1 Introduction and statement of the main results

Let \(X\) be a compact CR manifold of dimension \(2n-1\), \(n\ge 2\). When \(X\) is strongly pseudoconvex and dimension of \(X\) is greater than five, a classical theorem of Boutet de Monvel [3] asserts that \(X\) can be globally CR embedded into \(\mathbb {C}^N\), for some \(N\in {\mathbb {N}}\). For a strongly pseudoconvex CR manifold of dimension greater than five, the dimension of the kernel of the tangential Cauchy–Riemmann operator \(\overline{\partial }_b\) is infinite and we can find many CR functions to embed \(X\) into complex space. When the Levi form of \(X\) has negative eigenvalues, the dimension of the kernel of \(\overline{\partial }_b\) is finite and could be zero and in general, \(X\) can not be globally CR embedded into complex space. Inspired by Kodaira, we introduced in [9] (see also [12]) the idea of embedding CR manifolds by means of CR sections of tensor powers \(L^k\) of a CR line bundle \(L\rightarrow X\). If the dimension of the space \(H^0_b(X,L^k)\) of CR sections of \(L^k\) is large, when \(k\rightarrow \infty \), one should find many CR sections to embed \(X\) into projective space. In analogy to the Kodaira embedding theorem, it is natural to ask if \(X\) can be globally embedded into projective space when it carries a CR line bundle with positive curvature? To understand this question, it is crucial to be able to know if \(\mathrm{dim\,}H^0_b(X,L^k)\sim k^n\), for \(k\) large? The following conjecture was implicit in [12, p.47-48]

Conjecture 1.1

If \(L\) is positive and the Levi form of \(X\) has at least two negative and two positive eigenvalues\(,\) then

$$\begin{aligned} \mathrm{dim\,}H^0_b(X,L^k)\sim k^n, \end{aligned}$$

for \(k\) large.

The difficulty of this conjecture comes from the presence of positive eigenvalues of the curvature of the line bundle and negative eigenvalues of the Levi form of \(X\) and this causes the associated Kohn Laplacian to have no semi-classical spectral gap. This problem is also closely related to the fact that in the global \(L^2\)-estimates for the \(\overline{\partial }_b\)-operator of Kohn–Hörmander, there is a curvature term from the line bundle as well from the boundary and, in general, it is very difficult to control the sign of the total curvature contribution.

In complex geometry, Demailly’s holomorphic Morse inequalities [6] handled the corresponding analytical difficulties in a new way. Inspired by Demailly, we established analogs of the holomorphic Morse inequalities of Demailly for CR manifolds (see [9, Theorem 1.8])

Theorem 1.2

(Hsiao–Marinescu, 2009) We assume that the Levi form of \(X\) has at least two negative and two positive eigenvalues. Then\(,\) as \(k\rightarrow \infty ,\)

$$\begin{aligned}&-\mathrm{dim\,}H^0_b(X, L^k)+\mathrm{dim\,}H^1_b(X, L^k)\nonumber \\&\quad \leqslant \frac{k^{n}}{2(2\pi )^{n}}\left( -\int _X\int _{\mathbb {R}_{\phi (x),0}} \left| \det (M^\phi _x+s{\mathcal L}_x)\right| \mathrm{d}s\,\mathrm{d}v_X(x)\right. \nonumber \\&\left. \qquad +\int _X\int _{\mathbb {R}_{\phi (x),1}}\left| \det (M^\phi _x +s{\mathcal L}_x)\right| \mathrm{d}s\,\mathrm{d}v_X(x)\right) +o(k^n), \end{aligned}$$
(1.1)

where \(M^\phi _x\) is the associated curvature of \(L\) at \(x\in X\) \((\)see Definition 1.9), \(H^1_b(X, L^k)\) denotes the first \(\overline{\partial }_b\) cohomology group with values in \(L^k,\) \(\mathrm{d}v_X(x)\) is the volume form on \(X,\) \({\mathcal L}_x\) denotes the Levi form of \(X\) at \(x\in X,\) and for \(x\in X,\) \(q=0,1,\)

$$\begin{aligned}&\mathbb {R}_{\phi (x),q}=\{s\in \mathbb {R};\, M^\phi _x+s{\mathcal L}_x\, \mathrm{has\, exactly}\, q \,\mathrm{negative\, eigenvalues} \nonumber \\&\quad \quad \quad \quad \quad \quad \mathrm{and}\, n-1-q \,\mathrm{positive\, eigenvalues}\}. \end{aligned}$$
(1.2)

From (1.1), we see that if

$$\begin{aligned} \int _X\int _{\mathbb {R}_{\phi (x),0}}\left| \det (M^\phi _x+s{\mathcal L}_x)\right| \mathrm{d}s\,\mathrm{d}v_X(x)\!>\! \int _X\int _{\mathbb {R}_{\phi (x),1}}\left| \det (M^\phi _x+s{\mathcal L}_x)\right| \mathrm{d}s\,\mathrm{d}v_X(x)\qquad \end{aligned}$$
(1.3)

then \(L\) is big that is \(\mathrm{dim\,}H^0_b(X,L^k)\sim k^n\). This is a very general criterion and it is desirable to refine it in some cases where (1.3) is not easy to verify. The problem still comes from the presence of positive eigenvalues of \(M^\phi _x\) and negative eigenvalues of \({\mathcal L}_x\).

For the better understanding, let’s see a simple example. We consider compact analogs of the Heisenberg group \(H_n\). Let \(\lambda _1,\ldots ,\lambda _{n-1}\) be given non-zero integers. We assume that \(\lambda _1<0,\ldots ,\lambda _{n_-}<0,\lambda _{n_-+1}>0,\ldots ,\lambda _{n-1}>0\). Let \({\fancyscript{C}}H_n=(\mathbb {C}^{n-1}\times \mathbb {R})/_\sim \), where \((z, \theta )\sim (\widetilde{z}, \widetilde{\theta })\) if

$$\begin{aligned}&\widetilde{z}-z=(\alpha _1,\ldots ,\alpha _{n-1})\in \sqrt{2\pi }{\mathbb {Z}}^{n-1}+i\sqrt{2\pi }{\mathbb {Z}}^{n-1},\\&\widetilde{\theta }-\theta -i\sum ^{n-1}_{j=1}\lambda _j(z_j\overline{\alpha }_j-\overline{z}_j\alpha _j)\in \pi {\mathbb {Z}}. \end{aligned}$$

We can check that \(\sim \) is an equivalence relation and \({\fancyscript{C}}H_n\) is a compact manifold of dimension \(2n-1\). The equivalence class of \((z, \theta )\in \mathbb {C}^{n-1}\times \mathbb {R}\) is denoted by \([(z, \theta )]\). For a given point \(p=[(z, \theta )]\), we define the CR structure \(T^{1, 0}_p{\fancyscript{C}}H_n\) of \({\fancyscript{C}}H_n\) to be the space spanned by \(\left\{ \frac{\partial }{\partial z_j}+i\lambda _j\overline{z}_j\frac{\partial }{\partial \theta },\ \ j=1,\ldots ,n-1\right\} \). Then, \(({\fancyscript{C}}H_n,T^{1,0}_p{\fancyscript{C}}H_n)\) is a compact CR manifold of dimension \(2n-1\). With a suitable choose of a Hermitian metric on the complexified tangent bundle of \({\fancyscript{C}}H_n\), the Levi form of \({\fancyscript{C}}H_n\) at \(p\in {\fancyscript{C}}H_n\) is given by \({\mathcal L}_p=\sum ^{n-1}_{j=1}\lambda _j\mathrm{d}z_j\wedge \mathrm{d}\overline{z}_j\). Let \(L=(\mathbb {C}^{n-1}\times \mathbb {R}\times \mathbb {C})/_{\equiv }\) where \((z,\theta ,\eta )\equiv (\widetilde{z}, \widetilde{\theta }, \widetilde{\eta })\) if

$$\begin{aligned} (z,\theta )\sim (\widetilde{z}, \widetilde{\theta }),\quad \widetilde{\eta }=\eta \exp \left( \sum ^{n-1}_{j,t=1}\mu _{j,t}(z_j\overline{\alpha }_t+ \frac{1}{2}\alpha _j\overline{\alpha }_t)\right) ,\quad \text {for } (\alpha _1,\ldots ,\alpha _{n-1})=\widetilde{z}-z, \end{aligned}$$

where \(\mu _{j,t}=\mu _{t,j}\), \(j, t=1,\ldots ,n-1\), are given integers. We can check that \(\equiv \) is an equivalence relation and \(L\) is a CR line bundle over \({\fancyscript{C}}H_n\). For \((z, \theta , \eta )\in \mathbb {C}^{n-1}\times \mathbb {R}\times \mathbb {C}\), we denote \([(z, \theta , \eta )]\) its equivalence class. Take the pointwise norm \( \big |[(z, \theta , \eta )]\big |^2_{h^L}:=\left| \eta \right| ^2\exp \big (-\textstyle \sum ^{n-1}_{j,t=1}\mu _{j,t}z_j\overline{z}_t\big ) \) on \(L\). Then, the associated curvature of \(L\) is given by \(M^\phi _x=\sum ^{n-1}_{j,t=1}\mu _{j,t}\mathrm{d}z_j\wedge \mathrm{d}\overline{z}_t,\ \ \forall x\in {\fancyscript{C}}H_n\). In this simple example, Conjecture 1.1 becomes

Question 1.3

If \(n_-\ge 2\), \(n-1-n_-\ge 2\), and the matrix \(\left( \mu _{j,t}\right) ^{n-1}_{j,t=1}\) is positive definite, then \(\mathrm{dim\,}H^0_b({\fancyscript{C}}H_n,L^k)\sim k^n\)?

If \(\mu _{j,t}=\left| \lambda _j\right| \delta _{j,t}\), \(j,t=1,\ldots ,n-1\), and \(n_-\ge 2\), \(n-1-n_-\ge 2\), where \(\delta _{j,t}=1\) if \(j=t\), \(\delta _{j,t}=0\) if \(j\ne t\), then it is easy to see that \(\mathbb {R}_{\phi (x),1}=\emptyset \), where \(\mathbb {R}_{\phi (x),1}\) is given by (1.2). Combining this observation with Morse inequalities for CR manifolds [see (1.1)], we get

Theorem 1.4

If \(n_-\ge 2,\) \(n-1-n_-\ge 2,\) and \(\mu _{j,t}=\left| \lambda _j\right| \delta _{j,t},\) \(j,t=1,\ldots ,n-1,\) then \(\mathrm{dim\,}H^0_b({\fancyscript{C}}H_n,L^k)\sim k^n.\)

The assumptions in Theorem 1.4 are somehow restrictive. It is clear that we cannot go much further from Morse inequalities.Using Morse inequalities to approach Conjecture 1.1, we always have to impose extra conditions linking the Levi form and the curvature of the line bundle \(L\). Similar problems also appear in the works of Marinescu [12, 13] and Berman [2] where they studied the \(\overline{\partial }\)-Neumann cohomology groups associated to a high power of a given holomorphic line bundle on a compact complex manifold with boundary. To get many holomorphic sections, they also have to assume that, close to the boundary, the curvature of the line bundle is adapted to the Levi form of the boundary. In this work, by carefully studying semi-classical behavior of microlocal Fourier transforms of the extremal functions for the spaces of lower energy forms of the associated Kohn Laplacian, we could solve Conjecture 1.1 under rigidity conditions on \(X\) and \(L\) without any extra condition linking the Levi form of \(X\) and the curvature of \(L\). As an application, we solve Question 1.3 completely. The proof of our main result presents a new way to overcome the analytic difficulty mentioned in the discussion after Conjecture 1.1 under rigidity conditions. Using this new method, it is possible to remove the assumptions linking the curvatures of the line bundle and the boundary in the works of Marinescu [12, 13] and Berman [2] under rigidity conditions on the boundary and the line bundle.

The rigidity conditions we used in this work are inspired by the work of Baouendi–Rothschild–Treves [1]. They introduced rigidity condition on CR structure and proved that such a manifold can always be locally CR embedded in complex space as a generic submanifold. From their work, rigidity condition on CR structure seems suitable for our purpose. Initially, it is reasonable to first assume that \(X\) can be locally embedded and study global embeddability of \(X\). We can expect that the curvature of the line bundle and its transition functions have to satisfy some rigidity conditions (see Definition 1.7 and Definition 1.12). Moreover, with these geometric conditions, it is possible to establish a micolocal asymptotic expansion of the Szegö kernel and extend Kodaira embedding theorem to this situation.

The geometric objects introduced in this paper form large classes of CR manifolds and CR line bundles. We hope that these geometric objects will be interesting for CR geometers and will be useful in CR geometry.

1.1 Some standard notations

We shall use the following notations: \(\mathbb {R}\) is the set of real numbers, \(\overline{\mathbb {R}}_+:=\left\{ x\in \mathbb {R};\, x\ge 0\right\} \), \({\mathbb {N}}=\left\{ 1,2,\ldots \right\} \), \({\mathbb {N}}_0={\mathbb {N}}\bigcup \left\{ 0\right\} \). An element \(\alpha =(\alpha _1,\ldots ,\alpha _n)\) of \({\mathbb {N}}_0^n\) will be called a multiindex and the length of \(\alpha \) is: \(\left| \alpha \right| =\alpha _1+\cdots +\alpha _n\). We write \(x^\alpha =x_1^{\alpha _1}\cdots x^{\alpha _n}_n\), \(x=(x_1,\ldots ,x_n)\), \(\partial ^\alpha _x=\partial ^{\alpha _1}_{x_1}\cdots \partial ^{\alpha _n}_{x_n}\), \(\partial _{x_j}=\frac{\partial }{\partial x_j}\), \(D^\alpha _x=D^{\alpha _1}_{x_1}\cdots D^{\alpha _n}_{x_n}\), \(D_x=\frac{1}{i}\partial _x\), \(D_{x_j}=\frac{1}{i}\partial _{x_j}\). Let \(z=(z_1,\ldots ,z_n)\), \(z_j=x_{2j-1}+ix_{2j}\), \(j=1,\ldots ,n\), be coordinates of \(\mathbb {C}^n\). We write \(z^\alpha =z_1^{\alpha _1}\cdots z^{\alpha _n}_n\), \(\overline{z}^\alpha =\overline{z}_1^{\alpha _1}\cdots \overline{z}^{\alpha _n}_n\), \(\frac{\partial ^{\left| \alpha \right| }}{\partial z^\alpha }=\partial ^\alpha _z=\partial ^{\alpha _1}_{z_1}\cdots \partial ^{\alpha _n}_{z_n}\), \(\partial _{z_j}= \frac{\partial }{\partial z_j}=\frac{1}{2}\left( \frac{\partial }{\partial x_{2j-1}}-i\frac{\partial }{\partial x_{2j}}\right) \), \(j=1,\ldots ,n\). \(\frac{\partial ^{\left| \alpha \right| }}{\partial \overline{z}^\alpha }=\partial ^\alpha _{\overline{z}}=\partial ^{\alpha _1}_{\overline{z}_1}\cdots \partial ^{\alpha _n}_{\overline{z}_n}\), \(\partial _{\overline{z}_j}= \frac{\partial }{\partial \overline{z}_j}=\frac{1}{2}\left( \frac{\partial }{\partial x_{2j-1}}+i\frac{\partial }{\partial x_{2j}}\right) \), \(j=1,\ldots ,n\).

Let \(\Omega \) be a \(C^\infty \) paracompact manifold. We let \(T\Omega \) and \(T^*\Omega \) denote the tangent bundle of \(\Omega \) and the cotangent bundle of \(\Omega \), respectively. The complexified tangent bundle of \(\Omega \) and the complexified cotangent bundle of \(\Omega \) will be denoted by \(\mathbb {C}T\Omega \) and \(\mathbb {C}T^*\Omega \), respectively. We write \(\langle \,\cdot ,\cdot \,\rangle \) to denote the pointwise duality between \(T\Omega \) and \(T^*\Omega \). We extend \(\langle \,\cdot ,\cdot \,\rangle \) bilinearly to \(\mathbb {C}T\Omega \times \mathbb {C}T^*\Omega \). Let \(E\) be a \(C^\infty \) vector bundle over \(\Omega \). The fiber of \(E\) at \(x\in \Omega \) will be denoted by \(E_x\). Let \(F\) be another vector bundle over \(\Omega \). We write \(E\boxtimes F\) to denote the vector bundle over \(\Omega \times \Omega \) with fiber over \((x, y)\in \Omega \times \Omega \) consisting of the linear maps from \(E_x\) to \(F_y\).

1.2 Generalized Sasakian CR manifolds and generalized Sasakian CR line bundles

Let \((X,T^{1,0}X)\) be a CR manifold of dimension \(2n-1\), \(n\geqslant 2\), where \(T^{1,0}X\) is a CR structure of \(X\). That is, \(T^{1,0}X\) is a complex \(n-1\) dimensional subbundle of the complexified tangent bundle \(\mathbb {C}TX\), satisfying \(T^{1,0}X\bigcap T^{0,1}X=\left\{ 0\right\} \), where \(T^{0,1}X=\overline{T^{1,0}X}\), and \([{\mathcal {V}},{\mathcal {V}}]\subset {\mathcal {V}}\), where \({\mathcal {V}}=C^\infty (X,T^{1,0}X)\). In this section, we denote \(Y:=X\times \mathbb {R}\) and we write \(t\) to denote the standard coordinate of \(\mathbb {R}\). We need

Definition 1.5

We say that \((X,T^{1,0}X)\) is a generalized Sasakian CR manifold if there exists an integrable almost complex structure \(J:TY\rightarrow TY\), \(\mathbb {C}TY\rightarrow \mathbb {C}TY\), such that \(Ju=iu\), \(\forall u\in T^{1,0}X\).

Let \((X,T^{1,0}X)\) be a CR manifold of dimension \(2n-1\), \(n\geqslant 2\), and let \(J:TY\rightarrow TY\), \(\mathbb {C}TY\rightarrow \mathbb {C}TY\), be an almost complex structure. We say that \(J\) is a canonical complex structure on \(Y\) if \(J\) is integrable and \(Ju=iu\), \(\forall u\in T^{1,0}X\). Thus, \((X,T^{1,0}X)\) is a generalized Sasakian CR manifold if and only if there exists a canonical complex structure on \(Y\).

Let \((X,T^{1,0}X)\) be a generalized Sasakian CR manifold and let \(J:TY\rightarrow TY\), \(\mathbb {C}TY\rightarrow \mathbb {C}TY\) be any canonical complex structure on \(Y\). From the Newlander–Nirenberg theorem, \(J\) defines a complex structure \(T^{1,0}Y\supset T^{1,0}X\). Put \(T=J\frac{\partial }{\partial t}\). Then, \(T\in C^\infty (X,TX)\), \(T\) is a global real vector field on \(X\). Since \(J\) is integrable, it is easy to see that

$$\begin{aligned}&\mathbb {C}TX=T^{1,0}X\oplus T^{0,1}X\oplus \left\{ \lambda T;\, \lambda \in \mathbb {C}\right\} ,\nonumber \\&[T,{\mathcal {V}}]\subset {\mathcal {V}},\quad {\mathcal {V}}:=C^\infty (X,T^{1,0}X). \end{aligned}$$
(1.4)

Conversely, let \((X,T^{1,0}X)\) be a CR manifold of dimension \(2n-1\), \(n\geqslant 2\). We assume that there exists a global real vector field \(T\in C^\infty (X,\mathbb {C}TX)\) such that (1.4) hold. Then, one can define a canonical complex structure on \(Y\) by the rule:

$$\begin{aligned}&J:TY\rightarrow TY,\quad \mathbb {C}TY\rightarrow \mathbb {C}TY\\&Ju=iu,\quad \forall u\in T^{1,0}X,\qquad Jv=-iv,\quad \forall v\in T^{0,1}X,\qquad J\frac{\partial }{\partial t}=T. \end{aligned}$$

Thus, \((X,T^{1,0}X)\) is a generalized Sasakian CR manifold if and only if there exists a global real vector field \(T\in C^\infty (X,\mathbb {C}TX)\) such that (1.4) hold. We call \(T\) a rigid global real vector field.

