1 Introduction and statement of the main results

Let \((X, T^{1,0}X)\) be a CR manifold of dimension \(2n+1\), \(n\ge 1\). Let \(\Box ^{(q)}_b\) be the Kohn Lalpacian acting on (0, q) forms. The orthogonal projection \(S^{(q)}:L^2_{(0,q)}(X)\rightarrow \mathrm{Ker}\,\Box ^{(q)}_b\) onto \(\mathrm{Ker}\,\Box ^{(q)}_b\) is called the Szegő projection, while its distribution kernel \(S^{(q)}(x,y)\) is called the Szegő kernel. The study of the Szegő projection and kernel is a classical subject in several complex variables and CR geometry. A very important case is when X is a compact strictly pseudoconvex CR manifold. Assume first that X is the boundary of a strictly pseudoconvex domain. Boutet de Monvel-Sjöstrand [2] showed that \(S^{(0)}(x,y)\) is a complex Fourier integral operator.

The Boutet de Monvel-Sjöstrand description of the Szegő kernel had a profound impact in many research areas, especially through [4]: several complex variables, symplectic and contact geometry, geometric quantization, Kähler geometry, semiclassical analysis, quantum chaos, etc. cf. [6, 8, 11, 22, 29, 32], to quote just a few. These ideas also partly motivated the introduction of the recent direct approaches and their various extensions, see [18, 19, 21, 22].

Now, we consider a connected compact Lie group G acting on X. The study of G-invariant Szegő kernel is closely related to Mathematical physics and geometric quantization of CR manifolds. It is a fundamental problem to establish G-invariant Boutet de Monvel-Sjöstrand type theorems for G-invariant Szegő kernels and study the consequence of the G-invariant Szegő kernel. This is the motivation of this work. In this paper, we consider G-invariant Szegő kernel for (0, q) forms and we show that the G-invariant Szegő kernel for (0, q) forms is a complex Fourier integral operator. In particular, \(S^{(q)}(x,y)\) is smoothing outside \(\mu ^{-1}(0)\) and there is a precise description of the singularity near \(\mu ^{-1}(0)\), where \(\mu \) denotes the CR moment map. We apply our result to the case when X admits a transversal CR \(S^1\) action and deduce an asymptotic expansion for the mth Fourier component of the Szegő kernel for (0, q) forms as \(m\rightarrow +\infty \). As an application, we show that, if m large enough, quantization commutes with reduction.

In [20], Ma and Zhang have studied the asymptotic expansion of the invariant Bergman kernel of the \({\text {spin}}^c\) Dirac operator associated with high tensor powers of a positive line bundle on a symplectic manifold admitting a Hamiltonian action of a compact connected Lie group and its relation to the asymptotic expansion of Bergman kernel on symplectic reduced space, also the Toeplitz operator aspect in  [20, Section 4.5]. Their approach is inspired by the analytic localization techniques developed by Bismut and Lebeau [3]. About the quantization commutes with reduction problem in symplectic geometry, we refer the readers to [22]. In the second part of [22], Ma described how the G-invariant Bergman kernel concentrates on the Bergman kernel of the reduced space. Note that the “quantization commutes with reduction” in the situations in symplectic case is a very active subject. When the action connected Lie group is compact and the symplectic manifold is also compact, this question was solved finally by Meinrenken [24] and Tian-Zhang [31]. When the action connected Lie group is compact and the symplectic manifold is noncompact, this is a famous conjecture of Vergne and was solved by Ma-Zhang in [23].

It should be mentioned that in [7], Charles relates the Toeplitz operators on a compact complex manifold M with the Toeplitz operators on the “reduced” space for torus action, and in [26], Paoletti studied equivariant Szegő kernels on complex manifolds ( cf. [20, Remark 0.5]).

We now formulate the main results. We refer to Sect. 2 for some notations and terminology used here. Let \((X, T^{1,0}X)\) be a compact connected orientable CR manifold of dimension \(2n+1\), \(n\ge 1\), where \(T^{1,0}X\) denotes the CR structure of X. Fix a global non-vanishing real 1-form \(\omega _0\in C^\infty (X,T^*X)\) such that \(\langle \,\omega _0\,,\,u\,\rangle =0\), for every \(u\in T^{1,0}X\oplus T^{0,1}X\). The Levi form of X at \(x\in X\) is the Hermitian quadratic form on \(T^{1,0}_xX\) given by \(\mathcal {L}_x(U,\overline{V})=-\frac{1}{2i}\langle \,d\omega _0(x)\,,\,U\wedge \overline{V}\,\rangle \), for all \(U, V\in T^{1,0}_xX\). In this work, we assume that

Assumption 1.1

The Levi form is non-degenerate of constant signature \((n_-,n_+)\) on X. That is, the Levi form has exactly \(n_-\) negative and \(n_+\) positive eigenvalues at each point of X, where \(n_-+n_+=n\).

Let \(HX=\left\{ \mathrm{Re}\,u;\, u\in T^{1,0}X\right\} \) and let \(J:HX\rightarrow HX\) be the complex structure map given by \(J(u+\overline{u})=iu-i\overline{u}\), for every \(u\in T^{1,0}X\). In this work, we assume that X admits a d-dimensional connected compact Lie group G action. We assume throughout that

Assumption 1.2

The G action preserves \(\omega _0\) and J. That is, \(g^*\omega _0=\omega _0\) on X and \(g_*J=Jg_*\) on HX, for every \(g\in G\), where \(g^*\) and \(g_*\) denote the pull-back map and push-forward map of G, respectively.

Let \(\mathfrak {g}\) denote the Lie algebra of G. For any \(\xi \in \mathfrak {g}\), we write \(\xi _X\) to denote the vector field on X induced by \(\xi \). That is, \((\xi _X u)(x)=\frac{\partial }{\partial t}\left( u(\exp (t\xi )\circ x)\right) |_{t=0}\), for any \(u\in C^\infty (X)\).

Definition 1.3

The moment map associated to the form \(\omega _0\) is the map \(\mu :X \rightarrow \mathfrak {g}^*\) such that, for all \(x \in X\) and \(\xi \in \mathfrak {g}\), we have

$$\begin{aligned} \langle \mu (x), \xi \rangle = \omega _0(\xi _X(x)). \end{aligned}$$
(1.1)

In this work, we assume that

Assumption 1.4

0 is a regular value of \(\mu \), the action G on \(\mu ^{-1}(0)\) is freely and

$$\begin{aligned} \underline{\mathfrak {g}}_x\bigcap \underline{\mathfrak {g}}^{\perp _b}_x=\left\{ 0\right\} \text{ at } \text{ every } \text{ point } x\in Y, \end{aligned}$$
(1.2)

where \(\underline{\mathfrak {g}}=\mathrm{Span}\,(\xi _X;\, \xi \in \mathfrak {g})\), \(\underline{\mathfrak {g}}^{\perp _b}=\left\{ v\in HX;\, b(\xi _X,v)=0,\ \ \forall \xi _X\in \underline{\mathfrak {g}}\right\} \), b is the nondegenerate bilinear form on HX given by (2.4).

By Assumption 1.4, \(\mu ^{-1}(0)\) is a d-codimensional submanifold of X. Let \(Y:=\mu ^{-1}(0)\) and let \(HY:=HX\bigcap TY\). Note that if the Levi form is positive at Y, then (1.2) holds. Under the condition (1.2), in Sect. 2.5, we will show that \(\mathrm{dim}\,(HY\bigcap JHY)=2n-2d\) at every point of Y, \(\mu ^{-1}(0)/G=:Y_G\) is a CR manifold with natural CR structure induced by \(T^{1,0}X\) of dimension \(2n-2d+1\) and we can identify \(HY_G\) with \(HY\bigcap JHY\).

Fix a G-invariant smooth Hermitian metric \(\langle \, \cdot \,|\, \cdot \,\rangle \) on \(\mathbb {C}TX\) so that \(T^{1,0}X\) is orthogonal to \(T^{0,1}X\), \(\underline{\mathfrak {g}}\) is orthogonal to \(HY\bigcap JHY\) at every point of Y, \(\langle \, u \,|\, v \, \rangle \) is real if uv are real tangent vectors, \(\langle \,T\,|\,T\,\rangle =1\) and T is orthogonal to \(T^{1,0}X\oplus T^{0,1}X\), where T is given by (2.2). The Hermitian metric \(\langle \,\cdot \,|\, \cdot \, \rangle \) on \(\mathbb {C}TX\) induces, by duality, a Hermitian metric on \(\mathbb {C}T^*X\) and also on the bundles of (0, q) forms \(T^{*0,q}X, q=0, 1, \cdots , n\). We shall also denote all these induced metrics by \(\langle \, \cdot \,|\, \cdot \,\rangle \). Fix \(g\in G\). Let \(g^*:\Lambda ^r_x(\mathbb {C}T^*X)\rightarrow \Lambda ^r_{g^{-1}\circ x}(\mathbb {C}T^*X)\) be the pull-back map. Since G preserves J, we have \(g^*:T^{*0,q}_xX\rightarrow T^{*0,q}_{g^{-1}\circ x}X\), for all \(x\in X\). Thus, for \(u\in \Omega ^{0,q}(X)\), we have \(g^*u\in \Omega ^{0,q}(X)\). Put \(\Omega ^{0,q}(X)^G:=\left\{ u\in \Omega ^{0,q}(X);\, g^*u=u,\ \ \forall g\in G\right\} \). Since the Hermitian metric \(\langle \,\cdot \,|\,\cdot \,\rangle \) on \(\mathbb {C}TX\) is G-invariant, the \(L^2\) inner product \((\,\cdot \,|\,\cdot \,)\) on \(\Omega ^{0,q}(X)\) induced by \(\langle \,\cdot \,|\,\cdot \,\rangle \) is G-invariant. Let \(u\in L^2_{(0,q)}(X)\) and \(g\in G\), we can also define \(g^*u\) in the standard way (see the discussion in the beginning of Sect. 3.2). Put \(L^2_{(0,q)}(X)^G:=\left\{ u\in L^2_{(0,q)}(X);\, g^*u=u,\ \ \forall g\in G\right\} \). Let \(\Box ^{(q)}_b : \mathrm{Dom}\,\Box ^{(q)}_b\rightarrow L^2_{(0,q)}(X)\) be the Gaffney extension of Kohn Laplacian (see (3.1)). Put \((\mathrm{Ker}\,\Box ^{(q)}_b)^G:=\mathrm{Ker}\,\Box ^{(q)}_b\bigcap L^2_{(0,q)}(X)^G\). The G-invariant Szegő projection is the orthogonal projection \(S^{(q)}_G:L^2_{(0,q)}(X)\rightarrow (\mathrm{Ker}\,\Box ^{(q)}_b)^G\) with respect to \((\,\cdot \,|\,\cdot \,)\). Let \(S^{(q)}_G(x,y)\in D'(X\times X,T^{*0,q}X\boxtimes (T^{*0,q}X)^*)\) be the distribution kernel of \(S^{(q)}_G\). The first main result of this work is the following

Theorem 1.5

With the assumptions and notations above, suppose that \(\Box ^{(q)}_b : \mathrm{Dom}\,\Box ^{(q)}_b\rightarrow L^2_{(0,q)}(X)\) has closed range. If \(q\notin \left\{ n_-, n_+\right\} \), then \(S^{(q)}_G\equiv 0\) on X.

Suppose \(q\in \left\{ n_-, n_+\right\} \). Let D be an open set of X with \(D\bigcap \mu ^{-1}(0)=\emptyset \). Then, \(S^{(q)}_G\equiv 0\) on D.

Let \(p\in \mu ^{-1}(0)\) and let U be an open set of p and let \(x=(x_1,\ldots ,x_{2n+1})\) be local coordinates defined in U. Then, there exist continuous operators \(S^G_-, S^G_+:\Omega ^{0,q}_0(U)\rightarrow \Omega ^{0,q}(U)\) such that

$$\begin{aligned} S^{(q)}_G\equiv S^G_-+S^G_+\ \ \text{ on } U, \end{aligned}$$

and \(S^G_-(x,y)\), \(S^G_+(x,y)\) satisfy

$$\begin{aligned} S^G_{\mp }(x, y)\equiv \int ^{\infty }_{0}e^{i\Phi _{\mp }(x, y)t}a_{\mp }(x, y, t)dt\ \ \text{ on } U, \end{aligned}$$

with

$$\begin{aligned} \begin{array}{ll} &{}a_+(x, y, t), a_-(x, y, t)\in S^{n-\frac{d}{2}}_{\mathrm{cl}}\,(U\times U\times \mathbb {R}_+,T^{*0,q}X\boxtimes (T^{*0,q}X)^*), \\ &{}a_-(x,y,t)=0\ \ \text{ if } q\ne n_-,\ \ a_+(x,y,t)=0\ \ \text{ if } q\ne n_+,\\ &{}a^0_-(x,x)\ne 0,\ \ \forall x\in U, \ \ \text{ if } q=n_-,\ \ a^0_+(x,x)\ne 0,\ \ \forall x\in U, \ \ \text{ if } q=n_+, \end{array} \end{aligned}$$
(1.3)

\(a^0_-(x,x)\), \(a^0_+(x,x)\), \(x\in \mu ^{-1}(0)\bigcap U\), are given by (1.8) below, \(\Phi _-(x,y)\in C^\infty (U\times U)\),

$$\begin{aligned} \begin{array}{ll} &{}\mathrm{Im}\,\Phi _-(x,y)\ge 0,\\ &{}d_x\Phi _-(x,x)=-d_y\Phi _-(x,x)=-\omega _0(x),\ \ \forall x\in U\bigcap \mu ^{-1}(0), \end{array} \end{aligned}$$
(1.4)

there is a constant \(C\ge 1\) such that, for all \((x,y)\in U\times U\),

$$\begin{aligned} \begin{array}{ll} &{}|\Phi _-(x,y)|+\mathrm{Im}\,\Phi _-(x,y)\le C \left( \inf \left\{ d^2(g\circ x,y);\, g\in G\right\} +d^2(x,\mu ^{-1}(0))+d^2(y,\mu ^{-1}(0)) \right) , \\ &{}|\Phi _-(x,y)|+\mathrm{Im}\,\Phi _-(x,y)\ge \frac{1}{C} \left( \inf \left\{ d^2(g\circ x,y);\, g\in G\right\} +d^2(x,\mu ^{-1}(0))+d^2(y,\mu ^{-1}(0)) \right) ,\\ &{}Cd^2(x,\mu ^{-1}(0))\ge \mathrm{Im}\,\Phi _-(x,x)\ge \frac{1}{C}d^2(x,\mu ^{-1}(0)),\ \ \forall x\in U, \end{array} \end{aligned}$$
(1.5)

and \(\Phi _-(x,y)\) satisfies (1.18) below and (1.19) below, and \(\Phi _+(x,y)\in C^\infty (U\times U)\), \(-\overline{\Phi }_+(x,y)\) satisfies (1.4), (1.5), (1.18) below and (1.19) below.

We refer the reader to the discussion before (2.1) and Definition 3.1 for the precise meanings of \(A\equiv B\) and the symbol space \(S^{n-\frac{d}{2}}_{\mathrm{cl}}\,\), respectively.

Let \(\Phi \in C^\infty (U\times U)\). We assume that \(\Phi \) satisfies (1.4), (1.5), (1.18), (1.19). We will show in Theorem 5.2 that the functions \(\Phi \) and \(\Phi _-\) are equivalent on U in the sense of Definition 5.1 if and only if there is a function \(f\in C^\infty (U\times U)\) with \(f(x,x)=1\), for every \(x\in \mu ^{-1}(0)\), such that \(\Phi (x,y)-f(x,y)\Phi _-(x,y)\) vanishes to infinite order at \(\mathrm{diag}\,\Bigr ((\mu ^{-1}(0)\bigcap U)\times (\mu ^{-1}(0)\bigcap U)\Bigr )\). From this observation, we see that the leading term \(a^0_-(x,x)\), \(x\in \mu ^{-1}(0)\), is well-defined. To state the formula for \(a^0_-(x,x)\), we introduce some notations. For a given point \(x_0\in X\), let \(\{W_j\}_{j=1}^{n}\) be an orthonormal frame of \((T^{1,0}X,\langle \,\cdot \,|\,\cdot \,\rangle )\) near \(x_0\), for which the Levi form is diagonal at \(x_0\). Put

$$\begin{aligned} \mathcal {L}_{x_0}(W_j,\overline{W}_\ell )=\mu _j(x_0)\delta _{j\ell }\,,\;\; j,\ell =1,\ldots ,n\,. \end{aligned}$$

We will denote by

$$\begin{aligned} \det \mathcal {L}_{x_0}=\prod _{j=1}^{n}\mu _j(x_0)\,. \end{aligned}$$
(1.6)

Let \(\{T_j\}_{j=1}^{n}\) denote the basis of \(T^{*0,1}X\), dual to \(\{\overline{W}_j\}^{n}_{j=1}\). We assume that \(\mu _j(x_0)<0\) if  \(1\le j\le n_-\) and \(\mu _j(x_0)>0\) if  \(n_-+1\le j\le n\). Put

$$\begin{aligned} \begin{array}{ll} &{}\mathcal {N}(x_0,n_-):=\left\{ cT_1(x_0)\wedge \ldots \wedge T_{n_-}(x_0);\, c\in \mathbb {C}\right\} ,\\ &{}\mathcal {N}(x_0,n_+):=\left\{ cT_{n_-+1}(x_0)\wedge \ldots \wedge T_{n}(x_0);\, c\in \mathbb {C}\right\} \end{array} \end{aligned}$$

and let

$$\begin{aligned} \tau _{n_-}=\tau _{x_0,n_-}:T^{*0,q}_{x_0}X\rightarrow \mathcal {N}(x_0,n_-)\,,\quad \tau _{n_+}=\tau _{x_0,n_+}:T^{*0,q}_{x_0}X\rightarrow \mathcal {N}(x_0,n_+)\,, \end{aligned}$$
(1.7)

be the orthogonal projections onto \(\mathcal {N}(x_0,n_-)\) and \(\mathcal {N}(x_0,n_+)\) with respect to \(\langle \,\cdot \,|\,\cdot \,\rangle \), respectively.

Fix \(x\in \mu ^{-1}(0)\), consider the linear map

$$\begin{aligned} \begin{array}{rll} R_x:\underline{\mathfrak {g}}_x&{}\rightarrow &{}\underline{\mathfrak {g}}_x,\\ u&{}\rightarrow &{} R_xu,\ \ \langle \,R_xu\,|\,v\,\rangle =\langle \,d\omega _0(x)\,,\,Ju\wedge v\,\rangle . \end{array} \end{aligned}$$

Let \(\det R_x=\lambda _1(x)\cdots \lambda _d(x)\), where \(\lambda _j(x)\), \(j=1,2,\ldots ,d\), are the eigenvalues of \(R_x\).

Fix \(x\in \mu ^{-1}(0)\), put \(Y_x=\left\{ g\circ x;\, g\in G\right\} \). \(Y_x\) is a d-dimensional submanifold of X. The G-invariant Hermitian metric \(\langle \,\cdot \,|\,\cdot \,\rangle \) induces a volume form \(dv_{Y_x}\) on \(Y_x\). Put

$$\begin{aligned} V_{\mathrm{eff}}\,(x):=\int _{Y_x}dv_{Y_x}. \end{aligned}$$

Note that the function \(V_{\mathrm{eff}}\,(x)\) was already appeared in Ma-Zhang [23, (0,10)] as exactly the role in the expansion, cf. [23, (0.14)].

Theorem 1.6

With the notations used above, for \(a^0_-(x,y)\) and \(a^0_+(x,y)\) in (1.3), we have

$$\begin{aligned} a^0_{\mp }(x,x)=2^{d-1}\frac{1}{V_{\mathrm{eff}}\,(x)}\pi ^{-n-1+\frac{d}{2}}|\det R_x|^{-\frac{1}{2}}|\det \mathcal {L}_{x}|\tau _{x,n_{\mp }},\ \ \forall x\in \mu ^{-1}(0). \end{aligned}$$
(1.8)

We now assume that X admits an \(S^1\) action: \(S^1\times X\rightarrow X\). We write \(e^{i\theta }\) to denote the \(S^1\) action. Let \(T\in C^\infty (X, TX)\) be the global real vector field induced by the \(S^1\) action given by \((Tu)(x)=\frac{\partial }{\partial \theta }\left( u(e^{i\theta }\circ x)\right) |_{\theta =0}\), \(u\in C^\infty (X)\). We assume that the \(S^1\) action \(e^{i\theta }\) is CR and transversal (see Definition 4.1). We take \(\omega _0\in C^\infty (X,T^*X)\) to be the global real one form determined by \(\langle \,\omega _0\,,\,u\,\rangle =0\), for every \(u\in T^{1,0}X\oplus T^{0,1}X\) and \(\langle \,\omega _0\,,\,T\,\rangle =-1\). In this paper, we assume that

Assumption 1.7

$$\begin{aligned}&T \text{ is } \text{ transversal } \text{ to } \text{ the } \text{ space } \underline{\mathfrak {g}} \text{ at } \text{ every } \text{ point } p\in \mu ^{-1}(0),\nonumber \\&e^{i\theta }\circ g\circ x=g\circ e^{i\theta }\circ x,\ \ \forall x\in X,\ \ \forall \theta \in [0,2\pi [,\ \ \forall g\in G, \end{aligned}$$
(1.9)

and

$$\begin{aligned} G\times S^1 \text{ acts } \text{ freely } \text{ near } \mu ^{-1}(0). \end{aligned}$$

Let \(u\in \Omega ^{0,q}(X)\) be arbitrary. Define

$$\begin{aligned} Tu:=\frac{\partial }{\partial \theta }\bigr ((e^{i\theta })^*u\bigr )|_{\theta =0}\in \Omega ^{0,q}(X). \end{aligned}$$

For every \(m\in \mathbb {Z}\), let

$$\begin{aligned} \begin{array}{ll} &{}\Omega ^{0,q}_m(X):=\left\{ u\in \Omega ^{0,q}(X);\, Tu=imu\right\} ,\ \ q=0,1,2,\ldots ,n,\\ &{}\Omega ^{0,q}_{m}(X)^G=\left\{ u\in \Omega ^{0,q}(X)^G;\, Tu=imu\right\} ,\ \ q=0,1,2,\ldots ,n. \end{array} \end{aligned}$$

We denote \(C^\infty _m(X):=\Omega ^{0,0}_m(X)\), \(C^\infty _m(X)^G:=\Omega ^{0,0}_m(X)^G\). From the CR property of the \(S^1\) action and (1.9), it is not difficult to see that \(Tg^*\overline{\partial }_b=g^*T\overline{\partial }_b=\overline{\partial }_bg^*T=\overline{\partial }_bTg^*\) on \(\Omega ^{0,q}(X)\), for all \(g\in G\). Hence,

$$\begin{aligned} \overline{\partial }_b:\Omega ^{0,q}_m(X)^G\rightarrow \Omega ^{0,q+1}_m(X)^G,\ \ \forall m\in \mathbb {Z}. \end{aligned}$$

We now assume that the Hermitian metric \(\langle \,\cdot \,|\,\cdot \,\rangle \) on \(\mathbb {C}TX\) is \(G\times S^1\) invariant. Then the \(L^2\) inner product \((\,\cdot \,|\,\cdot \,)\) on \(\Omega ^{0,q}(X)\) induced by \(\langle \,\cdot \,|\,\cdot \,\rangle \) is \(G\times S^1\)-invariant. We then have

$$\begin{aligned} \begin{array}{ll} &{}Tg^*\overline{\partial }^*_b=g^*T\overline{\partial }^*_b=\overline{\partial }^*_bg^*T=\overline{\partial }^*_bTg^*\ \ \text{ on } \Omega ^{0,q}(X),\ \ \forall g\in G,\\ &{}Tg^*\Box ^{(q)}_b=g^*T\Box ^{(q)}_b=\Box ^{(q)}_bg^*T=\Box ^{(q)}_bTg^*\ \ \text{ on } \Omega ^{0,q}(X),\ \ \forall g\in G, \end{array} \end{aligned}$$

where \(\overline{\partial }^*_b\) is the \(L^2\) adjoint of \(\overline{\partial }_b\) with respect to \((\,\cdot \,|\,\cdot \,)\).

Let \(L^2_{(0,q), m}(X)^G\) be the completion of \(\Omega _m^{0,q}(X)^G\) with respect to \((\,\cdot \,|\,\cdot \,)\). We write \(L^2_m(X)^G:=L^2_{(0,0),m}(X)^G\). Put \(H^q_{b,m}(X)^G:=(\mathrm{Ker}\,\Box ^{(q)}_b)^G_m:=(\mathrm{Ker}\,\Box ^{(q)}_b)^G\bigcap L^2_{(0,q),m}(X)^G\). It is not difficult to see that, for every \(m\in \mathbb {Z}\), \((\mathrm{Ker}\,\Box ^{(q)}_b)^G_m\subset \Omega ^{0,q}_m(X)^G\) and \(\mathrm{dim}\,(\mathrm{Ker}\,\Box ^{(q)}_b)^G_m<\infty \). The mth G-invariant Szegő projection is the orthogonal projection \(S^{(q)}_{G,m}:L^2_{(0,q)}(X)\rightarrow (\mathrm{Ker}\,\Box ^{(q)}_b)^G_m\) with respect to \((\,\cdot \,|\,\cdot \,)\). Let \(S^{(q)}_{G,m}(x,y)\in C^\infty (X\times X,T^{*0,q}X\boxtimes (T^{*0,q}X)^*)\) be the distribution kernel of \(S^{(q)}_{G,m}\). The second main result of this work is the following

Theorem 1.8

With the assumptions and notations used above, if \(q\notin n_-\), then, as \(m\rightarrow +\infty \), \(S^{(q)}_{G,m}=O(m^{-\infty }) \ \text{ on } X\).

Suppose \(q=n_-\). Let D be an open set of X with \(D\bigcap \mu ^{-1}(0)=\emptyset \). Then, as \(m\rightarrow +\infty \), \(S^{(q)}_{G,m}=O(m^{-\infty })\ \ \text{ on } D\).

Let \(p\in \mu ^{-1}(0)\) and let U be an open set of p and let \(x=(x_1,\ldots ,x_{2n+1})\) be local coordinates defined in U. Then, as \(m\rightarrow +\infty \),

$$\begin{aligned} \begin{array}{ll} &{}S^{(q)}_{G,m}(x,y)=e^{im\Psi (x,y)}b(x,y,m)+O(m^{-\infty }),\\ &{}b(x,y,m)\in S^{n-\frac{d}{2}}_{\mathrm{loc}}\,(1; U\times U, T^{*0,q}X\boxtimes (T^{*0,q}X)^*),\\ &{}b(x,y,m)\sim \sum ^\infty _{j=0}m^{n-\frac{d}{2}-j}b_j(x,y) \text{ in } S^{n-\frac{d}{2}}_{\mathrm{loc}}\,(1; U\times U, T^{*0,q}X\boxtimes (T^{*0,q}X)^*),\\ &{}b_j(x,y)\in C^\infty (U\times U, T^{*0,q}X\boxtimes (T^{*0,q}X)^*),\ \ j=0,1,2,\ldots , \end{array} \end{aligned}$$

and

$$\begin{aligned} b_0(x,x)=2^{d-1}\frac{1}{V_{\mathrm{eff}}\,(x)}\pi ^{-n-1+\frac{d}{2}}|\det R_x|^{-\frac{1}{2}}|\det \mathcal {L}_{x}|\tau _{x,n_-},\ \ \forall x\in \mu ^{-1}(0), \end{aligned}$$
(1.10)

where \(\tau _{x,n_-}\) is given by (1.7), and \(\Psi (x,y)\in C^\infty (U\times U)\), \(d_x\Psi (x,x)=-d_y\Psi (x,x)=-\omega _0(x)\), for every \(x\in \mu ^{-1}(0)\), \(\Psi (x,y)=0\) if and only if \(x=y\in \mu ^{-1}(0)\) and there is a constant \(C\ge 1\) such that, for all \((x,y)\in U\times U\),

$$\begin{aligned} \begin{array}{ll} &{}\mathrm{Im}\,\Psi (x,y)\ge \frac{1}{C}\Bigr (d(x,\mu ^{-1}(0))^2+d(y,\mu ^{-1}(0))^2+\inf _{g\in G,\theta \in S^1}d(e^{i\theta }\circ g\circ x,y)^2\Bigr ),\\ &{}\mathrm{Im}\,\Psi (x,y)\le C\Bigr (d(x,\mu ^{-1}(0))^2+d(y,\mu ^{-1}(0))^2+\inf _{g\in G,\theta \in S^1}d(e^{i\theta }\circ g\circ x,y)^2\Bigr ). \end{array} \end{aligned}$$

(We refer the reader to Theorem 1.12 for more properties of the phase \(\Psi (x,y)\).)

We refer the reader to the discussion in the beginning of Sect. 2.2 and Definition 2.1 for the precise meanings of \(A=B+O(m^{-\infty })\) and the symbol space \(S^{n-\frac{d}{2}}_{\mathrm{loc}}\,\), respectively.

It is was proved in Theorem 1.12 in [15]) that when X admits a transversal and CR \(S^1\) action and the Levi form is non-degenerate of constant signature on X, then \(\Box ^{(q)}_b\) has \(L^2\) closed range.

Let \(Y_G:=\mu ^{-1}(0)/G\). In Theorem 2.5, we will show that \(Y_G\) is a CR manifold with natural CR structure induced by \(T^{1,0}X\) of dimension \(2n-2d+1\). Let \(\mathcal {L}_{Y_G,x}\) be the Levi form on \(Y_G\) at \(x\in Y_G\) induced naturally from \(\mathcal {L}\). Note that the bilinear form b is non-degenerate on \(\mu ^{-1}(0)\), where b is given by (2.4). Hence, on \((\underline{\mathfrak {g}}, \underline{\mathfrak {g}})\), b has constant signature on \(\mu ^{-1}(0)\). Assume that on \((\underline{\mathfrak {g}}, \underline{\mathfrak {g}})\), b has r negative eigenvalues and \(d-r\) positive eigenvalues on \(\mu ^{-1}(0)\). Hence \(\mathcal {L}_{Y_G}\) has \(q-r\) negative and \(n-d-q+r\) positive eigenvalues at each point of \(Y_G\). Let \(\Box ^{(q-r)}_{b,Y_G}\) be the Kohn Laplacian for \((0,q-r)\) forms on \(Y_G\). Fix \(m\in \mathbb {N}\). Let \(H^{q-r}_{b,m}(Y_G):=\left\{ u\in \Omega ^{0,q-r}(Y_G);\, \Box ^{(q-r)}_{b,Y_G}u=0,\ \ Tu=imu\right\} \). We will apply Theorem 1.8 to establish an isomorphism between \(H^{q}_{b,m}(X)^G\) and \(H^{q-r}_{b,m}(Y_G)\) if m large enough. We introduce some notations.

Since \(\underline{\mathfrak {g}}_x\) is orthogonal to \(H_xY\bigcap JH_xY\) and \(H_xY\bigcap JH_xY\subset \underline{\mathfrak {g}}^{\perp _b}_x\) (see Lemma 2.4 and (2.5) for the meaning of \(\underline{\mathfrak {g}}^{\perp _b}_x\)), for every \(x\in Y\), we can find a G-invariant orthonormal basis \(\left\{ Z_1,\ldots ,Z_n\right\} \) of \(T^{1,0}X\) on Y such that

$$\begin{aligned} \begin{array}{ll} &{}\mathcal {L}_x(Z_j(x),\overline{Z}_k(x))=\delta _{j,k}\lambda _j(x),\ \ j,k=1,\ldots ,n,\\ &{}Z_j(x)\in \underline{\mathfrak {g}}_x+iJ\underline{\mathfrak {g}}_x,\ \ j=1,2,\ldots ,d,\\ &{}Z_j(x)\in \mathbb {C}H_xY\bigcap J(\mathbb {C}H_xY),\ \ j=d+1,\ldots ,n. \end{array} \end{aligned}$$

Let \(\left\{ e_1,\ldots ,e_n\right\} \) denote the orthonormal basis of \(T^{*0,1}X\) on Y, dual to \(\left\{ \overline{Z}_1,\ldots ,\overline{Z}_n\right\} \). Fix \(s=0,1,2,\ldots ,n-d\). For \(x\in Y\), put

$$\begin{aligned} B^{*0,s}_xX=\left\{ \sum _{d+1\le j_1<\cdots<j_s\le n}a_{j_1,\ldots ,j_s}e_{j_1}\wedge \cdots \wedge e_{j_s};\, a_{j_1,\ldots ,j_s}\in \mathbb {C},\ \forall d+1\le j_1<\cdots <j_s\le n \right\} \end{aligned}$$

and let \(B^{*0,s}X\) be the vector bundle of Y with fiber \(B^{*0,s}_xX\), \(x\in Y\). Let \(C^\infty (Y,B^{*0,s}X)^G\) denote the set of all G-invariant sections of Y with values in \(B^{*0,s}X\). Let

$$\begin{aligned} \iota _G:C^\infty (Y,B^{*0,s}X)^G\rightarrow \Omega ^{0,s}(Y_G) \end{aligned}$$

be the natural identification.

