1 Introduction

Let X be the boundary of a strictly pseudo-convex domain D in \({\mathbb {C}}^{n+1}\). Then \((X,\,T^{1,0}X)\) is a contact manifold of dimension \(2n+1,\, n\ge 1\), where \(T^{1,0}X\) is the sub-bundle of \(TX\otimes {\mathbb {C}}\) defining the CR structure. We denote by \(\omega _0\in {\mathcal {C}}^{\infty }(X,T^*X)\) the contact 1-form whose kernel is the horizontal bundle \(HX\subset TX\); we refer to Sect. 2.1 for definitions. Associated with these data we can define the Hardy space H(X), it is the space of boundary values of holomorphic functions in D which lie in \(L^2(X)\), the Hilbert space of square integrable functions on X. Suppose a contact and CR action of a t-dimensional torus \({\mathbb {T}}\) is given; we denote by \(\mu : X\rightarrow {\mathfrak {t}}^*\) the associated CR moment map. Fix a weight \(i\,\nu \) in the lattice \(i\,{\mathbb {Z}}^t\subset {\mathfrak {t}}^*\), if \(0\in {\mathfrak {t}}^*\) does not lie in the image of the moment map, the isotypes

$$\begin{aligned}H(X)_{k\nu }=\{f\in H(X):\, (e^{i\, \theta }\cdot f)(x) = e^{ik\,\langle \nu ,\theta \rangle }f(x),\, \theta \in {\mathbb {R}}^t \}, \quad k\in {\mathbb {Z}}, \end{aligned}$$

are finite dimensional.

Suppose that the ray \(i\,{\mathbb {R}}_+\cdot \nu \in {\mathfrak {t}}^*\) is transversal to \(\mu \), then \(X_{\nu }:=\mu ^{-1}(i\,{\mathbb {R}}_+\cdot \nu )\) is a sub-manifold of X of codimension \(t-1\). There is a well-defined locally free action of \({\mathbb {T}}^{t-1}_{\nu }:=\exp _{{\mathbb {T}}}(i\,\ker \nu )\) on \(X_{\nu }\); the resulting orbifold \(X^{\textrm{red}}_{\nu }\) is called conic reduction of X with respect to the weight \(\nu \). Let \(\varphi \) be an Euclidean product on \({\mathfrak {t}}\), we shall also use the symbol \(\langle \cdot ,\,\cdot \rangle \) and denote by \({\lambda }^{\varphi }\in {\mathfrak {t}}\) be uniquely determined by \({\lambda }=\varphi ({\lambda }^{\varphi },\,\cdot )\) and \(\Vert {\lambda }\Vert \) the corresponding norm. By abuse of notation we write \(\lambda \) for \({\lambda }^{\varphi }\) and we identify \({\mathfrak {t}}\cong i {\mathbb {R}}^t\) with its dual. We set

$$\begin{aligned}\ker \nu = \nu ^{\perp }:=\left\{ {\lambda }\in {\mathfrak {t}}:\,\langle \nu ,\, {\lambda } \rangle =0 \right\} .\end{aligned}$$

The locus \(X_{\nu }\) is \({\mathbb {T}}\)-invariant; we will always assume that the action of \({\mathbb {T}}\) on \(X_{\nu }\) is locally free. After replacing \({\mathbb {T}}\) with its quotient by a finite subgroup, we may and will assume without loss of generality that the action is generically free. In Sect. 2.1, we show that \(X^{\textrm{red}}_{\nu }\) is CR manifold with positive definite Levi form of dimension \(2n-2t+3\).

Let us define \({\mathbb {T}}^1_{\nu }:=\exp _{{\mathbb {T}}}(i\,\nu )\), if \(\nu \) is coprime, we have a Lie group isomorphism

$$\begin{aligned}\kappa _{\nu }:\,S^1\rightarrow {\mathbb {T}}^1_{\nu },\quad e^{i\theta }\mapsto e^{i\theta \nu }\end{aligned}$$

between \({\mathbb {T}}^1_{\nu }:=\exp _{{\mathbb {T}}}(i\,\nu )\) and the circle \(S^1\). Let us denote by

$$\begin{aligned} \overline{{\mathbb {T}}^1_{\nu }}:={\mathbb {T}}/ {\mathbb {T}}^{t-1}_{\nu }\cong {\mathbb {T}}^1_{\nu }/ ({\mathbb {T}}^1_{\nu }\cap {\mathbb {T}}^{t-1}_{\nu }), \end{aligned}$$

then the character \(\chi _{\nu }\,:\,{\mathbb {T}}\rightarrow S^1,\,\chi _{\nu }(e^{i\,\theta }):=e^{ik\,\langle \nu ,\,\theta \rangle }\), being trivial on \({\mathbb {T}}^{t-1}_{\nu }\), descends to a character \(\chi _{\nu }':\,\overline{{\mathbb {T}}^1_{\nu }}\rightarrow S^1\) which is a Lie group isomorphism, see [14, Lemma 10]. Thus, we have a locally free circle action of \(\overline{{\mathbb {T}}^1_{\nu }}\) on \(X^{\textrm{red}}_{\nu }\), which induces an action on the Hardy space \(H(X^{\textrm{red}}_{\nu })\). Suppose that the action of \(\overline{{\mathbb {T}}^1_{\nu }}\) on X is transversal to the CR structure. We denote by \(H(X^{\textrm{red}}_{\nu })_k\) the corresponding k-th Fourier component, and we call the action of \(\overline{{\mathbb {T}}^1_{\nu }}\) on \(X^{\textrm{red}}_{\nu }\) residual circle action. The aim of this paper is to prove that \(H(X)_{k\nu }\) and \(H(X^{\textrm{red}}_{\nu })_k\) are isomorphic for k sufficiently large.

We prove the aforementioned result in the more general setting of CR manifolds for spaces of (0, q)-forms when k is large; more precisely we consider (0, q) forms with \(L^2\) coefficients and the corresponding projector \(S^{(q)}\) onto the kernel of the Kohn Laplacian \(H^q(X)\). Now, we make more precise the assumptions on the CR manifold X and on the group action.

Assumption 1.1

Let \((X,\, T^{1,0}X)\) be a compact connected orientable CR manifold of dimension \(2n+1\), \(n\ge 1\), and let \(\omega \) be the associated contact 1-form. The Levi form L 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 \(n_-+n_+=n\).

Concerning the group action, we always assume

Assumption 1.2

The action of \({\mathbb {T}}\) preserves the contact form \(\omega _0\) and the complex structure J. That is, \(g^*\omega _0=\omega _0\) on X and \(g_*J=Jg_*\) on the horizontal bundle HX for every \(g\in {\mathbb {T}}\) where \(g^*\) and \(g_*\) denote the pull-back map and push-forward map of \({\mathbb {T}}\), respectively.

Let \(X^{\textrm{red}}_{\nu }:=X_{\nu }/{\mathbb {T}}^{t-1}_{\nu }\), defined in the same way as before, more precisely we shall assume

Assumption 1.3

The moment map \(\mu \) is transverse to the ray \(i\,{\mathbb {R}}_+\cdot \nu \in {\mathfrak {t}}^*\), the action of \({\mathbb {T}}\) on \(X_{\nu }\) is locally free and for every \(x\in X_{\nu }\)

$$\begin{aligned}\textrm{val}_x(\nu ^{\perp })\cap \textrm{val}_x(\nu ^{\perp })^{\perp _b}=\{0\} \end{aligned}$$

where b is a bilinear form on \(H_xX\) such that

$$\begin{aligned} b(\cdot ,\, \cdot ) = \textrm{d}\omega _0(\cdot ,\, J\cdot ) \end{aligned}$$
(1)

and it is non-degenerate.