Let’s see some examples

Example 1.6

(I) Let \(M\) be an open subset with \(C^\infty \) boundary \(\partial M\) of a complex manifold \(M'\) of dimension \(n\). If for every \(x_0\in \partial M\), we can find local holomorphic coordinates \((z_1,\ldots ,z_n)\) defined in some neighborhood of \(x_0\), such that near \(x_0\), \(\partial M\) is given by the equation

$$\begin{aligned} \mathrm{Im\,}z_n=f(z_1,\ldots ,z_{n-1}),\quad f\in C^\infty \hbox { is real valued}, \end{aligned}$$

then \(\partial M\) is a generalized Sasakian CR manifold of dimension \(2n-1\).

(II) Let \(M\) be a complex manifold and \((E,h^E)\) be a holomorphic Hermitian line bundle on \(M\), where the Hermitian fiber metric on \(E\) is denoted by \(h^E\). Let \((E^*,h^{E^*})\) be the dual bundle of \(E\). We denote

$$\begin{aligned} G:=\left\{ v\in L^*;\, \left| v\right| _{h^{L^*}}<1\right\} ,\quad \partial G=\left\{ v\in L^*;\, \left| v\right| _{h^{L^*}}=1\right\} . \end{aligned}$$

The domain \(G\) is called Grauert tube associated to \(E\). It is easy to see that \(\partial G\) is a generalized Sasakian CR manifold.

(III) The hypersurface

$$\begin{aligned} \left\{ (z_1,\ldots ,z_n)\in \mathbb {C}^n;\, \sum ^n_{j=1}\lambda _j\left| z_j\right| ^2=R\right\} \end{aligned}$$

is a generalized Sasakian CR manifolds, where \(\lambda _j\in \mathbb {R}\), \(j=0,1,\ldots ,n\), \(R\in \mathbb {R}\).

(IV) Heisenberg groups and compact Heisenberg groups (see Sect. 9.1) are generalized Sasakian CR manifolds.

From now on, we assume that \((X,T^{1,0}X)\) is a compact generalized Sasakian CR manifold and we let \(\pi :Y\rightarrow X\) denote the standard projection.

Definition 1.7

Let \(L\) be a complex line bundle over \(X\). \((L,J)\) is a generalized Sasakian CR line bundle over \(X\), where \(J\) is a canonical complex structure on \(Y\) if the pull back \(\pi ^*L\) is a holomorphic line bundle over \(Y\) with respect to \(J\).

We need

Definition 1.8

Let \(T\in C^\infty (X,TX)\) be a rigid global real vector field on \(X\). Let \(U\Subset X\) be an open set. A function \(u\in C^\infty (U)\) is said to be a \(T\)-rigid CR function on \(U\) if \(Tu=0\) and \(Zu=0\) for all \(Z\in C^\infty (U,T^{0,1}X)\).

From now on, we let \((L,J)\) be a generalized Sasakian CR line bundle over \(X\) and we fix \(T=J\frac{\partial }{\partial t}\). \(T\) is a rigid global real vector field. Since \(\pi ^*L\) is a holomorphic line bundle over \(Y\) with respect to the canonical complex structure \(J\) on \(Y\), it is easy to see that \(X\) can be covered with open sets \(U_j\) with trivializing sections \(s_j\), \(j=1,2,\ldots \), such that the corresponding transition functions are \(T\)-rigid CR functions. In this paper, when trivializing sections \(s\) are used, we will assume that they are of this special form.

Fix a Hermitian fiber metric \(h^L\) on \(L\) and we will denote by \(\phi \) the local weights of the Hermitian metric \(h^L\). More precisely, if \(s\) is a local trivializing section of \(L\) on an open subset \(D\subset X\), then the local weight of \(h^L\) with respect to \(s\) is the function \(\phi \in C^\infty (D,\mathbb {R})\) for which

$$\begin{aligned} \left| s(x)\right| ^2_{h^L}=\mathrm{e}^{-\phi (x)},\quad x\in D. \end{aligned}$$
(1.5)

We write \(h^{\pi ^*L}\) to denote the pull back of \(h^L\) by the projection \(\pi \). Then, \(h^{\pi ^*L}\) is a Hermitian fiber metric on the holomorphic line bundle \(\pi ^*L\). Let \(R^{\pi ^*L}\) be the canonical curvature induced by \(h^{\pi ^*L}\). Let \(\overline{\partial }_J\) and \(\partial _J\) be the \((0,1)\) and \((1,0)\) part of the exterior differential operator \(d\) on functions with respect to \(J\). If \(s\) is a local trivializing section of \(L\) on an open subset \(D\subset X\), \(\left| s\right| ^2_{h^L}=\mathrm{e}^{-\phi (x)}\), then

$$\begin{aligned} R^{\pi ^*L}(y)=\partial _J\overline{\partial }_J\phi (\pi (y))\quad \hbox {on }D\times \mathbb {R}. \end{aligned}$$
(1.6)

We need

Definition 1.9

For \(p\in X\), we define the Hermitian quadratic form \(M^\phi _p\) on \(T^{1,0}_pX\) by

$$\begin{aligned} M^\phi _p(U, \overline{V})=\Big \langle U\wedge \overline{V}, R^{\pi ^*L}(y)\Big \rangle ,\quad \pi (y)=p,\ U, V\in T^{1,0}_pX. \end{aligned}$$
(1.7)

Remark 1.10

Let \(s\) be a local trivializing section of \(L\) on an open subset \(D\subset X\) and \(\phi \) the corresponding local weight as in (1.5). Let \(\overline{\partial }_b\) denote the tangential Cauchy–Riemann operator on functions (see [4, Chapter 7]). It is not difficult to see that for every \(p\in D\), we have

$$\begin{aligned} M^\phi _p(U, \overline{V})=\frac{1}{2}\Big \langle U\wedge \overline{V}, d\big (\overline{\partial }_b\phi -\partial _b\phi \big )(p)\Big \rangle ,\quad U, V\in T^{1,0}_pX, \end{aligned}$$
(1.8)

where \(d\) is the usual exterior derivative and \(\overline{\partial _b\phi }=\overline{\partial }_b\overline{\phi }\).

For \(p\in X\), let \({\mathcal L}_p\) be the Levi form (with respect to \(T\)) at \(p\) (see Definition 1.14, for the precise meaning).

Definition 1.11

We say that \(h^L\) is positive at \(x_0\in X\) if the Hermitian quadratic form \(M^\phi _{x_0}\) is positive, \(h^L\) is semi-positive if there is a positive constant \(\delta >0\) such that for every \(x\in X\) and \(s\in [-\delta , \delta ]\), the Hermitian quadratic form \(M^\phi _x+2s{\mathcal L}_x\) is semi-positive.

Since the transition functions are \(T\)-rigid CR functions, we can check that \(T\phi \) is a well-defined global smooth function on \(X\).

Definition 1.12

\(h^L\) is said to be a \(T\)-rigid Hermitian fiber metric on \((L,J)\) if

$$\begin{aligned} T\phi =C\hbox { on }X,\quad \hbox {for some constant } C, \end{aligned}$$
(1.9)

where \(\phi \) denotes the corresponding local weight as in (1.5).

Note that the constant \(C\) in (1.9) can be non-zero. (See Sect. 9.1).

Definition 1.13

We say that \((L,J,h^L)\) is a rigid generalized Sasakian CR line bundle over X if \((L,J)\) is a generalized Sasakian CR line bundle over \(X\) and \(h^L\) is a \(T\)-rigid Hermitian fiber metric on \((L,J)\), \(T=J\frac{\partial }{\partial t}\).

1.3 Hermitian CR geometry and the main results

Fix a smooth Hermitian metric \(\langle \,\cdot \,|\,\cdot \,\rangle \) on \(\mathbb {C}TX\) so that \(T^{1,0}X\) is pointwise orthogonal to \(T^{0,1}X\), \(T\) is pointwise orthogonal to \(T^{1,0}X\oplus T^{0,1}X\), \(\langle T|T\rangle :=\left\| T\right\| ^2=1\) and \(\langle u|v\rangle \) is real if \(u\), \(v\) are real tangent vectors.

Define

$$\begin{aligned}&T^{*1,0}X:=\left\{ e\in \mathbb {C}T^*X;\, \langle e,u\rangle =0,\forall u\in T^{0,1}X\oplus \left\{ \lambda T;\, \lambda \in \mathbb {C}\right\} \right\} ,\\&T^{*0,1}X:=\left\{ f\in \mathbb {C}T^*X;\, \langle f,v\rangle =0,\forall v\in T^{1,0}X\oplus \left\{ \lambda T;\, \lambda \in \mathbb {C}\right\} \right\} . \end{aligned}$$

\(T^{*1,0}X\) and \(T^{*0,1}X\) are subbundles of the complexified cotangent bundle \(\mathbb {C}T^*X\). Define the vector bundle of \((0, q)\) forms of \(X\) by \(\Lambda ^{0, q}T^*X:=\Lambda ^{q}T^{*0,1}X\). Let \(D\subset X\) be an open set. Let \(\Omega ^{0,q}(D)\) denote the space of smooth sections of \(\Lambda ^{0,q}T^*X\) over \(D\). Similarly, if \(E\) is a vector bundle over \(D\), then we let \(\Omega ^{0,q}(D, E)\) denote the space of smooth sections of \(\Lambda ^{0,q}T^*X\otimes E\) over \(D\). Let \(\Omega ^{0,q}_0(D, E)\) be the subspace of \(\Omega ^{0,q}(D, E)\) whose elements have compact support in \(D\). Let

$$\begin{aligned} \overline{\partial }_b:\Omega ^{0,q}(X)\rightarrow \Omega ^{0,q+1}(X) \end{aligned}$$
(1.10)

be the tangential Cauchy–Riemann operator (see [4, Chapter 7]).

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 \(\Lambda ^{0, q}T^*X\) the bundle of \((0, q)\) forms of \(X\). We shall also denote all these induced metrics by \(\langle \,\cdot \,|\,\cdot \,\rangle \). For \(f\in \Omega ^{0,q}(X)\), we denote the pointwise norm \(\left| f(x)\right| ^2:=\langle f(x)|f(x)\rangle \). Locally, there is a real \(1\)-form \(\omega _0\) of length one which is orthogonal to \(T^{*1,0}X\oplus T^{*0,1}X\). The form \(\omega _0\) is unique up to the choice of sign. We choose \(\omega _0\) so that \(\langle T, \omega _0\rangle =-1\). Therefore, \(\omega _0\) is uniquely determined. We call \(\omega _0\) the uniquely determined global real \(1\)-form. We have the pointwise orthogonal decompositions:

$$\begin{aligned} \mathbb {C}T^*X&= T^{*1,0}X\oplus T^{*0,1}X\oplus \left\{ \lambda \omega _0;\, \lambda \in \mathbb {C}\right\} , \nonumber \\ \mathbb {C}TX&= T^{1,0}X\oplus T^{0,1}X\oplus \left\{ \lambda T;\,\lambda \in \mathbb {C}\right\} . \end{aligned}$$
(1.11)

We recall

Definition 1.14

For \(p\in X\), the Levi form \({\mathcal L}_p\) is the Hermitian quadratic form on \(T^{1,0}_pX\) defined as follows. For any \(U,\ V\in T^{1,0}_pX\), pick \({\mathcal {U}},{\mathcal {V}}\in C^\infty (X, T^{1,0}X)\) such that \({\mathcal {U}}(p)=U\), \({\mathcal {V}}(p)=V\). Set

$$\begin{aligned} {\mathcal L}_p(U,\overline{V})=\frac{1}{2i}\big \langle \big [{\mathcal {U}},\overline{{\mathcal {V}}}\,\big ](p),\omega _0(p)\big \rangle , \end{aligned}$$
(1.12)

where \(\big [{\mathcal {U}},\overline{{\mathcal {V}}}\,\big ]={\mathcal {U}}\ \overline{{\mathcal {V}}}-\overline{{\mathcal {V}}}\ {\mathcal {U}}\) denotes the commutator of \({\mathcal {U}}\) and \(\overline{{\mathcal {V}}}\). Note that \({\mathcal L}_p\) does not depend on the choices of \({\mathcal {U}}\) and \({\mathcal {V}}\).

Since \({\mathcal L}_p\) is a Hermitian form there is a local orthonormal basis \(\{{\mathcal {U}}_1,\ldots ,{\mathcal {U}}_{n-1}\}\) of \(T^{1,0}X\) with respect to \(\langle \,\cdot \,|\,\cdot \,\rangle \) such that \({\mathcal L}_p\) is diagonal in this basis, \({\mathcal L}_p({\mathcal {U}}_j,\overline{{\mathcal {U}}}_t)=\delta _{j,t}\lambda _j(p)\), \(j,t=1,\ldots ,n-1\), \(\delta _{j,t}=1\) if \(j=t\), \(\delta _{j,t}=0\) if \(j\ne t\), \(\lambda _j(p)\in \mathbb {R}\), \(j=1,\ldots ,n-1\). The diagonal entries \(\{\lambda _1(p),\ldots ,\lambda _{n-1}(p)\}\) are called the eigenvalues of the Levi form at \(p\in X\) with respect to \(\langle \,\cdot \,|\,\cdot \,\rangle \).

Given \(q\in \{0,\ldots ,n-1\}\), the Levi form is said to satisfy condition \(Y(q)\) at \(p\in X\), if \({\mathcal L}_p\) has at least either \(\max {(q+1, n-q)}\) eigenvalues of the same sign or \(\min {(q+1,n-q)}\) pairs of eigenvalues with opposite signs. Note that the sign of the eigenvalues does not depend on the choice of the metric \(\langle \,\cdot \,|\,\cdot \,\rangle \).

Let \(L^k\), \(k>0\), be the \(k\)-th tensor power of the line bundle \(L\). We write \(\overline{\partial }_{b,k}\) to denote the tangential Cauchy–Riemann operator acting on forms with values in \(L^k\), defined locally by:

$$\begin{aligned} \overline{\partial }_{b,k}:\Omega ^{0,q}(X, L^k)\rightarrow \Omega ^{0,q+1}(X, L^k),\quad \overline{\partial }_{b,k}(s^ku):=s^k\overline{\partial }_bu, \end{aligned}$$
(1.13)

where \(s\) is a local trivialization of \(L\) on an open subset \(D\subset X\) and \(u\in \Omega ^{0,q}(D)\). We obtain a \(\overline{\partial }_{b,k}\)-complex \((\Omega ^{0,\bullet }(X, L^k),\overline{\partial }_{b,k})\) with cohomology

$$\begin{aligned} H^{\bullet }_b(X,L^k):=\ker \overline{\partial }_{b,k}/\mathrm{Im } \overline{\partial }_{b,k}. \end{aligned}$$
(1.14)

We assume that \(X\) is compact and \(Y(0)\) holds. By [11, 7.6-7.8], [7, 5.4.11-12], [4, Props. 8.4.8-9] and [8, Chapter 6], condition \(Y(0)\) implies that \(\dim H^0_b(X,L^k)<\infty \).

Our main result is the following

Theorem 1.15

Let \((X,T^{1,0}X)\) be a compact generalized Sasakian CR manifold of dimension \(2n-1,\) \(n\geqslant 2\) and let \((L,J,h^L)\) be a rigid generalized Sasakian CR line bundle over \(X\). Assume that \(h^L\) is semi-positive and positive at some point of \(X\). Suppose conditions \(Y(0)\) and \(Y(1)\) hold at each point of \(X\). Then\(,\) for \(k\) large\(,\) there is a constant \(c>0\) independent of \(k,\) such that

$$\begin{aligned} \mathrm{dim\,}H^0_b(X,L^k)\ge ck^n. \end{aligned}$$

It should be mentioned that the Levi curvature assumptions in Theorem 1.15 are a bit more general than the ones in Conjecture 1.1.

Remark 1.16

It should be mentioned that Theorem 1.15 implies the famous Grauert–Riemenschneider conjecture in complex geometry. Let \(M\) be a compact complex manifold of complex dimension \(n\) and let \(E\rightarrow M\) be a holomorphic line bundle with a Hermitian fiber metric \(h^E\). Let \(R^E\) denotes the canonical curvature on \(E\) induced by \(h^E\). We assume that \(R^E\) is semi-positive and positive at some point of \(M\). Then, Grauert–Riemenschneider conjecture claims that \(L\) is big, that is, \(\mathrm{dim\,}H^0(M, E^k)\sim k^n\), where \(H^0(M,E^k)\) denotes the space of global holomorphic sections of \(E^k\) the \(k\)-th power of \(E\). This conjecture was first solved by Siu [14]. Let’s see how to obtain this conjecture from Theorem 1.15. With the notations used above, let \((\widetilde{X},T^{1,0}\widetilde{X})\) be a compact generalized Sasakian CR manifold of dimension \(2m-1\), \(m\geqslant 2\), such that the Levi form of \(\widetilde{X}\) has at least two negative and two positive eigenvalues and let \((\widetilde{L},\widetilde{J},h^{\widetilde{L}})\) be a rigid generalized Sasakian CR line bundle over \(\widetilde{X}\) with \(h^{\widetilde{L}}\) is positive at every point of \(\widetilde{X}\). We can find such \((\widetilde{X},T^{1,0}\widetilde{X})\) and \((\widetilde{L},\widetilde{J},h^{\widetilde{L}})\) (see Sect. 9). Consider \(X=M\oplus \widetilde{X}\), \(T^{1,0}X:=T^{1,0}M\oplus T^{1,0}\widetilde{X}\), where \(T^{1,0}M\) denotes the holomorphic tangent bundle of \(M\). Then, \((X,T^{1,0}X)\) is a compact generalized Sasakian CR manifold of dimension \(2(m+n)-1\) and the Levi form of \(X\) has at least two negative and two positive eigenvalues. Thus, conditions \(Y(0)\) and \(Y(1)\) hold at each point of \(X\). Put \(L:=E\otimes \widetilde{L}\). Then, \(L\) is a complex line bundle over \(X\). Let \(J\) be the canonical complex structure on \(X\times \mathbb {R}\) induced by \(\widetilde{J}\) and the complex structure on \(M\). It is obviously that \((L,J)\) is a generalized Sasakian CR line bundle over \(X\). Put \(h^{L}=h^{E}\otimes h^{\widetilde{L}}\). Then, \(h^L\) is a Hermitian fiber metric on \(L\) and \((L,J,h^L)\) is a rigid generalized Sasakian CR line bundle over \(X\). Moreover, it is easy to check that \(h^L\) is semi-positive and positive at some point of \(X\). From Theorem 1.15, we conclude that for \(k\) large, there is a constant \(C_0>0\) such that

$$\begin{aligned} \mathrm{dim\,}H^0_b(X,L^k)\ge C_0k^{n+m}. \end{aligned}$$
(1.15)

We notice that \(\mathrm{dim\,}H^0_b(X,L^k)=\mathrm{dim\,}H^0(M,E^k)\times \mathrm{dim\,}H^0_b(\widetilde{X},\widetilde{L}^k)\) and it is well known that there is a constant \(C_1>0\) such that \(\mathrm{dim\,}H^0_b(\widetilde{X},\widetilde{L}^k)\le C_1k^{m}\) (see [9, Theorem 1.5]). Combining this observation and (1.15), we conclude that there is a constant \(c>0\) such that \(\mathrm{dim\,}H^0(M,E^k)\ge ck^{n}\).

We investigate Theorem 1.15 on generalized torus CR manifolds. Let \(\Phi ^t(x)\) be the \(T\)-flow. That is, \(\Phi ^t(x)\) is a differentiable mapping:

$$\begin{aligned} t\rightarrow \Phi ^t(x)\in X:I\rightarrow X, \end{aligned}$$

\(I\) is an open interval in \(\mathbb {R}\), \(0\in I\), such that \(\Phi ^0(x)=x\), \(\forall x\in X\), and \(\frac{\mathrm{d}\Phi ^t(x)}{\mathrm{d}t}=T(\Phi ^t(x))\). We need

Definition 1.17

We say that \((X,T^{1,0}X)\) is a generalized torus CR manifold if there is a constant \(\gamma _0>0\) such that \(\Phi ^t(x)\) is well defined, \(\forall \left| t\right| \le \gamma _0\), \(\forall x\in X\), and \(\Phi ^{\gamma _0}(x)=x\) for every \(x\in X\).