Assume that \(\lambda _1<0,\ldots ,\lambda _r<0\), and \(\lambda _{d+1}<0,\ldots ,\lambda _{n_--r+d}<0\). For \(x\in Y\), put

$$\begin{aligned} \hat{\mathcal {N}}(x,n_-)=\left\{ ce_{d+1}\wedge \cdots \wedge e_{n_--r+d};\, c\in \mathbb {C}\right\} , \end{aligned}$$

and let

$$\begin{aligned} \begin{array}{c} \hat{p}=\hat{p}_x:\mathcal {N}(x,n_-)\rightarrow \hat{\mathcal {N}}(x,n_-),\\ u=ce_1\wedge \cdots \wedge e_r\wedge e_{d+1}\wedge \cdots \wedge e_{n_--r+d}\rightarrow ce_{d+1}\wedge \cdots \wedge e_{n_--r+d}. \end{array} \end{aligned}$$

Let \(\iota :Y\rightarrow X\) be the natural inclusion and let \(\iota ^*:\Omega ^{0,q}(X)\rightarrow \Omega ^{0,q}(Y)\) be the pull-back of \(\iota \). Let \(q=n_-\). Let \(S^{(q-r)}_{Y_G,m}:L^2_{(0,q-r)}(Y_G)\rightarrow H^{q-r}_{b,m}(Y_G)\) be the orthogonal projection and let

$$\begin{aligned} f(x)=\sqrt{V_{\mathrm{eff}}\,(x)}|\det \,R_x|^{-\frac{1}{4}}\in C^\infty (Y)^G. \end{aligned}$$

Let

$$\begin{aligned} \begin{array}{rcl} \sigma _m:H^q_{b,m}(X)^G&{}\rightarrow &{} H^{q-r}_{b,m}(Y_G),\\ u&{}\rightarrow &{} m^{-\frac{d}{4}}S^{(q-r)}_{Y_G,m}\circ \iota _G\circ \hat{p}\circ \tau _{x,n_-}\circ f\circ \iota ^*\circ u. \end{array} \end{aligned}$$

In Sect. 6.2, we will show that

Theorem 1.9

With the notations and assumptions above, suppose that \(q=n_-\). If m is large, then \(\sigma _m:H^q_{b,m}(X)^G\rightarrow H^{q-r}_{b,m}(Y_G)\) is an isomorphism.

In particular, if m large enough, then \( \mathrm{dim}\,H^q_{b,m}(X)^G=\mathrm{dim}\,H^{q-r}_{b,m}(Y_G). \)

Remark 1.10

Let’s sketch the idea of the proof of Theorem 1.9. W can consider \(\sigma _m\) as a map from \(\Omega ^{0,q}(X)\rightarrow H^{q-r}_{b,m}(Y_G)\):

$$\begin{aligned} \begin{array}{rcl} \sigma _m:\Omega ^{0,q}(X)&{}\rightarrow &{} H^{q-r}_{b,m}(Y_G)\subset \Omega ^{0,q-r}(Y_G),\\ u&{}\rightarrow &{} m^{-\frac{d}{4}}S^{(q-r)}_{Y_G,m}\circ \iota _G\circ \hat{p}\circ \tau _{x,n_-}\circ f\circ \iota ^*\circ S^{(q)}_{G,m}u. \end{array} \end{aligned}$$

Let \(\sigma ^*_m:\Omega ^{0,q-r}(Y_G)\rightarrow \Omega ^{0,q}(X)\) be the adjoint of \(\sigma _m\). From Theorem 1.8 and some calculation of complex Fourier integral operators, we will show in Sect. 6.2 that \(F_m=\sigma ^*_m\sigma _m:\Omega ^{0,q}(X)\rightarrow \Omega ^{0,q}(X)\) is the same type of operator as \(S^{(q)}_{G,m}\) and

$$\begin{aligned} \frac{1}{C_0}F_m=(I+R_m)S^{(q)}_{G,m}, \end{aligned}$$
(1.11)

where \(C_0>0\) is a constant and \(R_m\) is also the same type of operator as \(S^{(q)}_{G,m}\), but the leading symbol of \(R_m\) vanishes at \(\mathrm{diag}\,(Y\times Y)\). By using the fact that the leading symbol of \(R_m\) vanishes at \(\mathrm{diag}\,(Y\times Y)\), we will show in Lemma 6.8 that \( \left\| R_mu\right\| \le \varepsilon _m\left\| u\right\| \), for all \(u\in \Omega ^{0,q}(X)\), for all \(m\in \mathbb {N}\), where \(\varepsilon _m\) is a sequence with \(\lim _{m\rightarrow \infty }\varepsilon _m=0\). In particular, if m is large enough, then the map

$$\begin{aligned} I+R_m:\Omega ^{0,q}(X)\rightarrow \Omega ^{0,q}(X) \end{aligned}$$
(1.12)

is injective. From (1.11) and (1.12), we deduce that, if m is large enough, then \(F_m:H^q_{b,m}(X)^G\rightarrow H^q_{b,m}(X)^G\) is injective. Hence \(\sigma _m:H^q_{b,m}(X)^G\rightarrow H^{q-r}_{b,m}(Y_G)\) is injective.

Similarly, we can repeat the argument above with minor change and deduce that if m is large enough, then the map \(\hat{F}_m=\sigma _m\sigma ^*_m: H^{q-r}_{b,m}(Y_G)\rightarrow H^{q-r}_{b,m}(Y_G)\) is injective. Hence, if m is large enough, then the map \(\sigma ^*_m:H^{q-r}_{b,m}(Y_G)\rightarrow H^{q}_{b,m}(X)^G\) is injective. Thus, \(\mathrm{dim}\,H^q_{b,m}(X)^G=\mathrm{dim}\,H^{q-r}_{b,m}(Y_G)\) and \(\sigma _m\) is an isomorphism if m large enough.

Let’s apply Theorem 1.9 to complex case. Let \((L,h^L)\) be a holomorphic line bundle over a connected compact complex manifold (MJ) with \(\mathrm{dim}\,_{\mathbb {C}}M=n\), where J denotes the complex structure map of M and \(h^L\) is a Hermitian fiber metric of L. Let \(R^L\) be the curvature of L induced by \(h^L\). Assume that \(R^L\) is non-degenerate of constant signature \((n_-,n_+)\) on M. Let K be a connected compact Lie group with Lie algebra \(\mathfrak {k}\). We assume that \(\mathrm{dim}\,_{\mathbb {R}}K=d\) and K acts holomorphically on (MJ), and that the action lifts to a holomorphic action on L. We assume further that \(h^L\) is preserved by the K-action. Then \(R^L\) is a K-invariant form. Let \(\omega =\frac{i}{2\pi }R^L\) and let \(\tilde{\mu }: M \rightarrow \mathfrak {k}^*\) be the moment map induced by \(\omega \). Assume that \(0 \in \mathfrak {k}^*\) is regular and the action of K on \(\tilde{\mu }^{-1}(0)\) is freely. The analogue of the Marsden-Weinstein reduction holds (see [10]). More precisely, the complex structure J on M induces a complex structure \(J_K\) on \(M_0:=\tilde{\mu }^{-1}(0)/K\), for which the line bundle \(L_0:=L/K\) is a holomorphic line bundle over \(M_0\).

For any \(\xi \in \mathfrak {k}\), we write \(\xi _M\) to denote the vector field on M induced by \(\xi \). Let \(\underline{\mathfrak {k}}=\mathrm{Span}\,(\xi _M;\, \xi \in \mathfrak {k})\). On \(\tilde{\mu }^{-1}(0)\), let \(b^L\) be the bilinear form on \(\underline{\mathfrak {k}}\times \underline{\mathfrak {k}}\) given by \(b^L(\,\cdot \,,\,\cdot \,)=\omega (\,\cdot \,,\,J\cdot \,)\). Assume that \(b^L\) has r negative eigenvalues and \(d-r\) positive eigenvalues on \(\tilde{\mu }^{-1}(0)\). Let \(q=n_-\). For \(m\in \mathbb {N}\), let \(H^q(M,L^m)^K\) denote the K-invariant qth Dolbeault cohomology group with values in \(L^m\) and let \(H^{q-r}(M_0,L^m_0)\) denote the \((q-r)\)th Dolbeault cohomology group with values in \(L^m_0\). Theorem 1.9 implies that, if m is large enough, then there is an isomorphism map: \(\widetilde{\sigma }_m:H^q(M,L^m)^K\rightarrow H^{q-r}(M_0,L^m_0)\). In particular, if m is large enough,then

$$\begin{aligned} \dim H^q(M,L^m)^K = \dim H^{q-r}(M_0,L_0^m). \end{aligned}$$
(1.13)

Note that when \(m=1\) and \(q=0\), the equality (1.13) was first proved in [10, §5]. For \(m=1\), the equality (1.13) was established in [30, 33] when L is positive. Zhang [33] combined the methods and results in [31] with Braverman’s idea [5] to construct a suitable quasi-isomorphisim to prove the equality (1.13). The proof of the equality (1.13) in [30] is completely algebraic, while the the proof of the equality (1.13) in [33] is purely analytic where different quasi-homomorphisms between Dolbeault complexes under considerations were constructed to prove the equality (1.13). If m is large enough and \(q=0\), an isomorphism map in (1.13) was also constructed in [20, (0.27), Corollary 4.13].

If m large enough and \(q=0\), an isomorphism map in (1.13) was also constructed in [20, (0.27), Corollary 4.13]. The point of [20, (0.27), Corollary 4.13] is to study the isometric aspect of this map, as an consequence of the asymptotic of G-invariant Bergman kernel of Ma-Zhang [20], they gave another proof that it is an isomorphism for m large, and this approaches of the isomorphism for m large is adopted in this paper. It should be mentioned that in this situation, a version of the full asymptotics of \(S^{(0)}_{G,m}(x,y)\) including (1.10) was established in  [20, Theorem 0.1, 0.2].

1.1 The phase functions \(\Phi _-(x,y)\) and \(\Psi (x,y)\)

In this section, we collect some properties of the phase functions \(\Phi _-(x,y)\), \(\Psi (x,y)\) in Theorem 1.5 and Theorem 1.8.

Let \(v=(v_1,\ldots ,v_d)\) be local coordinates of G defined in a neighborhood V of \(e_0\) with \(v(e_0)=(0,\ldots ,0)\). From now on, we will identify the element \(e\in V\) with v(e). Fix \(p\in \mu ^{-1}(0)\). In Theorem 3.7, we will show that there exist local coordinates \(v=(v_1,\ldots ,v_d)\) of G defined in a neighborhood V of \(e_0\) with \(v(e_0)=(0,\ldots ,0)\), local coordinates \(x=(x_1,\ldots ,x_{2n+1})\) of X defined in a neighborhood \(U=U_1\times U_2\) of p with \(0\leftrightarrow p\), where \(U_1\subset \mathbb {R}^d\) is an open set of \(0\in \mathbb {R}^d\), \(U_2\subset \mathbb {R}^{2n+1-d}\) is an open set of \(0\in \mathbb {R}^{2n+1-d} \) and a smooth function \(\gamma =(\gamma _1,\ldots ,\gamma _d)\in C^\infty (U_2,U_1)\) with \(\gamma (0)=0\in \mathbb {R}^d\) such that

$$\begin{aligned}&\begin{array}{ll} &{}(v_1,\ldots ,v_d)\circ (\gamma (x_{d+1},\ldots ,x_{2n+1}),x_{d+1},\ldots ,x_{2n+1})\\ &{}=(v_1+\gamma _1(x_{d+1},\ldots ,x_{2n+1}),\ldots ,v_d+\gamma _d(x_{d+1},\ldots ,x_{2n+1}),x_{d+1},\ldots ,x_{2n+1}),\\ &{}\forall (v_1,\ldots ,v_d)\in V,\ \ \forall (x_{d+1},\ldots ,x_{2n+1})\in U_2, \end{array} \nonumber \\\end{aligned}$$
(1.14)
$$\begin{aligned}&\begin{array}{ll} &{}\underline{\mathfrak {g}}=\mathrm{span}\,\left\{ \frac{\partial }{\partial x_1},\ldots ,\frac{\partial }{\partial x_d}\right\} ,\\ &{} \mu ^{-1}(0)\bigcap U=\left\{ x_{d+1}=\cdots =x_{2d}=0\right\} ,\\ &{}\text{ On } \mu ^{-1}(0)\bigcap U, \text{ we } \text{ have } J(\frac{\partial }{\partial x_j})=\frac{\partial }{\partial x_{d+j}}+a_j(x)\frac{\partial }{\partial x_{2n+1}}, j=1,2,\ldots ,d, \end{array} \end{aligned}$$
(1.15)

where \(a_j(x)\) is a smooth function on \(\mu ^{-1}(0)\bigcap U\), independent of \(x_1,\ldots ,x_{2d}\), \(x_{2n+1}\) and \(a_j(0)=0\), \(j=1,\ldots ,d\),

$$\begin{aligned} \begin{array}{ll} &{} T^{1,0}_pX=\mathrm{span}\,\left\{ Z_1,\ldots ,Z_n\right\} , \\ &{} Z_j=\frac{1}{2}(\frac{\partial }{\partial x_j}-i\frac{\partial }{\partial x_{d+j}})(p),\ \ j=1,\ldots ,d,\\ &{} Z_j=\frac{1}{2}(\frac{\partial }{\partial x_{2j-1}}-i\frac{\partial }{\partial x_{2j}})(p),\ \ j=d+1,\ldots ,n, \\ &{} \langle \,Z_j\,|\,Z_k\,\rangle =\delta _{j,k},\ \ j,k=1,2,\ldots ,n,\\ &{} \mathcal {L}_p(Z_j, \overline{Z}_k)=\mu _j\delta _{j,k},\ \ j,k=1,2,\ldots ,n \end{array} \end{aligned}$$
(1.16)

and

$$\begin{aligned} \begin{array}{clll} \omega _0(x)&{}=(1+O(|x|))dx_{2n+1}+\sum ^d_{j=1}4\mu _jx_{d+j}dx_j\\ &{}\quad +\sum ^n_{j=d+1}2\mu _jx_{2j}dx_{2j-1}-\sum ^n_{j=d+1}2\mu _jx_{2j-1}dx_{2j}\\ &{} +\sum ^{2n}_{j=d+1}b_jx_{2n+1}dx_j+O(|x|^2), \end{array} \end{aligned}$$
(1.17)

where \(b_{d+1}\in \mathbb {R},\ldots ,b_{2n}\in \mathbb {R}\). Put \(x''=(x_{d+1},\ldots ,x_{2n+1})\), \(\hat{x}''=(x_{d+1}, x_{d+2},\ldots ,x_{2d})\), \(\mathring{x}''=(x_{d+1},\ldots ,x_{2n})\). We have the following (see Theorem 3.11 and Theorem 3.12)

Theorem 1.11

With the notations above, the phase function \(\Phi _-(x,y)\in C^\infty (U\times U)\) is independent of \((x_1,\ldots ,x_d)\) and \((y_1,\ldots ,y_d)\). Hence, \(\Phi _-(x,y)=\Phi _-((0,x''),(0,y'')):=\Phi _-(x'',y'')\). Moreover, there is a constant \(c>0\) such that

$$\begin{aligned} \mathrm{Im}\,\Phi _-(x'',y'')\ge c\Bigr (|\hat{x}''|^2+|\hat{y}''|^2+|\mathring{x}''-\mathring{y}''|^2\Bigr ),\ \ \forall ((0,x''),(0,y''))\in U\times U. \end{aligned}$$
(1.18)

Furthermore,

$$\begin{aligned} \begin{array}{cl} \Phi _-(x'', y'')&{}=-x_{2n+1}+y_{2n+1}+2i\sum ^d_{j=1}|\mu _j|y^2_{d+j}+2i\sum ^d_{j=1}|\mu _j|x^2_{d+j}\\ &{}+i\sum ^{n}_{j=d+1}|\mu _j||z_j-w_j|^2 +\sum ^{n}_{j=d+1}i\mu _j(\overline{z}_jw_j-z_j\overline{w}_j)\\ &{}+\sum ^d_{j=1}(-b_{d+j}x_{d+j}x_{2n+1}+b_{d+j}y_{d+j}y_{2n+1})\\ &{}+\sum ^n_{j=d+1}\frac{1}{2}(b_{2j-1}-ib_{2j})(-z_jx_{2n+1}+w_jy_{2n+1})\\ &{}+\sum ^n_{j=d+1}\frac{1}{2}(b_{2j-1}+ib_{2j})(-\overline{z}_jx_{2n+1}+\overline{w}_jy_{2n+1})\\ &{}+(x_{2n+1}-y_{2n+1})f(x, y) +O(|(x, y)|^3), \end{array} \end{aligned}$$
(1.19)

where \(z_j=x_{2j-1}+ix_{2j}\), \(w_j=y_{2j-1}+iy_{2j}\), \(j=d+1,\ldots ,n\), \(\mu _j\), \(j=1,\ldots ,n\), and \(b_{d+1}\in \mathbb {R},\ldots ,b_{2n}\in \mathbb {R}\) are as in (1.17) and f is smooth and satisfies \(f(0,0)=0\), \(f(x, y)=\overline{f}(y, x)\).

We now assume that X admits an \(S^1\) action: \(S^1\times X\rightarrow X\). We will use the same notations as in Theorem 1.8. Recall that we work with Assumption 1.7. Let \(p\in \mu ^{-1}(0)\). We can repeat the proof of Theorem 3.7 with minor change and show that there exist local coordinates \(v=(v_1,\ldots ,v_d)\) of G defined in a neighborhood V of \(e_0\) with \(v(e_0)=(0,\ldots ,0)\), local coordinates \(x=(x_1,\ldots ,x_{2n+1})\) of X defined in a neighborhood \(U=U_1\times (\hat{U}_2\times ]-2\delta ,2\delta [)\) of p with \(0\leftrightarrow p\), where \(U_1\subset \mathbb {R}^d\) is an open set of \(0\in \mathbb {R}^d\), \(\hat{U}_2\subset \mathbb {R}^{2n-d}\) is an open set of \(0\in \mathbb {R}^{2n-d} \), \(\delta >0\), and a smooth function \(\gamma =(\gamma _1,\ldots ,\gamma _d)\in C^\infty (\hat{U}_2\times ]-2\delta ,2\delta [,U_1)\) with \(\gamma (0)=0\in \mathbb {R}^d\) such that \(T=-\frac{\partial }{\partial x_{2n+1}}\) and (1.14), (1.15), (1.16), (1.17) hold. We have the following

Theorem 1.12

With the notations above, the phase function \(\Psi \) satisfies \(\Psi (x,y)=-x_{2n+1}+y_{2n+1}+\hat{\Psi }(\mathring{x}'',\mathring{y}'')\), where \(\hat{\Psi }(\mathring{x}'',\mathring{y}'')\in C^\infty (U\times U)\), \(\mathring{x}''=(x_{d+1},\ldots ,x_{2n})\), \(\mathring{y}''=(y_{d+1},\ldots ,y_{2n})\), and \(\Psi \) satisfies (1.18) and (1.19).

2 Preliminaries

2.1 Standard notations

Let M be a \(C^\infty \) paracompact manifold. We let TM and \(T^*M\) denote the tangent bundle of M and the cotangent bundle of M, respectively. The complexified tangent bundle of M and the complexified cotangent bundle of M will be denoted by \(\mathbb {C}TM\) and \(\mathbb {C}T^*M\), respectively. Write \(\langle \,\cdot \,,\cdot \,\rangle \) to denote the pointwise duality between TM and \(T^*M\). We extend \(\langle \,\cdot \,,\cdot \,\rangle \) bilinearly to \(\mathbb {C}TM\times \mathbb {C}T^*M\). Let B be a \(C^\infty \) vector bundle over M. The fiber of B at \(x\in M\) will be denoted by \(B_x\). Let E be a vector bundle over a \(C^\infty \) paracompact manifold \(M_1\). We write \(B\boxtimes E^*\) to denote the vector bundle over \(M\times M_1\) with fiber over \((x, y)\in M\times M_1\) consisting of the linear maps from \(E_y\) to \(B_x\). Let \(Y\subset M\) be an open set. From now on, the spaces of distribution sections of B over Y and smooth sections of B over Y will be denoted by \(D'(Y, B)\) and \(C^\infty (Y, B)\), respectively. Let \(E'(Y, B)\) be the subspace of \(D'(Y, B)\) whose elements have compact support in Y.

We recall the Schwartz kernel theorem [12, Theorems 5.2.1, 5.2.6], [19, Thorem B.2.7]. Let B and E be \(C^\infty \) vector bundles over paracompact orientable \(C^\infty \) manifolds M and \(M_1\), respectively, equipped with smooth densities of integration. If \(A: C^\infty _0(M_1,E)\rightarrow D'(M,B)\) is continuous, we write \(K_A(x, y)\) or A(xy) to denote the distribution kernel of A. The following two statements are equivalent

  1. (1)

    A is continuous: \(E'(M_1,E)\rightarrow C^\infty (M,B)\),

  2. (2)

    \(K_A\in C^\infty (M\times M_1,B\boxtimes E^*)\).

If A satisfies (1) or (2), we say that A is smoothing on \(M \times M_1\). Let \(A,\hat{A}: C^\infty _0(M_1,E)\rightarrow D'(M,B)\) be continuous operators. We write

$$\begin{aligned} A\equiv \hat{A} (\text{ on } M\times M_1) \end{aligned}$$
(2.1)

if \(A-\hat{A}\) is a smoothing operator. If \(M=M_1\), we simply write “on M”.

Let \(H(x,y)\in D'(M\times M_1,B\boxtimes E^*)\). We write H to denote the unique continuous operator \(C^\infty _0(M_1,E)\rightarrow D'(M,B)\) with distribution kernel H(xy). In this work, we identify H with H(xy).

2.2 Some standard notations in semi-classical analysis

Let \(W_1\) be an open set in \(\mathbb {R}^{N_1}\) and let \(W_2\) be an open set in \(\mathbb {R}^{N_2}\). Let E and F be vector bundles over \(W_1\) and \(W_2\), respectively. An m-dependent continuous operator \(A_m:C^\infty _0(W_2,F)\rightarrow D'(W_1,E)\) is called m-negligible on \(W_1\times W_2\) if, for m large enough, \(A_m\) is smoothing and, for any \(K\Subset W_1\times W_2\), any multi-indices \(\alpha \), \(\beta \) and any \(N\in \mathbb {N}\), there exists \(C_{K,\alpha ,\beta ,N}>0\) such that

$$\begin{aligned} |\partial ^\alpha _x\partial ^\beta _yA_m(x, y)|\le C_{K,\alpha ,\beta ,N}m^{-N}\,\, \text {on } K,\ \ \forall m\gg 1. \end{aligned}$$

In that case we write

$$\begin{aligned} A_m(x,y)=O(m^{-\infty })\,\,\text {on } W_1\times W_2, \quad \text {or} \quad A_m=O(m^{-\infty })\,\,\text {on } W_1\times W_2. \end{aligned}$$

If \(A_m, B_m:C^\infty _0(W_2, F)\rightarrow D'(W_1, E)\) are m-dependent continuous operators, we write \(A_m= B_m+O(m^{-\infty })\) on \(W_1\times W_2\) or \(A_m(x,y)=B_m(x,y)+O(m^{-\infty })\) on \(W_1\times W_2\) if \(A_m-B_m=O(m^{-\infty })\) on \(W_1\times W_2\). When \(W=W_1=W_2\), we sometime write “on W”.

Let X and M be smooth manifolds and let E and F be vector bundles over X and M, respectively. Let \(A_m, B_m:C^\infty (M,F)\rightarrow C^\infty (X,E)\) be m-dependent smoothing operators. We write \(A_m=B_m+O(m^{-\infty })\) on \(X\times M\) if on every local coordinate patch D of X and local coordinate patch \(D_1\) of M, \(A_m=B_m+O(m^{-\infty })\) on \(D\times D_1\). When \(X=M\), we sometime write on X.

We recall the definition of the semi-classical symbol spaces

Definition 2.1

Let W be an open set in \(\mathbb {R}^N\). Let

$$\begin{aligned} \begin{array}{c} S(1;W):=\Big \{a\in C^\infty (W)\,|\, \forall \alpha \in \mathbb {N}^N_0: \sup _{x\in W}|\partial ^\alpha a(x)|<\infty \Big \},\\ S^0_{\mathrm{loc}}\,(1;W):=\Big \{(a(\cdot ,m))_{m\in \mathbb {R}}\,|\, \forall \alpha \in \mathbb {N}^N_0, \forall \chi \in C^\infty _0(W)\,:\,\sup _{m\in \mathbb {R}, m\ge 1}\sup _{x\in W}|\partial ^\alpha (\chi a(x,m))|<\infty \Big \}\,. \end{array} \end{aligned}$$

For \(k\in \mathbb {R}\), let

$$\begin{aligned} S^k_{\mathrm{loc}}(1):=S^k_{\mathrm{loc}}(1;W)=\Big \{(a(\cdot ,m))_{m\in \mathbb {R}}\,|\,(m^{-k}a(\cdot ,m))\in S^0_{\mathrm{loc}}\,(1;W)\Big \}\,. \end{aligned}$$

Hence \(a(\cdot ,m)\in S^k_{\mathrm{loc}}(1;W)\) if for every \(\alpha \in \mathbb {N}^N_0\) and \(\chi \in C^\infty _0(W)\), there exists \(C_\alpha >0\) independent of m, such that \(|\partial ^\alpha (\chi a(\cdot ,m))|\le C_\alpha m^{k}\) holds on W.

Consider a sequence \(a_j\in S^{k_j}_{\mathrm{loc}}\,(1)\), \(j\in \mathbb {N}_0\), where \(k_j\searrow -\infty \), and let \(a\in S^{k_0}_{\mathrm{loc}}\,(1)\). We say

$$\begin{aligned} a(\cdot ,m)\sim \sum \limits ^\infty _{j=0}a_j(\cdot ,m)\,\,\text {in } S^{k_0}_{\mathrm{loc}}\,(1), \end{aligned}$$

if, for every \(\ell \in \mathbb {N}_0\), we have \(a-\sum ^{\ell }_{j=0}a_j\in S^{k_{\ell +1}}_{\mathrm{loc}}\,(1)\). For a given sequence \(a_j\) as above, we can always find such an asymptotic sum a, which is unique up to an element in \(S^{-\infty }_{\mathrm{loc}}\,(1)=S^{-\infty }_{\mathrm{loc}}\,(1;W):=\cap _kS^k_{\mathrm{loc}}\,(1)\).

Similarly, we can define \(S^k_{\mathrm{loc}}\,(1;Y,E)\) in the standard way, where Y is a smooth manifold and E is a vector bundle over Y.

2.3 CR manifolds and bundles

Let \((X, T^{1,0}X)\) be a compact, connected and orientable CR manifold of dimension \(2n+1\), \(n\ge 1\), where \(T^{1,0}X\) is a CR structure of X, that is, \(T^{1,0}X\) is a subbundle of rank n of the complexified tangent bundle \(\mathbb {C}TX\), satisfying \(T^{1,0}X\cap T^{0,1}X=\{0\}\), where \(T^{0,1}X=\overline{T^{1,0}X}\), and \([\mathcal V,\mathcal V]\subset \mathcal V\), where \(\mathcal V=C^\infty (X, T^{1,0}X)\). There is a unique subbundle HX of TX such that \(\mathbb {C}HX=T^{1,0}X \oplus T^{0,1}X\), i.e. HX is the real part of \(T^{1,0}X \oplus T^{0,1}X\). Let \(J:HX\rightarrow HX\) be the complex structure map given by \(J(u+\overline{u})=iu-i\overline{u}\), for every \(u\in T^{1,0}X\). By complex linear extension of J to \(\mathbb {C}TX\), the i-eigenspace of J is \(T^{1,0}X \, = \, \left\{ V \in \mathbb {C}HX \, : \, JV \, = \, \sqrt{-1}V \right\} \). We shall also write (XHXJ) to denote a compact CR manifold.

We fix a real non-vanishing 1 form \(\omega _0\in C(X,T^*X)\) so that \(\langle \,\omega _0(x)\,,\,u\,\rangle =0\), for every \(u\in H_xX\), for every \(x\in X\). For each \(x \in X\), we define a quadratic form on HX by

$$\begin{aligned} \mathcal {L}_x(U,V) =\frac{1}{2}d\omega _0(JU, V), \forall \ U, V \in H_xX. \end{aligned}$$

We extend \(\mathcal {L}\) to \(\mathbb {C}HX\) by complex linear extension. Then, for \(U, V \in T^{1,0}_xX\),

$$\begin{aligned} \mathcal {L}_x(U,\overline{V}) = \frac{1}{2}d\omega _0(JU, \overline{V}) = -\frac{1}{2i}d\omega _0(U,\overline{V}). \end{aligned}$$

The Hermitian quadratic form \(\mathcal {L}_x\) on \(T^{1,0}_xX\) is called Levi form at x. We recall that in this paper, we always assume that the Levi form \(\mathcal {L}\) on \(T^{1,0}X\) is non-degenerate of constant signature \((n_-,n_+)\) on X, where \(n_-\) denotes the number of negative eigenvalues of the Levi form and \(n_+\) denotes the number of positive eigenvalues of the Levi form. Let \(T\in C^\infty (X,TX)\) be the non-vanishing vector field determined by

$$\begin{aligned} \omega _0(T)=-1,\quad d\omega _0(T,\cdot )\equiv 0\ \ \text{ on } TX. \end{aligned}$$
(2.2)

Note that X is a contact manifold with contact form \(\omega _0\), contact plane HX and T is the Reeb vector field.

Fix a smooth Hermitian metric \(\langle \, \cdot \,|\, \cdot \,\rangle \) on \(\mathbb {C}TX\) so that \(T^{1,0}X\) is orthogonal to \(T^{0,1}X\), \(\langle \, u \,|\, v \,\rangle \) is real if uv are real tangent vectors, \(\langle \,T\,|\,T\,\rangle =1\) and T is orthogonal to \(T^{1,0}X\oplus T^{0,1}X\). For \(u \in \mathbb {C}TX\), we write \(|u|^2 := \langle \, u\, |\, u\, \rangle \). Denote by \(T^{*1,0}X\) and \(T^{*0,1}X\) the dual bundles \(T^{1,0}X\) and \(T^{0,1}X\), respectively. They can be identified with subbundles of the complexified cotangent bundle \(\mathbb {C}T^*X\). Define the vector bundle of (0, q)-forms by \(T^{*0,q}X := \wedge ^qT^{*0,1}X\). The Hermitian metric \(\langle \, \cdot \, |\, \cdot \,\rangle \) on \(\mathbb {C}TX\) induces, by duality, a Hermitian metric on \(\mathbb {C}T^*X\) and also on the bundles of (0, q) forms \(T^{*0,q}X, q=0, 1, \cdots , n\). We shall also denote all these induced metrics by \(\langle \,\cdot \,|\, \cdot \,\rangle \). Note that we have the pointwise orthogonal decompositions:

$$\begin{aligned} \begin{array}{c} \mathbb {C}T^*X = T^{*1,0}X \oplus T^{*0,1}X \oplus \left\{ \lambda \omega _0: \lambda \in \mathbb {C} \right\} , \\ \mathbb {C}TX = T^{1,0}X \oplus T^{0,1}X \oplus \left\{ \lambda T: \lambda \in \mathbb {C} \right\} . \end{array} \end{aligned}$$

For \(x, y\in X\), let d(xy) denote the distance between x and y induced by the Hermitian metric \(\langle \cdot \mid \cdot \rangle \). Let A be a subset of X. For every \(x\in X\), let \(d(x,A):=\inf \left\{ d(x,y);\, y\in A\right\} \).

Let D be an open set of X. Let \(\Omega ^{0,q}(D)\) denote the space of smooth sections of \(T^{*0,q}X\) over D and let \(\Omega ^{0,q}_0(D)\) be the subspace of \(\Omega ^{0,q}(D)\) whose elements have compact support in D.

2.4 Contact reduction

Let G be a connected compact Lie group with Lie algebra \(\mathfrak {g}\) such that \(\dim _{\mathbb {R}}G = d\). We assume that the Lie group G acts on X preserving \(\omega _0\), i.e. \(g^*\omega _0 =\omega _0\), for any \(g \in G\). For any \(\xi \in \mathfrak {g}\), there is an induced vector field \(\xi _X\) on X given by \((\xi _X u)(x)=\frac{\partial }{\partial t}\left( u( \exp (t\xi )\circ x)\right) |_{t=0}\), for any \(u\in C^\infty (X)\).