We note that \(b(U,\,V)=2\,L(U,\,V)\) for every \(U,\,V\in HX\). By assumptions above, we will show that \(X^{\textrm{red}}_{\nu }\) is a CR manifold with natural CR structure induced by \(T^{1,0}X\) of dimension \(2n - 2(t-1) + 1\). Let \(L_{X^{\textrm{red}}_{\nu }}\) be the Levi form on \(X^{\textrm{red}}_{\nu }\) induced naturally from the Levi form L on X. For a given subspace \({\mathfrak {s}}\) of \({\mathfrak {t}}\), we denote \({\mathfrak {s}}_X\) the subspace of infinitesimal vector fields on X. Let us consider

$$\begin{aligned}B=\ker {\nu }_X\oplus J\ker {\nu }_X.\end{aligned}$$

Hence, b has constant signature on \(B\times B\), suppose b has r negative eigenvalues on \(B\times B\) where \(r \le n_-\) since L and b have the same number of negative eigenvalues on HX. Fix \(q=n_-\); hence, by Lemma 2.1, \(L_{X^{\textrm{red}}_{\nu }}\) has \(q-r\) negative eigenvalues at each point of \({X^{\textrm{red}}_{\nu }}\). We refer to Sect. 2.1 for definitions; we have:

Theorem 1.1

Suppose that \(\Box ^q_b\) has \(L^2\) closed range. Fix a maximal coprime weight \(\nu \ne 0\) in the lattice inside \({\mathfrak {t}}^*\) and assume that the circle action \(\overline{{\mathbb {T}}^1_{\nu }}\) is a transversal CR action. Fix \(q=n_-\), under the assumptions above, \(X^{\textrm{red}}_{\nu }\) is a compact CR manifold with non-degenerate Levi form having \(q-r\)-negative eigenvalues. There is a natural isomorphism of vector spaces \(\sigma _k\,:\, H^q(X)_{k\nu }\rightarrow H^{q-r}(X^{\textrm{red}}_{\nu })_k\) for k sufficiently large.

For strictly pseudoconvex domain, we have \(q=n_-=r=0\) and thus we have quantization commutes with reduction for spaces of functions for k large. We give a proof in Sect. 3, which is inspired from [7] (see [8] for the full extension of [7]), and it is a consequence of the microlocal properties of the projector \(S^{(q)}_{k\nu }\) described in Sect. 2.2 and calculus of Fourier integral operators of complex type, see [11]. Furthermore, we recall that the way to establish the isometry from kernel expansion for k large comes from [10].

The conic reductions defined above appear naturally in geometric quantization. In fact, given a Hamiltonian and holomorphic action with moment map \(\Phi \) of a compact Lie group G on a Hodge manifold \((M,\omega )\) with quantizing circle bundle \(\pi :X\rightarrow M\), one can always define an infinitesimal action of the Lie algebra \({\mathfrak {g}}\) on X. If it can be integrated to an action of the whole group G, then one has a representation of G on H(X). In [5] it was observed that associated to group actions one can define reductions \(M_{\textrm{red}}^{\Theta }\) by pulling back a G-invariant and proper sub-manifold \(\Theta \) of \({\mathfrak {g}}^*\) via the moment map \(\Phi \). When \(\Theta \) is chosen to be a cone through a co-adjoint orbit \(C({\mathcal {O}}_{\nu })\), one has associated reduction whose Hardy space of its quantization is

$$\begin{aligned} H(X_{\textrm{red}}^{C({\mathcal {O}}_{\nu })})= \bigoplus _{k\in {\mathbb {Z}}} H(X_{\textrm{red}}^{C({\mathcal {O}}_{\nu })})_k, \end{aligned}$$

where k labels an irreducible representation of the residual circle action described above. Now, it is natural to ask if this spaces are related to the decomposition induced by G on H(X). When \(G={\mathbb {T}}\) Theorem 1.1 states that they are isomorphic to \(H(X)_{k\nu }\); the canonical decomposition of the Hardy space H(X) of the quantitation by the built-in circle action does not play any role. The semi-classical parameter k is the one induced by the ladder \(k\,\nu \), \(k=0,1,2,\dots \), labeling unitary irreducible representations.

Further geometrical motivations for this theorem are explained in paper [14], where it is proved that \(\delta _k\) is an isomorphism for k large enough in the setting when X is the circle bundle of a polarized Hodge manifold whose Grauert tube is D. Thus, Theorem 1.1 generalizes the main theorem in [14] to compact “quantizable” pseudo-Kähler manifolds. We also refer to [12] for examples and the explicit expression of the leading term of the asymptotic expansion of \(\dim H(X)_{k\nu }\) as k goes to infinity. Along this line of research in [13] Toeplitz operators were studied for circle action; in [1] the study of asymptotics of compositions of Toeplitz operators with quantomopomorphism is addressed for torus actions.

2 Preliminaries

2.1 Geometric setting

We recall some notations concerning CR and contact geometry. 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. There is a unique sub-bundle HX of TX such that \(HX\otimes {\mathbb {C}}=T^{1,0}X \oplus T^{0,1}X\). Let \(J:HX\rightarrow HX\) be the complex structure map given by \(J(u+\overline{u})=i u-i\overline{u}\), for every \(u\in T^{1,0}X\). By complex linear extension of J to \(TX\otimes {\mathbb {C}}\), the i-eigenspace of J is \(T^{1,0}X\). We shall also write (XHXJ) to denote a CR manifold.

Since X is orientable, there always exists a real non-vanishing 1-form \(\omega _0\in {\mathcal {C}}^{\infty }(X,T^*X)\) so that \(\langle \,\omega _0(x),\,u\,\rangle =0\), for every \(u\in H_xX\), for every \(x\in X\); \(\omega _0\) is called contact form and it naturally defines a volume form on X. For each \(x \in X\), we define a quadratic form on HX by

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

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

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

The Hermitian quadratic form \({L}_x\) on \(T^{1,0}_xX\) is called Levi form at x. In the case when X is the circle bundle of an Hodge manifold \((M,\omega )\), the positivity of \(\omega \) implies that the number of negative eigenvalues of the Levi form is equal to n. The Reeb vector field \(R\in {\mathcal {C}}^\infty (X,TX)\) is defined to be the non-vanishing vector field determined by

$$\begin{aligned} \omega _0(R)\equiv 1,\quad \textrm{d}\omega _0(R,\cdot )\equiv 0\ \ \hbox { on}\ TX. \end{aligned}$$

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, \(\langle \,R\,|\,R\,\rangle =1\) and R 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 of \(T^{1,0}X\) and \(T^{0,1}X\), respectively. They can be identified with sub-bundles of the complexified cotangent bundle \({\mathbb {C}}T^*X\).

Assume that X admits an action of t-dimensional torus \({\mathbb {T}}\). In this work, we assume that the \({\mathbb {T}}\)-action preserves \(\omega _0\) and J; that is, \(t^*\omega _0=\omega _0\) on X and \(t_*J=Jt_*\) on HX. Let \({\mathfrak {t}}\) denote the Lie algebra of T, we identify \({\mathfrak {t}}\) with its dual \({\mathfrak {t}}^*\) by means of the scalar product \(\langle \cdot ,\cdot \rangle \). For any \(\xi \in {\mathfrak {t}}\), we write \(\xi _X\) to denote the vector field on X induced by \(\xi \). The moment map associated to the form \(\omega _0\) is the map \(\mu :X \rightarrow {\mathfrak {t}}^*\) such that, for all \(x \in X\) and \(\xi \in {\mathfrak {t}}\), we have

$$\begin{aligned} \langle \mu (x), \xi \rangle = \omega _0(\xi _X(x)). \end{aligned}$$
(2)

Fix a maximal weight \(\nu \ne 0\) in the lattice inside \({\mathfrak {t}}^*\). Suppose that \(i{\mathbb {R}}_+\cdot \nu \) is transversal to \(\mu \), so we have

$$\begin{aligned} {\mathfrak {t}}^*= i{\mathbb {R}}_+\cdot \nu \oplus \textrm{d}_p\mu (T_pX) \end{aligned}$$
(3)

and \(X_{\nu }\) is a sub-manifold of X of codimension \(2n+2-t\). We claim that the action of \({\mathbb {T}}^{t-1}_{\nu }\) on \(X_{\nu }\) is locally free. In fact, by the contrary suppose that there exists \(\xi \in \ker {\nu }\) such that \(\xi _X(x)=0\) on \(T_xX_\nu \), \(x\in X_{\nu }\). For each \(v\in T_xX\), we have

$$\begin{aligned}(\textrm{d}_x\mu (v))(\xi )=\textrm{d}_x\omega _0(\xi _X,v)=0 \end{aligned}$$

which contradicts (3).