Definition 1.18

We say that \((L,J)\) is an admissible generalized Sasakian CR line bundle over a compact generalized torus CR manifold \(X\) if we can find an open covering \(\left\{ U_j\right\} ^N_{j=1}\) of \(X\) such that \(L\) is trivial on \(U_j\), for each \(j\), and

$$\begin{aligned} \left\{ \Phi ^t(x);\, x\in U_j, \left| t\right| \le \gamma _0\right\} =U_j, \end{aligned}$$

for each \(j\), where \(\gamma _0>0\) is as in Definition 1.17.

Let \((L,J)\) be an admissible generalized Sasakian CR line bundle over a compact generalized torus CR manifold \((X,T^{1,0}X)\). Take any Hermitian fiber metric \(h^{L}\) on \(L\) and let \(\phi \) denotes the corresponding local weight as in (1.5). Let \(h^{L}_1\) be the Hermitian fiber metric on \(L\) locally given by \(\left| s\right| ^2_{h^{L}_1}=\mathrm{e}^{-\phi _1}\), where \(\phi _1=\frac{1}{\gamma _0}\int ^{\gamma _0}_0\phi (\Phi ^t(x))\mathrm{d}t\), \(\gamma _0>0\) is as in Definition 1.17, \(s\) is a local trivializing section of \(L\) with the special form in Definition 1.18. It is easy to check that \(h^{L}_1\) is well defined and \(T\phi _1=0\). Thus, \((L,J,h^{L}_1)\) is a rigid generalized Sasakian CR line bundle over \((X,T^{1,0}X)\). Moreover, we can show that if \(M^\phi _x\) is positive on \(X\) then \(M^{\phi _1}_x\) is positive on \(X\) (see Proposition 3.3, for the proof). Combining this with Theorem 1.15, we obtain

Theorem 1.19

Let \((X,T^{1,0}X)\) be a compact generalized torus CR manifold of dimension \(2n-1,\) \(n\geqslant 2\) and let \((L,J)\) be an admissible generalized Sasakian CR line bundle over \(X\) with a Hermitian fiber metric \(h^L\). We assume that \(h^L\) is positive on \(X\) and conditions \(Y(0)\) and \(Y(1)\) hold at each point of \(X.\) Then\(,\) for \(k\) large\(,\) there is a constant \(c>0\) independent of \(k,\) such that

$$\begin{aligned} \mathrm{dim\,}H^0_b(X,L^k)\ge ck^n. \end{aligned}$$

1.4 The outline of the proof of Theorem 1.15

Let \(\Box ^{(q)}_{b,k}\) denote the Kohn Laplacian with values in \(L^k\) (see Sect. 2). Fix \(q=0,1,\ldots ,n-1\). We assume that \(Y(q)\) holds. It is well known that \(\Box ^{(q)}_{b,k}\) has a discrete spectrum, each eigenvalues occurs with finite multiplicity and all eigenforms are smooth and \(\mathrm{Ker\,}\Box ^{(q)}_{b,k}:=\fancyscript{H}^q_b(X,L^k)\cong H^q_b(X,L^k)\). For \(\lambda \ge 0\), let \(\fancyscript{H}^q_{b,\le \lambda }(X,L^k)\) denote the space spanned by the eigenforms of \(\Box ^{(q)}_{b,k}\) whose eigenvalues are bounded by \(\lambda \). Now, we assume that \(Y(0)\) and \(Y(1)\) hold and \((L,h^L)\) is semi-positive and positive at some point of \(X\). Take \(\delta _0>0\) be a small constant so that \(M^\phi _x+2s{\mathcal L}_x\ge 0\), \(\forall \left| s\right| \le \delta _0\), \(\forall x\in X\). Take \(\psi (\eta )\in C^\infty _0(]-\delta _0,\delta _0[,\overline{\mathbb {R}}_+)\) so that \(\psi (\eta )=1\) if \(-\frac{\delta _0}{2}\le \eta \le \frac{\delta _0}{2}\). Take \(\chi (t)\in C^\infty _0(]-2,2[,\overline{\mathbb {R}}_+)\) so that \(0\le \chi (t)\le 1\) and \(\chi (t)=1\) if \(-1\le t\le 1\) and \(\chi (-t)=\chi (t)\) for all \(t\in \mathbb {R}\). Fix \(M>0\). Under the rigidity assumptions in Theorem 1.15, we can construct global continuous operators \(Q^{(0)}_{M,k}:C^\infty (X,L^k)\rightarrow C^\infty (X,L^k)\) and \(Q^{(1)}_{M,k}:\Omega ^{0,1}(X,L^k)\rightarrow \Omega ^{0,1}(X,L^k)\) such that

$$\begin{aligned} \overline{\partial }_{b,k}Q^{(0)}_{M,k}=Q^{(1)}_{M,k}\overline{\partial }_{b,k}\quad \hbox {on } C^\infty (X,L^k) \end{aligned}$$
(1.16)

and \(Q^{(0)}_{M,k}\), \(Q^{(1)}_{M,k}\) are formally given by the following. Let \(s\) be a local section of \(L\) on \(D\subset X\), \(\left| s\right| ^2_{h^L}=\mathrm{e}^{-\phi }\), and let \(\Phi ^t(x)\) be the \(T\)-flow. Then,

$$\begin{aligned} \left( Q^{(0)}_{M,k}f\right) (x)&= s^k\mathrm{e}^{\frac{k}{2}\phi (x)}\int \mathrm{e}^{-it\eta }\psi (\eta )\chi \left( \frac{t}{M}\right) \mathrm{e}^{-\frac{k}{2} \phi (\Phi ^{\frac{t}{k}}(x))}\widetilde{f}(\Phi ^{\frac{t}{k}}(x))\mathrm{d}t\mathrm{d}\eta \quad \hbox {on }D,\nonumber \\ \left( Q^{(1)}_{M,k}g\right) (x)&= s^k\mathrm{e}^{\frac{k}{2}\phi (x)}\int \mathrm{e}^{-it\eta }\psi (\eta )\chi \left( \frac{t}{M}\right) \mathrm{e}^{-\frac{k}{2} \phi (\Phi ^{\frac{t}{k}}(x))}\widetilde{g}(\Phi ^{\frac{t}{k}}(x))\mathrm{d}t\mathrm{d}\eta \quad \hbox {on }D,\nonumber \\ \end{aligned}$$
(1.17)

where \(f=s^k\widetilde{f}\in C^\infty _0(D,L^k)\), \(g=s^k\widetilde{g}\in \Omega ^{0,1}_0(D,L^k)\). (See Sect. 5, for the precise definitions of the operators \(Q^{(0)}_{M,k}\), \(Q^{(1)}_{M,k}\).) Let \(\langle \,\cdot \,|\,\cdot \,\rangle _{h^{L^k}}\) denote the Hermitian metric on \(\Lambda ^{0,q}T^*X\otimes L^k\) induced by \(h^{L}\) and \(\langle \,\cdot \,|\,\cdot \,\rangle \). Let \(\mathrm{d}v_X=\mathrm{d}v_X(x)\) be the volume form on \(X\) induced by \(\langle \,\cdot \,|\,\cdot \,\rangle \) and let \((\,\cdot \,|\,\cdot \,)_{h^{L^k}}\) be the \(L^2\) inner product on \(\Omega ^{0,q}(X,L^k)\) induced by \(\langle \,\cdot \,|\,\cdot \,\rangle _{h^{L^k}}\) and \(\mathrm{d}v_X\). For \(\lambda \ge 0\), define

$$\begin{aligned}&\left( Q^{(0)}_{M,k}\varPi ^{(0)}_{k,\,\le \lambda }\right) (x):=\sum ^{m_k}_{j=1}\langle Q^{(0)}_{M,k}f_j(x)| f_j(x)\rangle _{h^{L^k}}, \nonumber \\&\left( Q^{(1)}_{M,k}\varPi ^{(1)}_{k,\,\le \lambda }\overline{Q^{(1)}_{M,k}})(x\right) :=\sum ^{p_k}_{j=1}\langle Q^{(1)}_{M,k}g_j(x)| Q^{(1)}_{M,k}g_j(x)\rangle _{h^{L^k}}, \end{aligned}$$
(1.18)

where \(f_j(x)\in C^\infty (X, L^k)\), \(j=1,\ldots ,m_k\), is an orthonormal frame for the space \(\fancyscript{H}^0_{b,\,\le \lambda }(X, L^k)\) with respect to \((\,\cdot \,|\,\cdot \,)_{h^{L^k}}\), \(g_j(x)\in \Omega ^{0,1}(X, L^k)\), \(j=1,\ldots ,p_k\), is an orthonormal frame for \(\fancyscript{H}^1_{b,\,\le \lambda }(X, L^k)\) with respect to \((\,\cdot \,|\,\cdot \,)_{h^{L^k}}\). It is straightforward to see that the definitions (1.18) are independent of the choices of orthonormal frames. The point of our proof is that there exists a sequence \(\nu _k>0\) with \(\nu _k\rightarrow 0\) as \(k\rightarrow \infty \), such that

$$\begin{aligned}&\hbox {For each }x\in X, \lim _{k\rightarrow \infty }k^{-n}(Q^{(0)}_{M,k}\varPi ^{(0)}_{k,\le k\nu _k})(x) \hbox { exists and is real valued},\end{aligned}$$
(1.19)
$$\begin{aligned}&\lim _{k\rightarrow \infty }k^{-n}(Q^{(0)}_{M,k}\varPi ^{(0)}_{k,\le k\nu _k})(x)\nonumber \\&\quad \ge (2\pi )^{1-n}\int \psi (\xi )\det (M^\phi _x+2\xi {\mathcal L}_x){1\!\!1}_{\mathbb {R}_{x,0}}(\xi )\mathrm{d}\xi -\frac{C_1}{M^2}, \quad \forall x\in X,\end{aligned}$$
(1.20)
$$\begin{aligned}&\limsup _{k\rightarrow \infty }k^{-n}(Q^{(1)}_{M,k}\varPi ^{(1)}_{k,\le k\nu _k}\overline{Q^{(1)}_{M,k}})(x)\le \frac{C_1}{M^2},\quad \forall x\in X, \end{aligned}$$
(1.21)

and

$$\begin{aligned}&\sup \left\{ k^{-n}\left| (Q^{(0)}_{M,k}\varPi ^{(0)}_{k,\le k\nu _k})(x)\right| ;\, k>0, x\in X\right\} <\infty ,\nonumber \\&\sup \left\{ k^{-n}(Q^{(1)}_{M,k}\varPi ^{(0)}_{k,\le k\nu _k}\overline{Q^{(1)}_{M,k}})(x);\, k>0, x\in X\right\} <\infty , \end{aligned}$$
(1.22)

where \(\mathbb {R}_{x,0}\) is given by (2.15) and \({1\!\!1}_{\mathbb {R}_{x,0}}(\xi )=1\) if \(\xi \in \mathbb {R}_{x,0}\), \({1\!\!1}_{\mathbb {R}_{x,0}}(\xi )=0\) if \(\xi \notin \mathbb {R}_{x,0}\) and \(C_1>0\) is a constant independent of \(k\) and \(M\).

From (1.22), we can apply Lebesgue dominate theorem and Fatou’s lemma and we get using (1.20) and (1.21),

$$\begin{aligned}&\left| \int _X(Q^{(0)}_{M,k}\varPi ^{(0)}_{k,\le k\nu _k})(x)\mathrm{d}v_X(x)\right| \nonumber \\&\quad \ge k^n\Bigr ((2\pi )^{1-n}\int _X\left( \int \psi (\xi ) \det (M^\phi _x+2\xi {\mathcal L}_x){1\!\!1}_{\mathbb {R}_{x,0}}(\xi ) \mathrm{d}\xi \bigr )\mathrm{d}v_X(x)-\frac{C_2}{M^2}\right) +o(k^n),\end{aligned}$$
(1.23)
$$\begin{aligned}&\int _X(Q^{(1)}_{M,k}\varPi ^{(1)}_{k,\le k\nu _k}\overline{Q^{(1)}_{M,k}})(x)\mathrm{d}v_X(x)\le k^n\frac{C_2}{M^2}+o(k^n), \end{aligned}$$
(1.24)

where \(C_2>0\) is a constant independent of \(M\) and \(k\).

Let \(f_{1,k},f_{2,k},\ldots ,f_{d_k,k}\) be an orthonormal basis for \(\fancyscript{H}^0_b(X,L^k)\), where \(d_{k}=\mathrm{dim\,} \fancyscript{H}^0_{b}(X,L^k)\). Let \(\widetilde{f}_{1,k},\widetilde{f}_{2,k},\ldots ,\widetilde{f}_{n_k,k}\) be an orthonormal basis for the space \(\fancyscript{H}^0_{b,0<\mu \le k\nu _k} (X,L^k)\). From (1.23) and (1.18), we see that if \(M\) is large enough, then

$$\begin{aligned}&\sum ^{d_k}_{j=1}\left| \int _X\langle Q^{(0)}_{M,k}f_{j,k}|f_{j,k}\rangle _{h^{L^k}}(x)\mathrm{d}v_X(x)\right| +\sum ^{n_k}_{j=1}\left| \int _X\langle Q^{(0)}_{M,k}\widetilde{f}_{j,k}|\widetilde{f}_{j,k}\rangle _{h^{L^k}}(x)\mathrm{d}v_X(x)\right| \nonumber \\&\quad \ge \frac{k^n}{2}(2\pi )^{1-n}\int _X\Bigr (\int \psi (\xi )\det (M^\phi _x+2\xi {\mathcal L}_x) {1\!\!1}_{\mathbb {R}_{x,0}}(\xi )\mathrm{d}\xi \Bigr )\mathrm{d}v_X(x) \end{aligned}$$
(1.25)

for \(k\) large. From (1.16) and (1.18), it is not difficult to check that

$$\begin{aligned}&\sum ^{n_k}_{j=1}\left| \int _X\langle Q^{(0)}_{M,k}\widetilde{f}_{j,k}|\widetilde{f}_{j,k}\rangle _{h^{L^k}}(x)\mathrm{d}v_X(x)\right| \nonumber \\&\quad \le \Biggr (\int _X(Q^{(1)}_{M,k}\varPi ^{(1)}_{k,\le k\nu _k}\overline{Q^{(1)}_{M,k}})(x)\mathrm{d}v_X(x)\Biggr )^{\frac{1}{2}}\Biggr (\sum ^{n_k}_{j=1}\int _X\langle \widetilde{f}_{j,k}|\widetilde{f}_{j,k}\rangle _{h^{L^k}}(x)\mathrm{d}v_X(x)\Biggr )^{\frac{1}{2}}.\nonumber \\ \end{aligned}$$
(1.26)

It is well known (see [9, Theorem 1.4]) that

$$\begin{aligned} \sup \left\{ k^{-n}\sum ^{n_k}_{j=1}\langle \widetilde{f}_{j,k}|\widetilde{f}_{j,k}\rangle _{h^{L^k}}(x);\, k>0, x\in X\right\} <\infty . \end{aligned}$$
(1.27)

From (1.27), (1.24), (1.26) and (1.25), it is straightforward to see that if \(M\) is large enough, then

$$\begin{aligned}&\sum \limits _{j=1}^{d_k}\left| \int _X\langle Q^{(0)}_{M,k}f_{j,k}|f_{j,k}\rangle _{h^{L^k}}(x)\mathrm{d}v_X(x)\right| \nonumber \\&\quad \ge \frac{k^n}{4}(2\pi )^{1-n}\int _X\Biggr (\int \psi (\xi ) \det (M^\phi _x+2\xi {\mathcal L}_x) {1\!\!1}_{\mathbb {R}_{x,0}}(\xi )\mathrm{d}\xi \Biggr )\mathrm{d}v_X(x) \end{aligned}$$
(1.28)

for \(k\) large. Moreover, it is straightforward to see that there is a constant \(C_M>0\) independent of \(k\) such that \(\left| \int _X\langle Q^{(0)}_{M,k}u|u\rangle _{h^{L^k}}(x)\mathrm{d}v_X(x)\right| \le C_M\int _X\langle u|u\rangle _{h^{L^k}}(x)\mathrm{d}v_X(x)\), for all \(u\in C^\infty (X,L^k)\). Combining this with (1.28), we have

$$\begin{aligned}&C_Md_k=C_M\sum ^{d_k}_{j=1}\int _X\langle f_{j,k}|f_{j,k}\rangle _{h^{L^k}}(x)\mathrm{d}v_X(x) \ge \sum ^{d_k}_{j=1}\left| \int _X\langle Q^{(0)}_{M,k}f_{j,k}|f_{j,k}\rangle _{h^{L^k}}(x)\mathrm{d}v_X(x)\right| \\&\qquad \qquad \ge \frac{k^n}{4}(2\pi )^{1-n}\int _X\Bigr (\int \psi (\xi )\det (M^\phi _x+2\xi {\mathcal L}_x) {1\!\!1}_{\mathbb {R}_{x,0}}(\xi )\mathrm{d}\xi \Bigr )\mathrm{d}v_X(x). \end{aligned}$$

Theorem 1.15 follows.

The paper is organized as follows. In Sect. 2, we review the results in [9] about the asymptotic behavior of the Szegö kernel for lower energy forms to prove (1.19), (1.20) and (1.22). We introduce the extremal function for the space of lower energy forms with respect to a given continuous operator and relate it to the function \(Q^{(1)}_{M,k}\varPi ^{(1)}_{k,\le \lambda }\overline{Q^{(1)}_{M,k}}\) (see Lemma 2.2). This result will be used in the proof of (1.21). In Sect. 3, we introduce canonical coordinates on generalized Sasakian CR manifolds and prove that locally we can always find canonical coordinates and local section such that the corresponding local weight has a simple form (see Proposition 3.2). Canonical coordinates will be used in the constructions of the operators \(Q^{(0)}_{M,k}\) and \(Q^{(1)}_{M,k}\) and Proposition 3.2 will be used in Sect. 4 and the proofs of (1.19), (1.20) and (1.21). In Sect. 4, we modify the scaling technique developed in [9] and [10] to establish the semi-classical Kohn estimates (see Propositions 4.2) and a result about the asymptotic behavior of a sequence of forms with small energy (see Proposition 4.3). These results play important roles in the proofs of (1.19), (1.20) and (1.21). In Sect. 5, we construct the operators \(Q^{(0)}_{M,k}\) and \(Q^{(1)}_{M,k}\). In Sect. 6, we prove (1.22), (1.19), (1.20) and (1.23). In Sect. 7, we prove (1.21) and (1.24). In Sect. 8, we first prove the inequality (1.26) and then we complete the proof of Theorem 1.15. In Sect. 9, we exemplify our main result in two concrete examples, one of a quotient of the Heisenberg group and the other of a Grauert tube over the torus.

2 Szegö kernels for lower energy forms

We will use the same notations as Sect. 1. From now on, we assume that \((L,J,h^L)\) is a rigid generalized Sasakian CR line bundle over \(X\).