Definition 2.2

The contact moment map associated to the form \(\omega _0\) is the map \(\mu :X \rightarrow \mathfrak {g}^*\) such that, for all \(x \in X\) and \(\xi \in \mathfrak {g}\), we have

$$\begin{aligned} \langle \mu (x), \xi \rangle = \omega _0(\xi _X(x)). \end{aligned}$$
(2.3)

We now recall the contact reduction from [1, 9]. It was shown in [1, 9] that the contact moment map is G-equivariant, so G acts on \(Y:=\mu ^{-1}(0)\), where G acts on \(\mathfrak {g}^*\) through co-adjoint represent. Since we assume that the action of G on Y is freely, \(Y_G:=\mu ^{-1}(0)/G\) is a smooth manifold. Let \(\pi : Y \rightarrow Y_G\) and \(\iota :Y \hookrightarrow X\) be the natural quotient and inclusion, respectively, then there is a unique induced contact form \(\widetilde{\omega }_0\) on \(Y_G\) such \(\pi ^*\widetilde{\omega }_0 = \iota ^* \omega _0\). We denote by \(HY:= {\text {Ker}} \omega _0 \cap T(\mu ^{-1}(0)) = HX \cap TY\), then the induced contact plane on \(Y_G\) is \(HY_G := \pi _*HY\). In particular, \(\dim HY = 2n-d\) and \(\dim HY_G=2n-2d\).

2.5 CR reduction

In this subsection we study the reduction of CR manifolds with non-degenerate Levi curvature which is a CR analogue of the reduction on complex manifolds considered in [27, §2.1]. For the case of strictly pseudoconvex CR manifolds, the CR reduction was also studied in [17].

Recall that we work with Assumption 1.2. Let b be the nondegenerate bilinear form on HX such that

$$\begin{aligned} b(\cdot , \cdot ) = d\omega _0(\cdot , J\cdot ). \end{aligned}$$
(2.4)

We denote by \(\underline{\mathfrak {g}} := {\text {Span}} (\xi _X, \xi \in \mathfrak {g})\) the tangent bundle of the orbits in X. Let

$$\begin{aligned} \underline{\mathfrak {g}}^{\perp _b}=\left\{ v\in HX;\, b(\xi _X,v)=0,\ \ \forall \xi _X\in \underline{\mathfrak {g}}\right\} . \end{aligned}$$
(2.5)

Since we assume that \(\underline{\mathfrak {g}}_x \cap \underline{\mathfrak {g}}^{\perp _b}_x = \left\{ 0 \right\} \), for every \(x\in Y\), we immediately get

Lemma 2.3

When restricted to \( \underline{\mathfrak {g}} \times \underline{\mathfrak {g}}\), the bilinear form b is nondegenerate on Y.

For \(x \in Y, V \in H_xX\) and \(\xi \in \mathfrak {g}\), by (2.3) and (2.4), we have

$$\begin{aligned} b_x(\xi _X,JV) =-d\omega _0(x)(\xi _X, V) = -\left( d\mu (x)(V) \right) (\xi ). \end{aligned}$$

Therefore,

$$\begin{aligned} JV \in \underline{\mathfrak {g}}^{\perp _b} |_Y \iff d\mu (x)(V) =0. \end{aligned}$$
(2.6)

Since \(Y=\mu ^{-1}(0)\), we have

$$\begin{aligned} d\mu (x)(V) =0 \iff V \in T_xY. \end{aligned}$$
(2.7)

In particular, for \(x \in Y\),

$$\begin{aligned} \dim \underline{\mathfrak {g}}^{\perp _b}_x = \dim (H_xX \cap T_xY) = \dim H_xY =2n-d. \end{aligned}$$

By (2.2), (2.7) and the definition of \(\underline{\mathfrak {g}}\), we have \(\underline{\mathfrak {g}} \subset HX|_Y\). From Lemma 2.3, we can check that \(\underline{\mathfrak {g}}+\underline{\mathfrak {g}}^{\perp _b}=HX|_Y\). Since \(\underline{\mathfrak {g}}_x \cap \underline{\mathfrak {g}}^{\perp _b}_x = \left\{ 0 \right\} \), for every \(x\in Y\), this sum is a direct sum.

Let U be a small open G-invariant neighborhood of Y. Since G acts freely on Y, we can thus also assume that G acts freely on \(\overline{U}\). Since \(\underline{\mathfrak {g}}_x \cap \underline{\mathfrak {g}}^{\perp _b}_x = \left\{ 0 \right\} \), for \(x\in Y\), we have, for \(x \in Y\),

$$\begin{aligned} H_xU = \underline{\mathfrak {g}}_x \oplus \underline{\mathfrak {g}}^{\perp _b}_x. \end{aligned}$$
(2.8)

Then, by (2.8), we can choose the horizontal bundles of the fibrations \(U \rightarrow U_G:=U/G\) and \(Y \rightarrow Y_G\) to be

$$\begin{aligned} H^HU = \underline{\mathfrak {g}}^{\perp _b}|_U, \quad H^HY := H^HU|_Y \cap HY. \end{aligned}$$
(2.9)

Hence

$$\begin{aligned} HY = \underline{\mathfrak {g}}|_Y \oplus H^HY. \end{aligned}$$

Lemma 2.4

$$\begin{aligned}&\underline{\mathfrak {g}}^{\perp _b} |_Y =JHY. \end{aligned}$$
(2.10)
$$\begin{aligned}&HU|_Y = J\underline{\mathfrak {g}}|_Y \oplus HY = \underline{\mathfrak {g}}|_Y \oplus J\underline{\mathfrak {g}}|_Y \oplus H^HY. \end{aligned}$$
(2.11)

Proof

The identity (2.10) follows from (2.6) and (2.7). For \(x \in Y, V \in H_xY\) and \(\xi \in \mathfrak {g}\),

$$\begin{aligned} b_x(J\xi _X, V) = d\omega _0(x)(\xi _X, V) = \left( d\mu (x)(V) \right) (\xi )=0. \end{aligned}$$
(2.12)

Using (2.12), \(\dim H_xU = \dim H_xY + \dim J\underline{\mathfrak {g}}_x\), and the fact that b is nondegenerate on JHY, we obtain (2.11). \(\square \)

By (2.9), and (2.10), we have \(H^HY = JHY \cap HY\). In particular, \(H^HY\) is preserved by J, so we can define the homomorphism \(J_G\) on \(HY_G\) in the following way: For \(V \in HY_G\), we denote by \(V^H\) its lift in \(H^HY\), and we define \(J_G\) on \(Y_G\) by

$$\begin{aligned} (J_GV)^H = J(V^H). \end{aligned}$$
(2.13)

Hence, we have \(J_G: HY_G \rightarrow HY_G\) such that \(J_G^2 = -{\text {id}}\), where \({\text {id}}\) denotes the identity map \({\text {id}} \, : \, HY_G \rightarrow HY_G\). By complex linear extension of \(J_G\) to \(\mathbb {C}TY_G\), we can define the i-eigenspace of \(J_G\) is given by \(T^{1,0}Y_G \, = \, \left\{ V \in \mathbb {C}HY_G \, : \, J_GV \, = \, \sqrt{-1} V \right\} \).

Theorem 2.5

The subbundle \(T^{1,0}Y_G\) is a CR structure of \(Y_G\).

Proof

Let \(u, v \in C^\infty (Y_G, T^{1,0}Y_G)\), then we can find \(U, V \in C^\infty (Y_G, TY_G)\) such that

$$\begin{aligned} u= U - \sqrt{-1}J_GU, \qquad v=V-\sqrt{-1}J_GV. \end{aligned}$$

By (2.13), we have

$$\begin{aligned} u^H=U^H-\sqrt{-1}JU^H, \quad v = V^H - \sqrt{-1}JV^H \in T^{1,0}X \cap \mathbb {C}HY. \end{aligned}$$

Since \(T^{1,0}X\) is a CR structure and it is clearly that \([u^H, v^H] \in \mathbb {C}HY\), we have \([u^H, v^H] \in T^{1,0}X \cap \mathbb {C}HY\). Hence, there is a \(W \in C^\infty (X, HX)\) such that

$$\begin{aligned}{}[u^H, v^H] = W-\sqrt{-1}JW. \end{aligned}$$

In particular, \(W, JW \in HY\). Thus, \(W \in HY \cap JHY = H^HY\). Let \(X^H \in H^HY\) be a lift of \(X \in TY_G\) such that \(X^H=W\). Then we have

$$\begin{aligned}{}[u, v] = \pi _*[u^H, v^H] = \pi _*(X^H -\sqrt{-1}JX^H) = X - \sqrt{-1}J_GX \in T^{1,0}Y_G, \end{aligned}$$

i.e. we have \([C^\infty (Y_G, T^{1,0}Y_G), C^\infty (Y_G, T^{1,0}Y_G)] \subset C^\infty (Y_G, T^{1,0}Y_G)\). Therefore, \(T^{1,0}Y_G\) is a CR structure of \(Y_G\). \(\square \)

3 G-invariant Szegő kernel asymptotics

In this section, we will establish asymptotic expansion for the G-invariant Szegő kernel. We first review some known results for Szegő kernel.

3.1 Szegő kernel asymptotics

In this subsection, we don’t assume that our CR manifold admits a compact Lie group action but we still assume that the Levi form is non-degenerate of constant signature \((n_-,n_+)\). The Hermitian metric \(\langle \,\cdot \,|\,\cdot \,\rangle \) on \(\mathbb {C}TX\) induces, by duality, a Hermitian metric on \(\mathbb {C}T^*X\) and also on the bundles of (0, q) forms \(T^{*0,q}X\), \(q=0,1,\ldots ,n\). We shall also denote all these induced metrics by \(\langle \,\cdot \,|\,\cdot \,\rangle \). For \(u\in T^{*0,q}X\), we write \(|u|^2:=\langle \,u\,|\,u\,\rangle \). Let \(D\subset X\) be an open set. Let \(\Omega ^{0,q}(D)\) denote the space of smooth sections of \(T^{*0,q}X\) over D and let \(\Omega ^{0,q}_0(D)\) be the subspace of \(\Omega ^{0,q}(D)\) whose elements have compact support in D.

Let

$$\begin{aligned} \overline{\partial }_b:\Omega ^{0,q}(X)\rightarrow \Omega ^{0,q+1}(X) \end{aligned}$$

be the tangential Cauchy-Riemann operator. Let dv(x) be the volume form induced by the Hermitian metric \(\langle \,\cdot \,|\,\cdot \,\rangle \). The natural global \(L^2\) inner product \((\,\cdot \,|\,\cdot \,)\) on \(\Omega ^{0,q}(X)\) induced by dv(x) and \(\langle \,\cdot \,|\,\cdot \,\rangle \) is given by

$$\begin{aligned} (\,u\,|\,v\,):=\int _X\langle \,u(x)\,|\,v(x)\,\rangle \, dv(x)\,,\quad u,v\in \Omega ^{0,q}(X)\,. \end{aligned}$$

We denote by \(L^2_{(0,q)}(X)\) the completion of \(\Omega ^{0,q}(X)\) with respect to \((\,\cdot \,|\,\cdot \,)\). We write \(L^2(X):=L^2_{(0,0)}(X)\). We extend \((\,\cdot \,|\,\cdot \,)\) to \(L^2_{(0,q)}(X)\) in the standard way. For \(f\in L^2_{(0,q)}(X)\), we denote \(\left\| f\right\| ^2:=(\,f\,|\,f\,)\). We extend \(\overline{\partial }_{b}\) to \(L^2_{(0,r)}(X)\), \(r=0,1,\ldots ,n\), by

$$\begin{aligned} \overline{\partial }_{b}:\mathrm{Dom}\,\overline{\partial }_{b}\subset L^2_{(0,r)}(X)\rightarrow L^2_{(0,r+1)}(X)\,, \end{aligned}$$

where \(\mathrm{Dom}\,\overline{\partial }_{b}:=\{u\in L^2_{(0,r)}(X);\, \overline{\partial }_{b}u\in L^2_{(0,r+1)}(X)\}\) and, for any \(u\in L^2_{(0,r)}(X)\), \(\overline{\partial }_{b} u\) is defined in the sense of distributions. We also write

$$\begin{aligned} \overline{\partial }^{*}_{b}:\mathrm{Dom}\,\overline{\partial }^{*}_{b}\subset L^2_{(0,r+1)}(X)\rightarrow L^2_{(0,r)}(X) \end{aligned}$$

to denote the Hilbert space adjoint of \(\overline{\partial }_{b}\) in the \(L^2\) space with respect to \((\,\cdot \,|\,\cdot \, )\). Let \(\Box ^{(q)}_{b}\) denote the (Gaffney extension) of the Kohn Laplacian given by

$$\begin{aligned} \begin{array}{c} \mathrm{Dom}\,\Box ^{(q)}_{b}=\Big \{s\in L^2_{(0,q)}(X);\, s\in \mathrm{Dom}\,\overline{\partial }_{b}\cap \mathrm{Dom}\,\overline{\partial }^{*}_{b},\, \overline{\partial }_{b}s\in \mathrm{Dom}\,\overline{\partial }^{*}_{b}, \, \overline{\partial }^{*}_{b}s\in \mathrm{Dom}\,\overline{\partial }_{b}\Big \}\,,\\ \Box ^{(q)}_{b}s=\overline{\partial }_{b}\overline{\partial }^{*}_{b}s+\overline{\partial }^{*}_{b}\overline{\partial }_{b}s \,\,\text {for } s\in \mathrm{Dom}\,\Box ^{(q)}_{b}\,. \end{array} \end{aligned}$$
(3.1)

By a result of Gaffney, for every \(q=0,1,\ldots ,n\), \(\Box ^{(q)}_{b}\) is a positive self-adjoint operator (see [19, Proposition 3.1.2]). That is, \(\Box ^{(q)}_{b}\) is self-adjoint and the spectrum of \(\Box ^{(q)}_{b}\) is contained in \(\overline{\mathbb {R}}_+\), \(q=0,1,\ldots ,n\). Let

$$\begin{aligned} S^{(q)}:L^2_{(0,q)}(X)\rightarrow \mathrm{Ker}\,\Box ^{(q)}_b \end{aligned}$$
(3.2)

be the orthogonal projections with respect to the \(L^2\) inner product \((\,\cdot \,|\,\cdot \,)\) and let

$$\begin{aligned} S^{(q)}(x,y)\in D'(X\times X,T^{*0,q}X\boxtimes (T^{*0,q}X)^*) \end{aligned}$$

denote the distribution kernel of \(S^{(q)}\).

We recall Hörmander symbol space. Let \(D\subset X\) be a local coordinate patch with local coordinates \(x=(x_1,\ldots ,x_{2n+1})\).

Definition 3.1

For \(m\in \mathbb {R}\), \(S^m_{1,0}(D\times D\times \mathbb {R}_+,T^{*0,q}X\boxtimes (T^{*0,q}X)^*)\) is the space of all \(a(x,y,t)\in C^\infty (D\times D\times \mathbb {R}_+,T^{*0,q}X\boxtimes (T^{*0,q}X)^*)\) such that, for all compact \(K\Subset D\times D\) and all \(\alpha , \beta \in \mathbb {N}^{2n+1}_0\), \(\gamma \in \mathbb {N}_0\), there is a constant \(C_{\alpha ,\beta ,\gamma }>0\) such that

$$\begin{aligned} |\partial ^\alpha _x\partial ^\beta _y\partial ^\gamma _t a(x,y,t)|\le C_{\alpha ,\beta ,\gamma }(1+|t|)^{m-\gamma },\ \ \forall (x,y,t)\in K\times \mathbb {R}_+,\ \ t\ge 1. \end{aligned}$$

Put

$$\begin{aligned} S^{-\infty }(D\times D\times \mathbb {R}_+,T^{*0,q}X\boxtimes (T^{*0,q}X)^*) :=\bigcap _{m\in \mathbb {R}}S^m_{1,0}(D\times D\times \mathbb {R}_+,T^{*0,q}X\boxtimes (T^{*0,q}X)^*). \end{aligned}$$

Let \(a_j\in S^{m_j}_{1,0}(D\times D\times \mathbb {R}_+,T^{*0,q}X\boxtimes (T^{*0,q}X)^*)\), \(j=0,1,2,\ldots \) with \(m_j\rightarrow -\infty \), as \(j\rightarrow \infty \). Then there exists \(a\in S^{m_0}_{1,0}(D\times D\times \mathbb {R}_+,T^{*0,q}X\boxtimes (T^{*0,q}X)^*)\) unique modulo \(S^{-\infty }\), such that \(a-\sum ^{k-1}_{j=0}a_j\in S^{m_k}_{1,0}(D\times D\times \mathbb {R}_+,T^{*0,q}X\boxtimes (T^{*0,q}X)^*\big )\) for \(k=0,1,2,\ldots \).

If a and \(a_j\) have the properties above, we write \(a\sim \sum ^{\infty }_{j=0}a_j\) in \(S^{m_0}_{1,0}\big (D\times D\times \mathbb {R}_+,T^{*0,q}X\boxtimes (T^{*0,q}X)^*\big )\). We write

$$\begin{aligned} s(x, y, t)\in S^{m}_{\mathrm{cl}}\,\big (D\times D\times \mathbb {R}_+,T^{*0,q}X\boxtimes (T^{*0,q}X)^*\big ) \end{aligned}$$

if \(s(x, y, t)\in S^{m}_{1,0}\big (D\times D\times \mathbb {R}_+,T^{*0,q}X\boxtimes (T^{*0,q}X)^*\big )\) and

$$\begin{aligned} \begin{array}{lc} &{}s(x, y, t)\sim \sum ^\infty _{j=0}s^j(x, y)t^{m-j}\text { in }S^{m}_{1, 0} \big (D\times D\times \mathbb {R}_+\,,T^{*0,q}X\boxtimes (T^{*0,q}X)^*\big )\,,\\ &{}s^j(x, y)\in C^\infty \big (D\times D,T^{*0,q}X\boxtimes (T^{*0,q}X)^*\big ),\ j\in \mathbb {N}_0. \end{array} \end{aligned}$$

The following was proved in Theorem 4.8 in [15]

Theorem 3.2

Given \(q=0,1,2,\ldots ,n\). Assume that \(q\notin \left\{ n_-,n_+\right\} \). Then, \(S^{(q)}\equiv 0\) on X.

We have the following (see Theorem 1.2 in [13], Theorem 4.7 in [15] and see also [2] for \(q=0\))

Theorem 3.3

We recall that we work with the assumption that the Levi form is non-degenerate of constant signature \((n_-,n_+)\) on X. Let \(q=n_-\) or \(n_+\). Suppose that \(\Box ^{(q)}_b\) has \(L^2\) closed range. Then, \(S^{(q)}(x,y)\in C^\infty (X\times X\setminus {\mathrm{diag}\,(X\times X)},T^{*0,q}X\boxtimes (T^{*0,q}X)^*)\). Let \(D\subset X\) be any local coordinate patch with local coordinates \(x=(x_1,\ldots ,x_{2n+1})\). Then, there exist continuous operators \(S_-, S_+:\Omega ^{0,q}_0(D)\rightarrow D'(D,T^{*0,q}X)\) such that

$$\begin{aligned} S^{(q)}\equiv S_-+S_+\ \ \text{ on } D, \end{aligned}$$

and \(S_-(x,y)\), \(S_+(x,y)\) satisfy

$$\begin{aligned} S_{\mp }(x, y)\equiv \int ^{\infty }_{0}e^{i\varphi _{\mp }(x, y)t}s_{\mp }(x, y, t)dt\ \ \text{ on } D, \end{aligned}$$

with

$$\begin{aligned} \begin{array}{ll} &{}s_-(x, y, t), s_+(x,y,t)\in S^{n}_{\mathrm{cl}}\,(D\times D\times \mathbb {R}_+,T^{*0,q}X\boxtimes (T^{*0,q}X)^*), \\ &{}s_-(x,y,t)=0\ \ \text{ if } q\ne n_-,\ \ s_+(x,y,t)=0\ \ \text{ if } q\ne n_+,\\ &{}s^0_-(x,x)\ne 0,\ \ \forall x\in D,\ \ s^0_+(x,x)\ne 0,\ \ \forall x\in D, \end{array} \end{aligned}$$
(3.3)

and the phase functions \(\varphi _-\), \(\varphi _+\) satisfy

$$\begin{aligned} \begin{array}{ll} &{}\varphi _+(x,y), \varphi _-\in C^\infty (D\times D),\ \ \mathrm{Im}\,\varphi _-(x, y)\ge 0,\\ &{}\varphi _-(x, x)=0,\ \ \varphi _-(x, y)\ne 0\ \ \text{ if }\ \ x\ne y,\\ &{}d_x\varphi _-(x, y)\big |_{x=y}=-\omega _0(x), \ \ d_y\varphi _-(x, y)\big |_{x=y}=\omega _0(x), \\ &{}\varphi _-(x, y)=-\overline{\varphi }_-(y, x), \ \ -\overline{\varphi }_+(x, y)=\varphi _-(x,y). \end{array} \end{aligned}$$

Remark 3.4

It is well-known that for a strictly pseudoconvec CR manifold of dimension 3, \(\Box ^{(0)}_b\) does not have \(L^2\) closed range in general (see [28]). Kohn [16] proved that if \(q=n_-=n_+\) or \(|n_--n_+|>1\) then \(\Box ^{(q)}_b\) has \(L^2\) closed range.

The following result describes the phase function in local coordinates (see chapter 8 of part I in [13])

Theorem 3.5

For a given point \(p\in X\), let \(\{W_j\}_{j=1}^{n}\) be an orthonormal frame of \(T^{1, 0}X\) in a neighborhood of p such that the Levi form is diagonal at p, i.e. \(\mathcal {L}_{x_{0}}(W_{j},\overline{W}_{s})=\delta _{j,s}\mu _{j}\), \(j,s=1,\ldots ,n\). We take local coordinates \(x=(x_1,\ldots ,x_{2n+1})\), \(z_j=x_j+ix_{d+j}\), \(j=1,\ldots ,d\), \(z_j=x_{2j-1}+ix_{2j}\), \(j=d+1,\ldots ,n\), defined on some neighborhood of p such that \(\omega _0(p)=dx_{2n+1}\), \(x(p)=0\), and, for some \(c_j\in \mathbb {C}\), \(j=1,\ldots ,n\) ,

$$\begin{aligned} W_j=\frac{\partial }{\partial z_j}-i\mu _j\overline{z}_j\frac{\partial }{\partial x_{2n+1}}- c_jx_{2n+1}\frac{\partial }{\partial x_{2n+1}}+\sum ^{2n}_{k=1}a_{j,k}(x)\frac{\partial }{\partial x_k}+O(|x|^2),\ j=1,\ldots ,n\,, \end{aligned}$$
(3.4)

where \(a_{j,k}(x)\in C^\infty \), \(a_{j,k}(x)=O(|x|)\), for every \(j=1,\ldots ,n\), \(k=1,\ldots ,2n\). Set \(y=(y_1,\ldots ,y_{2n+1})\), \(w_j=y_j+iy_{d+j}\), \(j=1,\ldots ,d\), \(w_j=y_{2j-1}+iy_{2j}\), \(j=d+1,\ldots ,n\). Then, for \(\varphi _-\) in Theorem 3.3, we have

$$\begin{aligned} \mathrm{Im}\,\varphi _-(x,y)\ge c\sum ^{2n}_{j=1}|x_j-y_j|^2,\ \ c>0, \end{aligned}$$
(3.5)

in some neighbourhood of (0, 0) and

$$\begin{aligned} \begin{array}{lll} \varphi _-(x, y)=-x_{2n+1}+y_{2n+1}+i\sum ^{n}_{j=1}|\mu _j||z_j-w_j|^2 +\sum ^{n}_{j=1}\Bigr (i\mu _j(\overline{z}_jw_j-z_j\overline{w}_j)\\ \ +c_j(-z_jx_{2n+1}+w_jy_{2n+1})+\overline{c}_j(-\overline{z}_jx_{2n+1}+\overline{w}_jy_{2n+1})\Bigr )\\ +(x_{2n+1}-y_{2n+1})f(x, y) +O(|(x, y)|^3), \end{array} \end{aligned}$$
(3.6)

where f is smooth and satisfies \(f(0,0)=0\), \(f(x, y)=\overline{f}(y, x)\).

The following formula for the leading term \(s^0_-\) on the diagonal follows from [13, §9]. The formula for the leading term \(s^0_+\) on the diagonal follows similarly.

Theorem 3.6

We assume that the Levi form is non-degenerate of constant signature \((n_-,n_+)\) at each point of X. Suppose that \(\Box ^{(q)}_b\) has \(L^2\) closed range. If \(q=n_{\mp }\), then, for the leading term \(s^0_{\mp }(x,y)\) of the expansion (3.3) of \(s_{\mp }(x,y,t)\), we have

$$\begin{aligned} s^0_{\mp }(x_0, x_0)=\frac{1}{2}\pi ^{-n-1}|\det \mathcal {L}_{x_0}|\tau _{x_0,n_{\mp }}\,,\,\,x_0\in D, \end{aligned}$$

where \(\det \mathcal {L}_{x_0}\) is given by (1.6) and \(\tau _{x_0,n_{\mp }}\) is given by (1.7).

3.2 G-invariant Szegő kernel

Fix \(g\in G\). Let \(g^*:\Lambda ^r_x(\mathbb {C}T^*X)\rightarrow \Lambda ^r_{g^{-1}\circ x}(\mathbb {C}T^*X)\) be the pull-back map. Since G preserves J, we have \(g^*:T^{*0,q}_xX\rightarrow T^{*0,q}_{g^{-1}\circ x}X,\ \forall x\in X\). Thus, for \(u\in \Omega ^{0,q}(X)\), we have \(g^*u\in \Omega ^{0,q}(X)\) and we write \((g^*u)(x) := u(g\circ x)\). Put \(\Omega ^{0,q}(X)^G:=\left\{ u\in \Omega ^{0,q}(X);\, g^*u=u,\ \ \forall g\in G\right\} \). Now, we assume that the Hermitian metric \(\langle \,\cdot \,|\,\cdot \,\rangle \) on \(\mathbb {C}TX\) is G-invariant and \(\underline{\mathfrak {g}}\) is orthogonal to \(HY\bigcap JHY\) at every point of Y. The Hermitian metric is G-invariant means that, for any G-invariant vector fields U and V, \(\langle \,U\,|\,V\,\rangle \) is G-invariant. Then the \(L^2\) inner product \((\,\cdot \,|\,\cdot \,)\) on \(\Omega ^{0,q}(X)\) induced by \(\langle \,\cdot \,|\,\cdot \,\rangle \) is G-invariant, that is, \((\,u\,|\,v\,)=(\,g^*u\,|\,g^*v\,)\), for all \(u, v\in \Omega ^{0,q}(X)\), \(g\in G\). Let \(u\in L^2_{(0,q)}(X)\) and let \(g\in G\). Take \(u_j\in \Omega ^{0,q}(X)\), \(j=1,2,\ldots \), with \(u_j\rightarrow u\) in \(L^2_{(0,q)}(X)\) as \(j\rightarrow \infty \). Since \((\,\cdot \,|\,\cdot \,)\) is G-invariant, there is a \(v\in L^2_{(0,q)}(X)\) such that \(v=\lim _{j\rightarrow \infty }g^*u_j\). We define \(g^*u:=v\). It is clear that the definition is well-defined. We have \(g^*:L^2_{(0,q)}(X)\rightarrow L^2_{(0,q)}(X)\). Put \(L^2_{(0,q)}(X)^G:=\left\{ u\in L^2_{(0,q)}(X);\, g^*u=u,\ \forall g\in G\right\} \). It is not difficult to see that \(L^2_{(0,q)}(X)^G\) is the completion of \(\Omega ^{0,q}(X)^G\) with respect to \((\,\cdot \,|\,\cdot \,)\). We write \(L^2(X)^G:=L^2_{(0,0)}(X)^G\). Since G preserves J and \((\,\cdot \,|\,\cdot \,)\) is G-invariant, it is straightforward to see that

$$\begin{aligned} \begin{array}{ll} &{}g^*\overline{\partial }_b=\overline{\partial }_bg^*\ \ \text{ on } \mathrm{Dom}\,\overline{\partial }_b, \quad g^*\overline{\partial }^*_b=\overline{\partial }^*_bg^*\ \ \text{ on } \mathrm{Dom}\,\overline{\partial }^*_b,\\ &{}g^*\Box ^{(q)}_b=\Box ^{(q)}_bg^*\ \ \text{ on } \mathrm{Dom}\,\Box ^{(q)}_b. \end{array} \end{aligned}$$

Put \((\mathrm{Ker}\,\Box ^{(q)}_b)^G:=\mathrm{Ker}\,\Box ^{(q)}_b\bigcap L^2_{(0,q)}(X)^G\). The G-invariant Szegő projection is the orthogonal projection \(S^{(q)}_G:L^2_{(0,q)}(X)\rightarrow (\mathrm{Ker}\,\Box ^{(q)}_b)^G\) with respect to \((\,\cdot \,|\,\cdot \,)\). Let \(S^{(q)}_G(x,y)\in D'(X\times X,T^{*0,q}X\boxtimes (T^{*0,q}X)^*)\) be the distribution kernel of \(S^G\). Let \(d\mu \) be a Haar measure on G so that \(|G|_{d\mu }:=\int _Gd\mu =1\).Then,

$$\begin{aligned} S^{(q)}_G(x,y)=\int _GS^{(q)}(x,g\circ y)d\mu (g). \end{aligned}$$
(3.7)

Note that the integral (3.7) is defined in the sense of distribution.

3.3 G-invariant Szegő kernel asymptotics near \(\mu ^{-1}(0)\)

In this section, we will study G-invariant Szegő kernel near \(\mu ^{-1}(0)\).

Let \(e_0\in G\) be the identity element. Let \(v=(v_1,\ldots ,v_d)\) be the local coordinates of G defined in a neighborhood V of \(e_0\) with \(v(e_0)=(0,\ldots ,0)\). From now on, we will identify the element \(e\in V\) with v(e). We first need

Theorem 3.7

Let \(p\in \mu ^{-1}(0)\). There exist local coordinates \(v=(v_1,\ldots ,v_d)\) of G defined in a neighborhood V of \(e_0\) with \(v(e_0)=(0,\ldots ,0)\), local coordinates \(x=(x_1,\ldots ,x_{2n+1})\) of X defined in a neighborhood \(U=U_1\times U_2\) of p with \(0\leftrightarrow p\), where \(U_1\subset \mathbb {R}^d\) is an open set of \(0\in \mathbb {R}^d\), \(U_2\subset \mathbb {R}^{2n+1-d}\) is an open set of \(0\in \mathbb {R}^{2n+1-d} \) and a smooth function \(\gamma =(\gamma _1,\ldots ,\gamma _d)\in C^\infty (U_2,U_1)\) with \(\gamma (0)=0\in \mathbb {R}^d\) such that

$$\begin{aligned}&\begin{array}{ll} &{}(v_1,\ldots ,v_d)\circ (\gamma (x_{d+1},\ldots ,x_{2n+1}),x_{d+1},\ldots ,x_{2n+1})\\ &{}=(v_1+\gamma _1(x_{d+1},\ldots ,x_{2n+1}),\ldots ,v_d+\gamma _d(x_{d+1},\ldots ,x_{2n+1}),x_{d+1},\ldots ,x_{2n+1}),\\ &{}\forall (v_1,\ldots ,v_d)\in V,\ \ \forall (x_{d+1},\ldots ,x_{2n+1})\in U_2, \end{array} \qquad \end{aligned}$$
(3.8)
$$\begin{aligned}&\begin{array}{ll} &{}\underline{\mathfrak {g}}=\mathrm{span}\,\left\{ \frac{\partial }{\partial x_1},\ldots ,\frac{\partial }{\partial x_d}\right\} ,\\ &{} \mu ^{-1}(0)\bigcap U=\left\{ x_{d+1}=\cdots =x_{2d}=0\right\} ,\\ &{}\text{ On } \mu ^{-1}(0)\bigcap U\text{, } \text{ we } \text{ have } J(\frac{\partial }{\partial x_j})=\frac{\partial }{\partial x_{d+j}}+a_j(x)\frac{\partial }{\partial x_{2n+1}}\text{, } j=1,2,\ldots ,d, \end{array} \end{aligned}$$
(3.9)

where \(a_j(x)\) is a smooth function on \(\mu ^{-1}(0)\bigcap U\), independent of \(x_1,\ldots ,x_{2d}\), \(x_{2n+1}\) and \(a_j(0)=0\), \(j=1,\ldots ,d\),

$$\begin{aligned} \begin{array}{ll} &{}T^{1,0}_pX=\mathrm{span}\,\left\{ Z_1,\ldots ,Z_n\right\} ,\\ &{}Z_j=\frac{1}{2}(\frac{\partial }{\partial x_j}-i\frac{\partial }{\partial x_{d+j}})(p),\ \ j=1,\ldots ,d,\\ &{}Z_j=\frac{1}{2}(\frac{\partial }{\partial x_{2j-1}}-i\frac{\partial }{\partial x_{2j}})(p),\ \ j=d+1,\ldots ,n,\\ &{}\langle \,Z_j\,|\,Z_k\,\rangle =\delta _{j,k},\ \ j,k=1,2,\ldots ,n,\\ &{}\mathcal {L}_p(Z_j, \overline{Z}_k)=\mu _j\delta _{j,k},\ \ j,k=1,2,\ldots ,n \end{array} \end{aligned}$$
(3.10)

and

$$\begin{aligned} \begin{array}{ll} \omega _0(x)&{}=(1+O(|x|))dx_{2n+1}+\sum ^d_{j=1}4\mu _jx_{d+j}dx_j\\ &{}\quad +\sum ^n_{j=d+1}2\mu _jx_{2j}dx_{2j-1}-\sum ^n_{j=d+1}2\mu _jx_{2j-1}dx_{2j}\\ &{}\quad +\sum ^{2n}_{j=d+1}b_jx_{2n+1}dx_j+O(|x|^2), \end{array} \end{aligned}$$
(3.11)

where \(b_{d+1}\in \mathbb {R},\ldots ,b_{2n}\in \mathbb {R}\).

