}(({\sigma}(x,\cdot)^{w}).

In the case where Bi=1B_{i}=1 and V=0V=0 , they have been studied for themselves in [9], [26], and given by the formulas (12) and (13) respectively.

References

  • [1] Richard Beals and Peter Greiner. Calculus on Heisenberg manifolds, volume 119 of Ann. Math. Stud. Princeton, NJ: Princeton University Press, 1988.
  • [2] David Borthwick and Alejandro Uribe. Almost complex structures and geometric quantization. Math. Res. Lett., 3(6):845–861, 1996.
  • [3] L. Charles. Semiclassical Heisenberg pseudodifferential operators. in preparation.
  • [4] L. Charles. Semiclassical aspects of Geometric Quantization. PhD thesis, Paris-Dauphine University, 2000.
  • [5] L. Charles. Quantization of compact symplectic manifolds. J. Geom. Anal., 26(4):2664–2710, 2016.
  • [6] L. Charles. Landau levels of compact manifolds. Arxiv:2012:14190, 2020.
  • [7] L. Charles. On the spectrum of non degenerate magnetic Laplacians. Analysis & PDE, to appear, 2021.
  • [8] Y. Colin de Verdière. Méthodes semi-classiques et théorie spectrale. Lecture notes.
  • [9] Jan Dereziński and Maciej Karczmarczyk. On the Weyl symbol of the resolvent of the harmonic oscillator. Commun. Partial Differ. Equations, 42(10):1537–1548, 2017.
  • [10] Mouez Dimassi and Johannes Sjöstrand. Spectral asymptotics in the semi-classical limit, volume 268 of Lond. Math. Soc. Lect. Note Ser. Cambridge: Cambridge University Press, 1999.
  • [11] Semyon Dyatlov and Maciej Zworski. Mathematical theory of scattering resonances, volume 200 of Grad. Stud. Math. Providence, RI: American Mathematical Society (AMS), 2019.
  • [12] C.L. Epstein and R. Melrose. The Heisenberg algebra, index theory and homology. preprint, 2004.
  • [13] Frédéric Faure and Masato Tsujii. Prequantum transfer operator for symplectic Anosov diffeomorphism. Astérisque, (375):ix+222, 2015.
  • [14] Colin Guillarmou and Benjamin Küster. Spectral theory of the frame flow on hyperbolic 3-manifolds. Ann. Henri Poincaré, 22(11):3565–3617, 2021.
  • [15] Bernard Helffer. Théorie spectrale pour des opérateurs globalement elliptiques, volume 112 of Astérisque. Société Mathématique de France (SMF), Paris, 1984.
  • [16] Lars Hörmander. The analysis of linear partial differential operators. III: Pseudo-differential operators, volume 274 of Grundlehren Math. Wiss. Springer, Cham, 1985.
  • [17] Y. A. Kordyukov. Semiclassical asymptotic expansions for functions of the Bochner-Schrödinger operator. Russ. J. Math. Phys., 30(2):192–208, 2023.
  • [18] Yuri A. Kordyukov. Semiclassical spectral analysis of the Bochner-Schrödinger operator on symplectic manifolds of bounded geometry. Anal. Math. Phys., 12(1):Paper No. 22, 37, 2022.
  • [19] Max Lein. Semiclassical Dynamics and Magnetic Weyl Calculus. PhD thesis, Technische Universität München, 2011.
  • [20] Xiaonan Ma and George Marinescu. Holomorphic Morse inequalities and Bergman kernels, volume 254 of Progress in Mathematics. Birkhäuser Verlag, Basel, 2007.
  • [21] R. Melrose. Introduction to microlocal analysis. Lecture notes, 2003.
  • [22] Léo Morin. Review on spectral asymptotics for the semiclassical Bochner Laplacian of a line bundle. Confluentes Math., 14(1):65–79, 2022.
  • [23] Nicolas Raymond and San Vũ Ngọc. Geometry and spectrum in 2d magnetic wells. Ann. Inst. Fourier, 65(1):137–169, 2015.
  • [24] M. A. Shubin. Pseudodifferential operators and spectral theory. Springer-Verlag, Berlin, second edition, 2001. Translated from the 1978 Russian original by Stig I. Andersson.
  • [25] Michael E. Taylor. Noncommutative microlocal analysis. I, volume 313 of Mem. Am. Math. Soc. Providence, RI: American Mathematical Society (AMS), 1984.
  • [26] André Unterberger. Which pseudodifferential operators with radial symbols are non-negative? J. Pseudo-Differ. Oper. Appl., 7(1):67–90, 2016.
  • [27] Maciej Zworski. Semiclassical analysis, volume 138 of Grad. Stud. Math. Providence, RI: American Mathematical Society (AMS), 2012.