The action of \({\mathbb {T}}\) restricts to an action of \({\mathbb {T}}^{t-1}\) whose moment map is given by

$$\begin{aligned}\mu _{\vert {\mathbb {T}}^{t-1}}= p_{\nu }\circ \mu ,\quad \text { where } p_{\nu }:\, {\mathfrak {t}}^* \rightarrow {\mathfrak {t}}_{\nu }^* \end{aligned}$$

is the canonical projection onto the \(t-1\)-dimensional subspace \(\nu ^{\perp }\) in \({\mathfrak {t}}^*\). The transversality condition in 1.2 implies that \(0\in {\mathfrak {t}}_{\nu }^{t-1}\) is a regular value for the moment \(\mu _{\vert {\mathbb {T}}^{t-1}}\). Thus, we have

$$\begin{aligned}X_{\nu }/{\mathbb {T}}^{t-1}_{\nu }=\mu _{\vert {\mathbb {T}}^{t-1}}^{-1}(0)/{\mathbb {T}}^{t-1}_{\nu }\end{aligned}$$

and by assumptions 1.21.3 and [7, Section 2.5] (we shall also refer to [3] for definitions concerning CR structures on orbifolds) we have

Lemma 2.1

The space of orbits \(X_{\nu }/{\mathbb {T}}^{t-1}_{\nu }\) is a CR orbifold. Let us denote by \(\pi : X_{\nu } \rightarrow X^{\textrm{red}}_{\nu }\) and \(\iota : X_{\nu } \hookrightarrow X\) the natural projection and inclusion, respectively, then there is a unique induced contact form \(\omega _0^{\textrm{red}}\) on \(X_{\nu }/{\mathbb {T}}^{t-1}_{\nu }\) such that

$$\begin{aligned}\pi ^{*}\omega _0^{\textrm{red}}=\iota ^*\omega _0. \end{aligned}$$

In particular, set \(HX_{\nu }= TX_{\nu }\cap HX\), we have

$$\begin{aligned}HX= HX_{\nu }\oplus J\,i \nu ^{\perp }_X \quad \text { and }\quad HX_{\nu } = i \nu ^{\perp }_X\oplus \textrm{d}\pi ^{*}HX^{\textrm{red}}_{\nu }. \end{aligned}$$

We will also assume that \(\overline{{\mathbb {T}}^1_{\nu }}\)-action is transversal CR, that is, the infinitesimal vector field

$$\begin{aligned} (\nu _X u)(x)=\frac{\partial }{\partial t}\left( u(\exp (i t\,\nu )\circ x)\right) |_{t=0},\quad \text {for any }u\in C^\infty (X), \end{aligned}$$

preserves the CR structure \(T^{1,0}X\), so that \(\nu _X\) and \(T^{1,0}X\oplus T^{0,1}X\) generate the complex tangent bundle to X,

$$\begin{aligned} {\mathbb {C}}T_xX={\mathbb {C}}\nu _X(x)\oplus {\mathbb {C}}T_x^{1,0}X\oplus {\mathbb {C}}T_x^{0,1}X \qquad (x\in X). \end{aligned}$$

We define local coordinates that will be useful later. Recall that X admits a CR and transversal \(\overline{{\mathbb {T}}^1_{\nu }}\)-action which is locally free on \(X_{\nu }\), \(T\in {\mathcal {C}}^{\infty }(X,\,TX)\) denotes the global real vector field given by this infinitesimal circle action. We will take T to be our Reeb vector field R. In a similar way as in Theorem 3.6 in [7], there exist local coordinates \(v=(v_1,\dots ,v_{t-1})\) of \({\mathbb {T}}^{t-1}_{\nu }\) in a small neighborhood \(V_0\) of the identity e with \(v(e)=(0,\,\dots ,\,0)\), local coordinates \(x=(x_1\,\dots ,x_{2n+1})\) defined in a neighborhood \(U_1\times U_2\) of \(p\in X_{\nu }\), where \(U_1\subseteq {\mathbb {R}}^{t-1}\) (resp. \(U_{2}\subseteq {\mathbb {R}}^{2n+2-t}\)) is an open set of \(0\in {\mathbb {R}}^{t-1}\) (resp. \(0\in {\mathbb {R}}^{2n+2-t}\)) and \(p\equiv 0\in {\mathbb {R}}^{2n+1}\), and a smooth function \(\gamma =(\gamma _1,\dots ,\,\gamma _{t-1})\in {\mathcal {C}}^{\infty }(U_2,U_1)\) with \(\gamma (0)=0\) such that

$$\begin{aligned}&(v_1,\dots ,v_{t-1})\circ (\gamma (x_{t},\dots ,x_{2n+1}),x_{t},\dots ,x_{2n+1}) \\&=(v_1+\gamma _1(x_{t},\dots ,x_{2n+1}),\dots ,\,v_{t-1}+\gamma _d(x_{t},\dots ,x_{2n+1}),\,x_{t},\dots ,x_{2n+1}) \end{aligned}$$

for each \((v_1,\dots ,\,v_{t-1})\in V_0\) and \((x_{t},\dots ,x_{2n+1})\in U_2\). Furthermore, we have

$$\begin{aligned}{\mathfrak {t}}=\textrm{span}\left\{ \partial _{{x}_{j}}\right\} _{j=1,\dots , t-1},\quad \mu ^{-1}(i\,{{\mathbb {R}}}_{\nu }\cdot \nu )\cap U =\{x_{2d-t+1} = \dots =x_{2d}=0 \}, \end{aligned}$$

on \(\mu ^{-1}(i\,{{\mathbb {R}}}_{\nu }\cdot \nu )\cap U\) there exist smooth functions \(a_j\)’s with \(a_j(0)=0\) for every \(0\le j\le t-1\) and independent on \(x_1,\dots ,x_{2(t-1)},\,x_{2n+1}\) such that

$$\begin{aligned}J\left( \partial _{{x}_{j}}\right) =\partial _{{x}_{t-1+j}}+a_j(x)\partial _{{x}_{2n+1}}\qquad j=1,\dots ,t-1, \end{aligned}$$

the Levi form \({L}_p\), the Hermitian metric \(\langle \,\cdot \,|\,\cdot \,\rangle \) and the 1-form \(\omega _0\) can be written

$$\begin{aligned}{L}_p(Z_j,\,{\overline{Z}}_k)=\mu _j\,\delta _{j,k},\qquad \langle Z_j\vert \,{\overline{Z}}_k \rangle =\delta _{j,k}\qquad (1\le j,k\le n), \end{aligned}$$

and

$$\begin{aligned} \omega _0(x)=&(1+O(|x|))\textrm{d}x_{2n+1}+\sum _{j=1}^{t-1} 4\mu _jx_{t-1+j}\textrm{d}x_j +\sum _{j=t}^n 2\mu _jx_{2j}\textrm{d}x_{2j-1} \\&-\sum _{j=t}^n 2\mu _jx_{2j-1}\textrm{d}x_{2j}+\sum _{j=r}^{2n} b_jx_{2n+1}\textrm{d}x_j+O(|x|^2) \end{aligned}$$

where \(b_{r},\dots ,b_{2n}\in {\mathbb {R}}\),

$$\begin{aligned}T_p^{1,0}X=\textrm{span}\{Z_1,\dots ,Z_n\} \end{aligned}$$

and

$$\begin{aligned}Z_j&=\frac{1}{2}\,(\partial _{{x}_{j}}-i\, \partial _{{x}_{t-1+j}})(p)\qquad (j=1,\dots ,t-1)\,,\\ Z_j&=\frac{1}{2}\,(\partial _{{x}_{2j-1}}-i\, \partial _{{x}_{2j}})(p)\qquad (j=t,\dots ,n)\,. \end{aligned}$$

We need to define in local coordinates we just introduced the phase function of the \({\mathbb {T}}^{t-1}_{\nu }\)-invariant Szegő kernel \(\Phi _-(x,y)\in {\mathcal {C}}^\infty (U\times U)\) which is independent of \((x_1,\ldots ,x_{t-1})\) and \((y_1,\ldots ,y_{t-1})\). Hence, we write \(\Phi _-(x,y)=\Phi _-((0,x''),(0,y'')):=\Phi _-(x'',y'')\) and \(\mathring{x}'':=(x_{t},\ldots ,x_{2n})\), \(\mathring{y}'':=(y_{t},\ldots ,y_{2n})\). Moreover, there is a constant \(c>0\) such that

$$\begin{aligned} \mathrm{Im\,}\Phi _-(x'',y'')\ge c\left( |{\mathring{x}''}|^2+|{\mathring{y}''}|^2 +|{\mathring{x}''-\mathring{y}''}|^2\right) ,\ \ \text {for all } ((0,x''),(0,y''))\in U\times U. \end{aligned}$$
(4)

Furthermore,

$$\begin{aligned} \Phi _-&(x'', y'')=-x_{2n+1}+y_{2n+1}+2i\sum ^{t-1}_{j=1}\left| \mu _j\right| y^2_{t-1+j}+2i \sum ^{t-1}_{j=1}\left| \mu _j\right| x^2_{t-1+j} +i\sum ^{n}_{j=t}\left| \mu _j\right| \left| z_j-w_j\right| ^2\nonumber \\&+\sum ^{n}_{j=t}i\mu _j(\overline{z}_jw_j-z_j\overline{w}_j) +\sum ^d_{j=1}(-b_{t-1+j}x_{d+j}x_{2n+1}+b_{t-1+j}y_{t-1+j}y_{2n+1})\nonumber \\&+\sum ^n_{j=t}\frac{1}{2}(b_{2j-1}-ib_{2j})(-z_jx_{2n+1}+w_jy_{2n+1}) +\sum ^n_{j=t}\frac{1}{2}(b_{2j-1}+ib_{2j})(-\overline{z}_jx_{2n+1}+\overline{w}_jy_{2n+1}) \nonumber \\&+(x_{2n+1}-y_{2n+1})f(x, y) +O(\left| (x, y)\right| ^3), \end{aligned}$$
(5)

where \(z_j=x_{2j-1}+ix_{2j}\), \(w_j=y_{2j-1}+iy_{2j}\), \(j=t,\ldots ,n\), \(\mu _j\), \(j=1,\ldots ,n\), f is smooth and satisfies \(f(0,0)=0\), \(f(x, y)=\overline{f}(y, x)\).