The Hermitian fiber metric on \(L\) induces a Hermitian fiber metric on \(L^k\) that we shall denote by \(h^{L^k}\). If \(s\) is a local trivializing section of \(L\) then \(s^k\) is a local trivializing section of \(L^k\). The Hermitian metrics \(\langle \,\cdot \,|\,\cdot \,\rangle \) on \(\Lambda ^{0,q}T^*X\) and \(h^{L^k}\) induce Hermitian metrics on \(\Lambda ^{0,q}T^*X\otimes L^k\). We shall denote these induced metrics by \(\langle \,\cdot \,|\,\cdot \,\rangle _{h^{L^k}}\). For \(f\in \Omega ^{0,q}(X, L^k)\), we denote the pointwise norm \(\left| f(x)\right| ^2_{h^{L^k}}:=\langle f(x)|f(x)\rangle _{h^{L^k}}\). As (1.13), let

$$\begin{aligned} \overline{\partial }_{b,k}:\Omega ^{0,q}(X, L^k)\rightarrow \Omega ^{0,q+1}(X, L^k) \end{aligned}$$
(2.1)

denote the tangential Cauchy–Riemann operator acting on forms with values in \(L^k\). We denote by \(\mathrm{d}v_X=\mathrm{d}v_X(x)\) the volume form on \(X\) induced by the fixed Hermitian metric \(\langle \,\cdot \,|\,\cdot \,\rangle \) on \(\mathbb {C}TX\). Then, we get natural global \(L^2\) inner products \((\ |\ )_{h^{L^k}}\), \((\ |\ )\) on \(\Omega ^{0,q}(X, L^k)\) and \(\Omega ^{0,q}(X)\), respectively. We denote by \(L^2_{(0,q)}(X, L^k)\) the completion of \(\Omega ^{0,q}(X, L^k)\) with respect to \((\ |\ )_{h^{L^k}}\). For \(f\in \Omega ^{0,q}(X,L^k)\), we denote \(\left\| f\right\| ^2_{h^{L^k}}:=(f\ |\ f)_{h^{L^k}}\). Similarly, for \(f\in \Omega ^{0,q}(X)\), we denote \(\left\| f\right\| ^2:=(f\ |\ f)\). Let

$$\begin{aligned} \overline{\partial }^{\,*}_{b,k}:\Omega ^{0,q+1}(X, L^k)\rightarrow \Omega ^{0,q}(X, L^k) \end{aligned}$$
(2.2)

be the formal adjoint of \(\overline{\partial }_{b,k}\) with respect to \((\ |\ )_{h^{L^k}}\). The Kohn Laplacian with values in \(L^k\) is given by

$$\begin{aligned} \Box _{b,k}^{(q)}=\overline{\partial }^{\,*}_{b,k}\overline{\partial }_{b,k}+\overline{\partial }_{b,k}\overline{\partial }^{\,*}_{b,k}: \Omega ^{0,q}(X, L^k)\rightarrow \Omega ^{0,q}(X, L^k). \end{aligned}$$
(2.3)

We extend \(\overline{\partial }_{b,k}\) to \(L^2_{(0,r)}(X,L^k)\), \(r=0,1,\ldots ,n-1\), by

$$\begin{aligned} \overline{\partial }_{b,k}:\mathrm{Dom\,}\overline{\partial }_{b,k}\subset L^2_{(0,r)}(X, L^k)\rightarrow L^2_{(0,r+1)}(X, L^k), \end{aligned}$$
(2.4)

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

$$\begin{aligned} \overline{\partial }^{\,*}_{b,k}:\mathrm{Dom\,}\overline{\partial }^{\,*}_{b,k}\subset L^2_{(0,r+1)}(X, L^k)\rightarrow L^2_{(0,r)}(X, L^k) \end{aligned}$$
(2.5)

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

$$\begin{aligned} \mathrm{Dom\,}\Box ^{(q)}_{b,k}&= \{s\in L^2_{(0,q)}(X,L^k);\, s\in \mathrm{Dom\,}\overline{\partial }_{b,k}\cap \mathrm{Dom\,}\overline{\partial }^{\,*}_{b,k},\nonumber \\&\quad \overline{\partial }_{b,k}u\in \mathrm{Dom\,}\overline{\partial }^{\,*}_{b,k},\ \overline{\partial }^{\,*}_{b,k}u\in \mathrm{Dom\,}\overline{\partial }_{b,k}\}, \end{aligned}$$
(2.6)

and \(\Box ^{(q)}_{b,k}s=\overline{\partial }_{b,k}\overline{\partial }^{\,*}_{b,k}s+\overline{\partial }^{\,*}_{b,k}\overline{\partial }_{b,k}s\) for \(s\in \mathrm{Dom\,}\Box ^{(q)}_{b,k}\). We notice that \(\Box ^{(q)}_{b,k}\) is a positive self-adjoint operator. For a Borel set \(B\subset \mathbb {R}\), we denote by \(E(B)\) the spectral projection of \(\Box ^{(q)}_{b,k}\) corresponding to the set \(B\), where \(E\) is the spectral measure of \(\Box ^{(q)}_{b,k}\) (see [5, section 2], for the precise meanings of spectral projection and spectral measure). We notice that the spectrum of \(\Box ^{(q)}_{b,k}\) is contained in \(\overline{\mathbb {R}}_+\). For \(\lambda \ge 0\), we set

$$\begin{aligned}&\fancyscript{H}^q_{b,\le \lambda }(X, L^k):=\mathrm{Range }E\big ((-\infty ,\lambda ]\big )\subset L^2_{(0,q)}(X,L^k),\nonumber \\&\fancyscript{H}^q_{b,>\lambda }(X,L^k):=\mathrm{Range }E\big ((\lambda ,\infty )\big )\subset L^2_{(0,q)}(X,L^k). \end{aligned}$$
(2.7)

It is well known (see [5, section 2]) that for all \(\lambda >0\),

$$\begin{aligned} L^2_{(0,q)}(X,L^k)=\fancyscript{H}^q_{b,\le \lambda }(X, L^k)\oplus \fancyscript{H}^q_{b,>\lambda }(X, L^k) \end{aligned}$$
(2.8)

and

$$\begin{aligned} \left\| u\right\| ^2_{h^{L^k}}\le \frac{1}{\lambda }\left( \Box ^{(q)}_{b,k}u\ |\ u\right) _{h^{L^k}},\quad \forall u\in \fancyscript{H}^q_{b,>\lambda }(X,L^k)\bigcap \mathrm{Dom\,}\Box ^{(q)}_{b,k}. \end{aligned}$$
(2.9)

For \(\lambda =0\), we denote

$$\begin{aligned} \fancyscript{H}_b^q(X, L^k):=\fancyscript{H}^q_{b,\le 0}(X,L^k)={{\mathrm{Ker}}}\Box ^{(q)}_{b,k}. \end{aligned}$$
(2.10)

Now, fix \(q\in \left\{ 0,1,\ldots ,n-1\right\} \) and until further notice we assume that \(Y(q)\) holds. By [11, 7.6-7.8], [7, 5.4.11-12], [4, Props. 8.4.8-9] and [8, Chapter 6], we know that \(\Box ^{(q)}_{b,k}\) is hypoelliptic, has compact resolvent, the strong Hodge decomposition holds and \(\Box ^{(q)}_{b,k}\) has a discrete spectrum, each eigenvalues occurs with finite multiplicity and all eigenforms are smooth. Hence, for any \(\lambda \ge 0\),

$$\begin{aligned} \dim \fancyscript{H}_{b,\le \lambda }^q(X, L^k)<\infty ,\quad \fancyscript{H}_{b,\le \lambda }^q(X, L^k)\subset \Omega ^{0,q}(X, L^k),\quad \fancyscript{H}_b^q(X, L^k)\cong H_b^q(X, L^k).\nonumber \\ \end{aligned}$$
(2.11)

Let \(g_j(x)\in \Omega ^{0,q}(X, L^k)\), \(j=1,\ldots ,d_k\), \(d_k=\mathrm{dim\,}\fancyscript{H}^q_{b,\,\leqslant \lambda }(X,L^k)\), be any orthonormal frame for the space \(\fancyscript{H}_{b,\,\le \lambda }^q(X, L^k)\) with respect to \((\,\cdot \,|\,\cdot \,)_{h^{L^k}}\). The Szegö kernel function \(\varPi ^{(q)}_{k,\le \lambda }(x)\) of the space \(\fancyscript{H}^q_{b,\leqslant \lambda }(X,L^k)\) is given by

$$\begin{aligned} \varPi ^{(q)}_{k,\,\leqslant \lambda }(x):=\sum ^{d_k}_{j=1}\left| g_j(x)\right| ^2_{h^{L^k}}. \end{aligned}$$
(2.12)

Let

$$\begin{aligned} A:\Omega ^{0,q}(X,L^k)\rightarrow \Omega ^{0,q}(X,L^k) \end{aligned}$$

be a continuous operator. We define

$$\begin{aligned}&\left( A\varPi ^{(q)}_{k,\,\leqslant \lambda }\right) (x):=\sum ^{d_k}_{j=1}\langle Ag_j(x)| g_j(x)\rangle _{h^{L^k}},\end{aligned}$$
(2.13)
$$\begin{aligned}&\left( A\varPi ^{(q)}_{k,\,\leqslant \lambda }\overline{A}\right) (x):=\sum ^{d_k}_{j=1}\left| Ag_j(x)\right| ^2_{h^{L^k}}. \end{aligned}$$
(2.14)

It is straightforward to see that the definitions (2.12), (2.13) and (2.14) are independent of the choices of orthonormal frame \(g_j\), \(j=1,\ldots ,d_k\).

For \(q=0,1,\ldots ,n-1\) and \(x\in X\), set

$$\begin{aligned} \mathbb {R}_{x,q}&= \big \{s\in \mathbb {R};\, M^\phi _x+2s{\mathcal L}_x\text { has exactly } q \hbox { negative eigenvalues} \nonumber \\&\quad \text {and } n-1-q \hbox { positive eigenvalues}\big \}, \end{aligned}$$
(2.15)

where \(M^\phi _x\) is given by (1.8) and the eigenvalues of the Hermitian quadratic form \(M^\phi _x+2s{\mathcal L}_x\), \(s\in \mathbb {R}\), are calculated with respect to the Hermitian metric \(\langle \,\cdot \,|\,\cdot \,\rangle \). It is not difficult to see that if \(Y(q)\) holds at each point of \(X\) then there is a constant \(C>0\) such that

$$\begin{aligned} \mathbb {R}_{x,\,q}\subset [-C,C]\quad \hbox {for all }x\in X. \end{aligned}$$
(2.16)

Denote by \(\det (M^\phi _x+2s{\mathcal L}_x)\) the product of all the eigenvalues of \(M^\phi _x+2s{\mathcal L}_x\). Assuming (2.16) holds, the function

$$\begin{aligned} X\longrightarrow {\mathbb {R}},\quad x\longmapsto \int _{\mathbb {R}_{x,q}}\left| \det (M^\phi _x+2s{\mathcal L}_x)\right| \mathrm{d}s \end{aligned}$$
(2.17)

is well defined. Since \(M^\phi _x\) and \({\mathcal L}_x\) are continuous functions of \(x\in X\), we conclude that the function (2.17) is continuous.

The following is well known (see [9, Theorem 1.6])

Theorem 2.1

Assume that condition \(Y(q)\) holds at each point of \(X\). Then\(,\) for any sequence \(\nu _k>0\) with \(\nu _k\rightarrow 0\) as \(k\rightarrow \infty ,\) there is a constant \(C_0>0\) independent of \(k,\) such that

$$\begin{aligned} k^{-n}\varPi ^{(q)}_{k,\,\leqslant k\nu _k}(x)\leqslant C_0 \end{aligned}$$
(2.18)

for all \(x\in X\). Moreover\(,\) there is a sequence \(\mu _k>0,\) \(\mu _k\rightarrow 0\) as \(k\rightarrow \infty ,\) such that for any sequence \(\nu _k>0\) with \(\lim _{k\rightarrow \infty }\frac{\mu _k}{\nu _k}=0\) and \(\nu _k\rightarrow 0\) as \(k\rightarrow \infty ,\) we have

$$\begin{aligned} \lim _{k\rightarrow \infty }k^{-n}\varPi ^{(q)}_{k,\,\leqslant k\nu _k}(x)=(2\pi )^{-n}\int _{\mathbb {R}_{x,q}}\left| \det (M^\phi _x+2s{\mathcal L}_x)\right| \mathrm{d}s, \end{aligned}$$
(2.19)

for all \(x\in X.\)

We introduce some notations. For \(p\in X\), we can choose a smooth orthonormal frame \(e_1,\ldots ,e_{n-1}\) of \(T^{*0,1}X\) over a neighborhood \(U\) of \(p\). We say that a multiindex \(J=(j_1,\ldots ,j_q)\in \{1,\ldots ,n-1\}^q\) has length \(q\) and write \(\left| J\right| =q\). We say that \(J\) is strictly increasing if \(1\leqslant j_1<j_2<\cdots <j_q\leqslant n-1\). For \(J=(j_1,\ldots ,j_q)\) we define \(e_J:=e_{j_1}\wedge \cdots \wedge e_{j_q}\). Then, \(\left\{ e_J;\, \left| J\right| =q, J\hbox { strictly increasing}\right\} \) is an orthonormal frame for \(\Lambda ^{0,q}T^*X\) over \(U\).

For \(f\in \Omega ^{0,q}(X, L^k)\), we may write

$$\begin{aligned} f|_U={\mathop {\mathop {\sum }\nolimits ^{'}}\limits _{\left| J\right| =q}} f_Je_J,\quad \text {with } f_J=\langle f|e_J\rangle \in C^\infty (U,L^k), \end{aligned}$$

where \(\sum '\) means that the summation is performed only over strictly increasing multiindices. We call \(f_J\) the component of \(f\) along \(e_J\). It will be clear from the context what frame is being used. For \(q>0\), the extremal function \(S^{(q)}_{k,\le \lambda ,J}\) for the space \(\fancyscript{H}^q_{b,\le \lambda }(X,L^k)\) along the direction \(e_J\) is defined by

$$\begin{aligned} S^{(q)}_{k,\le \lambda ,J}(y)=\sup _{\alpha \in \,\fancyscript{H}_{b,\le \lambda }^q(X, L^k),\,\left\| \alpha \right\| _{h^{L^k}}=1}\left| \alpha _J(y)\right| ^2_{h^{L^k}}, \end{aligned}$$
(2.20)

where \(\alpha _J\) denotes the component of \(\alpha \) along \(e_J\). Let

$$\begin{aligned} A:\Omega ^{0,q}(X,L^k)\rightarrow \Omega ^{0,q}(X,L^k) \end{aligned}$$

be a continuous operator. For \(\left| J\right| =q\), \(J\) is strictly increasing, we define

$$\begin{aligned} (AS^{(q)}_{k,\,\leqslant \lambda ,J}\overline{A})(y):=\sup _{\alpha \in \,\fancyscript{H}_{b,\le \lambda }^q(X, L^k),\,\left\| \alpha \right\| _{h^{L^k}}=1}\left| (A\alpha )_J(y)\right| ^2_{h^{L^k}}, \end{aligned}$$
(2.21)

where \((A\alpha )_J\) denotes the component of \(A\alpha \) along \(e_J\). Similarly, when \(q=0\), we define

$$\begin{aligned}&S^{(0)}_{k,\le \lambda }(y)=\sup _{\alpha \in \,\fancyscript{H}_{b,\le \lambda }^0(X, L^k),\,\left\| \alpha \right\| _{h^{L^k}}=1}\left| \alpha (y)\right| ^2_{h^{L^k}}, \nonumber \\&(AS^{(0)}_{k,\,\leqslant \lambda }\overline{A})(y):=\sup _{\alpha \in \,\fancyscript{H}_{b,\le \lambda }^0(X, L^k),\,\left\| \alpha \right\| _{h^{L^k}}=1}\left| (A\alpha )(y)\right| ^2_{h^{L^k}}. \end{aligned}$$
(2.22)

We need the following

Lemma 2.2

Fix \(\lambda \ge 0\). Let \(A:\Omega ^{0,q}(X,L^k)\rightarrow \Omega ^{0,q}(X,L^k)\) be a continuous operator. For every local orthonormal frame

$$\begin{aligned} \left\{ e_J(y); \left| J\right| =q,\,J\, \mathrm{strictly \, increasing}\right\} \end{aligned}$$

of \(\Lambda ^{0,q}T^*X\) over an open set \(U\subset X,\) we have when \(q>0,\)

$$\begin{aligned} \left( A\varPi ^{(q)}_{k,\le \lambda }\overline{A}\right) (y)={\mathop {\mathop {\sum }\nolimits ^{'}}\limits _{\left| J\right| =q}}\left( AS^{(q)}_{k,\le \lambda ,J}\overline{A}\right) (y), \end{aligned}$$
(2.23)

for every \(y\in U\).

Similarly\(,\) when \(q=0,\) we have

$$\begin{aligned} \left( A\varPi ^{(0)}_{k,\le \lambda }\overline{A}\right) (y)=\left( AS^{(0)}_{k,\le \lambda }\overline{A}\right) (y), \end{aligned}$$
(2.24)

for every \(y\in U\).

We remind that \(A\varPi ^{(q)}_{k,\le \lambda }\overline{A}\) is given by (2.14).

Proof

Let \((f_j)_{j=1,\ldots ,d_k}\) be an orthonormal frame for the space \(\fancyscript{H}_{b,\le \lambda }^q(X, L^k)\). Let \(s\) be a local section of \(L\) on \(U\), \(\left| s\right| ^2_{h^L}=\mathrm{e}^{-\phi }\). On \(U\), we write

$$\begin{aligned}&Af_j=s^k\widetilde{g}_j,\ \ \widetilde{g}_j\in \Omega ^{0,q}(U),\quad j=1,\ldots ,d_k,\\&\widetilde{g}_j={\mathop {\mathop {\sum }\nolimits ^{'}}\limits _{\left| J\right| =q}}\widetilde{g}_{j,J}e_J,\quad j=1,\ldots ,d_k. \end{aligned}$$

On \(U\), we write

$$\begin{aligned} \left( A\varPi ^{(q)}_{k,\le \lambda }\overline{A}\right) (y)={\mathop {\mathop {\sum }\nolimits ^{'}}\limits _{\left| J\right| =q}}\left( A\varPi ^{(q)}_{k,\le \lambda ,J}\overline{A}\right) (y), \end{aligned}$$
(2.25)

where

$$\begin{aligned} (A\varPi ^{(q)}_{k,\le \lambda ,J}\overline{A})(y):=\mathrm{e}^{-\phi (y)}\sum _j\left| \widetilde{g}_{j,J}(y)\right| ^2. \end{aligned}$$

It is easy to see that \((A\varPi ^{(q)}_{k,\le \lambda ,J}\overline{A})(y)\) is independent of the choice of the orthonormal frame \((f_j)_{j=1,\ldots ,d_k}\). Take \(\alpha \in \fancyscript{H}^q_{b,\le \lambda }(X, L^k)\) of unit norm. Since \(\alpha \) is contained in an orthonormal base, obviously \(|(A\alpha )_J(y)|^2_{h^{L^k}}\leqslant (A\varPi ^{(q)}_{k,\le \lambda ,J}\overline{A})(y)\), where \((A\alpha )_J\) denotes the component of \(A\alpha \) along \(e_J\). Thus,

$$\begin{aligned} (AS^{(q)}_{k,\le \lambda ,J}\overline{A})(y)\leqslant (A\varPi ^{(q)}_{k,\le \lambda ,J}\overline{A})(y),\quad \text {for all strictly increasing }J, \left| J\right| =q. \end{aligned}$$
(2.26)

Fix a point \(p\in U\) and a strictly increasing multiindex \(J\) with \(\left| J\right| =q\). We may assume that \(\sum ^{d_k}_{j=1}\left| \widetilde{g}_{j,J}(p)\right| ^2\ne 0\). Put

$$\begin{aligned} \textstyle u(y)=\Big (\sum ^{d_k}_{j=1}\left| \widetilde{g}_{j,J}(p)\right| ^2\Big )^{-1/2}\cdot \sum ^N_{j=1}\overline{\widetilde{g}_{j,J}(p)}f_j(y). \end{aligned}$$

We can easily check that \(u\in \fancyscript{H}^q_{b,\le \lambda }(X, L^k)\) and \(\left\| u\right\| _{h^{L^k}}=1\). Hence,

$$\begin{aligned} \left| (Au)_{J}(p)\right| ^2_{h^{L^k}}\leqslant (AS^{(q)}_{k,\le \lambda ,J}\overline{A})(p), \end{aligned}$$

therefore,

$$\begin{aligned} \left( A\varPi ^{(q)}_{k,\le \lambda ,J}\overline{A}\right) (p)=\sum ^{d_k}_{j=1}\mathrm{e}^{-\phi (p)}\left| \widetilde{g}_{j,J}(p)\right| ^2=|(Au)_{J}(p)|^2_{h^{L^k}}\leqslant (AS^{(q)}_{k,\le \lambda ,J}\overline{A})(p). \end{aligned}$$

From this and (2.26), we conclude that \(A\varPi ^{(q)}_{k,\le \lambda ,J}\overline{A}=AS^{(q)}_{k,\le \lambda ,J}\overline{A}\) for all strictly increasing multiindices \(J\) with \(\left| J\right| =q\). Combining this with (2.25), (2.23) follows.