Proof

From the standard proof of Frobenius Theorem, it is not difficult to see that there exist local coordinates \(v=(v_1,\ldots ,v_d)\) of G defined in a neighborhood V of \(e_0\) with \(v(e_0)=(0,\ldots ,0)\) and local coordinates \(x=(x_1,\ldots ,x_{2n+1})\) of X defined in a neighborhood U of p with \(x(p)=0\) such that

$$\begin{aligned} \begin{array}{ll} &{}(v_1,\ldots ,v_d)\circ (0,\ldots ,0,x_{d+1},\ldots ,x_{2n+1})\\ &{}=(v_1,\ldots ,v_d,x_{d+1},\ldots ,x_{2n+1}),\ \ \forall (v_1,\ldots ,v_d)\in V,\ \ \forall (0,\ldots ,0,x_{d+1},\ldots ,x_{2n+1})\in U, \end{array} \end{aligned}$$
(3.12)

and

$$\begin{aligned} \underline{\mathfrak {g}}=\mathrm{span}\,\left\{ \frac{\partial }{\partial x_1},\ldots ,\frac{\partial }{\partial x_d}\right\} . \end{aligned}$$
(3.13)

Since \(p\in \mu ^{-1}(0)\), we have \(\omega _0(p)(\frac{\partial }{\partial x_j}(p))=0\), \(j=1,2,\ldots ,d\), and hence \( \frac{\partial }{\partial x_j}(p)\in H_pX\), \(j=1,2,\ldots ,d\). Consider the linear map

$$\begin{aligned} \begin{array}{rl} R:\underline{\mathfrak {g}}_p&{}\rightarrow \underline{\mathfrak {g}}_p,\\ u&{}\rightarrow Ru,\ \ \langle \,Ru\,|\,v\,\rangle =\langle \,d\omega _0\,,\,Ju\wedge v\,\rangle . \end{array} \end{aligned}$$

Since R is self-adjoint, by using linear transformation in \((x_1,\ldots ,x_d)\), we can take \((x_1,\ldots ,x_d)\) such that, for \(j, k = 1, 2, \ldots , d\),

$$\begin{aligned} \langle \,R\frac{\partial }{\partial x_j}(p)\,|\,\frac{\partial }{\partial x_k}(p)\,\rangle =4\mu _j\delta _{j,k}, \quad \langle \,\frac{\partial }{\partial x_j}(p)\,|\,\frac{\partial }{\partial x_k}(p)\,\rangle =2\delta _{j,k}. \end{aligned}$$
(3.14)

By taking linear transformation in \((v_1,\ldots ,v_d)\), (3.12) still hold.

Let \(\omega _0(\frac{\partial }{\partial x_j})=a_j(x)\in C^\infty (U)\), \(j=1,2,\ldots ,d\). Since \(a_j(x)\) is G-invariant, we have \(\frac{\partial a_j(x)}{\partial x_s}=0\), \(j,s=1,2,\ldots ,d\). By the definition of the moment map, we have

$$\begin{aligned} \mu ^{-1}(0)\bigcap U=\left\{ x\in U;\, a_1(x)=\cdots =a_d(x)=0\right\} . \end{aligned}$$

Since p is a regular value of the moment map \(\mu \), the matrix \(\left( \frac{\partial a_j}{\partial x_s}(p)\right) _{1\le j\le d,d+1\le s\le 2n+1}\) is of rank d. We may assume that the matrix \(\left( \frac{\partial a_j}{\partial x_s}(p)\right) _{1\le j\le d,d+1\le s\le 2d}\) is non-singular. Thus, \((x_1,\ldots ,x_d,a_1,\ldots ,a_d,x_{2d+1},\ldots ,x_{2n+1})\) are also local coordinates of X. Hence, we can take \(v=(v_1,\ldots ,v_d)\) and \(x=(x_1,\ldots ,x_{2n+1})\) such that (3.12), (3.13), (3.14) hold and

$$\begin{aligned} \mu ^{-1}(0)\bigcap U=\left\{ x=(x_1,\ldots ,x_{2n+1})\in U;\, x_{d+1}=\cdots =x_{2d}=0\right\} . \end{aligned}$$
(3.15)

On \(\mu ^{-1}(0)\bigcap U\), let

$$\begin{aligned} J(\frac{\partial }{\partial x_j})=b_{j,1}(x)\frac{\partial }{\partial x_1}+\cdots +b_{j,2n+1}(x)\frac{\partial }{\partial x_{2n+1}},\ \ j=1,2,\ldots ,d. \end{aligned}$$

Since we only work on \(\mu ^{-1}(0)\), \(b_{j,k}(x)\) is independent of \(x_{d+1},\ldots ,x_{2d}\), for all \(j=1,\ldots ,d\), \(k=1,\ldots ,2n+1\). Moreover, it is easy to see that \(b_{j,k}(x)\) is also independent of \(x_{1},\ldots ,x_{d}\), for all \(j=1,\ldots ,d\), \(k=1,\ldots ,2n+1\). Let \(\widetilde{x}''=(x_{2d+1},\ldots ,x_{2n+1})\). Hence, \(b_{j,k}(x)=b_{j,k}(\widetilde{x}'')\), \(j=1,\ldots ,d\), \(k=1,\ldots ,2n+1\). We claim that the matrix \(\left( b_{j,k}(\widetilde{x}'')\right) _{1\le j\le d,d+1\le k\le 2d}\) is non-singular near p. If not, it is easy to see that there is a non-zero vector \(u\in J\underline{\mathfrak {g}}\bigcap HY\), where \(Y=\mu ^{-1}(0)\). Let \(u=Jv\), \(v\in \underline{\mathfrak {g}}\). Then, \(v\in \underline{\mathfrak {g}}\bigcap JHY=\underline{\mathfrak {g}}\bigcap \underline{\mathfrak {g}}^{\perp _b}\) (see (2.10)). Since \(\underline{\mathfrak {g}} \cap \underline{\mathfrak {g}}^{\perp _b} = \left\{ 0 \right\} \) on \(\mu ^{-1}(0)\), we deduce that \(v=0\) and we get a contradiction. The claim follows. From the claim, we can use linear transformation in \((x_{d+1},\ldots ,x_{2d})\) (the linear transform depends smoothly on \(\widetilde{x}''\)) and we can take \((x_{d+1},\ldots ,x_{2d})\) such that on \(\mu ^{-1}(0)\),

$$\begin{aligned}&J(\frac{\partial }{\partial x_j})=b_{j,1}(\widetilde{x}'')\frac{\partial }{\partial x_1}+\cdots +b_{j,d}(\widetilde{x}'')\frac{\partial }{\partial x_{d}}+\frac{\partial }{\partial x_{d+j}}\\&+b_{j,2d+1}(\widetilde{x}'')\frac{\partial }{\partial x_{2d+1}}+\cdots +b_{j,2n+1}(\widetilde{x}'')\frac{\partial }{\partial x_{2n+1}}, \end{aligned}$$

where \(j=1,2,\ldots ,d\). Consider the coordinates change:

$$\begin{aligned} \begin{array}{ll} &{}x=(x_1,\ldots ,x_{2n+1})\rightarrow u=(u_1,\ldots ,u_{2n+1}),\\ &{}(x_1,\ldots ,x_{2n+1}) \rightarrow (x_1-\sum ^d_{j=1}b_{j,1}(\widetilde{x}'')x_{d+j},\ldots ,x_d-\sum ^d_{j=1}b_{j,d}(\widetilde{x}'')x_{d+j},x_{d+1},\ldots ,x_{2d},\\ &{}\quad \quad x_{2d+1}-\sum ^d_{j=1}b_{j,2d+1}(\widetilde{x}'')x_{d+j},\ldots ,x_{2n+1}-\sum ^d_{j=1}b_{j,2n+1}(\widetilde{x}'')x_{d+j}). \end{array} \end{aligned}$$

Then,

$$\begin{aligned} \begin{array}{ll} &{}\frac{\partial }{\partial x_j}\rightarrow \frac{\partial }{\partial u_j},\ \ j=1,\ldots ,d,2d+1,\ldots ,2n+1,\\ &{}\frac{\partial }{\partial x_{d+j}}\rightarrow -b_{j,1}\frac{\partial }{\partial u_1}-\cdots -b_{j,d}\frac{\partial }{\partial u_d}+\frac{\partial }{\partial u_{d+j}}\\ &{} -b_{j,2d+1}\frac{\partial }{\partial u_{2d+1}}-\cdots -b_{j,2n+1}\frac{\partial }{\partial u_{2n+1}},\ \ j=1,\ldots ,d. \end{array} \end{aligned}$$

Hence, on \(\mu ^{-1}(0)\bigcap U\), \(J(\frac{\partial }{\partial x_j})\rightarrow \frac{\partial }{\partial u_{d+j}}\), \(j=1,\ldots ,d\). Thus, we can take \(v=(v_1,\ldots ,v_d)\) and \(x=(x_1,\ldots ,x_{2n+1})\) such that (3.8), (3.13), (3.14), (3.15) hold and on \(\mu ^{-1}(0)\bigcap U\),

$$\begin{aligned} J(\frac{\partial }{\partial x_j})=\frac{\partial }{\partial x_{d+j}},\ \ j=1,2,\ldots ,d. \end{aligned}$$

Let \(Z_j=\frac{1}{2}(\frac{\partial }{\partial x_j}-i\frac{\partial }{\partial x_{d+j}})(p)\in T^{1,0}_pX\), \(j=1,\ldots ,d\). From (3.14), we can check that

$$\begin{aligned} \mathcal {L}_p(Z_j,\overline{Z}_k)=\mu _j\delta _{j,k}, \quad \langle \,Z_j\,|\,Z_k\,\rangle =\delta _{j,k}, \quad j, k = 1, \ldots , d. \end{aligned}$$

Since \(\underline{\mathfrak {g}}_p\) is orthogonal to \(H_pY\bigcap JH_pY\) and \(H_pY\bigcap JH_pY\subset \underline{\mathfrak {g}}^{\perp _b}_p\), we can find an orthonormal frame \(\left\{ Z_1,\ldots ,Z_d,V_1,\ldots ,V_{n-d}\right\} \) for \(T^{1,0}_pX\) such that the Levi form \(\mathcal {L}_p\) is diagonalized with respect to \(Z_1,\ldots ,Z_d,V_1,\ldots ,V_{n-d}\), where \(V_1\in \mathbb {C}H_pY\bigcap J\mathbb {C}H_pY ,\ldots ,V_{n-d}\in \mathbb {C}H_pY\bigcap J\mathbb {C}H_pY\). Write

$$\begin{aligned} \mathrm{Re}\,V_j=\sum ^{2n+1}_{k=1}\alpha _{j,k}\frac{\partial }{\partial x_k},\ \ \mathrm{Im}\,V_j=\sum ^{2n+1}_{k=1}\beta _{j,k}\frac{\partial }{\partial x_k},\ \ j=1,\ldots ,n-d. \end{aligned}$$

We claim that \(\alpha _{j,k}=\beta _{j,k}=0\), for all \(k=d+1,\ldots ,2d\), \(j=1,\ldots ,n-d\). Fix \(j=1,\ldots ,n-d\). Since \(\mathrm{Re}\,V_j\in \underline{\mathfrak {g}}^{\perp _b}_p\) and \(\mathrm{span}\,\left\{ \frac{\partial }{\partial x_{d+1}},\ldots ,\frac{\partial }{\partial x_{2d}}\right\} \in \underline{\mathfrak {g}}_p^{\perp _b}\), we conclude that

$$\begin{aligned} \sum ^d_{k=1}\alpha _{j,k}\frac{\partial }{\partial x_k}+\sum ^{2n+1}_{k=2d+1}\alpha _{j,k}\frac{\partial }{\partial x_k}\in \underline{\mathfrak {g}}^{\perp _b}_p\bigcap H_pY. \end{aligned}$$
(3.16)

From (2.10) and (3.16), we deduce that

$$\begin{aligned} \sum ^d_{k=1}\alpha _{j,k}\frac{\partial }{\partial x_k}+\sum ^{2n+1}_{k=2d+1}\alpha _{j,k}\frac{\partial }{\partial x_k}\in JH_pY\bigcap H_pY=\underline{\mathfrak {g}}^{\perp _b}_p\bigcap H_pY \end{aligned}$$

and hence

$$\begin{aligned} J\Bigr (\sum ^d_{k=1}\alpha _{j,k}\frac{\partial }{\partial x_k}+\sum ^{2n+1}_{k=2d+1}\alpha _{j,k}\frac{\partial }{\partial x_k}\Bigr )\in \underline{\mathfrak {g}}^{\perp _b}_p\bigcap H_pY. \end{aligned}$$
(3.17)

From (3.17) and notice that \(J(\mathrm{Re}\,V_j)\in \underline{\mathfrak {g}}^{\perp _b}_p\), we deduce that

$$\begin{aligned} J(\sum ^{2d}_{k=d+1}\alpha _{j,k}\frac{\partial }{\partial x_k})\in \underline{\mathfrak {g}}_p\bigcap \underline{\mathfrak {g}}^{\perp _b}_p=\left\{ 0\right\} . \end{aligned}$$

Thus, \(\alpha _{j,k}=0\), for all \(k=d+1,\ldots ,2d\), \(j=1,\ldots ,n-d\). Similarly, we can repeat the procedure above and deduce that \(\beta _{j,k}=0\), for all \(k=d+1,\ldots ,2d\), \(j=1,\ldots ,n-d\).

Since \(\mathrm{span}\,\left\{ \mathrm{Re}\,V_j, \mathrm{Im}\,V_j;\, j=1,\ldots ,n-d\right\} \) is transversal to \(\underline{\mathfrak {g}}_p\oplus J\underline{\mathfrak {g}}_p\), we can take linear transformation in \((x_{2d+1},\ldots ,x_{2n+1})\) so that

$$\begin{aligned} \begin{array}{ll} &{}\mathrm{Re}\,V_j=\alpha _{j,1}\frac{\partial }{\partial x_1}+\cdots +\alpha _{j,d}\frac{\partial }{\partial x_{d}}+\frac{\partial }{\partial x_{2j-1+2d}},\ \ j=1,2,\ldots ,n-d,\\ &{}\mathrm{Im}\,V_j=\beta _{j,1}\frac{\partial }{\partial x_1}+\cdots +\beta _{j,d}\frac{\partial }{\partial x_{d}}+\frac{\partial }{\partial x_{2j+2d}},\ \ j=1,2,\ldots ,n-d. \end{array} \end{aligned}$$

Consider the coordinates change:

$$\begin{aligned} \begin{array}{rcl} x=(x_1,\ldots ,x_{2n+1})&{}\rightarrow &{} u=(u_1,\ldots ,u_{2n+1}),\\ (x_1,\ldots ,x_{2n+1})&{} \rightarrow &{} (x_1-\sum ^d_{j=1}\alpha _{j,1}x_{2j-1+2d}-\sum ^{d}_{j=1}\beta _{j,1}x_{2j+2d},\ldots ,x_d \\ &{} &{} -\sum ^d_{j=1}\alpha _{j,d}x_{2j-1+2d} - \sum ^{d}_{j=1}\beta _{j,d}x_{2j+2d}, x_{d+1},\ldots , x_{2n+1}) \end{array} \end{aligned}$$

Then,

$$\begin{aligned} \begin{array}{rcl} \frac{\partial }{\partial x_j}&{} \rightarrow &{} \frac{\partial }{\partial u_j},\ \ j=1,\ldots ,2d,\\ \frac{\partial }{\partial x_{2j-1+2d}}&{} \rightarrow &{} -\alpha _{j,1}\frac{\partial }{\partial u_1}-\cdots -\alpha _{j,d}\frac{\partial }{\partial u_d}+\frac{\partial }{\partial u_{2j-1+2d}},\ \ j=1,\ldots ,n-d,\\ \frac{\partial }{\partial x_{2j+2d}}&{} \rightarrow &{} -\beta _{j,1}\frac{\partial }{\partial u_1}-\cdots -\beta _{j,d}\frac{\partial }{\partial u_d}+\frac{\partial }{\partial u_{2j+2d}},\ \ j=1,\ldots ,n-d. \end{array} \end{aligned}$$

Thus, we can take \(v=(v_1,\ldots ,v_d)\) and \(x=(x_1,\ldots ,x_{2n+1})\) such that (3.8), (3.9) and (3.10) hold.

Now, we can take linear transformation in \(x_{2n+1}\) so that \(\omega _0(p)=dx_{2n+1}\). Let \(W_j\), \(j=1,\ldots ,n\) be an orthonormal basis of \(T^{1,0}X\) such that \(W_j(p)=Z_j\), \(j=1,\ldots ,n\), where \(Z_j\in T^{1,0}_pX\), \(j=1,\ldots ,n\), are as in (3.10). Let \(\widetilde{x}=(\widetilde{x}_1,\ldots ,\widetilde{x}_{2n+1})\) be the coordinates as in Theorem 3.5. It is easy to see that

$$\begin{aligned} \begin{array}{ll} &{}\widetilde{x}_j=x_j+a_jx_{2n+1}+h_j(x),\ \ h_j(x)=O(|x|^2),\ \ a_j\in \mathbb {R},\ \ j=1,2,\ldots ,2n,\\ &{}\widetilde{x}_{2n+1}=x_{2n+1}+h_{2n+1}(x),\ \ h_{2n+1}(x)=O(|x|^2). \end{array} \end{aligned}$$
(3.18)

We may change \(x_{2n+1}\) be \(x_{2n+1}+h_{2n+1}(0,\ldots ,0,x_{d+1},\ldots ,x_{2n},0)\) and we have

$$\begin{aligned} \frac{\partial ^2\widetilde{x}_{2n+1}}{\partial x_j\partial x_k}(p)=0,\ \ j, k=\left\{ d+1,\ldots ,2n\right\} . \end{aligned}$$
(3.19)

Note that when we change \(x_{2n+1}\) to \(x_{2n+1}+h_{2n+1}(0,\ldots ,0,x_{d+1},\ldots ,x_{2n},0)\), \(\frac{\partial }{\partial x_j}\) will change to \(\frac{\partial }{\partial x_j}+\alpha _j(x)\frac{\partial }{\partial x_{2n+1}}\), \(j=d+1,\ldots ,2n\), where \(\alpha _j(x)\) is a smooth function on \(\mu ^{-1}(0)\bigcap U\), independent of \(x_1,\ldots ,x_{d}\), \(x_{2n+1}\) and \(\alpha _j(0)=0\), \(j=d+1,\ldots ,2n\). Hence, on \(\mu ^{-1}(0)\bigcap U\), we have \(J(\frac{\partial }{\partial x_j})=\frac{\partial }{\partial x_{d+j}}+a_j(x)\frac{\partial }{\partial x_{2n+1}}\), \(j=1,2,\ldots ,d\), where \(a_j(x)\) is a smooth function on \(\mu ^{-1}(0)\bigcap U\), independent of \(x_1,\ldots ,x_{2d}\), \(x_{2n+1}\) and \(a_j(0)=0\), \(j=1,\ldots ,d\).

From (3.4) and (3.18), it is straightforward to see that

$$\begin{aligned} \begin{array}{ll} &{}\omega _0(\widetilde{x})=(1+O(|\widetilde{x}|))d\widetilde{x}_{2n+1}+\sum ^d_{j=1}2\mu _j\widetilde{x}_{d+j}d\widetilde{x}_j+\sum ^d_{j=1}(-2\mu _j\widetilde{x}_j)d\widetilde{x}_{d+j}\\ &{}\quad +\sum ^n_{j=d+1}2\mu _jx_{2j}dx_{2j-1}+\sum ^n_{j=d+1}(-2\mu _j\widetilde{x}_{2j-1})d\widetilde{x}_{2j}+\sum ^{2n}_{j=1}\hat{b}_j\widetilde{x}_{2n+1}d\widetilde{x}_j+O(|x|^2)\\ &{}=(1+O(|x|))dx_{2n+1}+\sum ^d_{j=1}(2\mu _jx_{d+j}+\frac{\partial \widetilde{x}_{2n+1}}{\partial x_j})dx_j+\sum ^d_{j=1}(-2\mu _jx_j+\frac{\partial \widetilde{x}_{2n+1}}{\partial x_{d+j}})dx_{d+j}\\ &{}\quad +\sum ^n_{j=d+1}(2\mu _jx_{2j}+\frac{\partial \widetilde{x}_{2n+1}}{\partial x_{2j-1}})dx_{2j-1}+\sum ^n_{j=d+1}(-2\mu _jx_{2j-1}+\frac{\partial \widetilde{x}_{2n+1}}{\partial x_{2j}})dx_{2j}\\ &{}\quad +\sum ^{2n}_{j=1}\widetilde{b}_jx_{2n+1}dx_j+O(|x|^2), \end{array} \end{aligned}$$
(3.20)

where \(\widetilde{b}_j\in \mathbb {R}, \hat{b}_j\in \mathbb {R}\), \(j=1,\ldots ,2n\). Note that \(\omega _0\) is G-invariant. From this observation and (3.20), we deduce that

$$\begin{aligned} \begin{array}{ll} &{}\frac{\partial ^2\widetilde{x}_{2n+1}}{\partial x_j\partial x_k}(p)=0,\ \ j\in \left\{ 1,\ldots ,d\right\} , k\in \left\{ 1,\ldots ,d\right\} \bigcup \left\{ 2d+1,\ldots ,2n\right\} ,\\ &{}\frac{\partial ^2\widetilde{x}_{2n+1}}{\partial x_{d+j}\partial x_k}(p)=2\mu _j\delta _{j,k},\ \ j, k\in \left\{ 1,\ldots ,d\right\} . \end{array} \end{aligned}$$
(3.21)

From (3.21), (3.20) and (3.19), it is straightforward to see that

$$\begin{aligned} \begin{array}{ll} \omega _0(x)&{}=(1+O(|x|))dx_{2n+1}+\sum ^d_{j=1}4\mu _jx_{d+j}dx_j\\ &{}\quad +\sum ^n_{j=d+1}2\mu _jx_{2j}dx_{2j-1}-\sum ^n_{j=d+1}2\mu _jx_{2j-1}dx_{2j}+\sum ^{2n}_{j=1}b_jx_{2n+1}dx_j+O(|x|^2), \end{array} \end{aligned}$$
(3.22)

where \(b_{1}\in \mathbb {R},\ldots ,b_{2n}\in \mathbb {R}\). Since \(\omega _0(p)(\frac{\partial }{\partial x_j})=0\) on \(x_{d+1}=\cdots =x_{2d}=0\), \(j=1,2,\ldots ,d\), we deduce that \(b_1=\cdots =b_d=0\) and we get (3.11). The theorem follows. \(\square \)

We need

Theorem 3.8

Let \(p\in \mu ^{-1}(0)\) and take local coordinates \(x=(x_1,\ldots ,x_{2n+1})\) of X defined in an open set Uof p with \(0\leftrightarrow p\) such that (3.9), (3.10) and (3.11) hold. Let \(\varphi _-(x,y)\in C^\infty (U\times U)\) be as in Theorem 3.3. Then,

$$\begin{aligned} \begin{array}{ll} \varphi _-(x, y)&{}=-x_{2n+1}+y_{2n+1}-2\sum ^d_{j=1}\mu _jx_jx_{d+j}+2\sum ^d_{j=1}\mu _jy_jy_{d+j} +i\sum ^{n}_{j=1}|\mu _j||z_j-w_j|^2 \\ &{} +\sum ^{n}_{j=1}i\mu _j(\overline{z}_jw_j-z_j\overline{w}_j)+\sum ^d_{j=1}(-\frac{i}{2}b_{d+j})(-z_jx_{2n+1}+w_jy_{2n+1})\\ &{} +\sum ^d_{j=1}(\frac{i}{2}b_{d+j})(-\overline{z}_jx_{2n+1}+\overline{w}_jy_{2n+1})+\sum ^n_{j=d+1}\frac{1}{2}(b_{2j-1}-ib_{2j})(-z_jx_{2n+1}+w_jy_{2n+1})\\ &{} +\sum ^n_{j=d+1}\frac{1}{2}(b_{2j-1}+ib_{2j})(-\overline{z}_jx_{2n+1}+\overline{w}_jy_{2n+1})+(x_{2n+1}-y_{2n+1})f(x, y) +O(|(x, y)|^3), \end{array} \end{aligned}$$
(3.23)

where \(z_j=x_j+ix_{d+j}\), \(j=1,\ldots ,d\), \(z_j=x_{2j-1}+ix_{2j}\), \(j=2d+1,\ldots ,2n\), \(\mu _j\), \(j=1,\ldots ,n\), and \(b_{d+1}\in \mathbb {R},\ldots ,b_{2n}\in \mathbb {R}\) are as in (3.11) and f is smooth and satisfies \(f(0,0)=0\), \(f(x, y)=\overline{f}(y, x)\).

Proof

Let \(\widetilde{x}=(\widetilde{x}_1,\ldots ,\widetilde{x}_{2n+1})\) be the coordinates as in Theorem 3.5. It is easy to see that

$$\begin{aligned} \begin{array}{ll} &{}\widetilde{x}_j=x_j+a_jx_{2n+1}+h_j(x),\ \ h_j(x)=O(|x|^2),\ \ a_j\in \mathbb {R},\ \ j=1,2,\ldots ,2n,\\ &{}\widetilde{x}_{2n+1}=x_{2n+1}+h_{2n+1}(x),\ \ h_{2n+1}(x)=O(|x|^2). \end{array} \end{aligned}$$
(3.24)

From (3.4), it is straightforward to see that

$$\begin{aligned} \begin{array}{ll} &{}\omega _0(\widetilde{x})=(1+O(|\widetilde{x}|))d\widetilde{x}_{2n+1}+\sum ^d_{j=1}2\mu _j\widetilde{x}_{d+j}d\widetilde{x}_j+\sum ^d_{j=1}(-2\mu _j\widetilde{x}_j)d\widetilde{x}_{d+j}\\ &{}\quad +\sum ^n_{j=d+1}2\mu _jx_{2j}dx_{2j-1}+\sum ^n_{j=d+1}(-2\mu _j\widetilde{x}_{2j-1})d\widetilde{x}_{2j}+\sum ^{2n}_{j=1}\hat{b}_j\widetilde{x}_{2n+1}d\widetilde{x}_j+O(|x|^2), \end{array} \end{aligned}$$
(3.25)

where

$$\begin{aligned} \begin{array}{ll} &{}\hat{b}_j=c_j+\overline{c}_j,\ \ j\in \left\{ 1,\ldots ,d\right\} \bigcup \left\{ 2d+1,2d+3,\ldots ,2n-1\right\} ,\\ &{}\hat{b}_j=ic_j-i\overline{c}_j,\ \ j\in \left\{ d+1,\ldots ,2d\right\} \bigcup \left\{ 2d+2,\ldots ,2n\right\} . \end{array} \end{aligned}$$

From (3.25) and (3.11), it is not difficulty to see that (see also (3.20))

$$\begin{aligned} \begin{array}{ll} &{}\frac{\partial ^2\widetilde{x}_{2n+1}}{\partial x_j\partial x_k}(p)=0,\ \ j\in \left\{ 1,\ldots ,d\right\} , k\in \left\{ 1,\ldots ,d\right\} ,\\ &{}\frac{\partial ^2\widetilde{x}_{2n+1}}{\partial x_j\partial x_k}(p)=0,\ \ j\in \left\{ 1,\ldots ,2n\right\} , k\in \left\{ 2d+1,\ldots ,2n\right\} ,\\ &{}\frac{\partial ^2\widetilde{x}_{2n+1}}{\partial x_{d+j}\partial x_k}(p)=2\mu _j\delta _{j,k},\ \ j, k\in \left\{ 1,\ldots ,d\right\} . \end{array} \end{aligned}$$
(3.26)

From (3.24), (3.26) and (3.6), it is straightforward to check that

$$\begin{aligned} \begin{array}{ll} \varphi _-(x, y)&{}=-x_{2n+1}+y_{2n+1}-2\sum ^d_{j=1}\mu _jx_jx_{d+j}+2\sum ^d_{j=1}\mu _jy_jy_{d+j} +i\sum ^{n}_{j=1}|\mu _j||z_j-w_j|^2 \\ &{}+\sum ^{n}_{j=1}i\mu _j(\overline{z}_jw_j-z_j\overline{w}_j)+\sum ^n_{j=1}\beta _j(-z_jx_{2n+1}+w_jy_{2n+1})\\ &{}+\sum ^n_{j=1}\overline{\beta }_j(-\overline{z}_jx_{2n+1}+\overline{w}_jy_{2n+1})+(x_{2n+1}-y_{2n+1})f(x, y) +O(|(x, y)|^3), \end{array} \end{aligned}$$
(3.27)

where \(\beta _j\in \mathbb {C}\), \(j=1,\ldots ,n\) and f is smooth and satisfies \(f(0,0)=0\), \(f(x, y)=\overline{f}(y, x)\). We now determine \(\beta _j\), \(j=1,\ldots ,n\). We can compute that

$$\begin{aligned} \begin{array}{ll} &{}\frac{\partial \varphi _-}{\partial x_j}(x,x)=-4\mu _jx_{d+j}-(\beta _j+\overline{\beta }_j)x_{2n+1}+O(|x|^2),\ \ j=1,\ldots ,d,\\ &{}\frac{\partial \varphi _-}{\partial x_{d+j}}(x,x)=-i(\beta _j-\overline{\beta }_j)x_{2n+1}+O(|x|^2),\ \ j=1,\ldots ,d,\\ &{}\frac{\partial \varphi _-}{\partial x_{2j-1}}(x,x)=-2\mu _jx_{2j}-(\beta _j+\overline{\beta }_j)x_{2n+1}+O(|x|^2),\ \ j=d+1,\ldots ,n,\\ &{}\frac{\partial \varphi _-}{\partial x_{2j}}(x,x)=2\mu _jx_{2j-1}+(-i\beta _j+i\overline{\beta }_j)x_{2n+1}+O(|x|^2),\ \ j=d+1,\ldots ,n. \end{array} \end{aligned}$$
(3.28)

Note that \(d_x\varphi _-(x,x)=-\omega _0(x)\). From this observation and (3.11), we deduce that

$$\begin{aligned} \begin{array}{ll} &{}\frac{\partial \varphi _-}{\partial x_j}(x,x)=-4\mu _jx_{d+j}+O(|x|^2),\ \ j=1,\ldots ,d,\\ &{}\frac{\partial \varphi _-}{\partial x_{d+j}}(x,x)=-b_{d+j}x_{2n+1}+O(|x|^2),\ \ j=1,\ldots ,d,\\ &{}\frac{\partial \varphi _-}{\partial x_{2j-1}}(x,x)=-2\mu _jx_{2j}-b_{2j-1}x_{2n+1}+O(|x|^2),\ \ j=d+1,\ldots ,n,\\ &{}\frac{\partial \varphi _-}{\partial x_{2j}}(x,x)=2\mu _jx_{2j-1}-b_{2j}x_{2n+1}+O(|x|^2),\ \ j=d+1,\ldots ,n. \end{array} \end{aligned}$$
(3.29)

From (3.28) and (3.29), we deduce that

$$\begin{aligned} \begin{array}{ll}&\beta _j=-\frac{i}{2}b_{d+j},\ \ j=1,\ldots ,d,\quad \text {and} \quad \beta _j=\frac{1}{2}(b_{2j-1}-ib_{2j}),\ \ j=d+1,\ldots ,n. \end{array} \end{aligned}$$
(3.30)

From (3.30) and (3.27), we get (3.23). \(\square \)

We now work with local coordinates as in Theorem 3.7. From (3.23), we see that near \((p,p)\in U\times U\), we have \(\frac{\partial \varphi _-}{\partial y_{2n+1}}\ne 0\). Using the Malgrange preparation theorem [12, Th. 7.5.7], we have

$$\begin{aligned} \varphi _-(x,y)=g(x,y)(y_{2n+1}+\hat{\varphi }_-(x,\mathring{y})) \end{aligned}$$
(3.31)

in some neighborhood of (pp), where \(\mathring{y}=(y_1,\ldots ,y_{2n})\), \(g, \hat{\varphi }_-\in C^\infty \). Since \(\mathrm{Im}\,\varphi _-\ge 0\), it is not difficult to see that \(\mathrm{Im}\,\hat{\varphi }_-\ge 0\) in some neighborhood of (pp). We may take U small enough so that (3.31) holds and \(\mathrm{Im}\,\hat{\varphi }_-\ge 0\) on \(U\times U\). From [25, Th. 4.2], we see that since \(\varphi _-(x,y)\) and \(\hat{\varphi }_-(x,y)\) satisfy (3.31), \(\varphi _-(x,y)t\) and \((y_{2n+1}+\hat{\varphi }_-(x,\mathring{y}))t\) are equivalent in the sense of Melin–Sjöstrand. More precisely, for any \(k\in \mathbb {R}\) and any \(b_1(x,y,t)\in S^{k}_{\mathrm{cl}}\,\big (U\times U\times \mathbb {R}_+,T^{*0,q}X\boxtimes (T^{*0,q}X)^*\big )\), we can find \(b_2(x,y,t)\in S^{k}_{\mathrm{cl}}\,\big (U\times U\times \mathbb {R}_+,T^{*0,q}X\boxtimes (T^{*0,q}X)^*\big )\) such that

$$\begin{aligned} \int ^\infty _0e^{i\varphi _-(x,y)t}b_1(x,y,t)dt\equiv \int ^\infty _0e^{i\hat{\varphi }_-(x,y)t}b_2(x,y,t)dt\ \ \text{ on } U \end{aligned}$$

and vise versa. We can replace the phase \(\varphi _-\) by \(y_{2n+1}+\hat{\varphi }_-(x,\mathring{y})\). From now on, we assume that \(\varphi _-(x,y)\) has the form

$$\begin{aligned} \varphi _-(x,y)=y_{2n+1}+\hat{\varphi }_-(x,\mathring{y}). \end{aligned}$$
(3.32)

It is easy to check that \(\varphi _-(x,y)\) satisfies (3.5) and (3.23) with \(f(x,y)=0\).