We now consider \(\overline{{\mathbb {T}}^1_{\nu }}\) circle action on X. Let \(p\in \mu ^{-1}(i{\mathbb {R}}_+\cdot \nu )\), there exist local coordinates \(v=(v_1,\dots ,v_{t-1})\) of \({\mathbb {T}}^{t-1}\) in a small neighborhood \(V_0\) of e with \(v(e)=(0,\,\dots ,\,0)\), local coordinates \(x=(x_1\,\dots ,x_{2n+1})\) defined in a neighborhood \(U_1\times U_2\) of p, where \(U_1\subseteq {\mathbb {R}}^{t-1}\) (resp. \(U_{t-1}\subseteq {\mathbb {R}}^{2n+t}\)) is an open set of \(0\in {\mathbb {R}}^{t-1}\) (resp. \(0\in {\mathbb {R}}^{2n+t}\)) and \(p\equiv 0\in {\mathbb {R}}^{2n+1}\), and a smooth function \(\gamma =(\gamma _1,\dots ,\,\gamma _t)\in {\mathcal {C}}^{\infty }(U_2,U_1)\) with \(\gamma (0)=0\) such that \(T=-\frac{\partial }{\partial x_{2n+1}}\) and all the properties for the local coordinates defined before hold. The phase function \(\Psi \) satisfies \(\Psi (x,y)=-x_{2n+1}+y_{2n+1}+{\hat{\Psi }}(\mathring{x}'',\mathring{y}'')\), where \({\hat{\Psi }}(\mathring{x}'',\mathring{y}'')\in {\mathcal {C}}^\infty (U\times U)\) and \(\Psi \) satisfies (5).

2.2 Hardy spaces

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

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

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

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

to denote the Hilbert space adjoint of \({\overline{\partial }}_{b}\) in the \(L^2\) space with respect to \((\,\cdot \,|\,\cdot \, )\). There is a well-defined orthogonal projection

$$\begin{aligned} S^{(q)}:L^2_{(0,q)}(X)\rightarrow \mathrm{Ker\,}\Box ^q_b \end{aligned}$$
(6)

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

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

denote the distribution kernel of \(S^{(q)}\). We will write \(H^q(X)\) for \(\ker \Box ^q_b\), and for functions, we simply write H(X) to mean \(H^0(X)\). The distributional kernel \(S^{(q)}(x,y)\) was studied in [6], before recalling an explicit description describing the oscillatory integral defining the distribution \(S^{(q)}(x,y)\) we need to fix some notation. To begin, let us review the concept of the Hörmander symbol space. Let \(D\subset X\) be a local coordinate patch with local coordinates \(x=(x_1,\ldots ,x_{2n+1})\).

Definition 2.1

For every \(m\in {\mathbb {R}}\), we denote with

$$\begin{aligned} S^m_{1,0}(D\times D\times {\mathbb {R}}_+,T^{*0,q}X\boxtimes (T^{*0,q}X)^*)\subseteq {\mathcal {C}}^\infty (D\times D\times {\mathbb {R}}_+,T^{*0,q}X\boxtimes (T^{*0,q}X)^*)\nonumber \\ \end{aligned}$$
(7)

the space of all a 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

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

For simplicity we denote with \(S^m_{1,0}\) the spaces defined in (7); furthermore, we write

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

Let \(a_j\in S^{m_j}_{1,0}\), \(j=0,1,2,\ldots \) with \(m_j\rightarrow -\infty \), as \(j\rightarrow \infty \). Then there exists \(a\in S^{m_0}_{1,0}\) unique modulo \(S^{-\infty }\), such that

$$\begin{aligned} a-\sum ^{k-1}_{j=0}a_j\in S^{m_k}_{1,0}(D\times D\times {\mathbb {R}}_+,T^{*0,q}X\boxtimes (T^{*0,q}X)^*\big ) \end{aligned}$$

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 \(S^{m_0}_{1,0}\).

It is known that the characteristic set of \(\Box ^q_b\) is given by

$$\begin{aligned} \Sigma =\Sigma ^-\cup \Sigma ^+,\quad \Sigma ^-=\left\{ (x,\lambda \omega _0(x))\in T^*X;\,\lambda <0\right\} , \end{aligned}$$

and \(\Sigma ^+\) is defined similarly for \(\lambda >0\). We recall the following theorem (see [6, Theorem 1.2]).

Theorem 2.1

Suppose that the Levi form is non-degenerate and \(\Box ^q_b\) has \(L^2\) closed range. Then, there exist continuous operators \(S_-, S_+: L^2_{(0,q)}(X)\rightarrow \mathrm{Ker\,}\Box ^q_b\) such that

$$\begin{aligned} S^{(q)}=S_-+S_+,\quad S_+\equiv 0\ \ \hbox { if}\ q\ne n_+\end{aligned}$$

and

$$\begin{aligned}\mathrm{WF'\,}(S_-)=\mathrm{diag\,}(\Sigma ^-\times \Sigma ^-),\quad \mathrm{WF'\,}(S_+)=\mathrm{diag\,}(\Sigma ^+\times \Sigma ^+)\ \ \hbox { if}\ q=n_-=n_+, \end{aligned}$$

where \(\mathrm{WF'\,}(S_-)=\left\{ (x,\xi ,y,\eta )\in T^*X\times T^*X;\,(x,\xi ,y,-\eta )\in \mathrm{WF\,}(S_-)\right\} \), \(\mathrm{WF\,}(S_-)\) is the wave front set of \(S_-\) in the sense of Hörmander.

Moreover, consider any small local coordinate patch \(D\subset X\) with local coordinates \(x=(x_1,\ldots ,x_{2n+1})\), then \(S_-(x,y)\), \(S_+(x,y)\) satisfy

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

with

$$\begin{aligned}s_{\mp }(x,y,t)\sim \sum ^\infty _{j=0}s^j_{\mp }(x, y)t^{n-j}\text { in }S^{n}_{1, 0}(D\times D\times {\mathbb {R}}_+,T^{*0,q}X\boxtimes (T^{*0,q}X)^*),\end{aligned}$$

\(s_+(x,y,t)=0\) if \(q\ne n_+\) and \(s^0_-(x,x)\ne 0\) for all \(x\in D\). The phase functions \(\varphi _-\), \(\varphi _+\) satisfy

$$\begin{aligned}\varphi _+, \varphi _-\in {\mathcal {C}}^\infty (D\times D),\ \ \mathrm{Im\,}\varphi _{\mp }(x, y)\ge 0,\,\varphi _-(x, x)=0,\ \ \varphi _-(x, y)\ne 0\ \ \text{ if }\ \ x\ne y,\end{aligned}$$

and

$$\begin{aligned} d_x\varphi _-(x, y)\big |_{x=y}=-\omega _0(x), \ \ d_y\varphi _-(x, y)\big |_{x=y}=\omega _0(x), \,-\overline{\varphi }_+(x, y)=\varphi _-(x,y). \end{aligned}$$

Remark 2.1

Kohn [9] proved that if \(q=n_-=n_+\) or \(\left| n_--n_+\right| >1\) then \(\Box ^q_b\) has \(L^2\) closed range. For a description of the phase function in local coordinates see chapter 8 of part I in [6].