The proof of (2.24) is the same. The lemma follows. \(\square \)

3 Canonical coordinates of generalized Sasakian CR manifolds

In this work, we need the following beautiful result due to Baouendi–Rothschild–Treves [1, section1]

Theorem 3.1

We recall that we work with the assumption that \((X,T^{1,0}X)\) is a generalized Sasakian CR manifold and we fix a rigid global real vector field \(T=J\frac{\partial }{\partial t}\). For every point \(x_0\in X,\) there exists local coordinates \(x=(x_1,\ldots ,x_{2n-1})=(z,\theta )=(z_1,\ldots ,z_{n-1},\theta ),\) \(z_j=x_{2j-1}+ix_{2j},\) \(j=1,\ldots ,n-1,\) \(\theta =x_{2n-1},\) defined in some small neighborhood \(U\) of \(x_0\) such that

$$\begin{aligned}&T=\frac{\partial }{\partial \theta },\nonumber \\&\quad Z_j=\frac{\partial }{\partial z_j}+i\frac{\partial \varphi }{\partial z_j}(z)\frac{\partial }{\partial \theta },\quad j=1,\ldots ,n-1, \end{aligned}$$
(3.1)

where \(Z_j(x),\) \(j=1,\ldots ,n-1,\) form a basis of \(T^{1,0}_xX,\) for each \(x\in U,\) and \(\varphi (z)\in C^\infty (U,\mathbb {R})\) independent of \(\theta .\)

Let \(x=(x_1,\ldots ,x_{2n-1})\) be local coordinates of \(X\) defined in some open set in \(X\). In this paper, when we write \(x=(x_1,\ldots ,x_{2n-1})=(z,\theta )\) we mean that \(z=(z_1,\ldots ,z_{n-1})\), \(z_j=x_{2j-1}+ix_{2j}\), \(j=1,\ldots ,n-1\), \(\theta =x_{2n-1}\). We call \(x\) canonical coordinates if \(x\) satisfies (3.1).

We also need

Proposition 3.2

For a given point \(p\in X\), we can find canonical coordinates \(x=(x_1,\ldots ,x_{2n-1})=(z,\theta )\) and local section \(s\), \(\left| s\right| ^2_{h^L}=\mathrm{e}^{-\phi }\), defined in some small neighborhood \(D\) of \(p\) such that

$$\begin{aligned}&x(p)=0,\nonumber \\&Z_j=\frac{\partial }{\partial z_j}+i\lambda _j\overline{z}_j\frac{\partial }{\partial \theta }+O(\left| z\right| ^2)\frac{\partial }{\partial \theta },\quad j=1,\ldots ,n-1,\nonumber \\&\frac{\partial }{\partial z_1},\ldots ,\frac{\partial }{\partial z_{n-1}}\hbox { is an orthonormal frame for }T^{1,0}_pX,\\&\phi (z,\theta )=\beta \theta +\sum ^{n-1}_{j,t=1}\mu _{j,t}\overline{z}_j z_t+O(\left| z\right| \left| \theta \right| )+O(\left| \theta \right| ^2)+O(\left| (z,\theta )\right| ^3),\nonumber \end{aligned}$$
(3.2)

where \(Z_1(x),\ldots , Z_{n-1}(x)\) form a basis of \(T^{1, 0}_xX\) varying smoothly with \(x\) in a neighborhood of \(p\), \(\lambda _1,\ldots ,\lambda _{n-1}\) are the eigenvalues of \({\mathcal L}_p\) with respect to \(\langle \,\cdot \,|\,\cdot \,\rangle \), \(\beta \in \mathbb {R}\), \(\mu _{j,t}\in \mathbb {C}\), \(\mu _{j,t}=\overline{\mu _{t,j}}\), \(j,t=1,\ldots ,n-1\).

Proof

Fix \(p\in X\). Let \(\widetilde{x}=(\widetilde{x}_1,\ldots ,\widetilde{x}_{2n-1})=(\widetilde{z},\widetilde{\theta })=(\widetilde{z}_1,\ldots ,\widetilde{z}_{n-1},\widetilde{\theta })\), \(\widetilde{z}_j=\widetilde{x}_{2j-1}+i\widetilde{x}_{2j}\), \(j=1,\ldots ,n-1\), \(\widetilde{\theta }=\widetilde{x}_{2n-1}\) be canonical coordinates of \(X\) defined in some small neighborhood \(D\) of \(p\). We have

$$\begin{aligned}&T=\frac{\partial }{\partial \widetilde{\theta }},\nonumber \\&Z_j=\frac{\partial }{\partial \widetilde{z}_j}+i\frac{\partial \widetilde{\varphi }}{\partial \widetilde{z}_j}(\widetilde{z})\frac{\partial }{\partial \widetilde{\theta }},\quad j=1,\ldots ,n-1, \end{aligned}$$
(3.3)

where \(Z_j(\widetilde{x})\), \(j=1,\ldots ,n-1\), form a basis of \(T^{1,0}_{\widetilde{x}}X\), for each \(\widetilde{x}\in D\), and \(\widetilde{\varphi }(\widetilde{z})\in C^\infty (D,\mathbb {R})\) independent of \(\widetilde{\theta }\). It is easy to see that we can take \(\widetilde{x}\) so that \(\widetilde{x}(p)=0\). Near \(p\), we write

$$\begin{aligned} \widetilde{\varphi }(\widetilde{z})=a+\sum ^{n-1}_{j=1}(\alpha _j\widetilde{z}_j+\overline{\alpha }_j\overline{\widetilde{z}_j})+O(\left| \widetilde{z}\right| ^2), \end{aligned}$$
(3.4)

where \(a\in \mathbb {C}\), \(\alpha _j\in \mathbb {C}\), \(j=1,\ldots ,n-1\). Let \({\hat{z}}=\widetilde{z}\), \(\hat{\theta }=\widetilde{\theta }-\sum ^{n-1}_{j=1}(i\alpha _j\widetilde{z}_j-i\overline{\alpha }_j\overline{\widetilde{z}_j})\). Then, \(({\hat{z}},\hat{\theta })\) form canonical coordinates of \(X\) near \(p\) and we can check that

$$\begin{aligned}&\frac{\partial }{\partial \hat{\theta }}=\frac{\partial }{\partial \widetilde{\theta }},\nonumber \\&\frac{\partial }{\partial {\hat{z}}_j}=\frac{\partial }{\partial \widetilde{z}_j}+i\alpha _j\frac{\partial }{\partial \widetilde{\theta }},\quad j=1,\ldots ,n-1. \end{aligned}$$
(3.5)

From (3.5), (3.4) and (3.3), we see that

$$\begin{aligned}&T=\frac{\partial }{\partial \hat{\theta }},\nonumber \\&Z_j=\frac{\partial }{\partial {\hat{z}}_j}+i\frac{\partial \hat{\varphi }}{\partial {\hat{z}}_j}({\hat{z}})\frac{\partial }{\partial \hat{\theta }}, \quad j=1,\ldots ,n-1, \end{aligned}$$
(3.6)

where

$$\begin{aligned} \hat{\varphi }({\hat{z}})=\widetilde{\varphi }({\hat{z}})-\sum ^{n-1}_{j=1}\left( \alpha _j{\hat{z}}_j+\overline{\alpha }_j\overline{{\hat{z}}_j}\right) =a+O\left( \left| {\hat{z}}\right| ^2\right) . \end{aligned}$$

Thus, \(\frac{\partial }{\partial {\hat{z}}_1},\ldots ,\frac{\partial }{\partial {\hat{z}}_{n-1}}\) is a basis of \(T^{1,0}_pX\). By taking some linear transformation, we can take \({\hat{z}}\) so that \(\frac{\partial }{\partial {\hat{z}}_j}\), \(j=1,\ldots ,n-1\), is an orthonormal frame for \(T^{1,0}_pX\) and the Levi form is diagonal at \(p\) with respect to \(\frac{\partial }{\partial {\hat{z}}_j}\), \(j=1,\ldots ,n-1\). We write

$$\begin{aligned} \hat{\varphi }({\hat{z}})=\alpha +\sum ^{n-1}_{j,t=1}(\beta _{j,t} {\hat{z}}_j{\hat{z}}_t+\overline{\beta }_{j,t}\overline{{\hat{z}}_j}\,\overline{{\hat{z}}_t})+ \sum ^{n-1}_{j,t=1}\gamma _{j,t}\overline{{\hat{z}}_j}{\hat{z}}_t+O(\left| {\hat{z}}\right| ^3), \end{aligned}$$
(3.7)

where \(\beta _{j,t}\in \mathbb {C}\), \(\gamma _{j,t}\in \mathbb {C}\), \(\gamma _{j,t}=\overline{\gamma _{t,j}}\), \(j,t=1,\ldots ,n-1\). Since the Levi form is diagonal at \(p\) with respect to \(\frac{\partial }{\partial {\hat{z}}_j}\), \(j=1,\ldots ,n-1\), we can check that

$$\begin{aligned} \gamma _{j,t}=\lambda _j\delta _{j,t},\quad j,t=1,\ldots ,n-1, \end{aligned}$$
(3.8)

where \(\lambda _1,\ldots ,\lambda _{n-1}\) are the eigenvalues of \({\mathcal L}_p\) with respect to \(\langle \,\cdot \,|\,\cdot \,\rangle \). Let \(z={\hat{z}}\), \(\theta =\hat{\theta }-\sum ^{n-1}_{j,t=1}i(\beta _{j,t}{\hat{z}}_j{\hat{z}}_t-\overline{\beta }_{j,t}\overline{{\hat{z}}_j}\,\overline{{\hat{z}}_t})\). Then, \((z,\theta )\) form canonical coordinates of \(X\) near \(p\) and we can check that

$$\begin{aligned}&\frac{\partial }{\partial \theta }=\frac{\partial }{\partial \hat{\theta }},\nonumber \\&\frac{\partial }{\partial z_j}=\frac{\partial }{\partial {\hat{z}}_j}+i\sum ^{n-1}_{t=1}\beta _{j,t}{\hat{z}}_t\frac{\partial }{\partial \hat{\theta }}, \quad j=1,\ldots ,n-1. \end{aligned}$$
(3.9)

From (3.6), (3.7), (3.8) and (3.9), we can check that

$$\begin{aligned}&T=\frac{\partial }{\partial \theta },\\&Z_j=\frac{\partial }{\partial z_j}+i\lambda _j\overline{z}_j\frac{\partial }{\partial \theta }+O(\left| z\right| ^2)\frac{\partial }{\partial \theta },\quad j=1,\ldots ,n-1. \end{aligned}$$

Since \(\frac{\partial }{\partial z_j}\), \(j=1,\ldots ,n-1\), is an orthonormal frame of \(T^{1,0}_pX\), we conclude that \(x=(z,\theta )\) satisfies the first three properties in (3.2).

Let \(\hat{s}\) be a local section defined in some neighborhood of \(p\), \(\left| \hat{s}\right| ^2_{h^L}=\mathrm{e}^{-\hat{\phi }}\). Near \(p\), we write

$$\begin{aligned} \hat{\phi }(z,\theta )=c+\beta \theta +\sum ^{n-1}_{j=1}(a_jz_j+\overline{a}_j\overline{z}_j)+O(\left| (z,\theta )\right| ^2), \end{aligned}$$
(3.10)

where \(c\in \mathbb {R}\), \(\beta \in \mathbb {R}\) and \(a_j\in \mathbb {C}\), \(j=1,\ldots ,n-1\). Let

$$\begin{aligned} g(z)=\mathrm{e}^{\frac{c}{2}}\left( 1+\sum ^{n-1}_{j=1}a_jz_j\right) . \end{aligned}$$
(3.11)

Then, \(g(z)\) is a rigid CR function. We may replace \(\hat{s}\) by \(g\hat{s}:=\widetilde{s}\). We have

$$\begin{aligned} \left| \widetilde{s}\right| ^2_{h^L}=\mathrm{e}^{-\widetilde{\phi }}=\left| g\right| ^2\mathrm{e}^{-\hat{\phi }}=\mathrm{e}^{2\log \left| g\right| -\hat{\phi }}. \end{aligned}$$
(3.12)

From (3.11), we can check that

$$\begin{aligned} 2\log \left| g\right| =c+\sum ^{n-1}_{j=1}(a_jz_j+\overline{a}_j\overline{z}_j)+O(\left| z\right| ^2). \end{aligned}$$

Combining this with (3.12) and (3.10), we conclude that

$$\begin{aligned} \widetilde{\phi }(z,\theta )=\beta \theta +O(\left| (z,\theta )\right| ^2). \end{aligned}$$

Near \(p\), we write

$$\begin{aligned} \widetilde{\phi }(z,\theta )&= \beta \theta +\sum ^{n-1}_{j,t=1}(c_{j,t}z_jz_t+\overline{c_{j,t}}\,\overline{z}_j\overline{z}_t)+\sum ^n_{j,t=1}\mu _{j,t}\overline{z}_jz_t\nonumber \\&+O(\left| z\right| \left| \theta \right| )+O(\left| \theta \right| ^2)+O(\left| (z,\theta )\right| ^3), \end{aligned}$$
(3.13)

where \(c_{j,t}\in \mathbb {C}\), \(\mu _{j,t}\in \mathbb {C}\), \(\mu _{j,t}=\overline{\mu _{t,j}}\), \(j,t=1,\ldots ,n-1\). Let

$$\begin{aligned} g_1(z)=1+\sum ^{n-1}_{j,t=1}c_{j,t}z_jz_t. \end{aligned}$$
(3.14)

Then, \(g_1(z)\) is a rigid CR function. We may replace \(\widetilde{s}\) by \(g_1\widetilde{s}:=s\). We have

$$\begin{aligned} \left| s\right| ^2_{h^L}=\mathrm{e}^{-\phi }=\left| g_1\right| ^2\mathrm{e}^{-\widetilde{\phi }}=\mathrm{e}^{2\log \left| g_1\right| -\widetilde{\phi }}. \end{aligned}$$
(3.15)

From (3.14), we can check that

$$\begin{aligned} 2\log \left| g_1\right| =\sum ^{n-1}_{j,t=1}(c_{j,t}z_jz_t+\overline{c_{j,t}}\,\overline{z}_j\overline{z}_t)+O(\left| z\right| ^3). \end{aligned}$$

Combining this with (3.15) and (3.13), we conclude that

$$\begin{aligned} \phi (z,\theta )=\beta \theta +\sum ^{n-1}_{j,t=1}\mu _{j,t}\overline{z}_jz_t+O(\left| z\right| \left| \theta \right| )+O(\left| \theta \right| ^2)+O(\left| (z,\theta )\right| ^3). \end{aligned}$$

The proposition follows. \(\square \)

Proposition 3.3

We assume that \(X\) is a generalized torus CR manifold and \((L,J)\) is an admissible generalized Sasakian CR line bundle over \(X\) \((\)see Definition 1.17 and Definition 1.18). Let \(\phi \) and \(\phi _1\) be as in the discussion after Definition 1.18. If \(M^{\phi }_x\) is positive on \(X,\) then \(M^{\phi _1}_x\) is positive on \(X.\)

Proof

Let \(\left\{ W_1\subset W'_1,\ldots ,W_N\subset W'_N\right\} \) be open sets of \(X\) such that \(X=\bigcup ^N_{j=1}W_j\) and there exist canonical coordinates on \(W'_j\), for each \(j\) and there is a constant \(\epsilon _0>0\) such that for each \(x\in X\), \(\Phi ^t(x)\) is well defined, \(\forall \left| t\right| \le \epsilon _0\), and

$$\begin{aligned} \left\{ \Phi ^t(x);\, x\in W_j,\left| t\right| \le \epsilon _0\right\} \subset W'_j, \end{aligned}$$

for each \(j\), where \(\Phi ^t(x)\) is the \(T\)-flow. Fix \(t_0\in [-\epsilon _0,\epsilon _0]\). Put \(\widetilde{\phi }(x)=\phi (\Phi ^{t_0}x)\). It is obviously that \(\widetilde{\phi }(x)\) also define a Hermitian fiber metric on \(L\). Using canonical coordinates (3.1), we can check that

$$\begin{aligned} d(\overline{\partial }_b\widetilde{\phi }-\partial _b\widetilde{\phi })(x)=d(\overline{\partial }_b\phi -\partial \phi )(\Phi ^{t_0}x),\quad \forall x\in X. \end{aligned}$$

Thus,

$$\begin{aligned} M^{\widetilde{\phi }}_x=M^\phi _{\Phi ^{t_0}(x)},\quad \forall x\in X. \end{aligned}$$

Similarly, fix \(t_1\in [-\epsilon _0,\epsilon _0]\) and put \(\hat{\phi }(x)=\widetilde{\phi }(\Phi ^{t_1}x)=\phi (\Phi ^{t_0+t_1}x)\). We have

$$\begin{aligned} M^{\hat{\phi }}_{x}=M^{\widetilde{\phi }}_{\Phi ^{t_1}(x)}=M^{\phi }_{\Phi ^{t_0+t_1}(x)},\quad \forall x\in X. \end{aligned}$$

Continuing in this way, we obtain for any \(t\in [0,\gamma _0]\), \(M^{\phi (\Phi ^t(x))}_{x}=M^{\phi }_{\Phi ^t(x)}\), \(\forall x\in X\), where \(\gamma _0\) is as in Definition 1.17. Thus,

$$\begin{aligned} M^{\phi _1}_x=\frac{1}{\gamma _0}\int ^{\gamma _0}_0M^\phi _{\Phi ^t(x)}\mathrm{d}t,\quad \forall x\in X. \end{aligned}$$

The proposition follows. \(\square \)

4 The scaling technique

In this section, we modify the scaling technique developed in [9] and [10] to prove (1.19), (1.20) and (1.21).

Fix a point \(p\in X\). Let \(x=(x_1,\ldots ,x_{2n-1})=(z,\theta )\) be canonical coordinates of \(X\) defined in some small neighborhood \(D\) of \(p\) and let \(s\) be a local section of \(L\) on \(D\), \(\left| s\right| ^2_{h^L}=\mathrm{e}^{-\phi }\). We take \(x\) and \(s\) so that

$$\begin{aligned}&x(p)=0,\nonumber \\&Z_j=\frac{\partial }{\partial z_j}+i\lambda _j\overline{z}_j\frac{\partial }{\partial \theta }+O(\left| z\right| ^2)\frac{\partial }{\partial \theta },\quad j=1,\ldots ,n-1,\nonumber \\&\frac{\partial }{\partial z_1},\ldots ,\frac{\partial }{\partial z_{n-1}}\hbox { is an orthonormal frame for } T^{1,0}_pX,\\&\phi (z,\theta )=\beta \theta +\sum ^{n-1}_{j,t=1}\mu _{j,t}\overline{z}_j z_t+O(\left| z\right| \left| \theta \right| )+O(\left| \theta \right| ^2)+O(\left| (z,\theta )\right| ^3),\nonumber \end{aligned}$$
(4.1)

where \(Z_1(x),\ldots , Z_{n-1}(x)\) form a basis of \(T^{1, 0}_xX\) varying smoothly with \(x\) in a neighborhood of \(p\), \(\lambda _1,\ldots ,\lambda _{n-1}\) are the eigenvalues of \({\mathcal L}_p\) with respect to \(\langle \,\cdot \,|\,\cdot \,\rangle \), \(\beta \in \mathbb {R}\), \(\mu _{j,k}\in \mathbb {C}\), \(\mu _{j,t}=\overline{\mu _{t,j}}\), \(j,t=1,\ldots ,n-1\). By Proposition 3.2, this is always possible. Fix \(q\in \left\{ 0,1,\ldots ,n-1\right\} \). In this section, we work on \((0,q)\) forms and we work with this local coordinates \(x=(z,\theta )\).