We now study \(S^{(q)}_G(x,y)\). From Theorem 3.2, we get

Theorem 3.9

Assume that \(q\notin \left\{ n_-,n_+\right\} \).Then, \(S^{(q)}_G\equiv 0\) on X.

Assume that \(q=n_-\) and \(\Box ^{(q)}_b\) has \(L^2\) closed range. Fix \(p\in \mu ^{-1}(0)\) and let \(v=(v_1,\ldots ,v_d)\) and \(x=(x_1,\ldots ,x_{2n+1})\) be the local coordinates of G and X as in Theorem 3.7. Assume that \(d\mu =m(v)dv=m(v_1,\ldots ,v_d)dv_1\cdots dv_d\) on V, where V is an open neighborhood of \(e_0\in G\) as in Theorem 3.7. From (3.7), we have

$$\begin{aligned} S^{(q)}_G(x,y)=\int _G\chi (g)S^{(q)}(x,g\circ y)d\mu (g)+\int _G(1-\chi (g))S^{(q)}(x,g\circ y)d\mu (g), \end{aligned}$$

where \(\chi \in C^\infty _0(V)\), \(\chi =1\) near \(e_0\). Since G is freely on \(\mu ^{-1}(0)\), if U and V are small, there is a constant \(c>0\) such that

$$\begin{aligned} d(x,g\circ y)\ge c,\ \ \forall x, y\in U, g\in \mathrm{Supp}\,(1-\chi ), \end{aligned}$$
(3.33)

where U is an open set of \(p\in \mu ^{-1}(0)\) as in Theorem 3.7. From now on, we take U and V small enough so that (3.33) holds. In view of Theorem 3.3, we see that \(S^{(q)}(x,y)\) is smoothing away from diagonal. From this observation and (3.33), we conclude that \(\int _G(1-\chi (g))S^{(q)}(x,g\circ y)d\mu (g)\equiv 0\) on U and hence

$$\begin{aligned} S^{(q)}_G(x,y)\equiv \int _G\chi (g)S^{(q)}(x,g\circ y)d\mu (g) \text{ on } U. \end{aligned}$$
(3.34)

From Theorem 3.3 and (3.34), we have

$$\begin{aligned} \begin{array}{ll} &{}S^{(q)}_G(x,y)\equiv \hat{S}_{-}(x,y)+\hat{S}_{+}(x,y) \text{ on } U,\\ &{}\hat{S}_{\mp }(x,y)=\int _G\chi (g)S_{\mp }(x,g\circ y)d\mu (g), \end{array} \end{aligned}$$
(3.35)

Write \(x=(x',x'')=(x',\hat{x}'',\widetilde{x}'')\), \(y=(y',y'')=(y',\hat{y}'',\widetilde{y}'')\), where \(\hat{x}''=(x_{d+1},\ldots ,x_{2d})\), \(\hat{y}''=(y_{d+1},\ldots ,y_{2d})\), \(\widetilde{x}''=(x_{2d+1},\ldots ,x_{2n+1})\), \(\widetilde{y}''=(y_{2d+1},\ldots ,y_{2n+1})\). Since \(S^{(q)}_G(x,y)\) is G-invariant, we have \(S^{(q)}_G(x,y)=S^{(q)}_G((0,x''),(\gamma (y''),y''))\), where \(\gamma \in C^\infty (U_2,U_1)\) is as in Theorem 3.7. From this observation and (3.35), we have

$$\begin{aligned} S^{(q)}_G(x,y)\equiv \hat{S}_{-}((0,x''),(\gamma (y''),y''))+\hat{S}_{+}((0,x''),(\gamma (y''),y'')) \text{ on } U. \end{aligned}$$
(3.36)

Write \(\mathring{x}''=(x_{d+1},\ldots ,x_{2n})\), \(\mathring{y}''=(y_{d+1},\ldots ,y_{2n})\) From (3.32), (3.36), Theorem 3.7 and Theorem 3.3, we have

$$\begin{aligned} \begin{array}{ll} &{}\hat{S}_{-}((0,x''),(\gamma (y''),y''))\\ &{}\equiv \int e^{i(y_{2n+1}+\hat{\varphi }_-((0,x''),(v+\gamma (y''),\mathring{y}'')))t}s_-((0,x''),(v+\gamma (y''),y''),t)m(v)dvdt. \end{array} \end{aligned}$$
(3.37)

From (3.23), it is straightforward to see that

$$\begin{aligned} \mathrm{det}\,\Bigr (\left( \frac{\partial ^2\hat{\varphi }_-}{\partial v_k\partial v_j}(p,p)\right) ^d_{j,k=1}\Bigr )=(2i)^d|\mu _1|\cdots |\mu _d|\ne 0. \end{aligned}$$
(3.38)

We pause and introduce some notations. Let W be an open set of \(\mathbb {R}^N\), \(N\in \mathbb {N}\). From now on, we write \(W^\mathbb {C}\) to denote an open set in \(\mathbb {C}^N\) with \(W^\mathbb {C}\bigcap \mathbb {R}^N=W\) and for \(f\in C^\infty (W)\), from now on, we write \(\widetilde{f}\in C^\infty (W^\mathbb {C})\) to denote an almost analytic extension of f (see Section 2 in [25]). Let \(h(x'',y'')\in C^\infty (U\times U,\mathbb {C}^d)\) be the solution of the system

$$\begin{aligned} \frac{\partial \widetilde{\hat{\varphi }_-}}{\partial \widetilde{y}_j}((0,x''),(h(x'',y'')+\gamma (y''),\mathring{y}''))=0,\ \ j=1,2,\ldots ,d, \end{aligned}$$
(3.39)

and let

$$\begin{aligned} \Phi _-(x'',y''):=y_{2n+1}+\widetilde{\hat{\varphi }_-}((0,x''),(h(x'',y'')+\gamma (y''),\mathring{y}'')). \end{aligned}$$
(3.40)

It is known that (see page 147 in [25]) \(\mathrm{Im}\,\Phi _-(x'',y'')\ge 0\). Note that

$$\begin{aligned} \frac{\partial \hat{\varphi }_-}{\partial v_j}|_{\hat{x}''=\hat{y}''=0, \widetilde{x}''=\widetilde{y}'', x'=v+\gamma (y'')=0}=-\langle \,\omega _0(x)\,,\,\frac{\partial }{\partial x_j}\,\rangle =0, \end{aligned}$$

where \(x=(0,(0,\widetilde{x}''))\). We deduce that for \(\hat{x}''=\hat{y}''=0\), \(\widetilde{x}''=\widetilde{y}''\), \(v=-\gamma (y'')\) are real critical points. From this observation, we can calculate that

$$\begin{aligned} d_x\Phi _-|_{x''=y'', \hat{x}''=0}=-f(x'')\omega _0(x),\ \ d_y\Phi _-|_{x''=y'', \hat{x}''=0}=f(x'')\omega _0(x), \end{aligned}$$
(3.41)

where \(x=(0,\widetilde{x}'')\) and \(f\in C^\infty \) is a positive function with \(f(p)=1\). By using stationary phase formula of Melin–Sjöstrand [25], we can carry out the v integral in (3.37) and get

$$\begin{aligned} \hat{S}_{-}((0,x''),(\gamma (y''),y''))\equiv \int e^{i\Phi _-(x'',y'')t}a_-(x'',y'',t)dt\ \ \ \text{ on } U, \end{aligned}$$

where \(a_-(x'',y'',t)\sim \sum ^\infty _{j=0}t^{n-\frac{d}{2}-j}a^0_-(x'',y'')\) in \(S^{n-\frac{d}{2}}_{1,0}(U\times U\times \mathbb {R}_+, T^{*0,q}X\boxtimes (T^{*0,q}X)^*)\),

$$\begin{aligned}&a^j_-(x'',y'')\in C^\infty (U\times U,T^{*0,q}X\boxtimes (T^{*0,q}X)^*),\ \ j=0,1,\ldots , \nonumber \\&a^0_-(p,p)=\frac{1}{2}m(0)\pi ^{-n-1+\frac{d}{2}}|\mu _1|^{\frac{1}{2}}\cdots |\mu _d|^{\frac{1}{2}}|\mu _{d+1}|\cdots |\mu _n|\tau _{p,n_-}. \end{aligned}$$
(3.42)

We now study the property of the phase \(\Phi _-(x'',y'')\). We need the following which is known (see Section 2 in [25])

Theorem 3.10

There exist a constant \(c>0\) and an open set \(\Omega \in \mathbb {R}^d\) such that

$$\begin{aligned} \mathrm{Im}\,\Phi _-(x'',y'')\ge c\inf _{v\in \Omega }\left\{ \mathrm{Im}\,\hat{\varphi }_-((0,x''),(v+\gamma (y''),\mathring{y}''))+|d_v\hat{\varphi }_-((0,x''),(v+\gamma (y''),\mathring{y}''))|^2\right\} , \end{aligned}$$
(3.43)

for all \(((0,x''),(0,y''))\in U\times U\).

We can now prove

Theorem 3.11

If U is small enough, then there is a constant \(c>0\) such that

$$\begin{aligned} \mathrm{Im}\,\Phi _-(x'',y'')\ge c\Bigr (|\hat{x}''|^2+|\hat{y}''|^2+|\mathring{x}''-\mathring{y}''|^2\Bigr ),\ \ \forall ((0,x''),(0,y''))\in U\times U. \end{aligned}$$
(3.44)

Proof

From (3.5), we see that there is a constant \(c_1>0\) such that

$$\begin{aligned} \mathrm{Im}\,\hat{\varphi }_-((0,x''),(v+\gamma (y''),\mathring{y}''))\ge c_1(|v+\gamma (y'')|^2+|\mathring{x}''-\mathring{y}''|^2),\ \ \forall v\in \Omega , \end{aligned}$$
(3.45)

where \(\Omega \) is any open set of \(0\in \mathbb {R}^d\). From (3.45) and (3.43), we conclude that there is a constant \(c_2>0\) such that

$$\begin{aligned} \mathrm{Im}\,\Phi _-(x'',y'')\ge c_2(|\mathring{x}''-\mathring{y}''|^2+|d_{y'}\hat{\varphi }_-((0,x''),(0,\mathring{x}''))|^2). \end{aligned}$$
(3.46)

From (3.23), we see that the matrix

$$\begin{aligned}\left( \frac{\partial ^2\hat{\varphi }_-}{\partial x_j\partial x_k}(p,p)+\frac{\partial ^2\hat{\varphi }_-}{\partial y_j\partial y_k}(p,p)\right) _{1\le k\le d, d+1\le j\le 2d}\end{aligned}$$

is non-singular. From this observation and notice that \(d_{y'}\hat{\varphi }_-((0,x''),(0,\mathring{x}''))|_{\hat{x}''}=0\), we deduce that if U is small enough then there is a constant \(c_3>0\) such that

$$\begin{aligned} |d_{y'}\hat{\varphi }_-((0,x''),(0,x''))|\ge c_3|\hat{x}''|. \end{aligned}$$
(3.47)

From (3.47) and (3.46), the theorem follows. \(\square \)

From now on, we assume that U is small enough so that (3.44) holds.

We now determine the Hessian of \(\Phi _-(x'',y'')\) at (pp). Let \(\hat{h}(x'',y''):=h(x'',y'')+\gamma (y'')\). From (3.39), we have

$$\begin{aligned} \frac{\partial ^2\hat{\varphi }_-}{\partial x_{d+1}\partial y_1}(p,p)+\sum ^d_{j=1}\frac{\partial ^2\hat{\varphi }_-}{\partial y_1\partial y_j}(p,p)\frac{\partial \hat{h}_j}{\partial x_{d+1}}(p,p)=0. \end{aligned}$$
(3.48)

From (3.23), we can calculate that

$$\begin{aligned} \frac{\partial ^2\hat{\varphi }_-}{\partial x_{d+1}\partial y_1}(p,p)=2\mu _1,\ \ \frac{\partial ^2\hat{\varphi }_-}{\partial y_1\partial y_j}(p,p)=2i|\mu _1|\delta _{1,j},\ \ j=1,2,\ldots ,d. \end{aligned}$$
(3.49)

From (3.49) and (3.48), we obtain \(\frac{\partial \hat{h}_1}{\partial x_{d+1}}(p,p)=i\frac{\mu _1}{|\mu _1|}\). We can repeat the procedure above several times and deduce that

$$\begin{aligned} \frac{\partial \hat{h}_j}{\partial x_{d+k}}(p,p)=\frac{\partial \hat{h}_j}{\partial y_{d+k}}(p,p)=i\frac{\mu _j}{|\mu _j|}\delta _{j,k},\ \ j,k=1,2,\ldots ,d. \end{aligned}$$
(3.50)

From (3.50), (3.23), (3.40) and by some straightforward computation (we omit the details), we get

Theorem 3.12

With the notations used above, we have

$$\begin{aligned} \begin{array}{ll} \Phi _-(x'', y'')&{}=-x_{2n+1}+y_{2n+1}+2i\sum ^d_{j=1}|\mu _j|y^2_{d+j}+2i\sum ^d_{j=1}|\mu _j|x^2_{d+j}\\ &{}+i\sum ^{n}_{j=d+1}|\mu _j||z_j-w_j|^2 +\sum ^{n}_{j=d+1}i\mu _j(\overline{z}_jw_j-z_j\overline{w}_j)\\ &{}+\sum ^d_{j=1}(-b_{d+j}x_{d+j}x_{2n+1}+b_{d+j}y_{d+j}y_{2n+1})\\ &{}+\sum ^n_{j=d+1}\frac{1}{2}(b_{2j-1}-ib_{2j})(-z_jx_{2n+1}+w_jy_{2n+1})\\ &{}+\sum ^n_{j=d+1}\frac{1}{2}(b_{2j-1}+ib_{2j})(-\overline{z}_jx_{2n+1}+\overline{w}_jy_{2n+1})\\ &{}+(x_{2n+1}-y_{2n+1})f(x, y) +O(|(x, y)|^3), \end{array} \end{aligned}$$
(3.51)

where \(z_j=x_{2j-1}+ix_{2j}\), \(j=2d+1,\ldots ,2n\), \(\mu _j\), \(j=1,\ldots ,n\), and \(b_{d+1}\in \mathbb {R},\ldots ,b_{2n}\in \mathbb {R}\) are as in (3.11) and f is smooth and satisfies \(f(0,0)=0\), \(f(x, y)=\overline{f}(y, x)\).

We can change \(\Phi _-(x'',y'')\) be \(\Phi _-(x'',y'')\frac{1}{f(x'')}\), where \(f(x'')\) is as in (3.41). Thus,

$$\begin{aligned} d_x\Phi _-|_{x''=y'', \hat{x}''=0}=-\omega _0(x),\ \ d_y\Phi _-|_{x''=y'', \hat{x}''=0}=\omega _0(x), \end{aligned}$$
(3.52)

where \(x=(0,\widetilde{x}'')\). It is clear that \(\Phi _-(x'',y'')\) still satisfies (3.44) and (3.51).

We now determine the leading term \(a_-^0(p,p)\). In view of (3.42), we only need to calculate m(0). Put \(Y_p=\left\{ g\circ p;\, g\in G\right\} \). \(Y_p\) is a d-dimensional submanifold of X. The G-invariant Hermitian metric \(\langle \,\cdot \,|\,\cdot \,\rangle \) induces a volume form \(dv_{Y_p}\) on \(Y_p\). Put

$$\begin{aligned} V_{\mathrm{eff}}\,(p):=\int _{Y_p}dv_{Y_p}. \end{aligned}$$

For \(f(g)\in C^\infty (G)\), let \(\hat{f}(g\circ p):=f(g)\), \(\forall g\in G\). Then, \(\hat{f}\in C^\infty (Y_p)\). Let \(d\hat{\mu }\) be the measure on G given by \(\int _Gfd\hat{\mu }:=\int _{Y_p}\hat{f}dv_{Y_p}\), for all \(f\in C^\infty (G)\). It is not difficult to see that \(d\hat{\mu }\) is a Haar measure and

$$\begin{aligned} \int _Gd\hat{\mu }=V_{\mathrm{eff}}\,(p). \end{aligned}$$
(3.53)

Recall that we work with the local coordinates in Theorem 3.7. In view of (3.10), we see that \(\left\{ \frac{1}{\sqrt{2}}\frac{\partial }{\partial x_1},\ldots ,\frac{1}{\sqrt{2}}\frac{\partial }{\partial x_d}\right\} \) is an orthonormal basis for \(\underline{\mathfrak {g}}_p\). Hence \(m(0)=2^{\frac{d}{2}}\frac{1}{V_{\mathrm{eff}}\,(p)}\). From this observation, (3.53) and (3.42), we get

$$\begin{aligned} a^0_-(p,p)=2^{\frac{d}{2}-1}\frac{1}{V_{\mathrm{eff}}\,(p)}\pi ^{-n-1+\frac{d}{2}}|\mu _1|^{\frac{1}{2}}\cdots |\mu _d|^{\frac{1}{2}}|\mu _{d+1}|\cdots |\mu _n|\tau _{p,n_-}. \end{aligned}$$
(3.54)

Similarly, we can repeat the procedure above and deduce that

$$\begin{aligned} \hat{S}_{+}((0,x''),(\gamma (y''),y''))\equiv \int e^{i\Phi _+(x'',y'')t}a_+(x'',y'',t)dt\ \ \ \text{ on } U, \end{aligned}$$

where \(a_+(x'',y'',t)\sim \sum ^\infty _{j=0}t^{n-\frac{d}{2}-j}a^j_+(x'',y'')\) in \(S^{n-\frac{d}{2}}_{1,0}(U\times U\times \mathbb {R}_+, T^{*0,q}X\boxtimes (T^{*0,q}X)^*)\),

$$\begin{aligned}&a^j_+(x'',y'')\in C^\infty (U\times U,T^{*0,q}X\boxtimes (T^{*0,q}X)^*),\ \ j=0,1,\ldots ,\nonumber \\&a^0_+(p,p)=2^{\frac{d}{2}-1}\frac{1}{V_{\mathrm{eff}}\,(p)}\pi ^{-n-1+\frac{d}{2}}|\mu _1|^{\frac{1}{2}}\cdots |\mu _d|^{\frac{1}{2}}|\mu _{d+1}|\cdots |\mu _n|\tau _{p,n_+}, \end{aligned}$$
(3.55)

and \(\Phi _+(x'',y'')\in C^\infty (U\times U)\), \(\mathrm{Im}\,\Phi _+(x'',y'')\ge 0\), \(-\overline{\Phi }_+(x'',y'')\) satisfies (3.44), (3.51) and (3.52).

Summing up, we get one of the main result of this work

Theorem 3.13

We recall that we work with the assumption that the Levi form is non-degenerate of constant signature \((n_-,n_+)\) on X. Let \(q=n_-\) or \(n_+\). Suppose that \(\Box ^{(q)}_b\) has \(L^2\) closed range. Let \(p\in \mu ^{-1}(0)\) and let \(x=(x_1,\ldots ,x_{2n+1})\) be the local coordinates defined in an open set U of p such that \(x(p)=0\) and (3.8), (3.9), (3.10), (3.11) hold. Write \(x''=(x_{d+1},\ldots ,x_{2n+1})\). Then, there exist continuous operators \(S^G_-, S^G_+:\Omega ^{0,q}_0(U)\rightarrow \Omega ^{0,q}(U)\) such that

$$\begin{aligned} S^{(q)}_G\equiv S^G_-+S^G_+\ \ \text{ on } U, \end{aligned}$$

and \(S^G_-(x,y)\), \(S^G_+(x,y)\) satisfy

$$\begin{aligned} S^G_{\mp }(x, y)\equiv \int ^{\infty }_{0}e^{i\Phi _{\mp }(x'', y'')t}a_{\mp }(x'', y'', t)dt\ \ \text{ on } U, \end{aligned}$$

with

$$\begin{aligned} \begin{array}{ll} &{}a_-(x, y, t), a_+(x,y,t)\in S^{n-\frac{d}{2}}_{\mathrm{cl}}\,(U\times U\times \mathbb {R}_+,T^{*0,q}X\boxtimes (T^{*0,q}X)^*), \\ &{}a^0_-(x,x)\ne 0,\ \ \forall x\in U, \ \ a^0_+(x,x)\ne 0,\ \ \forall x\in U, \end{array} \end{aligned}$$

\(a^0_-(p,p)\) and \(a^0_+(p,p)\) are given by (3.54) and (3.55) respectively, \(\Phi _-(x'',y'')\in C^\infty (U\times U)\) satisfies (3.52), (3.44) and (3.51), \(\Phi _+(x'',y'')\in C^\infty (U\times U)\), \(-\overline{\Phi }_+(x'',y'')\) satisfies (3.52), (3.44) and (3.51).

3.4 G-invariant Szegő kernel asymptotics away \(\mu ^{-1}(0)\)

The goal of this section is to prove the following

Theorem 3.14

Let D be an open set of X with \(D\bigcap \mu ^{-1}(0)=\emptyset \). Then, \(S^{(q)}_G\equiv 0\ \ \text{ on } D\).

We first need

Lemma 3.15

Let \(p\notin \mu ^{-1}(0)\). Then, there are open sets U of p and V of \(e\in G\) such that for any \(\chi \in C^\infty _0(V)\), we have

$$\begin{aligned} \int _GS^{(q)}(x,g\circ y)\chi (g)d\mu (g)\equiv 0\ \ \text{ on } U. \end{aligned}$$
(3.56)

Proof

If \(q\notin \left\{ n_-,n_+\right\} \). By Theorem 3.2, we get (3.56). We may assume that \(q=n_-\). Take local coordinates \(v=(v_1,\ldots ,v_d)\) of G defined in a neighborhood V of \(e_0\) with \(v(e_0)=(0,\ldots ,0)\), local coordinates \(x=(x_1,\ldots ,x_{2n+1})\) of X defined in a neighborhood \(U=U_1\times U_2\) of p with \(0\leftrightarrow p\), where \(U_1\subset \mathbb {R}^d\) is an open set of \(0\in \mathbb {R}^d\), \(U_2\subset \mathbb {R}^{2n+1-d}\) is an open set of \(0\in \mathbb {R}^{2n+1-d}\), such that

$$\begin{aligned} \begin{array}{ll} &{}(v_1,\ldots ,v_d)\circ (\gamma (x_{d+1},\ldots ,x_{2n+1}),x_{d+1},\ldots ,x_{2n+1})\\ &{}=(v_1+\gamma _1(x_{d+1},\ldots ,x_{2n+1}),\ldots ,v_d+\gamma _d(x_{d+1},\ldots ,x_{2n+1}),x_{d+1},\ldots ,x_{2n+1}),\\ &{}\forall (v_1,\ldots ,v_d)\in V,\ \ \forall (x_{d+1},\ldots ,x_{2n+1})\in U_2, \end{array} \end{aligned}$$

and

$$\begin{aligned} \underline{\mathfrak {g}}=\mathrm{span}\,\left\{ \frac{\partial }{\partial x_1},\ldots ,\frac{\partial }{\partial x_d}\right\} , \end{aligned}$$

where \(\gamma =(\gamma _1,\ldots ,\gamma _d)\in C^\infty (U_2,U_1)\) with \(\gamma (0)=0\in \mathbb {R}^d\). Note that we don’t use the local coordinates in Theorem 3.7. It should be notice that G needs not act locally freely on near p, (3.33) need not be true. We can not use off-diagonal expansion for the Szegő kernel to get this lemma. We will use some properties of the phase \(\varphi _-\) and integrations by parts to obtain this lemma. From Theorem 3.3, we have

$$\begin{aligned} \int _GS^{(q)}(x,g\circ y)\chi (g)d\mu (g)\equiv \int _GS_-(x,g\circ y)\chi (g)d\mu (g)+\int _GS_+(x,g\circ y)\chi (g)d\mu (g) \text{ on } U. \end{aligned}$$
(3.57)

From Theorem 3.3, we have

$$\begin{aligned}&\int _GS_-(x,g\circ y)\chi (g)d\mu (g)\\&\equiv \int e^{i(\varphi _-(x,(v+\gamma (y''),y''))t}s_-(x,(v+\gamma (y''),y''),t)\chi (v)m(v)dvdt, \end{aligned}$$

where \(y''=(y_{d+1},\ldots ,y_{2n+1})\), \(m(v)dv=d\mu |_V\). Since \(p\notin \mu ^{-1}(0)\) and notice that \(d_y\varphi _-(x,x)=\omega _0(x,x)\), we deduce that if V and U are small then \(d_v(\varphi _-(x,(v+\gamma (y''),y'')))\ne 0\), for every \(v\in V\), \((x,y)\in U\times U\). Hence, by using integration by parts with respect to v, we get

$$\begin{aligned} \int _GS_-(x,g\circ y)\chi (g)d\mu (g)\equiv 0\ \ \text{ on } U. \end{aligned}$$
(3.58)

Similarly, we have

$$\begin{aligned} \int _GS_+(x,g\circ y)\chi (g)d\mu (g)\equiv 0\ \ \text{ on } U. \end{aligned}$$
(3.59)

From (3.57), (3.58) and (3.59), the lemma follows. \(\square \)

Lemma 3.16

Let \(p\notin \mu ^{-1}(0)\) and let \(h\in G\). We can find open sets U of p and V of h such that for every \(\chi \in C^\infty _0(V)\), we have \(\int _GS^{(q)}(x,g\circ y)\chi (g)d\mu (g)\equiv 0\ \ \text{ on } U\).

Proof

Let U and V be open sets as in Lemma 3.15. Let \(\hat{V}=hV\). Then, \(\hat{V}\) is an open set of G. Let \(\hat{\chi }\in C^\infty _0(\hat{V})\). We have

$$\begin{aligned}&\int _GS^{(q)}(x,g\circ y)\hat{\chi }(g)d\mu (g)=\int _GS^{(q)}(x,h\circ g\circ y)\hat{\chi }(h\circ g)d\mu (g)\nonumber \\&=\int _GS^{(q)}(x,g\circ y)\chi (g)d\mu (g), \end{aligned}$$
(3.60)

where \(\chi (g):=\hat{\chi }(h\circ g)\in C^\infty _0(V)\). From (3.60) and Lemma 3.15, we deduce that

$$\begin{aligned} \int _GS^{(q)}(x,g\circ y)\hat{\chi }(g)d\mu (g)\equiv 0\ \ \text{ on } U. \end{aligned}$$

The lemma follows. \(\square \)

Proof of Theorem 3.14

Fix \(p\in D\). We need to show that \(S^{(q)}_G\) is smoothing near p. Let \(h\in G\). By Lemma 3.16, we can find open sets \(U_h\) of p and \(V_h\) of h such that for every \(\chi \in C^\infty _0(V_h)\), we have

$$\begin{aligned} \int _GS^{(q)}(x,g\circ y)\chi (g)d\mu (g)\equiv 0\ \ \text{ on } U_h. \end{aligned}$$
(3.61)

Since G is compact, we can find open sets \(U_{h_j}\) and \(V_{h_j}\), \(j=1,\ldots ,N\), such that \(G=\bigcup ^N_{j=1}V_{h_j}\). Let \(U=D\bigcap \Bigr (\bigcap ^N_{j=1}U_{h_j}\Bigr )\) and let \(\chi _j\in C^\infty _0(V_{h_j})\), \(j=1,\ldots ,N\), with \(\sum ^N_{j=1}\chi _j=1\) on G. From (3.61), we have

$$\begin{aligned} S^{(q)}_G(x,y)=\int _GS^{(q)}(x,g\circ y)d\mu (g)=\sum ^N_{j=1}\int _GS^{(q)}(x,g\circ y)\chi _j(g)d\mu (g)\equiv 0\ \ \text{ on } U. \end{aligned}$$

The theorem follows. \(\square \)

From Theorems 3.9, 3.13 and 3.14, we get Theorem 1.5.

4 G-invariant Szegő kernel asymptotics on CR manifolds wit \(S^1\) action

Let \((X, T^{1,0}X)\) be a compact CR manifold of dimension \(2n+1\), \(n\ge 1\). We assume that X admits an \(S^1\) action: \(S^1\times X\rightarrow X\). We write \(e^{i\theta }\) to denote the \(S^1\) action. Let \(T\in C^\infty (X, TX)\) be the global real vector field induced by the \(S^1\) action given by \((Tu)(x)=\frac{\partial }{\partial \theta }\left( u(e^{i\theta }\circ x)\right) |_{\theta =0}\), \(u\in C^\infty (X)\). We recall

Definition 4.1

We say that the \(S^1\) action \(e^{i\theta }\) is CR if \([T, C^\infty (X, T^{1,0}X)]\subset C^\infty (X, T^{1,0}X)\) and the \(S^1\) action is transversal if for each \(x\in X\), \(\mathbb {C}T(x)\oplus T_x^{1,0}X\oplus T_x^{0,1}X=\mathbb {C}T_xX\). Moreover, we say that the \(S^1\) action is locally free if \(T\ne 0\) everywhere. It should be mentioned that transversality implies locally free.

We assume now that \((X, T^{1,0}X)\) is a compact connected CR manifold with a transversal CR locally free \(S^1\) action \(e^{i\theta }\) and we let T be the global vector field induced by the \(S^1\) action. Let \(\omega _0\in C^\infty (X,T^*X)\) be the global real one form determined by \(\langle \,\omega _0\,,\,u\,\rangle =0\), for every \(u\in T^{1,0}X\oplus T^{0,1}X\), and \(\langle \,\omega _0\,,\,T\,\rangle =-1\). Note that \(\omega _0\) and T satisfy (2.2). Assume that X admits a compact connected Lie group G action and the Lie group G acts on X preserving \(\omega _0\) and J. We recall that we work with Assumption 1.7.

We now assume that the Hermitian metric \(\langle \,\cdot \,|\,\cdot \,\rangle \) on \(\mathbb {C}TX\) is \(G\times S^1\) invariant. Then the \(L^2\) inner product \((\,\cdot \,|\,\cdot \,)\) on \(\Omega ^{0,q}(X)\) induced by \(\langle \,\cdot \,|\,\cdot \,\rangle \) is \(G\times S^1\)-invariant. We then have

$$\begin{aligned} \begin{array}{ll} &{}Tg^*\overline{\partial }^*_b=g^*T\overline{\partial }^*_b=\overline{\partial }^*_bg^*T=\overline{\partial }^*_bTg^*\ \ \text{ on } \Omega ^{0,q}(X),\ \ \forall g\in G,\\ &{}Tg^*\Box ^{(q)}_b=g^*T\Box ^{(q)}_b=\Box ^{(q)}_bg^*T=\Box ^{(q)}_bTg^*\ \ \text{ on } \Omega ^{0,q}(X),\ \ \forall g\in G. \end{array} \end{aligned}$$

Let \(L^2_{(0,q),m}(X)^G\) be the completion of \(\Omega ^{0,q}_m(X)^G\) with respect to \((\,\cdot \,|\,\cdot \,)\). We write \(L^2_m(X)^G:=L^2_{(0,0),m}(X)^G\). Put \((\mathrm{Ker}\,\Box ^{(q)}_b)^G_m:=(\mathrm{Ker}\,\Box ^{(q)}_b)^G\bigcap L^2_{(0,q),m}(X)^G\). It is not difficult to see that, for every \(m\in \mathbb {Z}\), \((\mathrm{Ker}\,\Box ^{(q)}_b)^G_m\subset \Omega ^{0,q}_m(X)^G\) and \(\mathrm{dim}\,(\mathrm{Ker}\,\Box ^{(q)}_b)^G_m<\infty \). The mth G-invariant Szegő projection is the orthogonal projection \(S^{(q)}_{G,m}:L^2_{(0,q)}(X)\rightarrow (\mathrm{Ker}\,\Box ^{(q)}_b)^G_m\) with respect to \((\,\cdot \,|\,\cdot \,)\). Let \(S^{(q)}_{G,m}(x,y)\in C^\infty (X\times X,T^{*0,q}X\boxtimes (T^{*0,q}X)^*)\) be the distribution kernel of \(S^{(q)}_{G,m}\). We can check that

$$\begin{aligned} S^{(q)}_{G,m}(x,y)=\frac{1}{2\pi }\int ^{\pi }_{-\pi }S^{(q)}_G(x,e^{i\theta }\circ y)e^{im\theta }d\theta . \end{aligned}$$
(4.1)

The goal of this section is to study the asymptotics of \(S^{(q)}_{G,m}\) as \(m\rightarrow +\infty \).