Now we focus on the decomposition induced by the group action on \(H^q(X)\). Fix \(t\in {\mathbb {T}}\) and let \(t^*:\Lambda ^r_x({\mathbb {C}}T^*X)\rightarrow \Lambda ^r_{t^{-1}\circ x}({\mathbb {C}}T^*X)\) be the pull-back map. Since \({\mathbb {T}}\) preserves J, we have \(t^*:T^{*0,q}_xX\rightarrow T^{*0,q}_{g^{-1}\circ x}X\), for all \(x\in X.\) Thus, for \(u\in \Omega ^{0,q}(X)\), we have \(t^*u\in \Omega ^{0,q}(X)\). Put

$$\begin{aligned} \Omega ^{0,q}(X)_{k\nu }:=\left\{ u\in \Omega ^{0,q}(X);\, (e^{i\theta })^*u=e^{i\,k\langle \nu ,\,\theta \rangle }u,\ \ \forall \theta \in {\mathbb {R}}^r\right\} . \end{aligned}$$

Since the Hermitian metric \(\langle \,\cdot \,|\,\cdot \,\rangle \) on \({\mathbb {C}}TX\) is \({\mathbb {T}}\)-invariant, the \(L^2\) inner product \((\,\cdot \,|\,\cdot \,)\) on \(\Omega ^{0,q}(X)\) induced by \(\langle \,\cdot \,|\,\cdot \,\rangle \) is \({\mathbb {T}}\)-invariant. Let \(u\in L^2_{(0,q)}(X)\) and \(t\in {\mathbb {T}}\), we can also define \(t^*u\) in the standard way. We introduce the following notation

$$\begin{aligned}L^2_{(0,q)}(X)_{k\nu }:=\left\{ u\in L^2_{(0,q)}(X);\, (e^{i\theta })^*u=e^{i\,k\langle \nu ,\,\theta \rangle }u,\ \ \forall \theta \in {\mathbb {R}}^r\right\} ,\end{aligned}$$

and put

$$\begin{aligned}(\mathrm{Ker\,}\Box ^q_b)_{k\nu }:=\mathrm{Ker\,}\Box ^q_b\cap L^2_{(0,q)}(X)_{k\nu }.\end{aligned}$$

The equivariant Szegő projection is the orthogonal projection

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

with respect to \((\,\cdot \,|\,\cdot \,)\). Let \(S^{(q)}_{k\nu }(x,y)\in {\mathcal {D}}'(X\times X,T^{*0,q}X\boxtimes (T^{*0,q}X)^*)\) be the distribution kernel of \(S^{(q)}_{k\nu }\). The asymptotic expansion for the distributional kernel of the projector \(S^{(0)}_{k\nu }\) was studied in [12] when X is the quantizing circle bundle of a given Hodge manifold in Heisenberg local coordinates. Using similar ideas as in [7] one can generalize the results for (0, q)-forms in the following theorem, we give a sketch of the proof, which is similar to the proof of [7, Theorem 1.8], here the group G in [7] is \({\mathbb {T}}^{t-1}_{\nu }\), and the circle action in [7] is given by the action of \(\overline{{\mathbb {T}}^1_{\nu }}\).

Since the action is locally free we need to recall some notations from [2]. Since the action is locally free \(\mu ^{-1}(0)/\overline{{\mathbb {T}}^1_{\nu }}\) is an orbifold, let us denote with \(\pi \,:\,\mu ^{-1}(0)\rightarrow \mu ^{-1}(0)/\overline{{\mathbb {T}}^1_{\nu }}\) the projection. Furthermore the action of \(\overline{{\mathbb {T}}^1_{\nu }}\) commutes with the one of \({\mathbb {T}}^{t-1}_{\nu }\); then, we have a smooth locally free action of \({\mathbb {T}}^{t-1}_{\nu }\) on \(\mu ^{-1}(0)/\overline{{\mathbb {T}}^1_{\nu }}\). Given \(y\in X\) and g in the stabilizer \(({\mathbb {T}}^{t-1}_{\nu })_{\pi (y)}\), there exist \(|(\overline{{\mathbb {T}}^1_{\nu }})_y^1 |\) elements \(e^{ i \,\theta _{g,j}}\in \overline{{\mathbb {T}}^1_{\nu }}\) (\(j=1,\dots ,|(\overline{{\mathbb {T}}^1_{\nu }})_y^1 |\)) such that

$$\begin{aligned}g\circ y= e^{- i \,\theta _{g,j}}\circ y. \end{aligned}$$

Thus, all the elements \((e^{ i \,\theta },\,g)\in \overline{{\mathbb {T}}^1_{\nu }}\times {\mathbb {T}}^{t-1}_{\nu } ={\mathbb {T}}\) satisfying

$$\begin{aligned} e^{ i \,\theta }\cdot g\circ y=y. \end{aligned}$$

are of the form \((e^{ i \theta _{g,j}}, g)\) for each \(g\in ({\mathbb {T}}^{t-1}_{\nu })_{\pi (y)}\).

Theorem 2.2

Suppose that \(\Box ^q_b\) has \(L^2\) closed range. Then, there exist continuous operators \(S^-_{k\nu }, S^+_{k\nu }: L^2_{(0,q)}(X)\rightarrow (\mathrm{Ker\,}\Box ^q_b)_{k\nu }\) such that

$$\begin{aligned} S^{(q)}_{k\nu }=S^-_{k\nu }+S^+_{k\nu },\quad S_+\equiv 0\ \ \hbox { if}\ q\ne n_+\end{aligned}$$

Suppose for simplicity that \(q\ne n_+\). If \(q\ne n_-\), then \(S^{(q)}_{k\nu }\equiv O(k^{-\infty })\) on X.

Suppose \(q= n_-\) and let D be an open set in X such that the intersection \(\mu ^{-1}(i\,{\mathbb {R}}_+\cdot \nu )\cap D= \emptyset \). Then \(S^{(q)}_{k\nu }\equiv O(k^{-\infty })\) on D.

Let \(p\in \mu ^{-1}(i\,{\mathbb {R}}_+\cdot \nu )\) and let U a local neighborhood of p with local coordinates \((x_1,\dots \,x_{2n+1})\). Then, if \(q= n_-\), for every fix \(y\in U\), we consider \(S^{(q)}_{k\nu }(x,y)\) as a k-dependent smooth function in x, then

$$\begin{aligned}S^{(q)}_{k\nu }(x,\,y)= \sum _{h\in G_{\pi (y)}} \sum _{j=1}^{|S^1_x|} e^{ik\, \theta _{h,j}} e^{i k\Vert \nu \Vert \,\Psi (x,\,y)}\,b(x,\,y,\,k\Vert \nu \Vert )+O(k^{-\infty }). \end{aligned}$$

for every \(x\in U_y\), where \(U_y\) is a small open neighborhood of y. The phase function \(\Psi (x,\,y)\) is defined in local coordinates in the end of Sect. 2.1; the symbol satisfies

$$\begin{aligned}b(x,\,y,\,k\Vert \nu \Vert )\in S^{n+(1-t)/2}_{\textrm{loc}}(1,\,U\times U,\,T^{*\,(0,q)}X\boxtimes (T^{*\,(0,q)})^*)\, \end{aligned}$$

and the leading term of \(b(x,x,\Vert \nu \Vert )\) is nonzero.

Proof

Suppose \(q=n_-\), on small local neighborhood D of a point \(p\in X_\nu \) we have

$$\begin{aligned} S_{k\nu }^{(q)}(x,\,y)=\frac{1}{(2\,\pi )^t}\int _{-\pi }^{\pi }\dots \int _{-\pi }^{\pi } e^{i k\,\langle \nu ,\,\theta \rangle }\,S^{(q)}( x,\,e^{i \,\theta }\cdot y) \,\textrm{d}\theta \, \end{aligned}$$

where \(\theta =(\theta _1,\dots ,\theta _t)\). It is easy to prove that the oscillatory integral has a rapidly decreasing asymptotic as \(k\rightarrow +\infty \) far away from a local neighborhood of those elements \((e^{ i \,\theta },\,g)\in \overline{{\mathbb {T}}^1_{\nu }}\times {\mathbb {T}}^{t-1}_{\nu } ={\mathbb {T}}\) such that

$$\begin{aligned} e^{ i \,\theta }\cdot g\circ y=y. \end{aligned}$$

We shall consider the case of a local neighborhood of the identity in \({\mathbb {T}}\); we set \(\theta _t\) the variable for circle action of \(\overline{{\mathbb {T}}^1_{\nu }}\) and \((\theta _1,\dots ,\theta _{t-1})\) the variables for \({\mathbb {T}}^{t-1}_{\nu }\). We can then use local coordinates defined in 2.1; we have

$$\begin{aligned} S_{k\nu }^{(q)}(x,\,y)=\frac{1}{2\,\pi }\int _{-\pi }^{\pi } e^{i k\, \Vert \nu \Vert \theta _t -i\,kx_{2n+1}+i\,ky_{2n+1}}\,S_{{\mathbb {T}}^{t-1}}^{(q)}(\mathring{x},\,e^{i \,\theta _t}\cdot \mathring{y}) \,\textrm{d}\theta _t \end{aligned}$$
(8)