Let \((\ |\ )_{k\phi }\) be the inner product on the space \(\Omega ^{0,q}_0(D)\) defined as follows:

$$\begin{aligned} (f\ |\ g)_{k\phi }=\int _D\!\langle f|g\rangle \mathrm{e}^{-k\phi }\mathrm{d}v_X, \end{aligned}$$

where \(f, g\in \Omega ^{0,q}_0(D)\). Let \(\overline{\partial }^{\,*,k\phi }_b:\Omega ^{0,q+1}(D)\rightarrow \Omega ^{0,q}(D)\) be the formal adjoint of \(\overline{\partial }_b\) with respect to \(( \ |\ )_{k\phi }\). Put

$$\begin{aligned} \Box ^{(q)}_{b,k\phi }=\overline{\partial }_b\overline{\partial }^{\,*,k\phi }_b+ \overline{\partial }^{\,*,k\phi }_b\overline{\partial }_b:\Omega ^{0,q}(D)\rightarrow \Omega ^{0,q}(D). \end{aligned}$$

Let \(u\in \Omega ^{0,q}(D, L^k)\). Then, there exists \(\widetilde{u}\in \Omega ^{0,q}(D)\) such that \(u=s^k\widetilde{u}\) and we have

$$\begin{aligned} \Box ^{(q)}_{b,k}u=s^k\Box ^{(q)}_{b,k\phi }\widetilde{u}. \end{aligned}$$
(4.2)

Let \(U_1(z,\theta ),\ldots ,U_{n-1}(z,\theta )\) be an orthonormal frame of \(T^{1,0}_{(z,\theta )}X\) varying smoothly with \((z,\theta )\) in a neighborhood of \(p\). We take \(U_1,\ldots ,U_{n-1}\) so that \(U_j(0,0)=\frac{\partial }{\partial z_j}\), \(j=1,\ldots ,n-1\). Put

$$\begin{aligned} U_j(z,\theta )=\sum ^{n-1}_{t=1}a_{j,t}(z,\theta )Z_t,\quad j=1,\ldots ,n-1, \end{aligned}$$
(4.3)

where \(a_{j,t}\in C^\infty \), \(j,t=1,\ldots ,n-1\), \(Z_1,\ldots ,Z_{n-1}\) are as in (4.1). Then, we have

$$\begin{aligned} a_{j,t}(z,\theta )=\delta _{j,t}+O(\left| (z,\theta )\right| ),\quad j,t=1,\ldots ,n-1. \end{aligned}$$
(4.4)

Let \((e_j(z, \theta ))_{j=1,\ldots ,n-1}\) denote the basis of \(T^{*0,1}_{(z,\theta )}X\), dual to \((\overline{U}_j(z,\theta ))_{j=1,\ldots ,n-1}\). If \(w\in T^{*0,1}_zX\), let \((w\wedge )^*: \Lambda ^{0,q+1}T^*_zX\rightarrow \Lambda ^{0,q}T^*_zX,\ q\geqslant 0\), be the adjoint of the left exterior multiplication \(w\wedge : \Lambda ^{0,q}T^*_zX\rightarrow \Lambda ^{0,q+1}T^*_zX\), \(u\mapsto w\wedge u\):

$$\begin{aligned} \langle w\wedge u|v\rangle =\langle u|(w\wedge )^*v\rangle , \end{aligned}$$
(4.5)

for all \(u\in \Lambda ^{0,q}T^*_zX\), \(v\in \Lambda ^{0,q+1}T^*_zX\). Notice that \((w\wedge )^*\) depends \(\mathbb {C}\)-anti-linearly on \(w\). It is easy to see that

$$\begin{aligned} \overline{\partial }_b=\sum ^{n-1}_{j=1}e_j\wedge \overline{U}_j +\sum ^{n-1}_{j=1}(\overline{\partial }_b e_j)\wedge (e_j\wedge )^* \end{aligned}$$
(4.6)

and correspondingly

$$\begin{aligned} \overline{\partial }^{\,*,k\phi }_b=\sum ^{n-1}_{j=1}(e_j\wedge )^*\overline{U}^{\,*,k\phi }_j +\sum ^{n-1}_{j=1}e_j\wedge (\overline{\partial }_b e_j\wedge )^*, \end{aligned}$$
(4.7)

where \(\overline{U}^{\,*,k\phi }_j\) is the formal adjoint of \(\overline{U}_j\) with respect to \((\ |\ )_{k\phi }\), \(j=1,\ldots ,n-1\). We can check that for \(j=1,\ldots ,n-1\),

$$\begin{aligned} \overline{U}_j^{\,*,k\phi }=-U_j+k(U_j\phi )+s_j(z, \theta ), \end{aligned}$$
(4.8)

where \(s_j\in C^\infty (D)\), \(s_j\) is independent of \(k\), \(j=1,\ldots ,n-1\).

For \(r>0\), let \(D_r=\left\{ x=(z, \theta )\in \mathbb {R}^{2n-1};\, \left| x_j\right| <r,\ j=1,\ldots ,2n-1\right\} \). Let \(F_k\) be the scaling map: \(F_k(z, \theta )=(\frac{z}{\sqrt{k}}, \frac{\theta }{k})\). From now on, we assume that \(k\) is large enough so that \(F_k(D_{\log k})\subset D\). We define the scaled bundle \(F^*_k\Lambda ^{0,q}T^*X\) on \(D_{\log k}\) to be the bundle whose fiber at \((z, \theta )\in D_{\log k}\) is

$$\begin{aligned} F^*_k\Lambda ^{0,q}T^*_{(z,\theta )}X:=\Bigr \{\textstyle {\sum \nolimits ^{'}}_{\left| J\right| =q}a_Je_J\left( \frac{z}{\sqrt{k}},\frac{\theta }{k}\right) ;\, a_J\in \mathbb {C},\left| J\right| =q\Bigl \}. \end{aligned}$$

We take the Hermitian metric \(\langle \,\cdot \,|\,\cdot \,\rangle _{F^*_k}\) on \(F^*_k\Lambda ^{0,q}T^*X\) so that at each point \((z, \theta )\in D_{\log k}\),

$$\begin{aligned} \Bigr \{e_J\big (\tfrac{z}{\sqrt{k}}\;,\tfrac{\theta }{k}\big ); \left| J\right| =q, J\text { strictly increasing}\Bigl \}, \end{aligned}$$

is an orthonormal basis for \(F^*_k\Lambda ^{0,q}T^*_{(z,\theta )}X\). For \(r>0\), let \(F^*_k\Omega ^{0,q}(D_r)\) denote the space of smooth sections of \(F^*_k\Lambda ^{0,q}T^*X\) over \(D_r\). Let \(F^*_k\Omega ^{0,q}_0(D_r)\) be the subspace of \(F^*_k\Omega ^{0,q}(D_r)\) whose elements have compact support in \(D_r\). Given \(f\in \Omega ^{0,q}(F_k(D_{\log k}))\), we write \(f=\sum '_{\left| J\right| =q}f_Je_J\). We define the scaled form \(F_k^*f\in F^*_k\Omega ^{0,q}(D_{\log k})\) by:

$$\begin{aligned} F_k^*f={\mathop {\mathop {\sum }\nolimits ^{'}}\limits _{\left| J\right| =q}}f_J\Big (\frac{z}{\sqrt{k}}, \frac{\theta }{k}\Big )e_J\Big (\frac{z}{\sqrt{k}},\frac{\theta }{k}\Big ). \end{aligned}$$
(4.9)

Let \(P\) be a partial differential operator of order one on \(F_k(D_{\log k})\) with \(C^\infty \) coefficients. We write \(P=a(z, \theta )\frac{\partial }{\partial \theta }+\sum ^{2n-2}_{j=1}a_j(z, \theta )\frac{\partial }{\partial x_j}\), \(a, a_j\in C^\infty (F_k(D_{\log k}))\), \(j=1,\ldots ,2n-2\). The partial differential operator \(P_{(k)}\) on \(D_{\log k}\) is given by

$$\begin{aligned} P_{(k)}=\sqrt{k}F^*_ka\frac{\partial }{\partial \theta }+\sum ^{2n-2}_{j=1}F^*_ka_j\frac{\partial }{\partial x_j} =\sqrt{k}a\Big (\frac{z}{\sqrt{k}},\frac{\theta }{k}\Big )\frac{\partial }{\partial \theta }+\sum ^{2n-2}_{j=1}a_j\Big (\frac{z}{\sqrt{k}}, \frac{\theta }{k}\Big )\frac{\partial }{\partial x_j}.\nonumber \\ \end{aligned}$$
(4.10)

Let \(f\in C^\infty (F_k(D_{\log k}))\). We can check that

$$\begin{aligned} P_{(k)}(F^*_kf)=\frac{1}{\sqrt{k}}F^*_k(Pf). \end{aligned}$$
(4.11)

The scaled differential operator \(\overline{\partial }_{b,(k)}:F^*_k\Omega ^{0,q}(D_{\log k})\rightarrow F^*_k\Omega ^{0,q+1}(D_{\log k})\) is given by (compare to the formula (4.6) for \(\overline{\partial }_b\)):

$$\begin{aligned}&\overline{\partial }_{b,(k)}=\sum ^{n-1}_{j=1}e_j\Big (\frac{z}{\sqrt{k}},\frac{\theta }{k}\Big )\wedge \overline{U}_{j(k)}\nonumber \\&\quad \quad \quad \quad +\sum ^{n-1}_{j=1}\frac{1}{\sqrt{k}}(\overline{\partial }_b e_j)\Big (\frac{z}{\sqrt{k}},\frac{\theta }{k}\Big )\wedge \Big (e_j\Big (\frac{z}{\sqrt{k}},\frac{\theta }{k}\Big )\wedge \Big )^*. \end{aligned}$$
(4.12)

From (4.6) and (4.11), we can check that if \(f\in \Omega ^{0,q}(F_k(D_{\log k}))\), then

$$\begin{aligned} \overline{\partial }_{b,(k)}F^*_kf=\frac{1}{\sqrt{k}}F^*_k(\overline{\partial }_bf). \end{aligned}$$
(4.13)

Let \((\ |\ )_{kF^*_k\phi }\) be the inner product on the space \(F^*_k\Omega ^{0,q}_0(D_{\log k})\) defined as follows:

$$\begin{aligned} (f\ |\ g)_{kF^*_k\phi }=\int _{D_{\log k}}\!\langle f|g\rangle _{F^*_k}\mathrm{e}^{-kF^*_k\phi }(F^*_km)(z, \theta )\mathrm{d}v(z)\mathrm{d}\theta , \end{aligned}$$

where \(\mathrm{d}v_X=m\mathrm{d}v(z)\mathrm{d}\theta \) is the volume form, \(\mathrm{d}v(z)=2^{n-1}\mathrm{d}x_1\cdots \mathrm{d}x_{2n-2}\). Note that \(m(0,0)=1\). Let \(\overline{\partial }^{\,*,k\phi }_{b,(k)}:F^*_k\Omega ^{0,q+1}(D_{\log k})\rightarrow F^*_k\Omega ^{0,q}(D_{\log k})\) be the formal adjoint of \(\overline{\partial }_{b,(k)}\) with respect to \((\ |\ )_{kF^*_k\phi }\). Then, we can check that [compare the formula for \(\overline{\partial }^{\,*,k\phi }_b\), see (4.7) and (4.8)]

$$\begin{aligned}&\overline{\partial }^{\,*,k\phi }_{b,(k)}=\sum ^{n-1}_{j=1}\Big (e_j\Big (\frac{z}{\sqrt{k}},\frac{\theta }{k}\Big )\wedge \Big )^*\Bigr (-U_{j(k)}+\sqrt{k}F^*_k(U_j\phi )+\frac{1}{\sqrt{k}}F^*_ks_j\Bigr )\nonumber \\&\quad \quad \quad +\sum ^{n-1}_{j=1}\frac{1}{\sqrt{k}}\,e_j\Big (\frac{z}{\sqrt{k}},\frac{\theta }{k}\Big )\wedge \Big ((\overline{\partial }_b e_j)\Big (\frac{z}{\sqrt{k}},\frac{\theta }{k}\Big )\wedge \Big )^*, \end{aligned}$$
(4.14)

where \(s_j\in C^\infty (D_{\log k})\), \(j=1,\ldots ,n-1\), are independent of \(k\). We also have

$$\begin{aligned} \overline{\partial }^{\,*,k\phi }_{b,(k)}F^*_kf=\frac{1}{\sqrt{k}}F^*_k(\overline{\partial }^{\,*,k\phi }_bf),\quad f\in \Omega ^{0,q+1}(F_k(D_{\log k})). \end{aligned}$$
(4.15)

We define now the scaled Kohn Laplacian:

$$\begin{aligned} \Box ^{(q)}_{b,k\phi ,(k)}:=\overline{\partial }^{\,*,k\phi }_{b,(k)}\overline{\partial }_{b,(k)}+\overline{\partial }_{b,(k)}\overline{\partial }^{\,*,k\phi }_{b,(k)}:F^*_k\Omega ^{0,q}(D_{\log k})\rightarrow F^*_k\Omega ^{0,q}(D_{\log k}). \end{aligned}$$
(4.16)

From (4.13) and (4.15), we see that if \(f\in \Omega ^{0,q}(F_k(D_{\log k}))\), then

$$\begin{aligned} \left( \Box ^{(q)}_{b,k\phi ,(k)}\right) F^*_kf=\frac{1}{k}F^*_k(\Box ^{(q)}_{b,k\phi }f). \end{aligned}$$
(4.17)

From (4.3), (4.4) and (4.1), we can check that

$$\begin{aligned} \overline{U}_{j(k)}=\frac{\partial }{\partial \overline{z}_j}-i\lambda _jz_j\frac{\partial }{\partial \theta }+\epsilon _kZ_{j,k},\quad j=1,\ldots ,n-1, \end{aligned}$$
(4.18)

on \(D_{\log k}\), where \(\epsilon _k\) is a sequence tending to zero with \(k\rightarrow \infty \) and \(Z_{j,k}\) is a first order differential operator and all the derivatives of the coefficients of \(Z_{j,k}\) are uniformly bounded in \(k\) on \(D_{\log k}\), \(j=1,\ldots ,n-1\). Similarly, from (4.3), (4.4) and (4.1), it is straightforward to see that

$$\begin{aligned}&-U_{t(k)}+\sqrt{k}F^*_k(U_t\phi )+\frac{1}{\sqrt{k}}F^*_ks_t\nonumber \\&\quad =-\frac{\partial }{\partial z_t}-i\lambda _t\overline{z}_t\frac{\partial }{\partial \theta }+i\lambda _t\overline{z}_t\beta \!+\!\sum ^{n-1}_{j=1}\mu _{j,\,t}\,\overline{z}_j\!+\!\delta _kV_{t,\,k},\quad t=1,\ldots ,n-1,\qquad \end{aligned}$$
(4.19)

on \(D_{\log k}\), where \(\delta _k\) is a sequence tending to zero with \(k\rightarrow \infty \) and \(V_{t,k}\) is a first order differential operator and all the derivatives of the coefficients of \(V_{t,k}\) are uniformly bounded in \(k\) on \(D_{\log k}\), \(t=1,\ldots ,n-1\). From (4.19), (4.18) and (4.16), (4.14), (4.12), it is straightforward to obtain the following

Proposition 4.1

We have that

$$\begin{aligned}&\Box ^{(q)}_{b,k\phi ,(k)}=\sum ^{n-1}_{j=1}\Biggr [\Biggr (-\frac{\partial }{\partial z_j}-i\lambda _j\overline{z}_j\frac{\partial }{\partial \theta }+i\lambda _j\overline{z}_j\beta +\sum ^{n-1}_{t=1}\mu _{t,\,j}\,\overline{z}_t\Biggr )\Bigr (\frac{\partial }{\partial \overline{z}_j}-i\lambda _jz_j\frac{\partial }{\partial \theta }\Bigr )\Biggr ]\nonumber \\&\qquad \quad \qquad +\sum ^{n-1}_{j,\,t=1}e_j\Bigg (\frac{z}{\sqrt{k}}, \frac{\theta }{k}\Bigg )\wedge \Big (e_t\Bigg (\frac{z}{\sqrt{k}},\frac{\theta }{k}\Bigg )\wedge \Big )^*\Bigr (\Bigr (\mu _{j,\,t}+i\lambda _j\delta _{j,\,t}\beta \Bigr )-2i\lambda _j\delta _{j,\,t}\,\frac{\partial }{\partial \theta }\Biggr )\nonumber \\&\qquad \quad \qquad +\varepsilon _kP_k, \end{aligned}$$

on \(D_{\log k},\) where \(\varepsilon _k\) is a sequence tending to zero with \(k\rightarrow \infty ,\) \(P_k\) is a second order differential operator and all the derivatives of the coefficients of \(P_k\) are uniformly bounded in \(k\) on \(D_{\log k}.\)

Let \(D\subset D_{\log k}\) be an open set and let \(W^s_{kF^*_k\phi }(D,F^*_k\Lambda ^{0, q}T^*X)\), \(s\in {\mathbb {N}}_0\), denote the Sobolev space of order \(s\) of sections of \(F^*_k\Lambda ^{0,q}T^*X\) over \(D\) with respect to the weight \(\mathrm{e}^{-kF^*_k\phi }\). The Sobolev norm on this space is given by

$$\begin{aligned} \left\| u\right\| ^2_{kF^*_k\phi ,s,D} ={\mathop {\mathop {\sum }\nolimits ^{'}}\limits _{\begin{array}{c} \alpha \in {\mathbb {N}}^{2n-1}_0,\;\left| \alpha \right| \leqslant s\;\\ {\left| J\right| =q} \end{array}}} \int _{D}\!\left| \partial ^\alpha _{x}u_J\right| ^2\mathrm{e}^{-kF^*_k\phi }(F^*_km)(z, \theta )\mathrm{d}v(z)\mathrm{d}\theta , \end{aligned}$$
(4.20)

where \(u=\sum '_{\left| J\right| =q}u_Je_J\big (\frac{z}{\sqrt{k}},\frac{\theta }{k}\big )\in W^s_{kF^*_k\phi } (D,F^*_k\Lambda ^{0,q}T^*X)\) and \(m\) is the volume form. If \(s=0\), we write \(\left\| \cdot \right\| _{kF^*_k\phi ,D}\) to denote \(\left\| \cdot \right\| _{kF^*_k\phi ,0,D}\). The following is well known (see [9, Proposition 2.4 and Lemma 2.5])

Proposition 4.2

Assume that \(Y(q)\) holds at each point of \(X\). For every \(r>0\) with \(D_{2r}\subset D_{\log k}\) and \(s\in {\mathbb {N}}_0,\) there are constants \(C_{r,s}>0,\) \(C_r>0,\) \(C_{r,s}\) and \(C_r\) are independent of \(k,\) such that

$$\begin{aligned} \left\| u\right\| ^2_{kF^*_k\phi ,s+1,D_{r}}\leqslant C_{r,s}\Bigr (\left\| u\right\| ^2_{kF^*_k\phi ,D_{2r}}+\big \Vert \Box ^{(q)}_{b,k\phi ,(k)}u\big \Vert ^2_{kF^*_k\phi ,s,D_{2r}}\Bigl ),\quad u\in F^*_k\Omega ^{0,q}(D_{\log k})\nonumber \\ \end{aligned}$$
(4.21)

and

$$\begin{aligned} \sup _{x\in D_r}\left| u(x)\right| ^2 \leqslant C_r\Bigr (\left\| u\right\| ^2_{kF^*_k\phi ,D_{2r}}+\sum ^n_{m=1}\big \Vert (\Box ^{(q)}_{b,k\phi ,(k)})^mu\big \Vert ^2_{kF^*_k\phi ,D_{2r}}\Bigl ),\quad u\in F^*_k\Omega ^{0,q}(D_{\log k}).\nonumber \\ \end{aligned}$$
(4.22)