From Theorem 3.14, (4.1) and by using integration by parts several times, we get

Theorem 4.2

Let \(D\subset X\) be an open set with \(D\bigcap \mu ^{-1}(0)=\emptyset \). Then, \(S^{(q)}_{G,m}=O(m^{-\infty })\ \ \text{ on } D\).

We now study \(S^{(q)}_{G,m}\) near \(\mu ^{-1}(0)\). We can repeat the proof of Theorem 3.7 with minor change and get

Theorem 4.3

Let \(p\in \mu ^{-1}(0)\). There exist local coordinates \(v=(v_1,\ldots ,v_d)\) of G defined in a neighborhood V of \(e_0\) with \(v(e_0)=(0,\ldots ,0)\), local coordinates \(x=(x_1,\ldots ,x_{2n+1})\) of X defined in a neighborhood \(U=U_1\times (\hat{U}_2\times ]-2\delta ,2\delta [)\) of p with \(0\leftrightarrow p\), where \(U_1\subset \mathbb {R}^d\) is an open set of \(0\in \mathbb {R}^d\), \(\hat{U}_2\subset \mathbb {R}^{2n-d}\) is an open set of \(0\in \mathbb {R}^{2n-d} \), \(\delta >0\), and a smooth function \(\gamma =(\gamma _1,\ldots ,\gamma _d)\in C^\infty (\hat{U}_2\times ]-2\delta ,2\delta [,U_1)\) with \(\gamma (0)=0\in \mathbb {R}^d\) such that

$$\begin{aligned} \begin{array}{ll} &{}(v_1,\ldots ,v_d)\circ (\gamma (x_{d+1},\ldots ,x_{2n+1}),x_{d+1},\ldots ,x_{2n+1})\\ &{}=(v_1+\gamma _1(x_{d+1},\ldots ,x_{2n+1}),\ldots ,v_d+\gamma _d(x_{d+1},\ldots ,x_{2n+1}),x_{d+1},\ldots ,x_{2n+1}),\\ &{}\forall (v_1,\ldots ,v_d)\in V,\ \ \forall (x_{d+1},\ldots ,x_{2n+1})\in \hat{U}_2\times ]-2\delta ,2\delta [, \end{array} \end{aligned}$$
$$\begin{aligned} \begin{array}{ll} &{}T=-\frac{\partial }{\partial x_{2n+1}},\quad \underline{\mathfrak {g}}=\mathrm{span}\,\left\{ \frac{\partial }{\partial x_1},\ldots ,\frac{\partial }{\partial x_d}\right\} ,\\ &{}\mu ^{-1}(0)\bigcap U=\left\{ x_{d+1}=\cdots =x_{2d}=0\right\} ,\\ &{}\text{ On } \mu ^{-1}(0)\bigcap U, \text{ we } \text{ have } J(\frac{\partial }{\partial x_j})=\frac{\partial }{\partial x_{d+j}}+a_j(x)\frac{\partial }{\partial x_{2n+1}}, j=1,2,\ldots ,d, \end{array} \end{aligned}$$
(4.2)

where \(a_j(x)\) is a smooth function on \(\mu ^{-1}(0)\bigcap U\), independent of \(x_1,\ldots ,x_{2d}\), \(x_{2n+1}\) and \(a_j(0)=0\), \(j=1,\ldots ,d\),

$$\begin{aligned} \begin{array}{ll} &{}T^{1,0}_pX=\mathrm{span}\,\left\{ Z_1,\ldots ,Z_n\right\} ,\\ &{}Z_j=\frac{1}{2}(\frac{\partial }{\partial x_j}-i\frac{\partial }{\partial x_{d+j}})(p),\ \ j=1,\ldots ,d,\\ &{}Z_j=\frac{1}{2}(\frac{\partial }{\partial x_{2j-1}}-i\frac{\partial }{\partial x_{2j}})(p),\ \ j=d+1,\ldots ,n,\\ &{}\langle \,Z_j\,|\,Z_k\,\rangle =\delta _{j,k},\ \ j,k=1,2,\ldots ,n,\\ &{}\mathcal {L}_p(Z_j, \overline{Z}_k)=\mu _j\delta _{j,k},\ \ j,k=1,2,\ldots ,n \end{array} \end{aligned}$$

and

$$\begin{aligned} \begin{array}{rl} \omega _0(x)&{}=(1+O(|x|))dx_{2n+1}+\sum ^d_{j=1}4\mu _jx_{d+j}dx_j\\ &{}\quad +\sum ^n_{j=d+1}2\mu _jx_{2j}dx_{2j-1}-\sum ^n_{j=d+1}2\mu _jx_{2j-1}dx_{2j}+O(|x|^2). \end{array} \end{aligned}$$

Remark 4.4

Let \(p\in \mu ^{-1}(0)\) and let \(x=(x_1,\ldots ,x_{2n+1})\) be the local coordinates as in Theorem 4.3. We can change \(x_{2n+1}\) be \(x_{2n+1}-\sum ^d_{j=1}a_j(x)x_{d+j}\), where \(a_j(x)\), \(j=1,\ldots ,d\), are as in (4.2). With this new local coordinates \(x=(x_1,\ldots ,x_{2n+1})\), on \(\mu ^{-1}(0)\bigcap U\), we have \(J(\frac{\partial }{\partial x_j})=\frac{\partial }{\partial x_{d+j}},\ \ j=1,2,\ldots ,d\). Moreover, it is clear that \(\Phi _-(x,y)+\sum ^d_{j=1}a_j(x)x_{d+j}-\sum ^{d=1}_{j=1}a_j(y)y_{d+j}\) satisfies (1.19). Note that \(a_j(x)\) is a smooth function on \(\mu ^{-1}(0)\bigcap U\), independent of \(x_1,\ldots ,x_{2d}\), \(x_{2n+1}\) and \(a_j(0)=0\), \(j=1,\ldots ,d\).

We now work with local coordinates as in Theorem 4.3. From (3.51), we see that near \((p,p)\in U\times U\), we have \(\frac{\partial \Phi _-}{\partial y_{2n+1}}\ne 0\). Using the Malgrange preparation theorem [12, Th. 7.5.7], we have

$$\begin{aligned} \Phi _-(x,y)=g(x,y)(y_{2n+1}+\hat{\Phi }_-(x'',\mathring{y}'')) \end{aligned}$$
(4.3)

in some neighborhood of (pp), where \(\mathring{y}''=(y_{d+1},\ldots ,y_{2n})\), \(g, \hat{\Phi }_-\in C^\infty \). Since \(\mathrm{Im}\,\Phi _-\ge 0\), it is not difficult to see that \(\mathrm{Im}\,\hat{\Phi }_-\ge 0\) in some neighborhood of (pp). We may take U small enough so that (4.3) holds and \(\mathrm{Im}\,\hat{\Phi }_-\ge 0\) on \(U\times U\). From [25, Th. 4.2], we see that since \(\Phi _-(x,y)\) and \(y_{2n+1}+\hat{\Phi }_-(x'',\mathring{y}'')\) satisfy (4.3), \(\Phi _-(x,y)t\) and \((y_{2n+1}+\hat{\Phi }_-(x'',\mathring{y}''))t\) are equivalent in the sense of Melin–Sjöstrand (see the discussion after (3.31), for the meaning of equivalent in the sense of Melin–Sjöstrand). We can replace the phase \(\Phi _-\) by \(y_{2n+1}+\hat{\Phi }_-(x,\mathring{y''})\). From now on, we assume that

$$\begin{aligned} \Phi _-(x,y)=y_{2n+1}+\hat{\Phi }_-(x'',\mathring{y}''). \end{aligned}$$
(4.4)

It is easy to check that \(\Phi _-(x,y)\) satisfies (3.44) and (3.51) with \(f(x,y)=0\). Similarly, from now on, we assume that

$$\begin{aligned} \Phi _+(x,y)=-y_{2n+1}+\hat{\Phi }_+(x'',\mathring{y}''). \end{aligned}$$
(4.5)

We now study \(S^{(q)}_{G,m}(x,y)\). From Theorem 3.9, we get

Theorem 4.5

Assume that \(q\notin \left\{ n_-,n_+\right\} \). Then, \(S^{(q)}_{G,m}=O(m^{-\infty })\) on X.

Assume that \(q=n_-\). It was proved in Theorem 1.12 in [15] that when X admits a transversal \(S^1\) action, then \(\Box ^{(q)}_b\) has \(L^2\) closed range. Fix \(p\in \mu ^{-1}(0)\) and let \(v=(v_1,\ldots ,v_d)\) and \(x=(x_1,\ldots ,x_{2n+1})\) be the local coordinates of G and X as in Theorem 4.3 and let U and V be open sets as in Theorem 4.3. We take U small enough so that there is a constant \(c>0\) such that

$$\begin{aligned} d(e^{i\theta }\circ g\circ x,y)\ge c,\ \ \forall (x,y)\in U\times U,\ \ \forall g\in G, \theta \in [-\pi ,-\delta ]\bigcup [\delta ,\pi ], \end{aligned}$$
(4.6)

where \(\delta >0\) is as in Theorem 4.3. We will study \(S^{(q)}_{G,m}(x,y)\) in U. From (4.1), we have

$$\begin{aligned} \begin{array}{rlll} S^{(q)}_{G,m}(x,y)&{}=\frac{1}{2\pi }\int ^{\pi }_{-\pi }S^{(q)}_G(x,e^{i\theta }\circ y)e^{im\theta }d\theta \\ &{} =\frac{1}{2\pi }\int ^{\pi }_{-\pi }e^{-imx_{2n+1}+imy_{2n+1}}S^{(q)}_G(\mathring{x},e^{i\theta }\circ \mathring{y})e^{im\theta }d\theta \\ &{}=I+II,\\ &{}I=\frac{1}{2\pi }\int ^{\pi }_{-\pi }e^{-imx_{2n+1}+imy_{2n+1}}\chi (\theta )S^{(q)}_G(\mathring{x},e^{i\theta }\circ \mathring{y})e^{im\theta }d\theta ,\\ &{}II=\frac{1}{2\pi }\int ^{\pi }_{-\pi }e^{-imx_{2n+1}+imy_{2n+1}}(1-\chi (\theta ))S^{(q)}_G(\mathring{x},e^{i\theta }\circ \mathring{y})e^{im\theta }d\theta , \end{array} \end{aligned}$$

where \(\mathring{x}=(x_1,\ldots ,x_{2n},0)\in U\), \(\mathring{y}=(y_1,\ldots ,y_{2n},0)\in U\), \(\chi \in C^\infty _0(]-2\delta ,2\delta [)\), \(\chi =1\) on \([-\delta , \delta ]\). We first study II. We have

$$\begin{aligned} II=\frac{1}{2\pi }\int ^{\pi }_{-\pi }\int _Ge^{-imx_{2n+1}+imy_{2n+1}}(1-\chi (\theta ))S^{(q)}(\mathring{x},e^{i\theta }\circ g\circ \mathring{y})e^{im\theta }d\mu (g)d\theta . \end{aligned}$$
(4.7)

From (4.7), (4.6) and notice that \(S^{(q)}\) is smoothing away from diagonal, we deduce that

$$\begin{aligned} II=O(m^{-\infty }). \end{aligned}$$

We now study I. From Theorem 3.13, (4.1), (4.4) and (4.5), we have

$$\begin{aligned} \begin{array}{ll} &{}I=I_0+I_1,\\ &{}I_0=\frac{1}{2\pi }\int ^\infty _0\int ^{\pi }_{-\pi }e^{-imx_{2n+1}+imy_{2n+1}}\chi (\theta )e^{i(-\theta +\hat{\Phi }_-(\mathring{x}'',\mathring{y}''))t+im\theta }a_-(\mathring{x}'', (\mathring{y}'',-\theta ),t)dtd\theta ,\\ &{}I_1=\frac{1}{2\pi }\int ^\infty _0\int ^{\pi }_{-\pi }e^{-imx_{2n+1}+imy_{2n+1}}\chi (\theta )e^{i(\theta +\hat{\Phi }_+(\mathring{x}'',\mathring{y}''))t+im\theta }a_+(\mathring{x}'', (\mathring{y}'',-\theta ),t)dtd\theta . \end{array} \end{aligned}$$

We first study \(I_1\). From \(\frac{\partial }{\partial \theta }\Bigr (i(\theta +\hat{\Phi }_+(\mathring{x}'',\mathring{y}''))t+im\theta \Bigr )\ne 0\), we can integrate by parts with respect to \(\theta \) several times and deduce that

$$\begin{aligned} I_1=O(m^{-\infty }). \end{aligned}$$

We now study \(I_0\). We have

$$\begin{aligned} I_0=\frac{1}{2\pi }\int ^\infty _0\int ^{\pi }_{-\pi }e^{-imx_{2n+1}+imy_{2n+1}}\chi (\theta )e^{im(-\theta t+\hat{\Phi }_-(\mathring{x}'',\mathring{y}'')t+\theta )}ma_-(\mathring{x}'', (\mathring{y}'',-\theta ),mt)dtd\theta . \end{aligned}$$
(4.8)

We can use the complex stationary phase formula of Melin–Sjöstrand [25, Theorem 2.3] to carry the \(dtd\theta \) integration in (4.8) and get (the calculation is similar as in the proof of Theorem 3.17 in [14], we omit the details)

$$\begin{aligned} \begin{array}{ll} &{}I_0=e^{im\Psi (x,y)}b(\mathring{x}'',\mathring{y}'',m)+O(m^{-\infty }),\\ &{}\Psi (x,y)=\hat{\Phi }_-(\mathring{x}'',\mathring{y}'')-x_{2n+1}+y_{2n+1},\\ &{}b(\mathring{x}'',\mathring{y}'',m)\in S^{n-\frac{d}{2}}_{\mathrm{loc}}\,(1; U\times U, T^{*0,q}X\boxtimes (T^{*0,q}X)^*),\\ &{}b(\mathring{x}'',\mathring{y}'',m)\sim \sum ^\infty _{j=0}m^{n-\frac{d}{2}-j}b_j(\mathring{x}'',\mathring{y}'') \text{ in } S^{n-\frac{d}{2}}_{\mathrm{loc}}\,(1; U\times U, T^{*0,q}X\boxtimes (T^{*0,q}X)^*),\\ &{}b_j(\mathring{x}'',\mathring{y}'')\in C^\infty (U\times U, T^{*0,q}X\boxtimes (T^{*0,q}X)^*),\ \ j=0,1,2,\ldots ,\\ &{}b_0(p,p)=a^0_-(p,p)=2^{\frac{d}{2}-1}\frac{1}{V_{\mathrm{eff}}\,(p)}\pi ^{-n-1+\frac{d}{2}}|\mu _1|^{\frac{1}{2}}\cdots |\mu _d|^{\frac{1}{2}}|\mu _{d+1}|\cdots |\mu _n|\tau _{p,n_-}. \end{array} \end{aligned}$$

Assume that \(q=n_+\ne n_-\). If \(m\rightarrow -\infty \), then the expansion for \(S^{(q)}_{G,m}(x,y)\) as \(m\rightarrow -\infty \) is similar to \(q=n_-\) case. When \(m\rightarrow +\infty \), we can repeat the method above with minor change and deduce that \(S^{(q)}_{G,m}(x,y)=O(m^{-\infty })\) on X. Summing up, we get Theorem 1.8.

5 Equivalent of the phase function \(\Phi _-(x,y)\)

Let \(p\in \mu ^{-1}(0)\) and let U be a small open set of p. We need

Definition 5.1

With the assumptions and notations used in Theorem 3.13, let \(\Phi _1, \Phi _2\in C^\infty (U\times U)\). We assume that \(\Phi _1\) and \(\Phi _2\) satisfy (3.52), (3.51) and (3.44). We say that \(\Phi _1\) and \(\Phi _2\) are equivalent on U if for any \(b_1(x,y,t)\in S^{n-\frac{d}{2}}_{\mathrm{cl}}\,\big (U\times U\times \mathbb {R}_+,T^{*0,q}X\boxtimes (T^{*0,q}X)^*\big )\) we can find \(b_2(x,y,t)\in S^{n-\frac{d}{2}}_{\mathrm{cl}}\,\big (U\times U\times \mathbb {R}_+,T^{*0,q}X\boxtimes (T^{*0,q}X)^*\big )\) such that

$$\begin{aligned} \int ^\infty _0e^{i\Phi _1(x,y)t}b_1(x,y,t)dt\equiv \int ^\infty _0e^{i\Phi _2(x,y)t}b_2(x,y,t)dt\ \ \text{ on } U \end{aligned}$$

and vise versa.

We characterize now the phase \(\Phi _-\).

Theorem 5.2

Let \(\Phi _-(x,y)\in C^\infty (U\times U)\) be as in Theorem 3.13. Let \(\Phi \in C^\infty (U\times U)\). We assume that \(\Phi \) satisfies (3.52), (3.51) and (3.44). The functions \(\Phi \) and \(\Phi _-\) are equivalent on U in the sense of Definition 5.1 if and only if there is a function \(f\in C^\infty (U\times U)\) with \(f(x,x)=1\) such that \(\Phi (x,y)-f(x,y)\Phi _-(x,y)\) vanishes to infinite order at \(\mathrm{diag}\,\Bigr ((\mu ^{-1}(0)\bigcap U)\times (\mu ^{-1}(0)\bigcap U)\Bigr )\).

Proof

The “\(\Leftarrow \)” part follows from global theory of complex Fourier integral operator of Melin–Sjöstrand [25]. We only need to prove the “\(\Rightarrow \)” part. Take local coordinates \(x=(x_1,\ldots ,x_{2n+1})\) defined in some small neighbourhood of p such that \(x(p)=0\) and \(\omega _0(p)=dx_{2n+1}\). Since \(d_y\Phi (x, y)|_{x=y\in \mu ^{-1}(0)}=d_y\Phi _-(x, y)|_{x=y\in \mu ^{-1}(0)}=\omega _0(x)\), we have \(\frac{\partial \Phi }{\partial y_{2n+1}}(p,p)=\frac{\partial \Phi _-}{\partial y_{2n+1}}(p,p)=1\). From this observation and the Malgrange preparation theorem [12, Theorem 7.5.7], we conclude that in some small neighborhood of (pp), we can find \(f(x,y), f_1(x,y)\in C^\infty \) such that

$$\begin{aligned} \Phi _-(x,y)=f(x,y)(y_{2n+1}+h(x,\mathring{y})),\quad \Phi (x,y)=f_1(x,y)(y_{2n+1}+h_1(x,\mathring{y})) \end{aligned}$$
(5.1)

in some small neighborhood of (pp), where \(\mathring{y}=(y_1,\ldots ,y_{2n})\). For simplicity, we assume that (5.1) hold on \(U\times U\). It is clear that \(\Phi _-(x,y)\) and \(y_{2n-1}+h(x,\mathring{y})\) are equivalent in the sense of Definition 5.1, \(\Phi (x,y)\) and \(y_{2n+1}+h_1(x,\mathring{y})\) are equivalent in the sense of Definition 5.1, we may assume that \(\Phi _-(x,y)=y_{2n+1}+h(x,\mathring{y})\) and \(\Phi (x,y)=y_{2n+1}+h_1(x,\mathring{y})\). Fix \(x_0\in \mu ^{-1}(0)\bigcap U\). We are going to prove that \(h(x,\mathring{y})-h_1(x,\mathring{y})\) vanishes to infinite order at \((x_0,x_0)\in (\mu ^{-1}(0)\bigcap U)\times (\mu ^{-1}(0)\bigcap U)\). Take

$$\begin{aligned} b(x,y,t)\sim \sum ^\infty _{j=0}b_j(x,y)t^{n-\frac{d}{2}-j}\in S^{n-\frac{d}{2}}_{\mathrm{cl}}\,\big (U\times U\times \mathbb {R}_+,T^{*0,q}X\boxtimes (T^{*0,q}X)^*\big ) \end{aligned}$$

with \(b_0(x,x)\ne 0\) at each \(x\in U\bigcap \mu ^{-1}(0)\). Since \(\Phi \) and \(\Phi _-\) are equivalent on U in the sense of Definition 5.1, we can find \(a(x,y,t)\in S^{n-\frac{d}{2}}_{\mathrm{cl}}\,\big (U\times U\times \mathbb {R}_+,T^{*0,q}X\boxtimes (T^{*0,q}X)^*\big )\) such that

$$\begin{aligned} \int ^\infty _0e^{i\Phi _-(x,y)t}b(x,y,t)dt\equiv \int ^\infty _0e^{i\Phi (x,y)t}a(x,y,t)dt\ \ \text{ on } U. \end{aligned}$$

Put \(x_0=(x^1_0,x^2_0,\ldots ,x^{2n+1}_0)\) and \(\mathring{x}_0=(x^1_0,\ldots ,x^{2n}_0)\). Take \(\tau \in C^\infty _0(\mathbb {R}^{2n+1})\), \(\tau _1\in C^\infty _0(\mathbb {R}^{2n})\), \(\chi \in C^\infty _0(\mathbb {R})\) so that \(\tau =1\) near \(x_0\), \(\tau _1=1\) near \(\mathring{x}_0\), \(\chi =1\) near \(x^{2n+1}_0\) and \(\mathrm{Supp}\,\tau \Subset U\), \(\mathrm{Supp}\,\tau _1\times \mathrm{Supp}\,\chi \Subset U'\times \mathrm{Supp}\,\chi \Subset U\), where \(U'\) is an open neighborhood of \(\mathring{x}_0\) in \(\mathbb {R}^{2n}\). For each \(k>0\), we consider the distributions

$$\begin{aligned} \begin{array}{ll} &{}A_k:u\mapsto \int ^\infty _0e^{i(y_{2n-1}+h(x,\mathring{y}))t-iky_{2n+1}}\tau (x)b(x,y,t)\tau _1(\mathring{y})\chi (y_{2n+1})u(\mathring{y})dydt,\\ &{}B_k:u\mapsto \int ^\infty _0e^{i(y_{2n+1}+h_1(x,\mathring{y}))t-iky_{2n+1}}\tau (x)a(x,y,t)\tau _1(\mathring{y})\chi (y_{2n+1})u(\mathring{y})dydt, \end{array} \end{aligned}$$

for \(u\in C^\infty _0(U',T^{*0,q}X)\). By using the stationary phase formula of Melin–Sjöstrand [25], we can show that (cf. the proof of [14, Theorem 3.12]) \(A_k\) and \(B_k\) are smoothing operators and

$$\begin{aligned} \begin{array}{ll} &{}A_k(x,\mathring{y})\equiv e^{ikh(x,\mathring{y})}g(x,\mathring{y},k)+O(k^{-\infty }),\\ &{}B_k(x,\mathring{y})\equiv e^{ikh_1(x,\mathring{y})}p(x,\mathring{y},k)+O(k^{-\infty }),\\ &{}g(x,\mathring{y},k), p(x,\mathring{y},k)\in S^{n-\frac{d}{2}}_{\mathrm{loc}}\,(1;U\times U',T^{*0,q}X\boxtimes (T^{*0,q}X)^*),\\ &{}g(x,\mathring{y},k)\sim \sum ^\infty _{j=0}g_j(x,\mathring{y})k^{n-\frac{d}{2}-j} \text{ in } S^{n-\frac{d}{2}}_{\mathrm{loc}}\,(1;U\times U',T^{*0,q}X\boxtimes (T^{*0,q}X)^*),\\ &{}p(x,\mathring{y},k)\sim \sum ^\infty _{j=0}p_j(x,y')k^{n-\frac{d}{2}-j} \text{ in } S^{n-\frac{d}{2}}_{\mathrm{loc}}\,(1;U\times U',T^{*0,q}X\boxtimes (T^{*0,q}X)^*),\\ &{}g_j(x,\mathring{y}), p_j(x,\mathring{y})\in C^\infty (U\times U',T^{*0,q}X\boxtimes (T^{*0,q}X)^*),\ \ j=0,1,\ldots ,\\ &{}g_0(x_0,\mathring{x}_0)\ne 0. \end{array} \end{aligned}$$

Since

$$\begin{aligned} \int ^\infty _0e^{i(y_{2n+1}+h(x,\mathring{y}))t}b(x,y,t)dt-\int ^\infty _0e^{i(y_{2n+1}+h_1(x,\mathring{y}))t}a(x,y,t)dt \end{aligned}$$

is smoothing, by using integration by parts with respect to \(y_{2n+1}\), it is easy to see that \(A_k-B_k=O(k^{-\infty })\) (see [14, Section 3]). Thus,

$$\begin{aligned} \begin{array}{ll} &{}e^{ikh(x,\mathring{y})}g(x,\mathring{y},k)=e^{ikh_1(x,\mathring{y})}p(x,\mathring{y},k)+F_k(x,\mathring{y}),\\ &{}F_k(x, \mathring{y}')=O(k^{-\infty }). \end{array} \end{aligned}$$
(5.2)

Now, we are ready to prove that \(h(x,\mathring{y})-h_1(x,\mathring{y})\) vanishes to infinite order at \((x_0,\mathring{x}_0)\). We assume that there exist \(\alpha _0\in \mathbb {N}^{2n+1}_0\), \(\beta _0\in \mathbb {N}^{2n}_0\), \(|\alpha _0|+|\beta _0|\ge 1\) such that

$$\begin{aligned} |\partial ^{\alpha _0}_x\partial ^{\beta _0}_{\mathring{y}}(ih(x,\mathring{y})-ih_1(x,\mathring{y}))|_{(x_0,\mathring{x}_0)}=C_{\alpha _0,\beta _0}\ne 0 \end{aligned}$$

and

$$\begin{aligned} |\partial ^{\alpha }_x\partial ^{\beta }_{\mathring{y}}(ih(x,\mathring{y})-ih_1(x,\mathring{y}))|_{(x_0,\mathring{x}_0)}=0\ \ \text{ if } |\alpha |+|\beta |<|\alpha _0|+|\beta _0|. \end{aligned}$$

From (5.2), we have

$$\begin{aligned} \begin{array}{ll} &{}|\partial ^{\alpha _0}_x\partial ^{\beta _0}_{\mathring{y}}\Bigr (e^{ikh(x,\mathring{y})-ikh_1(x,\mathring{y})}g(x,\mathring{y},k)-p(x,\mathring{y},k)\Bigr )|_{(x_0,\mathring{x}_0)}\\ &{}=-|\partial ^{\alpha _0}_x\partial ^{\beta _0}_{\mathring{y}}\Bigr (e^{-ikh_1(x,\mathring{y})}F_k(x,\mathring{y})\Bigr )|_{(x_0,\mathring{x}_0)}. \end{array} \end{aligned}$$
(5.3)

Since \(h_1(x_0,\mathring{x}_0)=-x^{2n+1}_0\) and \(F_k(x,\mathring{y})=O(k^{-\infty })\), we have

$$\begin{aligned} \lim _{k\rightarrow \infty }k^{-n+\frac{d}{2}-1}|\partial ^{\alpha _0}_x\partial ^{\beta _0}_{\mathring{y}}\Bigr (e^{-ikh_1(x,\mathring{y})}F_k(x,\mathring{y})\Bigr )|_{(x_0,\mathring{x}_0)}=0. \end{aligned}$$
(5.4)

On the other hand, we can check that

$$\begin{aligned} \begin{array}{ll} &{}\lim _{k\rightarrow \infty }k^{-n+\frac{d}{2}-1} |\partial ^{\alpha _0}_x\partial ^{\beta _0}_{\mathring{y}}\Bigr (e^{ikh(x,\mathring{y})-ikh_1(x,\mathring{y})}g(x,\mathring{y},k)- p(x,\mathring{y},k)\Bigr )|_{(x_0,\mathring{x}_0)}\\ &{}=C_{\alpha _0,\beta _0}g_0(x_0,\mathring{x}_0)\ne 0 \end{array} \end{aligned}$$
(5.5)

since \(g_0(x_0,\mathring{x}_0)\ne 0\). From (5.3), (5.4) and (5.5), we get a contradiction. Thus, \(h(x,\mathring{y})-h_1(x,\mathring{y})\) vanishes to infinite order at \((x_0,\mathring{x}_0)\). Since \(x_0\) is arbitrary, the theorem follows. \(\square \)

6 The proof of Theorem 1.9

6.1 Preparation

Fix \(p\in \mu ^{-1}(0)\) and let \(x=(x_1,\ldots ,x_{2n+1})\) be the local coordinates as in Remark 4.4 defined in an open set U of p. We may assume that \(U=\Omega _1\times \Omega _2\times \Omega _3\times \Omega _4\), where \(\Omega _1\subset \mathbb {R}^d\), \(\Omega _2\subset \mathbb {R}^d\) are open sets of \(0\in \mathbb {R}^d\), \(\Omega _3\subset \mathbb {R}^{2n-2d}\) is an open set of \(0\in \mathbb {R}^{2n-2d}\) and \(\Omega _4\) is an open set of \(0\in \mathbb {R}\). From now on, we identify \(\Omega _2\) with

$$\begin{aligned} \left\{ (0,\ldots ,0,x_{d+1},\ldots ,x_{2d},0,\ldots ,0)\in U;\, (x_{d+1},\ldots ,x_{2d})\in \Omega _2\right\} , \end{aligned}$$

\(\Omega _3\) with \(\left\{ (0,\ldots ,0,x_{2d+1},\ldots ,x_{2n},0)\in U;\, (x_{d+1},\ldots ,x_{2n})\in \Omega _3\right\} \), \(\Omega _2\times \Omega _3\) with

$$\begin{aligned} \left\{ (0,\ldots ,0,x_{d+1},\ldots ,x_{2n},0)\in U;\, (x_{d+1},\ldots ,x_{2n})\in \Omega _2\times \Omega _3\right\} . \end{aligned}$$

For \(x=(x_1,\ldots ,x_{2n+1})\), we write \(x''=(x_{d+1},\ldots ,x_{2n+1})\), \(\mathring{x}''=(x_{d+1},\ldots ,x_{2n})\), \(\hat{x}''=(x_{d+1},\ldots ,x_{2d})\),

$$\begin{aligned} \widetilde{x}''=(x_{2d+1},\ldots ,x_{2n+1}),\ \ \widetilde{\mathring{x}}''=(x_{2d+1},\ldots ,x_{2n}). \end{aligned}$$

From now on, we identify \(x''\) with \((0,\ldots ,0,x_{d+1},\ldots ,x_{2n+1})\in U\), \(\mathring{x}''=(x_{d+1},\ldots ,x_{2n})\) with \((0,\ldots ,0,x_{d+1},\ldots ,x_{2n},0)\in U\), \(\hat{x}''\) with \((0,\ldots ,0,x_{d+1},\ldots ,x_{2d},0,\ldots ,0)\in U\), \(\widetilde{x}''\) with \((0,\ldots ,0,x_{2d+1},\ldots ,x_{2n+1})\in U\), \(\widetilde{\mathring{x}}''\) with \((0,\ldots ,0,x_{2d+1},\ldots ,x_{2n},0)\). Since \(G\times S^1\) acts freely on \(\mu ^{-1}(0)\), we take \(\Omega _2\) and \(\Omega _3\) small enough so that if \(x, x_1\in \Omega _2\times \Omega _3\) and \(x\ne x_1\), then

$$\begin{aligned} g\circ e^{i\theta }\circ x\ne g_1\circ e^{i\theta _1}\circ x_1,\ \ \forall (g,e^{i\theta })\in G\times S^1, \ \ \forall (g_1,e^{i\theta _1})\in G\times S^1. \end{aligned}$$
(6.1)

We now assume that \(q=n_-\) and let \(\Psi (x,y)\in C^\infty (U\times U)\) be as in Theorem 1.8. From \(S^{(q)}_{G,m}=(S^{(q)}_{G,m})^*\), we get

$$\begin{aligned} e^{im\Psi (x,y)}b(x,y,m)=e^{-im\overline{\Psi }(y,x)}b^*(x,y,m)+O(m^{-\infty }), \end{aligned}$$
(6.2)

where \((S^{(q)}_{G,m})^*:L^2_{(0,q)}(X)\rightarrow L^2_{(0,q)}(X)\) is the adjoint of \(S^{(q)}_{G,m}:L^2_{(0,q)}(X)\rightarrow L^2_{(0,q)}(X)\) with respect to \((\,\cdot \,|\,\cdot \,)\) and \(b^*(x,y,m):T^{*0,q}_xX\rightarrow T^{*0,q}_yX\) is the adjoint of \(b(x,y,m):T^{*0,q}_yX\rightarrow T^{*0,q}_xX\) with respect to \(\langle \,\cdot \,|\,\cdot \,\rangle \). From (6.2), we can repeat the proof of Theorem 5.2 with minor change and deduce that

$$\begin{aligned} \Psi (x,y)+\overline{\Psi }(y,x) \text{ vanishes } \text{ to } \text{ infinite } \text{ order } \text{ at } \mathrm{diag}\,\Bigr ((\mu ^{-1}(0)\bigcap U)\times (\mu ^{-1}(0)\bigcap U)\Bigr ). \end{aligned}$$
(6.3)