where

$$\begin{aligned} S_{{\mathbb {T}}^{t-1}}^{(q)}({x},\,{y})=\frac{1}{(2\,\pi )^{t-1}}\int _{-\pi }^{\pi }\dots \int _{-\pi }^{\pi } \,S^{(q)}( x,\,e^{i \,\textrm{diag}(\theta _1,\dots ,\theta _{t-1})}\cdot y) \,\textrm{d}\theta _1\,\dots \textrm{d}\theta _{t-1}, \end{aligned}$$
(9)

and \(\mathring{x}=(x_1,\dots ,x_n,0)\) and \(\mathring{y}=(y_1,\dots ,y_n,0)\). The action of \({\mathbb {T}}\) restricts to an action of \({\mathbb {T}}^{t-1}\) whose moment map is given by

$$\begin{aligned}\mu _{\vert {\mathbb {T}}^{t-1}}= p_{\nu }\circ \mu ,\quad \text { where } p_{\nu }:\, {\mathfrak {t}}^* \rightarrow {\mathfrak {t}}_{\nu }^* \end{aligned}$$

is the canonical projection onto the \(t-1\)-dimensional subspace \(\nu ^{\perp }\) in \({\mathfrak {t}}^*\). Since we are assuming \(X_{\nu } \ne \emptyset \), we have that \(0\in {\mathfrak {t}}_{t-1}^*\) does lie in the image of the moment map \(\mu _{\vert {\mathbb {T}}^{t-1}}\). The proof follows in a similar way as in Theorem 1.8 in [7] where here we have the action of \(G={\mathbb {T}}^{t-1}\) with moment map \(\mu _{{\mathbb {T}}^{t-1}}\) and the \(k\nu \) Fourier components are the ones induced by the transversal \(\overline{{\mathbb {T}}^1_{\nu }}\)-action; notice that [7, Assumption 1.7] is satisfied. We shall also refer to [2, Theorem 5.4] for the locally free action case. \(\square \)

For ease of notation we introduce a definition to say that an operator behaves microlocally as equivariant Szegő projector we just studied. For simplicity we assume \(q=n_-\).

Definition 2.2

(Equivariant Szegő-type operator) Suppose that \(q=n_-\) and consider \(H: \Omega ^{0,q}(X)\rightarrow \Omega ^{0,q}(X)\) be a continuous operator with distribution kernel

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

We say that H is a complex Fourier integral operator of equivariant Szegő type of order \(n+(1-t)/2 \in {\mathbb {Z}}\) if H is smoothing away \(\mu ^{-1}(i\,{\mathbb {R}}_+\cdot \nu )\) and for given \(p\in \mu ^{-1}(i\,{\mathbb {R}}_+\cdot \nu )\) let D a local neighborhood of p with local coordinates \((x_1,\dots \,x_{2n+1})\). Then, the distributional kernel of H satisfies

$$\begin{aligned}H_{k}(x,y)= e^{i k\,\Psi (x,\,y)}\,a(x,\,y,\,k)+O(k^{-\infty })\quad \text { on }D\end{aligned}$$

where \(a\in S^{n+(1-t)/2}_{1,0}(D\times D\times {\mathbb {R}}_+,T^{*0,q}X\boxtimes (T^{*0,q}X)^*)\), and \(\Psi \) is as in the end of Sect. 2.1.

For \(k=0\), \(H^{(q)}(X)_0\) is the space of \({\mathbb {T}}\)-fixed vectors, by Theorem 1.5 in [7] we have \(\dim H_0^{(q)}(X)< +\infty \) since when \(0\notin \mu (X)\) the projector \(S^{(q)}_{{\mathbb {T}}}\) is smoothing. Now, given \(k_1, k_2\in {\mathbb {Z}}\) with \(k_1, k_2\ne 0\), consider \(f_1\in H^{(q)}(X)_{k_1\nu }\) and \(f_2\in H^{(q)}(X)_{k_2\nu }\), we have \(f_1\cdot f_2 \in H^{(q)}(X)_{(k_1+k_2)\nu }\); we can study \(\dim H(X)_{k\nu }\) for k large; we have

$$\begin{aligned}\dim H^{(q)}(X)_{k\nu } = \int _X S_{k\nu }^{(q)}(x,\,x)\,\textrm{dV}_X(x). \end{aligned}$$

These dimensions can be studied as \(k\rightarrow +\infty \) by using the microlocal properties of the Szegő kernel and Stationary Phase Lemma in a similar way as in Corollary 1.3 in [12] we have \(\dim H^{(q)}(X)_{k\nu }= O(k^{d+1-t})\). So, we have the following generalization of Theorem A.3 in [4], where it is proved for spaces of functions \(H(X)_{k\nu }\).

Lemma 2.2

If \(0\notin \mu (X)\), then \(H^q(X)_{k\nu }\) are finite dimensional.

3 Proof of Theorem 1.1

In this section we shall explain how to prove asymptotic commutativity for quantization and reduction for spaces of (0, q)-forms. We recall that \(X^{\textrm{red}}_{\nu }\) is a CR orbifold whose Levi form is non-degenerate, with \(n_--r\) negative eigenvalues; let us denote with \(S^{(q)}_{\textrm{red}}\) the corresponding Szegő kernel for (0, q)-forms whose k-th Fourier components are the one induced by the \(\overline{{\mathbb {T}}^1_{\nu }}\)-action on \(X^{\textrm{red}}_{\nu }\). We shall recall briefly its microlocal expression by [3, Theorem 1.2]. Now, let us denote by \(e^{i\theta }\cdot \) the transversal and CR locally free \(\overline{{\mathbb {T}}^1_{\nu }}\)-action and we take the Reeb vector field R to be the vector field on X induced by it.

Let \(q=n_--r\), and consider an open set \(U\subset X\), \(p\in U\), and an orbifold chart \(({\widetilde{U}},G_U)\rightarrow U\); we denote by \({\widetilde{x}}\) the coordinates on \({\widetilde{U}}\). For every \(\ell \in {\mathbb {N}}\), put

$$\begin{aligned}X_\ell :=\{x\in X;\, e^{i\theta }x\ne x, \theta \in [0,{2\pi }/{\ell }[, e^{i\,{2\pi }/{\ell }}x=x\}.\end{aligned}$$

With the assumptions and notations used above, assume that \(p\in X_\ell \), for some \(\ell \in {\mathbb {N}}\). We have as \(k\rightarrow +\infty \),

$$\begin{aligned} S^{(q)}_{\textrm{red},k}(x,y)=\sum ^{\ell -1}_{j=0}\sum _{g\in G_U} e^{\frac{2\pi kj}{\ell }}\, e^{ik\, \Psi ({\widetilde{x}}, e^{i\frac{2\pi j}{\ell }}\cdot g\cdot {\widetilde{y}})}b({\widetilde{x}},\,e^{i\frac{2\pi j}{\ell }}\cdot g\cdot {\widetilde{y}},k)+O(k^{-\infty }) \end{aligned}$$
(10)

where the phase function

$$\begin{aligned} \begin{aligned}&\Psi \in \,{C}^\infty ({{\widetilde{U}}}\times {{\widetilde{U}}})\,,\qquad \Psi ({{\widetilde{x}}},{{\widetilde{x}}})=0,\ \ \hbox { for all}\ {{\widetilde{x}}}\in {{\widetilde{U}}},\\&\inf _{e^{i\theta }\in S^1}\left\{ \textrm{dist}^2({{\widetilde{x}}},e^{i\theta }\,{{\widetilde{y}}})\right\} /C\le \mathrm{Im\,}\Psi ({{\widetilde{x}}},{{\widetilde{y}}})\le C\inf _{e^{i\theta }\in S^1}\left\{ \textrm{dist}^2({{\widetilde{x}}},e^{i\theta }\,{{\widetilde{y}}})\right\} \, \end{aligned} \end{aligned}$$

for each \(({{\widetilde{x}}},{{\widetilde{y}}})\in {{\widetilde{U}}}\times {{\widetilde{U}}}\), \(C>1\) is a constant, and the symbol satisfies

$$\begin{aligned} b({{\widetilde{x}}},{{\widetilde{y}}},k)\sim \sum ^{+\infty }_{j=0}b_j({{\widetilde{x}}},{{\widetilde{y}}})\,k^{n-j} \quad \text {in} \quad S^{n-(t-1)}(1;{{\widetilde{U}}}\times {{\widetilde{U}}},T^{*0,q}X\boxtimes (T^{*0,q}X)^*)\, \end{aligned}$$

and \(b_0({{\widetilde{x}}},{{\widetilde{x}}})\) is nonzero.