We pause and introduce some notations. We identify \(\mathbb {R}^{2n-1}\) with the Heisenberg group \(H_n:=\mathbb {C}^{n-1}\times \mathbb {R}\). We also write \((z, \theta )\) to denote the coordinates of \(H_n\), \(z=(z_1,\ldots ,z_{n-1})\in \mathbb {C}^{n-1}\), \(z_j=x_{2j-1}+ix_{2j}\), \(j=1,\ldots ,n-1\), and \(\theta \in \mathbb {R}\). Then,

$$\begin{aligned}&\Big \{U_{j,H_n}=\frac{\partial }{\partial z_j}+i\lambda _j\overline{z}_j\frac{\partial }{\partial \theta };\, j=1,\ldots ,n-1\Big \},\,\\&\Big \{U_{j,H_n},\ \overline{U}_{j,H_n},\ T=\frac{\partial }{\partial \theta };\,j=1,\ldots ,n-1\Big \} \end{aligned}$$

are orthonormal bases for the bundles \(T^{1, 0}H_n\) and \(\mathbb {C}TH_n\), respectively. Then,

$$\begin{aligned} \Big \{\mathrm{d}z_j,\ \mathrm{d}\overline{z}_j,\ \omega _0=-\mathrm{d}\theta +\sum ^{n-1}_{j=1}(i\lambda _j\overline{z}_j\mathrm{d}z_j-i\lambda _jz_j\mathrm{d}\overline{z}_j);j=1,\ldots ,n-1\Big \} \end{aligned}$$

is the basis of \(\mathbb {C}T^*H_n\) which is dual to \(\{U_{j,H_n},\overline{U}_{j,H_n}, -T;j=1,\ldots ,n-1\}\). We take the Hermitian metric \(\langle \,\cdot \,|\,\cdot \,\rangle \) on \(\Lambda ^{0,q}T^*H_n\) such that

$$\begin{aligned} \{\mathrm{d}\overline{z}_J;\, \left| J\right| =q, J\text { strictly increasing}\} \end{aligned}$$

is an orthonormal basis of \(\Lambda ^{0,q}T^*H_n\). The Cauchy–Riemann operator \(\overline{\partial }_{b,H_n}\) on \(H_n\) is given by

$$\begin{aligned} \overline{\partial }_{b,H_n}=\sum ^{n-1}_{j=1}\mathrm{d}\overline{z}_j\wedge \overline{U}_{j,H_n}:\Omega ^{0,q}(H_n)\rightarrow \Omega ^{0,q+1}(H_n). \end{aligned}$$
(4.23)

Put \(\phi _0(z, \theta )=\beta \theta +\sum ^{n-1}_{j,t=1}\mu _{j,\,t}\overline{z}_jz_t\in C^\infty (H_n,\mathbb {R})\), where \(\beta \) and \(\mu _{j,\,t}\), \(j,t=1,\ldots ,n-1\), are as in (4.1). Note that

$$\begin{aligned} \sup _{(z, \theta )\in D_{\log k}}\left| kF^*_k\phi -\phi _0\right| \rightarrow 0,\quad \hbox {as } k\rightarrow \infty . \end{aligned}$$
(4.24)

Let \((\ |\ )_{\phi _0}\) be the inner product on \(\Omega ^{0,q}_0(H_n)\) defined as follows:

$$\begin{aligned} (f\ |\ g)_{\phi _0}=\int _{H_n}\!\langle f|g\rangle \mathrm{e}^{-\phi _0}\mathrm{d}v(z)\mathrm{d}\theta , \quad f, g\in \Omega ^{0,q}_0(H_n), \end{aligned}$$

where \(\mathrm{d}v(z)=2^{n-1}\mathrm{d}x_1\mathrm{d}x_2\cdots \mathrm{d}x_{2n-2}\). Let \(\overline{\partial }^{\,*,\phi _0}_{b,H_n}:\Omega ^{0,q+1}(H_n)\rightarrow \Omega ^{0,q}(H_n)\) be the formal adjoint of \(\overline{\partial }_{b,H_n}\) with respect to \((\ |\ )_{\phi _0}\). We have

$$\begin{aligned} \overline{\partial }^{\,*,\phi _0}_{b,H_n}=\sum ^{n-1}_{t=1}(\mathrm{d}\overline{z}_t\wedge )^*\,\overline{U}_{t,H_n}^{\,*,\phi _0}:\Omega ^{0,q+1}(H_n)\rightarrow \Omega ^{0,q}(H_n), \end{aligned}$$
(4.25)

where

$$\begin{aligned} \overline{U}_{t,H_n}^{\,*,\phi _0}=-U_{t,H_n}+U_{t,H_n}\phi _0=-U_{t,H_n}+\sum ^{n-1}_{j=1}\mu _{j,\,t}\overline{z}_j+i\lambda _t\overline{z}_t\beta . \end{aligned}$$
(4.26)

The Kohn Laplacian on \(H_n\) is given by

$$\begin{aligned} \Box ^{(q)}_{b,H_n}=\overline{\partial }_{b,H_n}\overline{\partial }^{\,*,\phi _0}_{b,H_n}+\overline{\partial }^{\,*,\phi _0}_{b,H_n}\overline{\partial }_{b,H_n}:\Omega ^{0,q}(H_n)\rightarrow \Omega ^{0,q}(H_n). \end{aligned}$$
(4.27)

From (4.23), (4.25) and (4.26), we can check that

$$\begin{aligned} \Box ^{(q)}_{b,H_n}&= \sum ^{n-1}_{j=1}\overline{U}^{\,*,\phi _0}_{j,H_n}\overline{U}_{j,H_n} +\sum ^{n-1}_{j,\,t=1}\mathrm{d}\overline{z}_j\wedge (\mathrm{d}\overline{z}_t\wedge )^*\Bigr [\Bigr (\mu _{j,\,t}+i\lambda _j\delta _{j,\,t}\beta \Bigr )-2i\lambda _j\delta _{j,\,t}\,\frac{\partial }{\partial \theta }\Bigr ]\nonumber \\&= \sum ^{n-1}_{j=1}\Bigr [\Biggr (-\frac{\partial }{\partial z_j}-i\lambda _j\overline{z}_j\frac{\partial }{\partial \theta }+i\lambda _j\overline{z}_j\beta +\sum ^{n-1}_{t=1}\mu _{t,j}\overline{z}_t\Biggr )\Biggr (\frac{\partial }{\partial \overline{z}_j}-i\lambda _jz_j\frac{\partial }{\partial \theta }\Biggr )\Bigr ]\nonumber \\&+\sum ^{n-1}_{j,t=1}\mathrm{d}\overline{z}_j\wedge (\mathrm{d}\overline{z}_t\wedge )^*\Bigr [\Bigr (\mu _{j,\,t}+i\lambda _j\delta _{j,\,t}\beta \Bigr )-2i\lambda _j\delta _{j,\,t}\,\frac{\partial }{\partial \theta }\Bigr ]. \end{aligned}$$
(4.28)

Now, we can prove

Proposition 4.3

Assume that \(Y(q)\) holds at each point of \(X\). For each \(k,\) let \(\alpha _k\in F^*_k\Omega ^{0,q}(D_{\log k})\). We assume that \(\left\| \alpha _k\right\| _{kF^*_k\phi ,D_{\log k}}\leqslant 1\) for each \(,k\) and there is a sequence \(\nu _k>0,\) \(\nu _k\rightarrow 0\) as \(k\rightarrow \infty ,\) such that for each \(k,\)

$$\begin{aligned} \left\| (\Box ^{(q)}_{b,k\phi ,(k)})^m\alpha _k\right\| _{kF^*_k\phi ,D_{\log k}}\le \nu _k^m,\quad \forall m\in {\mathbb {N}}. \end{aligned}$$

Identify \(\alpha _k\) with a form on \(H_n\) by extending it with zero and write

$$\begin{aligned} \alpha _k={\mathop {\mathop {\sum }\nolimits ^{'}}\limits _{\left| J\right| =q}}\alpha _{k,J}e_J \left( \frac{z}{\sqrt{k}},\frac{\theta }{k}\right) . \end{aligned}$$

Then\(,\) there is a subsequence \(\left\{ \alpha _{k_j}\right\} \) of \(\left\{ \alpha _k\right\} \) such that for each strictly increasing multiindex \(J,\) \(\left| J\right| =q,\) \(\alpha _{k_j,\,J}\) converges uniformly with all its derivatives on any compact subset of \(H_n\) to a smooth function \(\alpha _J.\) Furthermore\(,\) if we put \(\alpha =\sum '_{\left| J\right| =q}\alpha _J\mathrm{d}\overline{z}_J,\) then \(\Box ^{(q)}_{b,H_n}\alpha =0.\)

Proof

From (4.21) and using induction, we get for any \(r>0\) and for every \(s\in {\mathbb {N}}_0\), there is a constant \(C_{r,s}>0\) independent of \(k\), such that

$$\begin{aligned} \left\| \alpha _k\right\| ^2_{kF^*_k\phi ,s+1,D_r}&\leqslant C_{r,s}\Bigr (\left\| \alpha _k\right\| ^2_{kF^*_k\phi ,D_{2r}}+\sum ^{s+1}_{m=1} \left\| (\Box ^{(q)}_{b,k\phi ,(k)})^m\alpha _k\right\| ^2_{kF^*_k\phi ,D_{2r}}\,\Bigr )\nonumber \\&\le C_{r,s}\Bigr (1+\sum ^{\infty }_{m=1}\nu _k^m\Bigr )\le \widetilde{C}_{r,s} \end{aligned}$$
(4.29)

for \(k\) large, where \(\widetilde{C}_{r,s}>0\) is independent of \(k\). Fix a strictly increasing multiindex \(J\), \(\left| J\right| =q\), and \(r>0\). Combining (4.29) with Rellich’s compactness theorem [15, p. 281], we conclude that there is a subsequence of \(\left\{ \alpha _{k,J}\right\} \), which converges in all Sobolev spaces \(W^s(D_r)\) for \(s>0\). From the Sobolev embedding theorem [15, p. 170], we see that the sequence converges in all \(C^l(D_r)\), \(l\geqslant 0\), \(l\in {\mathbb {Z}}\), locally uniformly. Choosing a diagonal sequence, with respect to a sequence of \(D_r\) exhausting \(H_n\), we get a subsequence \(\left\{ \alpha _{k_j,J}\right\} \) of \(\left\{ \alpha _{k,J}\right\} \) such that \(\alpha _{k_j,J}\) converges uniformly with all derivatives on any compact subset of \(H_n\) to a smooth function \(\alpha _J\).

Let \(J'\) be another strictly increasing multiindex, \(\left| J'\right| =q\). We can repeat the procedure above and get a subsequence \(\left\{ \alpha _{k_{j_s},J'}\right\} \) of \(\left\{ \alpha _{k_j,J'}\right\} \) such that \(\alpha _{k_{j_s},J'}\) converges uniformly with all derivatives on any compact subset of \(H_n\) to a smooth function \(\alpha _{J'}\). Continuing in this way, we get the first statement of the proposition.

Now, we prove the second statement of the proposition. Let \(P=(p_1,\ldots ,p_q)\), \(R=(r_1,\ldots ,r_q)\) be multiindices, \(\left| P\right| =\left| R\right| =q\). Define

$$\begin{aligned} \varepsilon ^P_R=\left\{ \begin{array}{l} 0,\quad \hbox {if } \left\{ p_1,\ldots ,p_q\right\} \ne \left\{ r_1,\ldots ,r_q\right\} , \\ \hbox {the sign of permutation } \left( \begin{array}{c} P \\ R \end{array}\right) ,\quad \hbox {if } \left\{ p_1,\ldots ,p_q\right\} =\left\{ r_1,\ldots ,r_q\right\} . \end{array}\right. \end{aligned}$$

For \(j, t=1,\ldots ,n-1\), define

$$\begin{aligned} \sigma ^{jtP}_{R}=\left\{ \begin{array}{ll} 0,&{}\quad \hbox {if } \mathrm{d}\overline{z}_j\wedge (\mathrm{d}\overline{z}_t\wedge )^*(\mathrm{d}\overline{z}^P)=0, \\ \varepsilon ^Q_R,&{}\quad \hbox {if } \mathrm{d}\overline{z}_j\wedge (\mathrm{d}\overline{z}_t\wedge )^*(\mathrm{d}\overline{z}^P)=\mathrm{d}\overline{z}^Q, \left| Q\right| =q. \end{array}\right. \end{aligned}$$

We may assume that \(\alpha _{k,J}\) converges uniformly with all derivatives on any compact subset of \(H_n\) to a smooth function \(\alpha _J\), for all strictly increasing \(J\), \(\left| J\right| =q\). As (4.29), we have for any \(r>0\) and for every \(s\in {\mathbb {N}}_0\), there is a constant \(C_{r,s}>0\) independent of \(k\), such that

$$\begin{aligned}&\left\| \Box ^{(q)}_{k,k\phi ,(k)}\alpha _k\right\| ^2_{kF^*_k\phi ,s+1,D_r}\nonumber \\&\quad \leqslant C_{r,s}\Biggr (\left\| \Box ^{(q)}_{k,k\phi ,(k)}\alpha _k\right\| ^2_{kF^*_k\phi ,D_{2r}}+\sum ^{s+1}_{m=1} \left\| (\Box ^{(q)}_{b,k\phi ,(k)})^{m+1}\alpha _k\right\| ^2_{kF^*_k\phi ,D_{2r}}\,\Biggr )\nonumber \\&\quad \le C_{r,s}\sum ^{\infty }_{m=1}\nu _k^m\rightarrow 0\quad \hbox {as } k\rightarrow \infty . \end{aligned}$$
(4.30)

Put

$$\begin{aligned} \beta _k:=\Box ^{(q)}_{k,k\phi ,(k)}\alpha _k={\mathop {\mathop {\sum }\nolimits ^{'}}\limits _{\left| J\right| =q}}\beta _{k,J}e_J \left( \frac{z}{\sqrt{k}},\frac{\theta }{k}\right) \in F^*_k\Omega ^{0,q}(D_{\log k}). \end{aligned}$$

Combining (4.30) with Sobolev embedding theorem [15, p. 170], we conclude that

$$\begin{aligned}&\beta _{k,J}\hbox { converges uniformly with all derivatives on any compact subset of } H_n\nonumber \\&\quad \hbox { to zero, for all strictly increasing } J, \left| J\right| =q. \end{aligned}$$
(4.31)

From the explicit formula of \(\Box ^{(q)}_{b,k\phi ,(k)}\) (see Proposition 4.1), it is not difficult to see that for all strictly increasing \(J\), \(\left| J\right| =q\), we have

$$\begin{aligned}&\sum ^{n-1}_{j=1}\overline{U}^{\,*,\phi _0}_{j,H_n}\overline{U}_{j,H_n}\alpha _{k,J}=-{\mathop {\mathop {\sum }\nolimits ^{'}}\limits _{\begin{array}{c} \left| P\right| =q,\\ 1\leqslant \, j,\,t\leqslant n-1 \end{array}}} \sigma ^{jtP}_{J} \Bigr [\left( \mu _{j,\,t}+i\lambda _j\delta _{j,\,t}\beta \Bigr )-2i\lambda _j\delta _{j,\,t}\frac{\partial }{\partial \theta }\right] \alpha _{k,P}\nonumber \\&\qquad \qquad \qquad \qquad \qquad \qquad \qquad \quad +\epsilon _kP_{k,J}\alpha _k+\beta _{k,J} \end{aligned}$$
(4.32)

on \(D_{\log k}\), where \(\epsilon _k\) is a sequence tending to zero with \(k\rightarrow \infty \) , \(P_{k,J}\) is a second order differential operator and all the derivatives of the coefficients of \(P_{k,J}\) are uniformly bounded in \(k\) on \(D_{\log k}\) and \(\beta _{k,J}\) is as in (4.31). By letting \(k\rightarrow \infty \) in (4.32) we get

$$\begin{aligned} \sum ^{n-1}_{j=1}\overline{U}^{\,*,\phi _0}_{j,H_n}\overline{U}_{j,H_n}\alpha _{J}=-{\mathop {\mathop {\sum }\nolimits ^{'}}\limits _{\begin{array}{c} \left| P\right| =q,\\ 1\leqslant \, j,\,t\leqslant n-1 \end{array}}} \sigma ^{jtP}_{J} \Bigr [\Bigr (\mu _{j,\,t}+i\lambda _j\delta _{j,\,t}\beta \Bigr )-2i\lambda _j\delta _{j,\,t}\frac{\partial }{\partial \theta }\Bigr ]\alpha _{P}\qquad \end{aligned}$$
(4.33)

on \(H_n\), for all strictly increasing \(J\), \(\left| J\right| =q\). From this and the explicit formula of \(\Box ^{(q)}_{b,H_n}\) [see (4.28)], we conclude that \(\Box ^{(q)}_{b,H_n}\alpha =0\). The proposition follows. \(\square \)

5 The operators \(Q^{(0)}_{M,k}\) and \(Q^{(1)}_{M,k}\)

From now on, we assume that \(h^L\) is semi-positive on \(X\) and positive at some point of \(X\) and conditions \(Y(0)\) and \(Y(1)\) hold at each point of \(X\).