From \(\overline{\partial }_bS^{(q)}_{G,m}=0\), we can check that

$$\begin{aligned} \overline{\partial }_b\Psi (x,y) \text{ vanishes } \text{ to } \text{ infinite } \text{ order } \text{ at } \mathrm{diag}\,\Bigr ((\mu ^{-1}(0)\bigcap U)\times (\mu ^{-1}(0)\bigcap U)\Bigr ). \end{aligned}$$
(6.4)

From (6.3), (6.4) and notice that \(\frac{\partial }{\partial x_j}-i\frac{\partial }{\partial x_{d+j}}\in T^{0,1}_xX\), \(j=1,\ldots ,d\), where \(x\in \mu ^{-1}(0)\) (see Remark 4.4), and \(\frac{\partial }{\partial x_j}\Psi (x,y)=\frac{\partial }{\partial y_j}\Psi (x,y)=0\), \(j=1,\ldots ,d\), we conclude that

$$\begin{aligned} \begin{array}{ll} &{}\frac{\partial }{\partial x_{d+j}}\Psi (x,y)|_{x_{d+1}=\cdots =x_{2d}=0} \text{ and } \frac{\partial }{\partial y_{d+j}}\Psi (x,y)|_{y_{d+1}=\cdots =y_{2d}=0} \text{ vanish } \text{ to } \text{ infinite } \text{ order } \text{ at }\\ &{}\mathrm{diag}\,\Bigr ((\mu ^{-1}(0)\bigcap U)\times (\mu ^{-1}(0)\bigcap U)\Bigr ). \end{array} \end{aligned}$$

Let \(G_j(x,y):=\frac{\partial }{\partial y_{d+j}}\Psi (x,y)|_{y_{d+1}=\cdots =y_{2d}=0}\), \(H_j(x,y):=\frac{\partial }{\partial x_{d+j}}\Psi (x,y)|_{x_{d+1}=\cdots =x_{2d}=0}\). Put

$$\begin{aligned} \Psi _1(x,y):=\Psi (x,y)-\sum ^d_{j=1}y_{d+j}G_j(x,y),\quad \Psi _2(x,y):=\Psi (x,y)-\sum ^d_{j=1}x_{d+j}H_j(x,y). \end{aligned}$$

Then, for \(j=1, 2, \ldots , d\),

$$\begin{aligned} \frac{\partial }{\partial y_{d+j}}\Psi _1(x,y)|_{y_{d+1}=\cdots =y_{2d}=0}=0 \quad \text {and} \quad \frac{\partial }{\partial x_{d+j}}\Psi _2(x,y)|_{x_{d+1}=\cdots =x_{2d}=0}=0, \end{aligned}$$
(6.5)

and, for \(j=1,2\),

$$\begin{aligned} \Psi (x,y)-\Psi _j(x,y) \text{ vanishes } \text{ to } \text{ infinite } \text{ order } \text{ at } \mathrm{diag}\,\Bigr ((\mu ^{-1}(0)\bigcap U)\times (\mu ^{-1}(0)\bigcap U)\Bigr ). \end{aligned}$$
(6.6)

We also write \(u=(u_1,\ldots ,u_{2n+1})\) to denote the local coordinates of U. Recall that for any smooth function \(f\in C^\infty (U)\), we write \(\widetilde{f}\in C^\infty (U^{\mathbb {C}})\) to denote an almost analytic extension of f (see the discussion after (3.38)). We consider the following two systems

$$\begin{aligned} \frac{\partial \widetilde{\Psi }_1}{\partial \widetilde{u}_{2d+j}}(\widetilde{x},\widetilde{\widetilde{u}''})+\frac{\partial \widetilde{\Psi }_2}{\partial \widetilde{x}_{2d+j}}(\widetilde{\widetilde{u}''},\widetilde{y})=0,\ \ j=1,2,\ldots ,2n-2d, \end{aligned}$$
(6.7)

and

$$\begin{aligned} \frac{\partial \widetilde{\Psi }_1}{\partial \widetilde{u}_{d+j}}(\widetilde{x},\widetilde{u''})+\frac{\partial \widetilde{\Psi }_2}{\partial \widetilde{x}_{d+j}}(\widetilde{u''},\widetilde{y})=0,\ \ j=1,2,\ldots ,2n-d, \end{aligned}$$
(6.8)

where \(\widetilde{\widetilde{u}''}=(0,\ldots ,0,\widetilde{u}_{2d+1},\ldots ,\widetilde{u}_{2n+1})\), \(\widetilde{u''}=(0,\ldots ,0,\widetilde{u}_{d+1},\ldots ,\widetilde{u}_{2n+1})\). From (6.5) and Theorem 1.12, we can take \(\widetilde{\Psi }_1\) and \(\widetilde{\Psi }_2\) so that for every \(j=1,2,\ldots ,d\),

$$\begin{aligned} \frac{\partial \widetilde{\Psi }_1}{\partial \widetilde{u}_{d+j}}(\widetilde{x},\widetilde{u''})=0 \quad \text {and} \quad \frac{\partial \widetilde{\Psi }_2}{\partial \widetilde{x}_{d+j}}(\widetilde{u''},\widetilde{y})=0,\ \ \text{ if } \widetilde{u}_{d+1}=\cdots =\widetilde{u}_{2d}=0, \end{aligned}$$
(6.9)

and, for \(j=1, 2\),

$$\begin{aligned} \widetilde{\Psi }_j(\widetilde{x}, \widetilde{y})=-\widetilde{x}_{2n+1}+\widetilde{y}_{2n+1}+\widetilde{\hat{\Psi }_j}(\widetilde{\mathring{x}''},\widetilde{\mathring{y}''}),\ \ \widetilde{\hat{\Psi }_j}\in C^\infty (U^\mathbb {C}\times U^\mathbb {C}), \end{aligned}$$
(6.10)

where \(\widetilde{\mathring{x}''}=(0,\ldots ,0,\widetilde{x}_{d+1},\ldots ,\widetilde{x}_{2n},0)\), \(\widetilde{\mathring{y}''}=(0,\ldots ,0,\widetilde{y}_{d+1},\ldots ,\widetilde{y}_{2n},0)\).

From Theorem 1.12, (1.19) and \(d_x\Psi (x,x)=-d_y\Psi (x,x)=-\omega _0(x),\ \forall x\in \mu ^{-1}(0)\), it is not difficult to see that

$$\begin{aligned} \frac{\partial \widetilde{\Psi }_1}{\partial \widetilde{u}_{d+j}}(\widetilde{x}'',\widetilde{x}'')+\frac{\partial \widetilde{\Psi }_2}{\partial \widetilde{x}_{d+j}}(\widetilde{x}'',\widetilde{x}'')=0,\ \ j=1,2,\ldots ,2n-d, \end{aligned}$$

and the matrices

$$\begin{aligned}&\left( \frac{\partial ^2\Psi }{\partial u_{2d+j}\partial u_{2d+k}}(p,p)+\frac{\partial ^2\Psi }{\partial x_{2d+j}\partial x_{2d+k}}(p,p)\right) ^{2n-2d}_{j,k=1},\\&\left( \frac{\partial ^2\Psi }{\partial u_{d+j}\partial u_{d+k}}(p,p)+\frac{\partial ^2\Psi }{\partial x_{d+j}\partial x_{d+k}}(p,p)\right) ^{2n-d}_{j,k=1} \end{aligned}$$

are non-singular. Moreover,

$$\begin{aligned} \begin{array}{ll} &{}\det \left( \frac{\partial ^2\Psi }{\partial u_{2d+j}\partial u_{2d+k}}(p,p)+\frac{\partial ^2\Psi }{\partial x_{2d+j}\partial x_{2d+k}}(p,p)\right) ^{2n-2d}_{j,k=1}=(4i|\mu _{d+1}|\cdots 4i|\mu _n|)^2,\\ &{}\det \left( \frac{\partial ^2\Psi }{\partial u_{d+j}\partial u_{d+k}}(p,p)+\frac{\partial ^2\Psi }{\partial x_{d+j}\partial x_{d+k}}(p,p)\right) ^{2n-d}_{j,k=1}=(8i|\mu _1|\cdots 8i|\mu _d|)(4i|\mu _{d+1}|\cdots 4i|\mu _n|)^2. \end{array} \end{aligned}$$

Hence, near (pp), we can solve (6.7) and (6.8) and the solutions are unique. Let \(\alpha (x,y)=(\alpha _{2d+1}(x,y),\ldots ,\alpha _{2n}(x,y))\in C^\infty (U\times U,\mathbb {C}^{2n-2d})\) and \(\beta (x,y)=(\beta _{d+1}(x,y),\ldots ,\beta _{2n}(x,y))\in C^\infty (U\times U,\mathbb {C}^{2n-d})\) be the solutions of (6.7) and (6.8), respectively. From (6.9), it is easy to see that

$$\begin{aligned} \beta (x,y)=(\beta _{d+1}(x,y),\ldots ,\beta _{2n}(x,y))=(0,\ldots ,0,\alpha _{2d+1}(x,y),\ldots ,\alpha _{2n}(x,y)). \end{aligned}$$
(6.11)

From (6.11), we see that the value of \(\widetilde{\Psi }_1(x,\widetilde{\widetilde{u}''})+\widetilde{\Psi }_2(\widetilde{\widetilde{u}''},y)\) at critical point \(\widetilde{\widetilde{u}''}=\alpha (x,y)\) is equal to the value of \(\widetilde{\Psi }_1(x,\widetilde{u''})+\widetilde{\Psi }_2(\widetilde{u''},y)\) at critical point \(\widetilde{u''}=\beta (x,y)\). Put

$$\begin{aligned} \Psi _3(x,y):=\widetilde{\Psi }_1(x,\alpha (x,y))+\widetilde{\Psi }_2(\alpha (x,y),y)=\widetilde{\Psi }_1(x,\beta (x,y))+\widetilde{\Psi }_2(\beta (x,y),y). \end{aligned}$$
(6.12)

\(\Psi _3(x,y)\) is a complex phase function. From (6.10), we have

$$\begin{aligned} \Psi _3(x,y)=-x_{2n+1}+y_{2n+1}+\hat{\Psi }_3(\mathring{x}'',\mathring{y}''),\ \ \hat{\Psi }_3(\mathring{x}'',\mathring{y}'')\in C^\infty (U\times U). \end{aligned}$$

Moreover, we have the following

Theorem 6.1

The function \(\Psi _3(x,y)-\Psi (x,y)\) vanishes to infinite order at \(\mathrm{diag}\,\Bigr ((\mu ^{-1}(0)\bigcap U)\times (\mu ^{-1}(0)\bigcap U)\Bigr )\).

Proof

We consider the kernel \(S^{(q)}_{G,m}\circ S^{(q)}_{G,m}\) on U. Let \(V\Subset U\) be an open set of p. Let \(\chi (\mathring{x}'')\in C^\infty _0(\Omega _2\times \Omega _3)\). From (6.1), we can extend \(\chi (\mathring{x}'')\) to \(W:=\left\{ g\circ e^{i\theta }\circ x;\, (g,e^{i\theta })\in G\times S^1, x\in \Omega _2\times \Omega _3\right\} \) by \(\chi (g\circ e^{i\theta }\circ \mathring{x}''):=\chi (\mathring{x}'')\), for every \((g,e^{i\theta })\in G\times S^1\). Assume that \(\chi =1\) on some neighborhood of V. Let \(\chi _1\in C^\infty _0(U)\) with \(\chi _1=1\) on some neighborhood of V and \(\mathrm{Supp}\,\chi _1\subset \left\{ x\in X;\, \chi (x)=1\right\} \). We have

$$\begin{aligned} \chi _1S^{(q)}_{G,m}\circ S^{(q)}_{G,m}=\chi _1S^{(q)}_{G,m}\chi \circ S^{(q)}_{G,m}+\chi _1S^{(q)}_{G,m}(1-\chi )\circ S^{(q)}_{G,m}. \end{aligned}$$
(6.13)

Let’s first consider \(\chi _1S^{(q)}_{G,m}(1-\chi )\circ S^{(q)}_{G,m}\). We have

$$\begin{aligned} (\chi _1S^{(q)}_{G,m}(1-\chi ))(x,u) =\frac{1}{2\pi }\chi _1(x)\int ^{\pi }_{-\pi }\int _GS^{(q)}(x,g\circ e^{i\theta }\circ u)(1-\chi (u))e^{im\theta }d\mu (g)d\theta . \end{aligned}$$
(6.14)

If \(u\notin \left\{ x\in X;\, \chi (x)=1\right\} \). Since \(\mathrm{Supp}\,\chi _1\subset \left\{ x\in X;\, \chi (x)=1\right\} \) and \(\chi (x)=\chi (g\circ e^{i\theta }\circ x)\), for every \((g,e^{i\theta })\in G\times S^1\), for every \(x\in X\), we conclude that \(g\circ e^{i\theta }\circ u\notin \mathrm{Supp}\,\chi _1\), for every \((g,e^{i\theta })\in G\times S^1\). From this observation and notice that \(S^{(q)}\) is smoothing away from diagonal, we can integrate by parts with respect to \(\theta \) in (6.14) and deduce that \(\chi _1S^{(q)}_{G,m}\circ (1-\chi )=O(m^{-\infty })\) and hence

$$\begin{aligned} \chi _1S^{(q)}_{G,m}(1-\chi )\circ S^{(q)}_{G,m}=O(m^{-\infty }). \end{aligned}$$
(6.15)

From (6.13) and (6.15), we get

$$\begin{aligned} \chi _1S^{(q)}_{G,m}\circ S^{(q)}_{G,m}=\chi _1S^{(q)}_{G,m}\chi \circ S^{(q)}_{G,m}+O(m^{-\infty }). \end{aligned}$$
(6.16)

We can check that on U,

$$\begin{aligned} \begin{array}{ll} &{}(\chi _1S^{(q)}_{G,m}\chi \circ S^{(q)}_{G,m})(x,y)\\ &{}=(2\pi )\int e^{im\Psi (x,u'')+im\Psi (u'',y)}\chi _1(x)b(x,\mathring{u}'',m)\chi (\mathring{u}'')b(\mathring{u}'',y,m)dv(\mathring{u}'')+O(m^{-\infty })\\ &{}=(2\pi )\int e^{im\Psi _1(x,u'')+im\Psi _2(u'',y)}\chi _1(x)b(x,\mathring{u}'',m)\chi (\mathring{u}'')b(\mathring{u}'',y,m)dv(\mathring{u}'')+O(m^{-\infty })\\ &{}(\text{ here } \text{ we } \text{ use } (6.6)), \end{array} \end{aligned}$$
(6.17)

where \(d\mu (g)d\theta dv(\mathring{u}'')=dv(x)\) on U. We use complex stationary phase formula of Melin–Sjöstrand [25] to carry out the integral (6.17) and get

$$\begin{aligned} \begin{array}{ll} &{}(\chi _1S^{(q)}_{G,m}\chi \circ S^{(q)}_{G,m})(x,y)=e^{im\Psi _3(x,y)}a(x,y,m)+O(m^{-\infty })\ \ \text{ on } U, \\ &{}a(x,y,m)\in S^{n-\frac{d}{2}}_{\mathrm{loc}}\,(1; U\times U, T^{*0,q}X\boxtimes (T^{*0,q}X)^*),\\ &{}a(x,y,m)\sim \sum ^\infty _{j=0}m^{n-\frac{d}{2}-j}a_j(x,y) \text{ in } S^{n-\frac{d}{2}}_{\mathrm{loc}}\,(1; U\times U, T^{*0,q}X\boxtimes (T^{*0,q}X)^*),\\ &{}a_j(x,y)\in C^\infty (U\times U, T^{*0,q}X\boxtimes (T^{*0,q}X)^*),\ \ j=0,1,2,\ldots ,\\ &{}a_0(p,p)\ne 0. \end{array} \end{aligned}$$
(6.18)

From (6.16), (6.18) and notice that \((\chi _1S^{(q)}_{G,m}\circ S^{(q)}_{G,m})(x,y)=(\chi _1S^{(q)}_{G,m})(x,y)\), we deduce that

$$\begin{aligned} e^{im\Psi _3(x,y)}a(x,y,m)=e^{im\Psi (x,y)}\chi _1(x)b(x,y,m)+O(m^{-\infty })\ \ \text{ on } U. \end{aligned}$$
(6.19)

From (6.19), we can repeat the proof of Theorem 5.2 with minor change and deduce that \(\Psi _3(x,y)-\Psi (x,y)\) vanishes to infinite order at \(\mathrm{diag}\,\Bigr ((\mu ^{-1}(0)\bigcap U)\times (\mu ^{-1}(0)\bigcap U)\Bigr )\). \(\square \)

The following two theorems follow from (6.6), (6.12), Theorem 6.1, complex stationary phase formula of Melin–Sjöstrand [25] and some straightforward computation. We omit the details.

Theorem 6.2

With the notations used above, let

$$\begin{aligned} \begin{array}{ll} &{}A_m(x,y)=e^{im\Psi (x,y)}a(x,y,m),\quad B_m(x,y)=e^{im\Psi (x,y)}b(x,y,m),\\ &{}a(x,y,m)\in S^{k}_{\mathrm{loc}}\,(1; U\times U, H\boxtimes F^*),\ \quad b(x,y,m)\in S^{\ell }_{\mathrm{loc}}\,(1; U\times U, F\boxtimes E^*),\\ &{}a(x,y,m)\sim \sum ^\infty _{j=0}m^{k-j}a_j(x,y) \text{ in } S^{k}_{\mathrm{loc}}\,(1; U\times U, H\boxtimes F^*),\\ &{}b(x,y,m)\sim \sum ^\infty _{j=0}m^{\ell -j}b_j(x,y) \text{ in } S^{\ell }_{\mathrm{loc}}\,(1; U\times U, F\boxtimes E^*),\\ &{}a_j(x,y)\in C^\infty (U\times U, H\boxtimes F^*),\quad b_j(x,y)\in C^\infty (U\times U, F\boxtimes E^*),\ \ j=0,1,2,\ldots , \end{array} \end{aligned}$$

where E, F and H are vector bundles over X. Let \(\chi (\mathring{x}'')\in C^\infty _0(\Omega _2\times \Omega _3)\). Then, we have

$$\begin{aligned} \begin{array}{ll} &{}\int A_m(x,u)\chi (\mathring{u}'')B_m(u,y)dv(\mathring{u}'')=e^{im\Psi (x,y)}c(x,y,m)+O(m^{-\infty }),\\ &{}c(x,y,m)\in S^{k+\ell -(n-\frac{d}{2})}_{\mathrm{loc}}\,(1; U\times U, H\boxtimes E^*),\\ &{}c(x,y,m)\sim \sum ^\infty _{j=0}m^{k+\ell -(n-\frac{d}{2})-j}c_j(x,y) \text{ in } S^{k+\ell -(n-\frac{d}{2})}_{\mathrm{loc}}\,(1; U\times U, H\boxtimes E^*),\\ &{}c_0(x,x)=2^{-n-\frac{d}{2}}\pi ^{n-\frac{d}{2}}|\det \mathcal {L}_{x}|^{-1}|\det R_x|^{\frac{1}{2}}a_0(x,x)b_0(x,x)\chi (\mathring{x}''),\ \ \forall x\in \mu ^{-1}(0)\bigcap U, \end{array} \end{aligned}$$

where \(|\det R_x|\) is in the discussion before Theorem 1.6.

Moreover, if there are \(N_1, N_2\in \mathbb {N}\), such that \(|a_0(x,y)|\le C|(x,y)-(x_0,x_0)|^{N_1}\), \(|b_0(x,y)|\le C|(x,y)-(x_0,x_0)|^{N_2}\), for all \(x_0\in \mu ^{-1}(0)\bigcap U\), where \(C>0\) is a constant, then,

$$\begin{aligned} |c_0(x,y)|\le \hat{C}|(x,y)-(x_0,x_0)|^{N_1+N_2}, \end{aligned}$$

for all \(x_0\in \mu ^{-1}(0)\bigcap U\), where \(\hat{C}>0\) is a constant.

Theorem 6.3

With the notations used above, let

$$\begin{aligned} \begin{array}{ll} &{}\mathcal {A}_m(x,\widetilde{y}'')=e^{im\Psi (x,\widetilde{y}'')}\alpha (x,\widetilde{y}'',m),\quad \mathcal {B}_m(\widetilde{x}'',y)=e^{im\Psi (\widetilde{x}'',y)}\beta (\widetilde{x}'',y,m),\\ &{}\alpha (x,\widetilde{y}'',m)\in S^{k}_{\mathrm{loc}}\,(1; U\times (\Omega _3\times \Omega _4), H\boxtimes F^*),\quad \beta (\widetilde{x}'',y,m)\\ &{} \in S^{\ell }_{\mathrm{loc}}\,(1; (\Omega _3\times \Omega _4)\times U, F\boxtimes E^*),\\ &{}\alpha (x,\widetilde{y}'',m)\sim \sum ^\infty _{j=0}m^{k-j}\alpha _j(x,\widetilde{y}'') \text{ in } S^{k}_{\mathrm{loc}}\,(1; U\times (\Omega _3\times \Omega _4), H\boxtimes F^*),\\ &{}\beta (\widetilde{x}'',y,m)\sim \sum ^\infty _{j=0}m^{\ell -j}\beta _j(\widetilde{x}'',y) \text{ in } S^{\ell }_{\mathrm{loc}}\,(1; (\Omega _3\times \Omega _4)\times U, F\boxtimes E^*),\\ &{}\alpha _j(x,\widetilde{y}'')\in C^\infty (U\times (\Omega _3\times \Omega _4), H\boxtimes F^*),\ \beta _j(\widetilde{x}'',y)\\ &{} \in C^\infty ((\Omega _3\times \Omega _4)\times U, F\boxtimes E^*),\ j=0,1,2,\ldots , \end{array} \end{aligned}$$

where E, F and H are vector bundles over X. Let \(\chi _1(\widetilde{\mathring{x}}'')\in C^\infty _0(\Omega _3)\). Then, we have

$$\begin{aligned} \begin{array}{ll} &{}\int \mathcal {A}_m(x,\widetilde{u}'')\chi _1(\widetilde{\mathring{u}}'')\mathcal {B}_m(\widetilde{u}'',y)dv(\widetilde{\mathring{u}})=e^{im\Psi (x,y)}\gamma (x,y,m)+O(m^{-\infty }),\\ &{}\gamma (x,y,m)\in S^{k+\ell -(n-d)}_{\mathrm{loc}}\,(1; U\times U, H\boxtimes E^*),\\ &{}\gamma (x,y,m)\sim \sum ^\infty _{j=0}m^{k+\ell -(n-d)-j}\gamma _j(x,y) \text{ in } S^{k+\ell -(n-d)}_{\mathrm{loc}}\,(1; U\times U, H\boxtimes E^*),\\ &{}\gamma _0(x,x)=2^{-n}\pi ^{n-d}|\det \mathcal {L}_{x}|^{-1}|\det R_x|\alpha _0(x,\widetilde{x}'')\beta _0(\widetilde{x}'',x)\chi _1(\widetilde{\mathring{x}}''),\ \ \forall x\in \mu ^{-1}(0)\bigcap U, \end{array} \end{aligned}$$

where \(|\det R_x|\) is in the discussion before Theorem 1.6.

Moreover, if there are \(N_1, N_2\in \mathbb {N}\), such that \(|\alpha _0(x,\widetilde{y}'')|\le C|(x,\widetilde{y}'')-(x_0,x_0)|^{N_1}\), \(|\beta _0(x,\widetilde{y}'')|\le C|(x,\widetilde{y}'')-(x_0,x_0)|^{N_2}\), for all \(x_0\in \mu ^{-1}(0)\bigcap U\), where \(C>0\) is a constant, then,

$$\begin{aligned} |\gamma _0(x,y)|\le \hat{C}|(x,y)-(x_0,x_0)|^{N_1+N_2}, \end{aligned}$$

for all \(x_0\in \mu ^{-1}(0)\bigcap U\), where \(\hat{C}>0\) is a constant.

6.2 The proof of Theorem 1.9

Since \(\underline{\mathfrak {g}}_x\) is orthogonal to \(H_xY\bigcap JH_xY\) and \(H_xY\bigcap JH_xY\subset \underline{\mathfrak {g}}^{\perp _b}_x\), for every \(x\in Y\), we can find a G-invariant orthonormal basis \(\left\{ Z_1,\ldots ,Z_n\right\} \) of \(T^{1,0}X\) on Y such that

$$\begin{aligned} \begin{array}{ll} &{}\mathcal {L}_x(Z_j(x),\overline{Z}_k(x))=\delta _{j,k}\lambda _j(x),\ \ j,k=1,\ldots ,n,\ \ x\in Y,\\ &{}Z_j(x)\in \underline{\mathfrak {g}}_x+iJ\underline{\mathfrak {g}}_x,\ \ \forall x\in Y,\ \ j=1,2,\ldots ,d,\\ &{}Z_j(x)\in \mathbb {C}H_xY\bigcap J(\mathbb {C}H_xY),\ \ \forall x\in Y,\ \ j=d+1,\ldots ,n. \end{array} \end{aligned}$$

Let \(\left\{ e_1,\ldots ,e_n\right\} \) denote the orthonormal basis of \(T^{*0,1}X\) on Y, dual to \(\left\{ \overline{Z}_1,\ldots ,\overline{Z}_n\right\} \). Fix \(s=0,1,2,\ldots ,n-d\). For \(x\in Y\), put

$$\begin{aligned} B^{*0,s}_xX=\left\{ \sum _{d+1\le j_1<\cdots<j_s\le n}a_{j_1,\ldots ,j_s}e_{j_1}\wedge \cdots \wedge e_{j_s};\, a_{j_1,\ldots ,j_s}\in \mathbb {C},\ \forall d+1\le j_1<\cdots <j_s\le n\right\} \end{aligned}$$

and let \(B^{*0,s}X\) be the vector bundle of Y with fiber \(B^{*0,s}_xX\), \(x\in Y\). Let \(C^\infty (Y,B^{*0,s}X)^G\) denote the set of all G-invariant sections of Y with values in \(B^{*0,s}X\). Let

$$\begin{aligned} \iota _G:C^\infty (Y,B^{*0,s}X)^G\rightarrow \Omega ^{0,s}(Y_G) \end{aligned}$$

be the natural identification.

Assume that \(\lambda _1<0,\ldots ,\lambda _r<0\), and \(\lambda _{d+1}<0,\ldots ,\lambda _{n_--r+d}<0\). For \(x\in Y\), put

$$\begin{aligned} \hat{\mathcal {N}}(x,n_-)=\left\{ ce_{d+1}\wedge \cdots \wedge e_{n_--r+d};\, c\in \mathbb {C}\right\} , \end{aligned}$$

and let

$$\begin{aligned} \begin{array}{ll} &{}\hat{p}=\hat{p}_x:\mathcal {N}(x,n_-)\rightarrow \hat{\mathcal {N}}(x,n_-),\\ &{}u=ce_1\wedge \cdots \wedge e_r\wedge e_{d+1}\wedge \cdots \wedge e_{n_--r+d}\rightarrow ce_{d+1}\wedge \cdots \wedge e_{n_--r+d}. \end{array} \end{aligned}$$

Let \(\iota :Y\rightarrow X\) be the natural inclusion and let \(\iota ^*:\Omega ^{0,q}(X)\rightarrow \Omega ^{0,q}(Y)\) be the pull-back of \(\iota \). Recall that we work with the assumption that \(q=n_-\). Let \(\Box ^{(q-r)}_{b,Y_G}\) be the Kohn Laplacian for \((0,q-r)\) forms on \(Y_G\). Fix \(m\in \mathbb {N}\). Let \(H^{q-r}_{b,m}(Y_G):=\left\{ u\in \Omega ^{0,q-r}(Y_G);\, \Box ^{(q-r)}_{b,Y_G}u=0,\ \ Tu=imu\right\} \). Let \(S^{(q-r)}_{Y_G,m}:L^2_{(0,q-r)}(Y_G)\rightarrow H^{q-r}_{b,m}(Y_G)\) be the orthogonal projection and let \(S^{(q-r)}_{Y_G,m}(x,y)\) be the distribution kernel of \(S^{(q-r)}_{Y_G,m}\). Let

$$\begin{aligned} f(x)=\sqrt{V_{\mathrm{eff}}\,(x)}|\det \,R_x|^{-\frac{1}{4}}\in C^\infty (Y)^G. \end{aligned}$$

Let

$$\begin{aligned} \begin{array}{cl} \sigma _m:\Omega ^{0,q}(X)&{}\rightarrow H^{q-r}_{b,m}(Y_G),\\ u&{}\rightarrow m^{-\frac{d}{4}}S^{(q-r)}_{Y_G,m}\circ \iota _G\circ \hat{p}\circ \tau _{x,n_-}\circ f\circ \iota ^*\circ S^{(q)}_{G,m}u. \end{array} \end{aligned}$$

Recall that \(\tau _{x,n_-}\) is given by (1.7). Let \(\sigma ^*_m:\Omega ^{0,q-r}(Y_G)\rightarrow \Omega ^{0,q}(X)\) be the adjoints of \(\sigma _m\). It is easy to see that \(\sigma ^*_mu\in H^q_{b,m}(X)^G:=(\mathrm{Ker}\,\Box ^{(q)}_b)^G_m,\ \forall u\in \Omega ^{0,q-r}(Y_G)\). Let \(\sigma _m(x,y)\) and \(\sigma ^*_m(x,y)\) denote the distribution kernels of \(\sigma _m\) and \(\sigma ^*_m\), respectively.

Let’s pause and recall some results for \(S^{(q-r)}_{Y_G,m}\). We first introduce some notations. Let \(\mathcal {L}_{Y_G,x}\) be the Levi form on \(Y_G\) at \(x\in Y_G\) induced naturally from \(\mathcal {L}\). The Hermitian metric \(\langle \,\cdot \,|\,\cdot \,\rangle \) on \(T^{1,0}X\) induces a Hermitian metric \(\langle \,\cdot \,|\,\cdot \,\rangle \) on \(T^{1,0}Y_G\). Let \(\det \,\mathcal {L}_{Y_G,x}=\lambda _1\ldots \lambda _{n-d}\), where \(\lambda _j\), \(j=1,\ldots ,n-d\), are the eigenvalues of \(\mathcal {L}_{Y_G,x}\) with respect to the Hermitian metric \(\langle \,\cdot \,|\,\cdot \,\rangle \). For \(x\in Y_G\), let \(\hat{\tau }_x:T^{*0,q-r}_xY_G\rightarrow \hat{\mathcal {N}}(x,n_-)\) be the orthogonal projection.