Let us recall briefly why we have a local chart \(({\widetilde{U}},\,G_U)\rightarrow U\), for ease of notation let us put \(G={\mathbb {T}}^{t-1}_{\nu }\). We recall that, for every \(x\in X_{\nu }\), by the slice theorem a neighborhood of any orbit \({\mathbb {T}}^{t-1}_{\nu }\cdot x=x_0\) is equivariantly diffeomorphic to a neighborhood of the zero section of the associated principal bundle

$$\begin{aligned} G\times _{G_x} N_x, \end{aligned}$$

where \(N_x\) is the normal space to \(G\cdot x\) in \(X_{\nu }\) and \(G_x\) is the stabilizer of x for the action of G, which is finite. Therefore, for some \(\epsilon >0\) and for an open ball \(B_{2{e}}(\epsilon )\subseteq N_x\), one has a homeomorphism \(B_{2{e}}(\epsilon )/G_x \cong {\overline{U}}\) onto some neighborhood of \(x_0\) in \(X^{\textrm{red}}_{\nu }\).

Since \(\nu ^{\perp }_X:=\textrm{val}(\nu ^{\perp })\) is orthogonal to

$$\begin{aligned} H_xX_{\nu }\cap J_x H_xX_{\nu }\quad \text { and } \quad H_xX_{\nu }\cap J_x H_xX_{\nu }\subset (\nu ^{\perp }_X)^{\perp _b} \end{aligned}$$

for every \(x\in X_{\nu }\), we can find a \({\mathbb {T}}\)-invariant orthonormal basis \(\{Z_1,\,\dots ,\, Z_n\}\) of \(T^{*\,0,1}X\) on \(X_{\nu }\) such that for each \(j,\,k=1,\dots n\),

$$\begin{aligned}L_x(Z_j(x),\,{\overline{Z}}_j(x))= \delta _{jk}\,\lambda _j(x), \end{aligned}$$

where

$$\begin{aligned}{} & {} Z_j(x)\in (\nu ^{\perp }_{X,x}+i J \nu ^{\perp }_{X,x})\hbox { for each}\ j=1,\,\dots ,\, t-1, \\{} & {} Z_j(x)\in H_xX_{\nu }\cap J_x(H_xX_{\nu }) \hbox { for each}\ j=r,\,\dots ,\, n. \end{aligned}$$

Let \(\{Z_1^*,\,\dots ,\, Z_n^*\}\) denote the orthonormal basis of \(T^{*\,0,1}X\) on \(X_{\nu }\), dual to \(Z_1,\,\dots ,\, Z_n\). Fix \(s = 0,\, 1,\, 2,\dots ,\, n-r+1\). For \(x \in X_{\nu }\), put

$$\begin{aligned}B_x^{*\,0,s}X=\left\{ \sum _{r\le j_1<\dots <j_s\le n} a_{j_1,\dots ,j_s}Z_{j_1}^*,\,\dots , Z_{j_s}^*;\,a_{j_1,\dots ,j_s}\in {\mathbb {C}},\, \right\} \end{aligned}$$

and let \(B^{*\,0,s}X\) be the vector bundle of \(X_{\nu }\) with fiber \(B_x^{*\,0,s}\), \(x\in X_{\nu }\). Let \(C^{\infty }(X_{\nu }, B^{*\,0,s}X)^{{\mathbb {T}}}\) denote the set of all \({\mathbb {T}}\)-invariant sections of \(X_{\nu }\) with values in \(B_x^{*\,0,s}X\); let

$$\begin{aligned}\iota _{\textrm{red}} :\,C^{\infty }(X_{\nu }, B^{*\,0,s}X)^{{\mathbb {T}}} \rightarrow \Omega ^{0,s}(X^{\textrm{red}}_{\nu }) \end{aligned}$$

be the natural identification. Before defining the map between \(H^q(X)_{k\nu }\) and \(H^{q-r}(X^{\textrm{red}}_{\nu })_{k}\) we need one more piece of notation. We can assume that \(\lambda _1< 0, \cdots , \lambda _{r} < 0\) and also \(\lambda _{t}< 0,\dots ,\, \lambda _{n_--r+t-1} < 0\). For \(x\in X_{\nu }\), set

$$\begin{aligned}\hat{{\mathcal {N}}}(x,n_-)=\{c\,Z_t^*\wedge \,\cdots \,\wedge Z_{n_--r+t-1}^*,c\in {\mathbb {C}} \} \end{aligned}$$

and let \({\hat{p}}\,:\,{\mathcal {N}}(x,n_-)\rightarrow \hat{{\mathcal {N}}}(x,n_-)\) to be

$$\begin{aligned}{\hat{p}}(c\, Z_1^*\wedge \,\cdots \,\wedge Z_r^*\wedge Z_t^*\wedge \,\cdots \,\wedge Z_{n_--r+t-1}^* ):= c\,Z_t^*\wedge \,\cdots \,\wedge Z_{n_--r+t-1}^* \end{aligned}$$

where c is a complex number. Put \(\iota _{\nu } \,:\, X_{\nu } \rightarrow X\) be the natural inclusion and let \(\iota _{\nu }^* \,:\, \Omega ^{0,q}(X)\rightarrow \Omega ^{0,q}(X_{\nu })\) be the pull-back of \(\iota _{\nu }\).

Let \(q = n_-\); now, inspired by [7], we define the map

$$\begin{aligned}\sigma _{k\nu }:\,H^{q}(X)_{k\nu }\rightarrow H^{q-r}(X^{\textrm{red}}_{\nu })_k \end{aligned}$$

given by

$$\begin{aligned} \sigma _{k\nu }(u):=(k\nu )^{(t-1)/4}\,S^{(q-r)}_{\textrm{red},k} \circ \iota _{\textrm{red}}\circ {\hat{p}} \circ \tau _{x,n_-}\circ e \circ \iota ^*_{\nu } \circ S^{(q)}_{k\nu }(u) , \end{aligned}$$
(11)

here, e(x) is a \({\mathbb {T}}\)-invariant smooth function on \(X_{\nu }\) that can be found explicitly. Notice that operators \(S^{(q-r)}_{\textrm{red},k}\) and \(S^{(q)}_{k\nu }\) appearing in (11) are known explicitly. \(S^{(q-r)}_{\textrm{red},k}\) is the k-th Fourier component of the standard Szegő kernel for the CR manifold \(X^{\textrm{red}}_{\nu }\); on the other hand, \(S^{(q)}_{k\nu }\) is described in Sect. 2.2.

Before stating the next theorem we shall specialize the local coordinates defined in Sect. 2.1. Let us consider \(p\in X_{\nu }\) and let \(x=(x_1,\,\dots ,\,x_{2n+1})\) be the local coordinates in an open neighborhood U of p defined in Sect. 2.1. We may assume that \(U=U_1\times U_2 \times U_3\times U_4\), where \(U_1\subset {\mathbb {R}}^{t-1}\), \(U_2\subset {\mathbb {R}}^{t-1}\) are open sets of \(0\in {\mathbb {R}}^{t-1}\), \(U_3\subset {\mathbb {R}}^{2n-2(t-1)}\) is an open set of \(0\in {\mathbb {R}}^{2n-2(t-1)}\) and \(U_4\) is an open set of \(0\in {\mathbb {R}}\). From now on, we can identify \(U_2\) with

$$\begin{aligned} \{(0,\,\dots ,\,0,\,x_{t},\dots ,\,x_{2(t-1)},\,0,\,\dots ,\,0)\in U:\,(x_t,\,\dots ,\,x_{2(t-1)})\in U_2\}\end{aligned}$$

and \(U_3\) with

$$\begin{aligned}\{(0,\,\dots ,\,0,\,x_{2t-1},\dots ,\,x_{2n},\,0)\in U:\,(x_t,\,\dots ,\,x_{2n})\in U_3\}.\end{aligned}$$

For a given orbifold chart \(({\widetilde{U}},\,G_U)\rightarrow U\) we write \({\tilde{U}}_i\) for the corresponding open sets in \({\tilde{U}}\). Eventually we recall that \({\widetilde{x}}'':=(x_{2t-1},\,\dots ,\,x_{2n+1})\).

Theorem 3.1

Under the assumptions above, if \(x\notin X_{\nu }\), then for every sufficiently small open set D of x with \({\overline{D}}\cap X_{\nu }=\emptyset \), we have \(\sigma _{k\nu }=O(k^{-\infty })\) on \(X^{\textrm{red}}_{\nu }\times D\).

Let \(\pi :\,X_{\nu } \rightarrow X^{\textrm{red}}_{\nu }\) the projection. If \(x,y\in X_{\nu }\) and \(\pi (x)\ne \pi (e^{i\theta }\cdot y)\) for every \(e^{i\theta }\in \overline{{\mathbb {T}}^1_{\nu }}\), then there exist open set U of \(\pi (x)\) in \(X^{\textrm{red}}_{\nu }\) and V of y in X such that \(\sigma _{k\nu }=O(k^{-\infty })\) on \(U\times V\).