Let \(\pi :Y\rightarrow Y_G\) be the natural quotient. Let \(S^{(q-r)}_{Y_G}:L^2_{(0,q-r)}(Y_G)\rightarrow \mathrm{Ker}\,\Box ^{(q-r)}_{b,Y_G}\) be the Szegő projection as (3.2). Since \(S^{(q-r)}_{Y_G}\) is smoothing away from diagonal (see Theorem 3.3), it is easy to see that for any \(x, y\in Y\), if \(\pi (e^{i\theta }\circ x)\ne \pi (e^{i\theta }\circ y)\), for every \(\theta \in [0,2\pi [\), then there are open sets U of \(\pi (x)\) in \(Y_G\) and V of \(\pi (y)\) in \(Y_G\) such that for all \(\hat{\chi }\in C^\infty _0(U)\), \(\widetilde{\chi }\in C^\infty _0(V)\), we have

$$\begin{aligned} \hat{\chi } S^{(q-r)}_{Y_G,m}\widetilde{\chi }=O(m^{-\infty })\ \ \text{ on } Y_G. \end{aligned}$$
(6.20)

Fix \(p\in Y\) and let \(x=(x_1,\ldots ,x_{2n+1})\) be the local coordinates as in Remark 4.4. We will use the same notations as in the beginning of Sect. 6.1. From now on, we identify \(\widetilde{x}''\) as local coordinates of \(Y_G\) near \(\pi (p)\in Y_G\) and we identify \(W:=\Omega _3\times \Omega _4\) with an open set of \(\pi (p)\) in \(Y_G\). It was proved in Theorem 4.11 in [14] that as \(m\rightarrow +\infty \),

$$\begin{aligned} \begin{array}{ll} &{}S^{(q-r)}_{Y_G,m}(\widetilde{x}'',\widetilde{y}'')=e^{im\phi (\widetilde{x}'',\widetilde{y}'')}b(\widetilde{x}'',\widetilde{y}'',m)+O(m^{-\infty })\ \ \text{ on } W,\\ &{}\beta (\widetilde{x}'',\widetilde{y}'',m)\in S^{n-d}_{\mathrm{loc}}\,(1; W\times W, T^{*0,q-r}Y_G\boxtimes (T^{*0,q-r}Y_G)^*),\\ &{}\beta (\widetilde{x}'',\widetilde{y}'',m)\sim \sum ^\infty _{j=0}m^{n-d-j}b_j(\widetilde{x}'',\widetilde{y}'') \text{ in } S^{n-d}_{\mathrm{loc}}\,(1; W\times W, T^{*0,q-r}Y_G\boxtimes (T^{*0,q-r}Y_G)^*),\\ &{}\beta _j(\widetilde{x}'',\widetilde{y}'')\in C^\infty (W\times W, T^{*0,q-r}Y_G\boxtimes (T^{*0,q-r}Y_G)^*),\ \ j=0,1,2,\ldots ,\\ &{}\beta _0(\widetilde{x}'',\widetilde{x}'')=\frac{1}{2}\pi ^{-(n-d)-1}|\det \,\mathcal {L}_{Y_G,\widetilde{x}''}|\hat{\tau }_{\widetilde{x}''},\ \ \forall \widetilde{x}''\in W, \end{array} \end{aligned}$$

and

$$\begin{aligned} \begin{array}{ll} &{}\phi (\widetilde{x}'',\widetilde{y}'')=-x_{2n+1}+y_{2n+1}+\hat{\phi }(\widetilde{\mathring{x}}'',\widetilde{\mathring{y}}'')\in C^\infty (W\times W),\\ &{}d_{\widetilde{x}''}\phi (\widetilde{x}'',\widetilde{y}'')=-d_{\widetilde{y}''}\phi (\widetilde{x}'',\widetilde{x}'')=-\omega _0(\widetilde{x}''),\\ &{}\mathrm{Im}\,\hat{\phi }(\widetilde{\mathring{x}}'',\widetilde{\mathring{y}}'')\ge c|\widetilde{\mathring{x}}''-\widetilde{\mathring{y}}''|^2,\ \ \text{ where } c>0 \text{ is } \text{ a } \text{ constant }, \\ &{}p_0(\widetilde{x}'', d_{\widetilde{x}''}\phi (\widetilde{x}'',\widetilde{y}'')) \text{ vanishes } \text{ to } \text{ infinite } \text{ order } \text{ at } \widetilde{\mathring{x}}''=\widetilde{\mathring{y}}'',\\ &{}\phi (\widetilde{x}'', \widetilde{y}'')=-x_{2n+1}+y_{2n+1}+i\sum ^{n}_{j=d+1}|\mu _j||z_j-w_j|^2 \\ &{}\quad +\sum ^{n}_{j=d+1}i\mu _j(\overline{z}_jw_j-z_j\overline{w}_j)+O(|(\widetilde{\mathring{x}}'', \widetilde{\mathring{y}}'')|^3), \end{array} \end{aligned}$$
(6.21)

where \(p_0\) denotes the principal symbol of \(\Box ^{(q-r)}_{b,Y_G}\), \(z_j=x_{2j-1}+ix_{2j}\), \(j=d+1,\ldots ,n\), and \(\mu _{d+1},\ldots ,\mu _n\) are the eigenvalues of \(\mathcal {L}_{Y_G,p}\).

Note that for any \(\phi _1(\widetilde{x}'',\widetilde{y}'')\in C^\infty (W\times W)\), if \(\phi _1\) satisfies (6.21), then \(\phi _1-\phi \) vanishes to infinite order at \(\widetilde{\mathring{x}}''=\widetilde{\mathring{y}}''\) (see Remark 3.6 in [14]). It is not difficult to see that the phase function \(\Psi (\widetilde{x}'',\widetilde{y}'')\) satisfies (6.21). Hence, we can replace the phase \(\phi (\widetilde{x}'',\widetilde{y}'')\) by \(\Psi (\widetilde{x}'',\widetilde{y}'')\) and we have

$$\begin{aligned} S^{(q-r)}_{Y_G,m}(\widetilde{x}'',\widetilde{y}'')=e^{im\Psi (\widetilde{x}'',\widetilde{y}'')}\beta (\widetilde{x}'',\widetilde{y}'',m)+O(m^{-\infty })\ \ \text{ on } W. \end{aligned}$$
(6.22)

We can now prove

Theorem 6.4

With the notations used above, if \(y\notin Y\), then for any open set D of y with \(\overline{D}\bigcap Y=\emptyset \), we have

$$\begin{aligned} \sigma _m=O(m^{-\infty })\ \ \text{ on } Y_G\times D. \end{aligned}$$
(6.23)

Let \(x, y\in Y\). If \(\pi (x)\ne \pi (e^{i\theta }\circ y)\), for every \(\theta \in [0,2\pi [\), then there are open sets \(U_G\) of \(\pi (x)\) in \(Y_G\) and V of y in X such that

$$\begin{aligned} \sigma _m=O(m^{-\infty })\ \ \text{ on } U_G\times V. \end{aligned}$$
(6.24)

Let \(p\in \mu ^{-1}(0)\) and let \(x=(x_1,\ldots ,x_{2n+1})\) be the local coordinates as in Remark 4.4. Then,

$$\begin{aligned}&\begin{array}{ll} &{}\sigma _m(\widetilde{x}'',y)=e^{im\Psi (\widetilde{x}'',y'')}\alpha (\widetilde{x}'',y'',m)+O(m^{-\infty })\ \ \text{ on } W\times U,\\ &{}\alpha (\widetilde{x}'',y'',m)\in S^{n-\frac{3}{4}d}_{\mathrm{loc}}\,(1; W\times U, T^{*0,q-r}Y_G\boxtimes (T^{*0,q}X)^*),\\ &{}\alpha (\widetilde{x}'',y'',m)\sim \sum ^\infty _{j=0}m^{n-\frac{3}{4}d-j}\alpha _j(\widetilde{x}'',y'') \text{ in } S^{n-\frac{3}{4}d}_{\mathrm{loc}}\,(1; W\times U, T^{*0,q-r}Y_G\boxtimes (T^{*0,q}X)^*),\\ &{}\alpha _j(\widetilde{x}'',y'')\in C^\infty (W\times U, T^{*0,q-r}Y_G\boxtimes (T^{*0,q}X)^*),\ \ j=0,1,2,\ldots , \end{array} \end{aligned}$$
(6.25)
$$\begin{aligned}&\alpha _0(\widetilde{x}'',\widetilde{x}'')=2^{-n+2d-1}\pi ^{\frac{d}{2}-n-1}\frac{1}{\sqrt{V_{\mathrm{eff}}\,(\widetilde{x}'')}}|\det \,\mathcal {L}_{\widetilde{x}''}||\det \,R_x|^{-\frac{3}{4}}\hat{\tau }_{\widetilde{x}''}\tau _{\widetilde{x}'',n_-},\ \ \forall \widetilde{x}''\in W, \end{aligned}$$
(6.26)

where U is an open set of p, \(W=\Omega _3\times \Omega _4\), \(\Omega _3\) and \(\Omega _4\) are open sets as in the beginning of Sect. 6.1.

Proof

Note that \(S^{(q)}_{G,m}=O(m^{-\infty })\) away Y. From this observation, we get (6.23). Let \(x, y\in Y\). Assume that \(\pi (x)\ne \pi (e^{i\theta }\circ y)\), for every \(\theta \in [0,2\pi [\). Since

$$\begin{aligned} S^{(q)}_{G,m}(x,y)=\frac{1}{2\pi |G|_{d\mu }}\int ^{\pi }_{-\pi }\int _GS^{(q)}(x,e^{i\theta }\circ g\circ y)e^{im\theta }d\mu (g)d\theta \end{aligned}$$

and \(S^{(q)}\) is smoothing away from diagonal, we can integrate by parts with respect to \(\theta \) and deduce that there are open sets \(U_1\) of x in X and \(V_1\) of y in X such that

$$\begin{aligned} S^{(q)}_{G,m}=O(m^{-\infty })\ \ \text{ on } U_1\times V_1. \end{aligned}$$
(6.27)

From (6.20), we see that there are open sets \(\hat{U}_G\) of \(\pi (x)\) in \(Y_G\) and \(\hat{V}_G\) of \(\pi (y)\) in \(Y_G\) such that

$$\begin{aligned} S^{(q-r)}_{Y_G,m}=O(m^{-\infty })\ \ \text{ on } \hat{U}_G\times \hat{V}_G. \end{aligned}$$
(6.28)

From (6.27) and (6.28), we get (6.24).

Fix \(u=(u_1,\ldots ,u_{2n+1})\in Y\bigcap U\). From (6.23) and (6.24), we only need to show that (6.25) and (6.26) hold near u and we may assume that \(u=(0,\ldots ,0,u_{2d+1},\ldots ,u_{2n},0)=\widetilde{\mathring{u}}''\). Let V be a small neighborhood of u. Let \(\chi (\widetilde{\mathring{x}}'')\in C^\infty _0(\Omega _3)\). From (6.1), we can extend \(\chi (\widetilde{\mathring{x}}'')\) to

$$\begin{aligned} Q=\left\{ g\circ e^{i\theta }\circ x;\, (g,e^{i\theta })\in G\times S^1, x\in \Omega _3\right\} \end{aligned}$$

by \(\chi (g\circ e^{i\theta }\circ \widetilde{\mathring{x}}''):=\chi (\widetilde{\mathring{x}}'')\), for every \((g,e^{i\theta })\in G\times S^1\). Assume that \(\chi =1\) on some neighborhood of V. Let \(V_G=\left\{ \pi (x);\, x\in V\right\} \). Let \(\chi _1\in C^\infty _0(Y_G)\) with \(\chi _1=1\) on some neighborhood of \(V_G\) and \(\mathrm{Supp}\,\chi _1\subset \left\{ \pi (x)\in Y_G;\, x\in Y, \chi (x)=1\right\} \). We have

$$\begin{aligned} \begin{array}{cl} \chi _1\sigma _m&{}=m^{-\frac{d}{4}}\chi _1S^{(q-r)}_{Y_G,m}\circ \iota _G\circ \hat{p}\circ \tau _{x,n_-}\circ f\circ \iota ^*\circ S^{(q)}_{G,m}\\ &{}=m^{-\frac{d}{4}}\chi _1S^{(q-r)}_{Y_G,m}\circ \iota _G\circ \hat{p}\circ \tau _{x,n_-}\circ f\circ \iota ^*\circ \chi S^{(q)}_{G,m}\\ &{}\quad +m^{-\frac{d}{4}}\chi _1S^{(q-r)}_{Y_G,m}\circ \iota _G\circ \hat{p}\circ \tau _{x,n_-}\circ f\circ \iota ^*\circ (1-\chi )S^{(q)}_{G,m}. \end{array} \end{aligned}$$
(6.29)

If \(u\in Y\) but \(u\notin \left\{ x\in X;\, \chi (x)=1\right\} \). Since \(\mathrm{Supp}\,\chi _1\subset \left\{ \pi (x)\in X;\, x\in Y, \chi (x)=1\right\} \) and \(\chi (x)=\chi (g\circ e^{i\theta }\circ x)\), for every \((g,e^{i\theta })\in G\times S^1\), for every \(x\in X\), we conclude that \(\pi (e^{i\theta }\circ u)\notin \mathrm{Supp}\,\chi _1\), for every \(e^{i\theta }\in S^1\). From this observation and (6.20), we get

$$\begin{aligned} m^{-\frac{d}{4}}\chi _1S^{(q-r)}_{Y_G,m}\circ \iota _G\circ \hat{p}\circ \tau _{x,n_-}\circ f\circ \iota ^*\circ (1-\chi )S^{(q)}_{G,m}=O(m^{-\infty })\ \ \text{ on } Y_G\times X. \end{aligned}$$
(6.30)

From (6.29) and (6.30), we get

$$\begin{aligned} \chi _1\sigma _m=m^{-\frac{d}{4}}\chi _1S^{(q-r)}_{Y_G,m}\circ \iota _G\circ \hat{p}\circ \tau _{x,n_-}\circ f\circ \iota ^*\circ \chi S^{(q)}_{G,m}+O(m^{\infty })\ \ \text{ on } Y_G\times X. \end{aligned}$$

From (6.22) and Theorem 1.8, we can check that on U,

$$\begin{aligned} \chi _1\sigma _m(\widetilde{x}'',y)=(2\pi )\int e^{im\Psi (\widetilde{x}'',\widetilde{v}'')+im\Psi (v'',y)}\chi _1(\widetilde{x})\beta (\widetilde{x}'',\widetilde{\mathring{v}}'',m)\hat{b}(\widetilde{\mathring{v}}'',y,m)dv(\widetilde{\mathring{v}}'')+O(m^{-\infty }), \end{aligned}$$
(6.31)

where \(\hat{b}(\widetilde{\mathring{v}}'',y,m)=\Bigr (\iota _G\circ \hat{p}\circ \tau _{x,n_-}\circ f\circ \iota ^*\circ \chi (\widetilde{\mathring{v}}'')\circ b\Bigr )(\widetilde{\mathring{v}}'',y,m)\). From (6.31) and Theorem 6.3, we see that (6.25) and (6.26) hold near u. The theorem follows. \(\square \)

Let

$$\begin{aligned} F_m:=\sigma _m^*\sigma _m:\Omega ^{0,q}(X)\rightarrow H^{q}_{b,m}(X)^G,\quad \hat{F}_m:=\sigma _m\sigma ^*_m:\Omega ^{0,q-r}(Y_G)\rightarrow H^{q-r}_{b,m}(Y_G). \end{aligned}$$

Let \(F_m(x,y)\) and \(\hat{F}_m(x,y)\) be the distribution kernels of \(F_m\) and \(\hat{F}_m\) respectively. From Theorems 6.2 and 6.3, we can repeat the proof of Theorem 6.4 with minor change and deduce the following two theorems

Theorem 6.5

With the notations used above, if \(y\notin Y\), then for any open set D of y with \(\overline{D}\bigcap Y=\emptyset \), we have \( F_m=O(m^{-\infty })\ \ \text{ on } X\times D\).

Let \(x, y\in Y\). If \(\pi (x)\ne \pi (e^{i\theta }\circ y)\), for every \(\theta \in [0,2\pi [\), then there are open sets \(D_1\) of x in X and \(D_2\) of y in X such that \( F_m=O(m^{-\infty })\ \ \text{ on } D_1\times D_2\).

Let \(p\in \mu ^{-1}(0)\) and let \(x=(x_1,\ldots ,x_{2n+1})\) be the local coordinates as in Remark 4.4. Then,

$$\begin{aligned} \begin{array}{ll} &{}F_m(x,y)=e^{im\Psi (x'',y'')}a(x'',y'',m)+O(m^{-\infty })\ \ \text{ on } U\times U,\\ &{}a(x'',y'',m)\in S^{n-\frac{d}{2}}_{\mathrm{loc}}\,(1; U\times U, T^{*0,q}X\boxtimes (T^{*0,q}X)^*),\\ &{}a(x'',y'',m)\sim \sum ^\infty _{j=0}m^{n-\frac{d}{2}-j}a_j(\widetilde{x}'',y'') \text{ in } S^{n-\frac{d}{2}}_{\mathrm{loc}}\,(1; U\times U, T^{*0,q}X\boxtimes (T^{*0,q}X)^*),\\ &{}a_j(x'',y'')\in C^\infty (U\times U, T^{*0,q}X\boxtimes (T^{*0,q}X)^*),\ \ j=0,1,2,\ldots , \end{array} \end{aligned}$$

and

$$\begin{aligned} a_0(\widetilde{x}'',\widetilde{x}'')=2^{-3n+4d-1}\pi ^{-n-1}\frac{1}{V_{\mathrm{eff}}\,(\widetilde{x}'')}|\det \,\mathcal {L}_{\widetilde{x}''}||\det \,R_x|^{-\frac{1}{2}}\tau _{\widetilde{x}'',n_-},\ \ \forall \widetilde{x}''\in U\bigcap Y, \end{aligned}$$
(6.32)

where U is an open set of p.

Theorem 6.6

Let \(x, y\in Y\). If \(\pi (x)\ne \pi (e^{i\theta }\circ y)\), for every \(\theta \in [0,2\pi [\), then there are open sets \(D_G\) of \(\pi (x)\) in \(Y_G\) and \(V_G\) of \(\pi (y)\) in \(Y_G\) such that \( \hat{F}_m=O(m^{-\infty })\ \ \text{ on } D_G\times V_G\).

Let \(p\in \mu ^{-1}(0)\) and let \(x=(x_1,\ldots ,x_{2n+1})\) be the local coordinates as in Remark 4.4. Then,

$$\begin{aligned} \begin{array}{ll} &{}\hat{F}_m(x,y)=e^{im\Psi (\widetilde{x}'',\widetilde{y}'')}\hat{a}(\widetilde{x}'',\widetilde{y}'',m)+O(m^{-\infty })\ \ \text{ on } W\times W,\\ &{}\hat{a}(\widetilde{x}'',\widetilde{y}'',m)\in S^{n-d}_{\mathrm{loc}}\,(1; W\times W, T^{*0,q-r}Y_G\boxtimes (T^{*0,q-r}Y_G)^*),\\ &{}\hat{a}(\widetilde{x}'',\widetilde{y}'',m)\sim \sum ^\infty _{j=0}m^{n-d-j}\hat{a}_j(\widetilde{x}'',\widetilde{y}'') \text{ in } S^{n-d}_{\mathrm{loc}}\,(1; W\times W, T^{*0,q-r}Y_G\boxtimes (T^{*0,q-r}Y_G)^*),\\ &{}\hat{a}_j(\widetilde{x}'',\widetilde{y}'')\in C^\infty (W\times W, T^{*0,q-r}Y_G\boxtimes (T^{*0,q-r}Y_G)^*),\ \ j=0,1,2,\ldots ,\\ &{}\hat{a}_0(\widetilde{x}'',\widetilde{x}'')=2^{-3n+\frac{5}{2}d-1}\pi ^{-n+\frac{d}{2}-1}|\det \,\mathcal {L}_{Y_G,\widetilde{x}''}|\hat{\tau }_{\widetilde{x}''},\ \ \forall \widetilde{x}''\in W \end{array} \end{aligned}$$

where \(W=\Omega _3\times \Omega _4\), \(\Omega _3\) and \(\Omega _4\) are open sets as in the beginning of Sect. 6.1.

Let \(R_m:=\frac{1}{C_0}F_m-S^{(q)}_{G,m}:\Omega ^{0,q}(X)\rightarrow H^q_{b,m}(X)^G\), where \(C_0=2^{-3d+3n}\pi ^{\frac{d}{2}}\). Since \(F_m=F_mS^{(q)}_{G,m}\), it is clear that

$$\begin{aligned} \frac{1}{C_0}F_m=S^{(q)}_{G,m}+R_m=S^{(q)}_{G,m}+R_mS^{(q)}_{G,m}=(I+R_m)S^{(q)}_{G,m}. \end{aligned}$$
(6.33)

Our next goal is to show that for m large, \(I+R_m:\Omega ^{0,q}(X)\rightarrow \Omega ^{0,q}(X)\) is injective. From Theorem 6.5 and Theorem 1.8, we see that if \(y\notin Y\), then for any open set D of y with \(\overline{D}\bigcap Y=\emptyset \), we have

$$\begin{aligned} R_m=O(m^{-\infty })\ \ \text{ on } X\times D. \end{aligned}$$
(6.34)

Let \(x, y\in Y\). If \(\pi (x)\ne \pi (e^{i\theta }\circ y)\), for every \(\theta \in [0,2\pi [\), then there are open sets \(D_1\) of x in X and \(D_2\) of y in X such that

$$\begin{aligned} R_m=O(m^{-\infty })\ \ \text{ on } D_1\times D_2. \end{aligned}$$
(6.35)

Let \(p\in \mu ^{-1}(0)\) and let \(x=(x_1,\ldots ,x_{2n+1})\) be the local coordinates as in Remark 4.4. Then,

$$\begin{aligned} \begin{array}{ll} &{}R_m(x,y)=e^{im\Psi (x'',y'')}r(x'',y'',m)+O(m^{-\infty })\ \ \text{ on } U\times U,\\ &{}r(x'',y'',m)\in S^{n-\frac{d}{2}}_{\mathrm{loc}}\,(1; U\times U, T^{*0,q}X\boxtimes (T^{*0,q}X)^*),\\ &{}r(x'',y'',m)\sim \sum ^\infty _{j=0}m^{n-\frac{d}{2}-j}r_j(x'',y'') \text{ in } S^{n-\frac{d}{2}}_{\mathrm{loc}}\,(1; U\times U, T^{*0,q}X\boxtimes (T^{*0,q}X)^*),\\ &{}r_j(x'',y'')\in C^\infty (U\times U, T^{*0,q}X\boxtimes (T^{*0,q}X)^*),\ \ j=0,1,2,\ldots . \end{array} \end{aligned}$$
(6.36)

Moreover, from (6.32) and (1.10), it is easy to see

$$\begin{aligned} |r_0(x,y)|\le C|(x,y)-(x_0,x_0)|, \end{aligned}$$
(6.37)

for all \(x_0\in \mu ^{-1}(0)\bigcap U\), where \(C>0\) is a constant. We need

Lemma 6.7

Let \(p\in \mu ^{-1}(0)\) and let \(x=(x_1,\ldots ,x_{2n+1})\) be the local coordinates as in Remark 4.4 defined in an open set U of p. Let

$$\begin{aligned} \begin{array}{ll} &{}H_m(x,y)=e^{im\Psi (x'',y'')}h(x,y,m)\ \ \text{ on } U\times U,\\ &{}h(x,y,m)\in S^{n-1-\frac{d}{2}}_{\mathrm{loc}}\,(1; U\times U, T^{*0,q}X\boxtimes (T^{*0,q}X)^*),\\ &{}h(x,y,m)\sim \sum ^\infty _{j=0}m^{n-1-\frac{d}{2}-j}h_j( x,y) \text{ in } S^{n-1-\frac{d}{2}}_{\mathrm{loc}}\,(1; U\times U, T^{*0,q}X\boxtimes (T^{*0,q}X)^*),\\ &{}h_j(x,y)\in C^\infty _0(U\times U, T^{*0,q}X\boxtimes (T^{*0,q}X)^*),\ \ j=0,1,2,\ldots . \end{array} \end{aligned}$$

Assume that \(h(x,y,m)\in C^\infty _0(U\times U, T^{*0,q}X\boxtimes (T^{*0,q}X)^*)\). Then, there is a constant \(\hat{C}>0\) independent of m such that

$$\begin{aligned} \left\| H_mu\right\| \le \delta _m\left\| u\right\| ,\ \ \forall u\in \Omega ^{0,q}(X),\ \ \forall m\in \mathbb {N}, \end{aligned}$$
(6.38)

where \(\delta _m\) is a sequence with \(\lim _{m\rightarrow \infty }\delta _m=0\).

Proof

Fix \(N\in \mathbb {N}\). It is not difficult to see that

$$\begin{aligned} \left\| H_mu\right\| \le \left\| (H^*_mH_m)^{2^N}u\right\| ^{\frac{1}{2^{N+1}}}\left\| u\right\| ^{1-\frac{1}{2^{N+1}}},\ \ \forall u\in \Omega ^{0,q}(X), \end{aligned}$$
(6.39)

where \(H^*_m\) denotes the adjoint of \(H_m\). From Theorem 6.2, we can repeat the proof of Theorem 6.4 with minor change and deduce that

$$\begin{aligned} \begin{array}{ll} &{}(H^*_mH_m)^{2^N}(x,y)=e^{im\Psi (x'',y'')}p (x,y,m)+O(m^{-\infty })\ \ \text{ on } U\times U,\\ &{}p(x,y,m)\in S^{n-2^{N+1}-\frac{d}{2}}_{\mathrm{loc}}\,(1; U\times U, T^{*0,q}X\boxtimes (T^{*0,q}X)^*),\\ &{}p(x,y,m)\in C^\infty _0(U\times U, T^{*0,q}X\boxtimes (T^{*0,q}X)^*). \end{array} \end{aligned}$$

Hence,

$$\begin{aligned} |(H^*_mH_m)^{2^N}(x,y)|\le \hat{C}m^{n-2^{N+1}-\frac{d}{2}},\ \ \forall (x,y)\in U\times U, \end{aligned}$$
(6.40)

where \(\hat{C}>0\) is a constant independent of m. Take N large enough so that \(n-2^{N+1}-\frac{d}{2}<0\). From (6.40) and (6.39), we get (6.38). \(\square \)

We also need

Lemma 6.8

Let \(p\in \mu ^{-1}(0)\) and let \(x=(x_1,\ldots ,x_{2n+1})\) be the local coordinates as in Remark 4.4 defined in an open set U of p. Let

$$\begin{aligned} \begin{array}{ll} &{}B_m(x,y)=e^{im\Psi (x'',y'')}g(x,y,m)\ \ \text{ on } U\times U,\\ &{}g(x,y,m)\in S^{n-\frac{d}{2}}_{\mathrm{loc}}\,(1; U\times U, T^{*0,q}X\boxtimes (T^{*0,q}X)^*),\\ &{}g(x,y,m)\sim \sum ^\infty _{j=0}m^{n-\frac{d}{2}-j}g_j( x,y) \text{ in } S^{n-\frac{d}{2}}_{\mathrm{loc}}\,(1; U\times U, T^{*0,q}X\boxtimes (T^{*0,q}X)^*),\\ &{}g_j(x,y)\in C^\infty _0(U\times U, T^{*0,q}X\boxtimes (T^{*0,q}X)^*),\ \ j=0,1,2,\ldots ,\\ &{}g(x,y)\in C^\infty _0(U\times U, T^{*0,q}X\boxtimes (T^{*0,q}X)^*). \end{array} \end{aligned}$$

Suppose that

$$\begin{aligned} |g_0(x,y)|\le C|(x,y)-(x_0,x_0)|, \end{aligned}$$

for all \(x_0\in \mu ^{-1}(0)\bigcap U\), where \(C>0\) is a constant. Then,

$$\begin{aligned} \left\| B_mu\right\| \le \varepsilon _m\left\| u\right\| ,\ \ \forall u\in \Omega ^{0,q}(X),\ \ \forall m\in \mathbb {N}, \end{aligned}$$
(6.41)

where \(\varepsilon _m\) is a sequence with \(\lim _{m\rightarrow \infty }\varepsilon _m=0\).

Proof

Fix \(N\in \mathbb {N}\). It is not difficult to see that

$$\begin{aligned} \left\| B_mu\right\| \le \left\| (B^*_mB_m)^{2^N}u\right\| ^{\frac{1}{2^{N+1}}}\left\| u\right\| ^{1-\frac{1}{2^{N+1}}},\ \ \forall u\in \Omega ^{0,q}(X), \end{aligned}$$
(6.42)

where \(B^*_m\) denotes the adjoint of \(B_m\). From Theorem 6.2, we can repeat the proof of Theorem 6.4 with minor change and deduce that

$$\begin{aligned} \begin{array}{ll} &{}(B^*_mB_m)^{2^N}(x,y)=e^{im\Psi (x'',y'')}\hat{g}(x,y,m)+O(m^{-\infty })\ \ \text{ on } U\times U,\\ &{}\hat{g}(x,y,m)\in S^{n-\frac{d}{2}}_{\mathrm{loc}}\,(1; U\times U, T^{*0,q}X\boxtimes (T^{*0,q}X)^*),\\ &{}\hat{g}(x,y,m)\sim \sum ^\infty _{j=0}m^{n-\frac{d}{2}-j}\hat{g}_j(x,y) \text{ in } S^{n-\frac{d}{2}}_{\mathrm{loc}}\,(1; U\times U, T^{*0,q}X\boxtimes (T^{*0,q}X)^*),\\ &{}\hat{g}_j(x,y)\in C^\infty _0(U\times U, T^{*0,q}X\boxtimes (T^{*0,q}X)^*),\ \ j=0,1,2,\ldots , \\ &{}\hat{g}(x,y,m)\in C^\infty _0(U\times U, T^{*0,q}X\boxtimes (T^{*0,q}X)^*), \end{array} \end{aligned}$$

and

$$\begin{aligned} |\hat{g}_0(x,y)|\le C|(x,y)-(x_0,x_0)|^{2^{N+1}}, \end{aligned}$$
(6.43)

for all \(x_0\in \mu ^{-1}(0)\bigcap U\), where \(C>0\) is a constant. Let

$$\begin{aligned} (B^*_mB_m)^{2^N}_0(x,y)=e^{im\Psi (x'',y'')}\hat{g}_0(x,y,m),\quad (B^*_mB_m)^{2^N}_1(x,y)=e^{im\Psi (x'',y'')}h(x,y,m), \end{aligned}$$

where \(h(x,y,m)=\hat{g}(x,y,m)-\hat{g}_0(x,y,m)\). It is clear that \(h(x,y,m)\in S^{n-1-\frac{d}{2}}_{\mathrm{loc}}\,(1; U\times U, T^{*0,q}X\boxtimes (T^{*0,q}X)^*)\). From Lemma 6.7, we see that

$$\begin{aligned} \left\| (B^*_mB_m)^{2^N}_1u\right\| \le \delta _m\left\| u\right\| ,\ \ \forall u\in \Omega ^{0,q}(X),\ \ \forall m\in \mathbb {N}, \end{aligned}$$
(6.44)

where \(\delta _m\) is a sequence with \(\lim _{m\rightarrow \infty }\delta _m=0\).

From (6.43), we see that

$$\begin{aligned} |\hat{g}_0(x,y)|\le C_1\Bigr (|\hat{x}''|+|\hat{y}''|+|\widetilde{\mathring{x}}''-\widetilde{\mathring{y}}''|\Bigr )^{2^{N+1}}, \end{aligned}$$
(6.45)

where \(C_1>0\) is a constant. From (3.44), we see that

$$\begin{aligned} |\mathrm{Im}\,\Psi (x,y)|\ge c\Bigr (|\hat{x}''|^2+|\hat{y}''|^2+|\widetilde{\mathring{x}}''-\widetilde{\mathring{y}}''|^2\Bigr ), \end{aligned}$$
(6.46)

where \(c>0\) is a constant. From (6.45) and (6.46), we conclude that

$$\begin{aligned} |(B^*_mB_m)^{2^N}_0(x,y)|\le \hat{C}m^{-2^N+n-\frac{d}{2}},\ \ \forall (x,y)\in U\times U, \end{aligned}$$
(6.47)

where \(\hat{C}>0\) is a constant independent of m. From (6.47), we see that if N large enough, then

$$\begin{aligned} \left\| (B^*_mB_m)^{2^N}_0u\right\| \le \hat{\delta }_m\left\| u\right\| ,\ \ \forall u\in \Omega ^{0,q}(X),\ \ \forall m\in \mathbb {N}, \end{aligned}$$
(6.48)

where \(\hat{\delta }_m\) is a sequence with \(\lim _{m\rightarrow \infty }\hat{\delta }_m=0\).

From (6.42), (6.44) and (6.48), we get (6.41). \(\square \)

From (6.34), (6.35), (6.36), (6.37) and Lemma 6.8, we get

Theorem 6.9

With the notations above, we have \(\left\| R_mu\right\| \le \varepsilon _m\left\| u\right\| ,\ \ \forall u\in \Omega ^{0,q}(X),\ \forall m\in \mathbb {N}\), where \(\varepsilon _m\) is a sequence with \(\lim _{m\rightarrow \infty }\varepsilon _m=0\).

In particular, if m is large enough, then the map \(I+R_m:\Omega ^{0,q}(X)\rightarrow \Omega ^{0,q}(X)\) is injective.

Proof of Theorem 1.9

From (6.33) and Theorem 6.9, we see that if m is large enough, then the map \(F_m=\sigma ^*_m\sigma _m: H^q_{b,m}(X)^G\rightarrow H^q_{b,m}(X)^G\) is injective. Hence, if m is large enough, then the map \(\sigma _m:H^q_{b,m}(X)^G\rightarrow H^{q-r}_{b,m}(Y_G)\) is injective and \(\mathrm{dim}\,H^q_{b,m}(X)^G\le \mathrm{dim}\,H^{q-r}_{b,m}(Y_G)\).

Similarly, we can repeat the proof of Theorem 6.9 with minor change and deduce that, if m is large enough, then the map \(\hat{F}_m=\sigma _m\sigma ^*_m: H^{q-r}_{b,m}(Y_G)\rightarrow H^{q-r}_{b,m}(Y_G)\) is injective. Hence, if m is large enough, then the map \(\sigma ^*_m:H^{q-r}_{b,m}(Y_G)\rightarrow H^{q}_{b,m}(X)^G\) is injective. Thus, \(\mathrm{dim}\,H^q_{b,m}(X)^G=\mathrm{dim}\,H^{q-r}_{b,m}(Y_G)\) and \(\sigma _m\) is an isomorphism if m large enough. \(\square \)