Eventually, let \(p\in X_{\nu }\cap X_{\ell }\), using the local coordinates defined above, we have

$$\begin{aligned}\sigma _{k\nu }({\tilde{x}}'',y'')= \sum ^{\ell -1}_{j=0}\sum _{g\in G_U} e^{\frac{2\pi kj}{\ell }} e^{ik\, \Psi ({\widetilde{x}}, \,e^{\frac{2\pi j}{\ell }}\cdot g\cdot {\widetilde{y}})} \,\alpha ({\widetilde{x}},\, e^{\frac{2\pi j}{\ell }}\cdot g\cdot {\widetilde{y}},k)+O(k^{-\infty })\quad \text {on }({\tilde{U}}_3\times {\tilde{U}}_4)\times {\tilde{U}}\end{aligned}$$

where

$$\begin{aligned}\alpha ({\tilde{x}}'',y'',k)\in S^{n-\frac{3}{4}(t-1)}_{\textrm{loc}}(1;\,({\tilde{U}}_3\times {\tilde{U}}_4)\times {\tilde{U}},\,T^{*,(0,q-r)}X^{\textrm{red}}_{\nu }\boxtimes (T^{*,(0,q)}X)^*) \end{aligned}$$

and the leading term \(\alpha _0\) in the expansion of \(\alpha \) can be computed explicitly along the diagonal.

Proof

Since, by Theorem 2.2, \(S^{(q)}_{k\nu }\) has rapidly decreasing asymptotics as k goes to infinity away \(X_{\nu }\), we obtain that \(\sigma _k=O(k^{-\infty })\) on \(X^{\textrm{red}}_{\nu }\times D\).

Now, let us prove the second statement. If \(x,y\in X_{\nu }\) and \(\pi (x)\ne \pi (e^{i\theta }\cdot y)\) for every \(e^{i\theta }\in \overline{{\mathbb {T}}^1_{\nu }}\), since \(S^{(q-r)}_{\textrm{red},k}\) is smoothing away from diagonal, then there exist open set U of \(\pi (x)\) in \(X^{\textrm{red}}_{\nu }\) and V of y in X such that \(\chi \,S^{(q-r)}_{\textrm{red},k}\eta = O(k^{-\infty })\) on \(X^{\textrm{red}}_{\nu }\), where \({\chi }\in C^{\infty }_0(U)\) and \(\eta \in C^{\infty }_0(V)\). Furthermore by (8), we see that, for sufficiently small neighborhoods of x and y, we can integrate by parts in \(\textrm{d}\theta _t\) and see that \(S^{(q)}_{k\nu }\) has rapidly decreasing asymptotic. The second statements follow.

Eventually, let \(p\in X_{\nu }\) and consider a small open neighborhood U with coordinates defined as above. Let \(\chi \in C^{\infty }_0({\tilde{U}}_3)\) be a \(G_U\) invariant bump function and assume \(\chi =1\) on some neighborhood of p, it extends naturally to a function on \({\mathbb {T}}\cdot {\tilde{U}}_3\). Let us put \(U^{\sharp } = \{\pi (x):\,x\in U\}\). Let \(\eta \in C^{\infty }_0(X^{\textrm{red}}_{\nu })\) such that \(\eta =1\) on some neighborhood of \(U^{\sharp }\) and

$$\begin{aligned} \textrm{supp}(\eta )\subset \{\pi (x)\in X^{\textrm{red}}_{\nu } :\, x\in X_{\nu },\,\chi (x)=1 \}. \end{aligned}$$

Thus, we get

$$\begin{aligned} \eta \,\sigma _{k\nu }\sim (k\nu )^{(t-1)/4}\,\eta \,S^{(q-r)}_{\textrm{red},k}\circ \iota _{\textrm{red}}\circ {\hat{p}} \circ \tau _{x,n_-}\circ e \circ \iota ^*_{\nu } \circ \chi \,S^{(q)}_{k\nu } . \end{aligned}$$

Now, we can compose the operators and we can use the complex stationary phase formula of [11]; we get the theorem. \(\square \)

In fact, \(\sigma _{k\nu }\) is an isomorphism for k large if we can prove that

$$\begin{aligned}\sigma _{k\nu }:\, H^{q}(X)_{k\nu }\rightarrow H^{q-r}(X^{\textrm{red}}_{\nu })_k\quad \text { and }\quad \sigma _{k\nu }^*:\, H^{q-r}(X^{\textrm{red}}_{\nu })_k\rightarrow H^{q}(X)_{k\nu } \end{aligned}$$

are injective for k large. Notice that, by Theorem 3.1, we have

$$\begin{aligned} \sigma _{k\nu }^*({x}'',{\tilde{y}}'')= \sum ^{\ell -1}_{j=0}\sum _{g\in G_U} e^{\frac{2\pi kj}{\ell }} e^{-i k\,\overline{\Psi ({x}'',e^{\frac{2\pi j}{\ell }}\cdot g\cdot {\tilde{y}}'')}} \,\beta ({x}'',\,e^{\frac{2\pi j}{\ell }}\cdot g\cdot {\tilde{y}}'',k)+O(k^{-\infty }) \end{aligned}$$
(12)

on \({\tilde{U}}\times ({\tilde{U}}_3\times {\tilde{U}}_4) \) where we can check \(\beta _0({\tilde{x}}'',{\tilde{x}}'')=\alpha _0({\tilde{x}}'',{\tilde{x}}'')\).

The injectivity of the map \(\sigma _{k\nu }\) is a consequence of the following theorem which is an adaptation of proof of the main theorem in [7].

Theorem 3.2

There exists a Fourier integral operator \(R_k\) of equivariant Szegő type of degree \(n-(t-1)/2\) such that

$$\begin{aligned} \sigma _{k\nu }^*\sigma _{k\nu } \equiv c_0 \,(1+R_{k})\,S^{(q)}_{k\nu } \end{aligned}$$
(13)

where \(c_0\) is a positive constant and \(1+R_{k}:\,\Omega ^{0,q}(X)\rightarrow \Omega ^{0,q}(X)\) is an injective Fourier integral operator of equivariant Szegő type.

Proof

By Theorem 3.1 and equation (12) we can compose \(\sigma _{k\nu }^*\) and \(\sigma _{k\nu }\) using the complex stationary phase formula of [11]. We can pick e in the definition of \(\sigma _{k\nu }\) so that the leading term of the symbol of \(\sigma _{k\nu }^*\sigma _{k\nu }\) agrees with the one of \(S^{(q)}_{k\nu }\) along the diagonal. Thus we can find an operator

$$\begin{aligned}R_k(x,\,y)= \sum ^{\ell -1}_{j=0}\sum _{g\in G_U} e^{\frac{2\pi kj}{\ell }} e^{ik\,\Psi (x'',e^{\frac{2\pi j}{\ell }}\cdot g\cdot y'')}\,r(x'',\,e^{\frac{2\pi j}{\ell }}\cdot g\cdot y'',k)+O(k^{-\infty })\quad \text {on }{\tilde{U}}\times {\tilde{U}} \end{aligned}$$

such that

$$\begin{aligned}r\in S^{n-(t-1)/2}(1;\,{\tilde{U}}\times {\tilde{U}},\,T^{*(0,q)}X\boxtimes (T^{*(0,q)}X)^*) \end{aligned}$$

and

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

for all \(x_0\in X_{\nu }\cap {\tilde{U}}\), where \(C>0\) is a constant.

The injectivity of the operator \(1+R_{k}\) follows from direct computations of the leading symbol of \(R_k\) which vanishes at \(\textrm{diag}(X_{\nu } \times X_{\nu } )\). In fact, as a consequence of this, by Lemma 6.7 and Lemma 6.8 in [7] we have that \(\Vert R_k u \Vert \le \epsilon _k \Vert u\Vert \) for all \(u \in \Omega ^{0,q}(X)\), for all \(k\in {\mathbb {N}}\) where \(\epsilon _k\) is a sequence with \(\lim _{k\rightarrow +\infty } \epsilon _k = 0\); this in turn implies, if k is large enough, that the map \(1+R_{k}\) is injective. \(\square \)

Analogously, the injectivity of the map \(\sigma _{k\nu }^*\) follows by studying \(\sigma _{k\nu }\sigma _{k\nu }^*\). We do not repeat the proof here since it is an application of the stationary phase formula for Fourier integral operators of complex type, see [11], and it follows in a similar way as above.