Take \(\delta _0>0\) be a small constant so that

$$\begin{aligned} M^\phi _x+2s{\mathcal L}_x\ge 0,\quad \forall \left| s\right| \le \delta _0,\ \forall x\in X. \end{aligned}$$
(5.1)

Take \(\psi (\eta )\in C^\infty _0(]-\delta _0,\delta _0[,\overline{\mathbb {R}}_+)\) so that \(\psi (\eta )=1\) if \(-\frac{\delta _0}{2}\le \eta \le \frac{\delta _0}{2}\). Let \(\hat{\psi }(t)=\int \mathrm{e}^{-it\eta }\psi (\eta )\mathrm{d}\eta \) be the Fourier transform of \(\psi \). Put

$$\begin{aligned} C_0:=\sup _{t\in \mathbb {R}}t^2\left| \hat{\psi }(t)\right| . \end{aligned}$$
(5.2)

Let \(E>0\) be a small constant so that

$$\begin{aligned}&\sqrt{\int _XE^2\mathrm{d}v_X(x)}\sqrt{2(2\pi )^{-n}\int _X\Biggr (\int \det (M^\phi _x+2\xi {\mathcal L}_x){1\!\!1}_{\mathbb {R}_{x,0}}(\xi )\mathrm{d}\xi \Biggr )\mathrm{d}v_X(x)}\nonumber \\&\quad \le \frac{(2\pi )^{1-n}}{4} \int _X\Biggr (\int \psi (\xi )\det (M^\phi _x+2\xi {\mathcal L}_x){1\!\!1}_{\mathbb {R}_{x,0}}(\xi )\mathrm{d}\xi \Biggr )\mathrm{d}v_X(x). \end{aligned}$$
(5.3)

Fix \(M>0\) be a large constant so that

$$\begin{aligned}&\frac{2C_0}{M}(2\pi )^{-n}\int _X\Biggr (\int \det (M^\phi _x+2\xi {\mathcal L}_x){1\!\!1}_{\mathbb {R}_{x,0}}(\xi )\mathrm{d}\xi \Biggr )\mathrm{d}v_X(x)\nonumber \\&\quad <\frac{(2\pi )^{1-n}}{2} \int _X\Bigr (\int \psi (\xi )\det (M^\phi _x+2\xi {\mathcal L}_x){1\!\!1}_{\mathbb {R}_{x,0}}(\xi )\mathrm{d}\xi \Bigr )\mathrm{d}v_X(x) \end{aligned}$$
(5.4)

and

$$\begin{aligned} \frac{2C_0}{M}\Biggr ((2\pi )^{-n}\int \left| \det (M^\phi _x+2\xi {\mathcal L}_x)\right| {1\!\!1}_{\mathbb {R}_{x,1}}(\xi )\mathrm{d}\xi \Biggr )^{\frac{1}{2}}<\frac{E}{\sqrt{n-1}},\quad \forall x\in X, \end{aligned}$$
(5.5)

where \({1\!\!1}_{\mathbb {R}_{x,1}}(\xi )=1\) if \(\xi \in \mathbb {R}_{x,1}\), \({1\!\!1}_{\mathbb {R}_{x,1}}(\xi )=0\) if \(\xi \notin \mathbb {R}_{x,1}\). Take \(\chi (t)\in C^\infty _0(]-2,2[,\overline{\mathbb {R}}_+)\) so that \(0\le \chi (t)\le 1\) and \(\chi (t)=1\) if \(-1\le t\le 1\) and \(\chi (-t)=\chi (t)\), for all \(t\in \mathbb {R}\). Put

$$\begin{aligned} \chi _M(t):=\chi \left( \frac{t}{M}\right) . \end{aligned}$$
(5.6)

As before, let \(\Phi ^t(x)\) be the \(T\)-flow. The operator \(Q^{(0)}_{M,k}\) is a continuous operator \(C^\infty (X,L^k)\rightarrow C^\infty (X,L^k)\) defined as follows. Let \(u\in C^\infty (X,L^k)\). Let \(D\Subset D'\Subset X\) be open sets of \(X\) and let \(s\) be a local section of \(L\) on \(D'\), \(\left| s\right| ^2_{h^L}=\mathrm{e}^{-\phi }\). On \(D'\), we write \(u=s^k\widetilde{u}\), \(\widetilde{u}\in C^\infty (D')\). Then,

$$\begin{aligned} \left( Q^{(0)}_{M,k}u\right) (x):=s^k\mathrm{e}^{\frac{k}{2}\phi (x)}\int \mathrm{e}^{-it\eta }\psi (\eta )\chi _M(t)\mathrm{e}^{-\frac{k}{2}\phi (\Phi ^{\frac{t}{k}}(x))}\widetilde{u}(\Phi ^{\frac{t}{k}}(x))\mathrm{d}t\mathrm{d}\eta \quad \hbox {on } D. \end{aligned}$$
(5.7)

We first notice that for \(k\) large, \(\Phi ^{\frac{t}{k}}(x)\) is well defined for all \(t\in \mathrm{Supp\,}\chi _M\), every \(x\in X\) and \(\Phi ^{\frac{t}{k}}(x)\in D'\) for all \(t\in \mathrm{Supp\,}\chi _M\), every \(x\in D\). We may assume that \(\Phi ^{\frac{t}{k}}(x)\) is well defined for all \(t\in \mathrm{Supp\,}\chi _M\), every \(x\in X\) and \(\Phi ^{\frac{t}{k}}(x)\in D'\) for all \(t\in \mathrm{Supp\,}\chi _M\), every \(x\in D\). Now, we check that the definition above is independent of the choice of local sections. Let \(\hat{s}\) be another local section of \(L\) on \(D'\), \(\left| \hat{s}\right| ^2_{h^L}=\mathrm{e}^{-\hat{\phi }}\). Then, we have \(\hat{s}=gs\) for some non-zero rigid CR function \(g\). We can check that

$$\begin{aligned}&\hat{\phi }=\phi -2\log \left| g\right| ,\nonumber \\&\mathrm{e}^{-\frac{k}{2}\hat{\phi }}=\mathrm{e}^{-\frac{k}{2}\phi }\left| g\right| ^k. \end{aligned}$$
(5.8)

Let \(u\in C^\infty (X,L^k)\). On \(D\), we write \(u=s^k\widetilde{u}=\hat{s}^k\hat{u}\). We have

$$\begin{aligned} \hat{u}=g^{-k}\widetilde{u}. \end{aligned}$$
(5.9)

From (5.8) and (5.9), we can check that

$$\begin{aligned} \mathrm{e}^{-\frac{k}{2}\hat{\phi }}\hat{u}=\mathrm{e}^{-\frac{k}{2}\phi }\left| g\right| ^kg^{-k}\widetilde{u}. \end{aligned}$$
(5.10)

Since \(Tg=0\), we have \((\left| g\right| ^kg^{-k})(\Phi ^{\frac{t}{k}}x)=(\left| g\right| ^kg^{-k})(x)\) for all \(t\in \mathrm{Supp\,}\chi _M\), \(x\in D\). From this observation and (5.10), it is easy to see that

$$\begin{aligned}&\int \mathrm{e}^{-it\eta }\psi (\eta )\chi _M(t)\mathrm{e}^{-\frac{k}{2}\hat{\phi }(\Phi ^{\frac{t}{k}}(x))}\hat{u}(\Phi ^{\frac{t}{k}}(x))\mathrm{d}t\mathrm{d}\eta \nonumber \\&\quad =\left( \left| g\right| ^kg^{-k}\right) (x)\int \mathrm{e}^{-it\eta }\psi (\eta )\chi _M(t)\mathrm{e}^{-\frac{k}{2}\phi (\Phi ^{\frac{t}{k}}(x))}\widetilde{u}(\Phi ^{\frac{t}{k}}(x))\mathrm{d}t\mathrm{d}\eta \quad \hbox {on } D. \end{aligned}$$
(5.11)

Furthermore, we can check that

$$\begin{aligned} \hat{s}^k\mathrm{e}^{\frac{k}{2}\hat{\phi }}=\left| g\right| ^{-k}g^k s^k\mathrm{e}^{\frac{k}{2}\phi }. \end{aligned}$$

Combining this with (5.11), we obtain

$$\begin{aligned}&\hat{s}^k\mathrm{e}^{\frac{k}{2}\hat{\phi }(x)}\int \mathrm{e}^{-it\eta }\psi (\eta )\chi _M(t)\mathrm{e}^{-\frac{k}{2} {\hat{\phi }}(\Phi ^{\frac{t}{k}}(x))}\hat{u}(\Phi ^{\frac{t}{k}}(x))\mathrm{d}t\mathrm{d}\eta \\&\quad =s^k\mathrm{e}^{\frac{k}{2}\phi (x)}\int \mathrm{e}^{-it\eta }\psi (\eta )\chi _M(t)\mathrm{e}^{-\frac{k}{2}\phi (\Phi ^{\frac{t}{k}}(x))}\widetilde{u}(\Phi ^{\frac{t}{k}}(x))\mathrm{d}t\mathrm{d}\eta \quad \hbox {on } D. \end{aligned}$$

Thus, the definition of \(Q^{(0)}_{M,k}\) is well defined.

We consider \((0,1)\) forms. The operator \(Q^{(1)}_{M,k}\) is a continuous operator

$$\begin{aligned} Q^{(1)}_{M,k}:\Omega ^{0,1}(X,L^k)\rightarrow \Omega ^{0,1}(X,L^k) \end{aligned}$$

defined as follows. Let \(D\) be an open set of \(X\). We assume that there exist canonical coordinates \(x\) defined in some neighborhood \(W\) of \(\overline{D}\) and \(L\) is trivial on \(W\). Let \(\psi (\eta )\) and \(\chi _M\) be as in (5.7). For \(k\) large, we have

$$\begin{aligned} \left\{ \Phi ^{\frac{t}{k}}(x)\in W;\, \forall x\in D,t\in \mathrm{Supp\,}\chi _M\right\} . \end{aligned}$$

Let \(s\) be a local section of \(L\) on \(W\), \(\left| s\right| ^2_{h^L}=\mathrm{e}^{-\phi }\). Let \(x=(x_1,\ldots ,x_{2n-1})=(z,\theta )\) be canonical coordinates on \(W\). Then,

$$\begin{aligned}&T=\frac{\partial }{\partial \theta },\nonumber \\&Z_j=\frac{\partial }{\partial z_j}+i\frac{\partial \varphi }{\partial z_j}(z)\frac{\partial }{\partial \theta },\quad j=1,\ldots ,n-1, \end{aligned}$$
(5.12)

where \(Z_j(x)\), \(j=1,\ldots ,n-1\), form a basis of \(T^{1,0}_xX\), for each \(x\in D\), and \(\varphi (z)\in C^\infty (D,\mathbb {R})\) independent of \(\theta \). We can check that \(\mathrm{d}\overline{z}_j\), \(j=1,\ldots ,n-1\), is the basis of \(T^{*0,1}X\), dual to \(\overline{Z}_j\), \(j=1,\ldots ,n-1\). Let \(u\in \Omega ^{0,1}(X,L^k)\). On \(W\), we write

$$\begin{aligned} u=s^k\sum ^{n-1}_{j=1}\widetilde{u}_j(x)\mathrm{d}\overline{z}_j,\quad \widetilde{u}_j\in C^\infty (D),\ j=1,\ldots ,n-1. \end{aligned}$$

Then,

$$\begin{aligned} (Q^{(1)}_{M,k}u)(x)\!:=\!s^ke^{\frac{k}{2}\phi (x)}\sum ^{n-1}_{j=1}\!\left( \int \mathrm{e}^{-it\eta }\psi (\eta )\!\chi _M(t)\mathrm{e}^{-\frac{k}{2}\!\phi (\Phi ^{\frac{t}{k}}(x))}\widetilde{u}_j(\Phi ^{\frac{t}{k}}(x))\mathrm{d}t\mathrm{d}\eta \right) \mathrm{d}\overline{z}_j\,\hbox {on } D.\nonumber \\ \end{aligned}$$
(5.13)

As before, we can show that the definition (5.13) is independent of the choices of local sections. Now, we check that the definition (5.13) is independent of the choice of canonical coordinates. Let \(y=(y_1,\ldots ,y_{2n-1})=(w,\gamma )\), \(w_j=y_{2j-1}+iy_{2j}\), \(j=1,\ldots ,n-1\), \(\gamma =y_{2n-1}\), be another canonical coordinates on \(W\). Then,

$$\begin{aligned}&T=\frac{\partial }{\partial \gamma },\nonumber \\&\widetilde{Z}_j=\frac{\partial }{\partial w_j}+i\frac{\partial \widetilde{\varphi }}{\partial w_j}(w)\frac{\partial }{\partial \gamma },\quad j=1,\ldots ,n-1, \end{aligned}$$
(5.14)

where \(\widetilde{Z}_j(y)\), \(j=1,\ldots ,n-1\), form a basis of \(T^{1,0}_yX\), for each \(y\in D\), and \(\widetilde{\varphi }(w)\in C^\infty (D,\mathbb {R})\) independent of \(\gamma \). From (5.14) and (5.12), it is not difficult to see that on \(W\), we have

$$\begin{aligned}&w=(w_1,\ldots ,w_{n-1})=(H_1(z),\ldots ,H_{n-1}(z))=H(z),\quad H_j(z)\in C^\infty ,\ \forall j,\nonumber \\&\gamma =\theta +G(z),\quad G(z)\in C^\infty , \end{aligned}$$
(5.15)

where for each \(j=1,\ldots ,n-1\), \(H_j(z)\) is holomorphic. From (5.15), we can check that

$$\begin{aligned} \mathrm{d}\overline{w}_j=\sum ^{n-1}_{l=1}\overline{\left( \frac{\partial H_j}{\partial z_l}\right) }\mathrm{d}\overline{z}_l,\quad j=1,\ldots ,n-1. \end{aligned}$$
(5.16)

From this observation, we have for \(u\in \Omega ^{0,1}(X,L^k)\),

$$\begin{aligned}&u=s^k\sum ^{n-1}_{j=1}\widetilde{u}_j(x)\mathrm{d}\overline{z}_j=s^k\sum ^{n-1}_{j=1}\hat{u}_j(y)\mathrm{d}\overline{w}_j\quad \hbox {on } W,\nonumber \\&\widetilde{u}_l(x)=\sum ^{n-1}_{j=1}\hat{u}_j(H(z),\theta +G(z))\overline{\frac{\partial H_j}{\partial z_l}(z)},\quad l=1,\ldots ,n-1. \end{aligned}$$
(5.17)

On \(D\), we have \(\Phi ^{\frac{t}{k}}(x)=(z,\frac{t}{k}+\theta )\), \(\Phi ^{\frac{t}{k}}(y)=(w,\frac{t}{k}+\gamma )\) and \(\frac{\partial H_j}{\partial z_l}(\Phi ^{\frac{t}{k}}(z))=\frac{\partial H_j}{\partial z_l}(z)\), \(j,l=1,\ldots ,n-1\), \(t\in \mathrm{Supp\,}\chi _M\). From this observation and (5.17), (5.16), it is straightforward to see that

$$\begin{aligned}&s^k\mathrm{e}^{\frac{k}{2}\phi (x)}\sum ^{n-1}_{l=1}\left( \int \mathrm{e}^{-it\eta }\psi (\eta )\chi _M(t)\mathrm{e}^{-\frac{k}{2}\phi (\Phi ^{\frac{t}{k}}(x))}\widetilde{u}_l(\Phi ^{\frac{t}{k}}(x))\mathrm{d}t\mathrm{d}\eta \right) \mathrm{d}\overline{z}_l\\&\quad =s^k\mathrm{e}^{\frac{k}{2}\phi (y)}\sum ^{n-1}_{j=1}\left( \int \mathrm{e}^{-it\eta }\psi (\eta )\chi _M(t)\mathrm{e}^{-\frac{k}{2}\phi (\Phi ^{\frac{t}{k}}(y))}\hat{u}_j(\Phi ^{\frac{t}{k}}(y))\mathrm{d}t\mathrm{d}\eta \right) \mathrm{d}\overline{w}_j. \end{aligned}$$

Thus, the definition (5.13) is independent of the choice of canonical coordinates. The operator \(Q^{(1)}_{M,k}\) is well defined.

Now, we claim that

$$\begin{aligned} Q^{(1)}_{M,k}\overline{\partial }_{b,k}u=\overline{\partial }_{b,k}Q^{(0)}_{M,k}u,\quad \forall u\in C^\infty (X,L^k). \end{aligned}$$
(5.18)

We work with canonical coordinates \(x=(z,\theta )\) as (5.12). For \(u\in C^\infty (X,L^k)\), we can check that

$$\begin{aligned} \overline{\partial }_{b,k}u=s^k\sum ^n_{j=1}(\overline{Z}_j\widetilde{u})\mathrm{d}\overline{z}_j=s^k\sum ^{n-1}_{j=1}\left( \frac{\partial \widetilde{u}}{\partial \overline{z}_j}-i\frac{\partial \varphi }{\partial \overline{z}_j}(z)\frac{\partial \widetilde{u}}{\partial \theta }\right) \mathrm{d}\overline{z}_j \end{aligned}$$
(5.19)

on \(W\), where \(u=s^k\widetilde{u}\) on \(W\). Combining (5.19) with (5.13), (5.7) and notice that \(\frac{\partial \widetilde{u}}{\partial \theta }(\Phi ^t(x))=\frac{\partial }{\partial \theta }\bigr (\widetilde{u}(\Phi ^t(x))\bigr )\), it is easy to see that

$$\begin{aligned}&Q^{(1)}_{M,k}\overline{\partial }_{b,k}u-\overline{\partial }_{b,k}Q^{(0)}_{M,k}u\nonumber \\&\quad =-\frac{k}{2}s^k\mathrm{e}^{\frac{k}{2}\phi (x)}\nonumber \\&\qquad \times \sum ^{n-1}_{j=1}\left( \int \mathrm{e}^{-it\eta }\psi (\eta )\chi _M(t)\mathrm{e}^{-\frac{k}{2}\phi (\Phi ^{\frac{t}{k}}(x))}\overline{Z}_j\bigr (\phi (x)-\phi (\Phi ^{\frac{t}{k}}(x))\bigr )\widetilde{u}(\Phi ^{\frac{t}{k}}(x))\mathrm{d}t\mathrm{d}\eta \right) \mathrm{d}\overline{z}_j.\nonumber \\ \end{aligned}$$
(5.20)

Since \(\overline{\partial }_bT\phi =0\), we have \(\overline{Z}_j\phi (x)=\overline{Z}_j\phi (\Phi ^{\frac{t}{k}}(x))\), \(j=1,\ldots ,n-1\), \(t\in \mathrm{Supp\,}\chi _M\). From this and (5.20), (5.18) follows.

6 The asymptotic behavior of \((Q^{(0)}_{M,k}\varPi ^{(0)}_{k,\le k\nu _k})(x)\)

We will use the same notations as before. We recall that we work with the assumption that \(Y(0)\) and \(Y(1)\) hold at each point of \(X\). We first need

Theorem 6.1

For any sequence \(\nu _k>0\) with \(\nu _k\rightarrow 0\) as \(k\rightarrow \infty ,\) there is a constant \(C>0\) independent of \(k,\) such that

$$\begin{aligned} \left| \left( Q^{(0)}_{M,k}\varPi ^{(0)}_{k,\,\leqslant k\nu _k}\overline{Q^{(0)}_{M,k}}\right) (x)\right| \leqslant Ck^n \end{aligned}$$
(6.1)

and

$$\begin{aligned} \left| \left( Q^{(0)}_{M,k}\varPi ^{(0)}_{k,\,\leqslant k\nu _k}\right) (x)\right| \leqslant Ck^n, \end{aligned}$$
(6.2)

for all \(x\in X,\) \(k>0\). Recall that \((Q^{(0)}_{M,k}\varPi ^{(0)}_{k,\le k\nu _k})(x)\) and \((Q^{(0)}_{M,k}\varPi ^{(0)}_{k,\le k\nu _k}\overline{Q^{(0)}_{M,k}})(x)\) are given by (2.13) and (2.14)\(,\) respectively.

Proof

Let \(\nu _k>0\) be any sequence with \(\nu _k\rightarrow 0\) as \(k\rightarrow \infty \). Let \(f_j\in C^\infty (X,L^k)\), \(j=1,\ldots ,d_k\), be an orthonormal frame for \(\fancyscript{H}^0_{b,\le k\nu _k}(X,L^k)\). From (2.24), we see that for each \(x\in X\),

$$\begin{aligned} (Q^{(0)}_{M,k}\varPi ^{(0)}_{k,\le k\nu _k}\overline{Q^{(0)}_{M,k}})(x)&= \sum ^{d_k}_{j=1}\left| (Q^{(0)}_{M,k}f_j)(x)\right| ^2_{h^{L^k}}\nonumber \\&= \sup _{\alpha \in \fancyscript{H}^0_{b,\le k\nu _k}(X,L^k),\left\| \alpha \right\| _{h^{L^k}}=1}\left| (Q^{(0)}_{M,k}\alpha )(x)\right| ^2_{h^{L^k}}. \end{aligned}$$
(6.3)

In view of (2.18), we see that there is a constant \(C>0\) independent of \(k\) such that

$$\begin{aligned} \varPi ^{(0)}_{k,\le k\nu _k}(x)=\sum ^{d_k}_{j=1}\left| f_j(x)\right| ^2_{h^{L^k}}\le Ck^n,\quad \forall x\in X. \end{aligned}$$
(6.4)

For \(\alpha \in \fancyscript{H}^0_{b,\le k\nu _k}(X,L^k)\), \(\left\| \alpha \right\| _{h^{L^k}}=1\), we have

$$\begin{aligned} \left| \alpha (x)\right| ^2_{h^{L^k}}\le \varPi ^{(0)}_{k,\le k\nu _k}(x)\le Ck^n,\quad \forall x\in X, \end{aligned}$$
(6.5)

where \(C>0\) is a constant independent of \(k\) and \(\alpha \). From (6.5) and (5.7), it is easy to see that there is a constant \(C_1>0\) independent of \(k\) such that

$$\begin{aligned} \left| (Q^{(0)}_{M,k}\alpha )(x)\right| ^2_{h^{L^k}}\le C_1k^n,\quad \forall x\in X,\ \forall \alpha \in \fancyscript{H}^0_{b,\le k\nu _k}(X,L^k), \left\| \alpha \right\| _{h^{L^k}}=1. \end{aligned}$$
(6.6)

From (6.6) and (6.3), (6.1) follows.

We have

$$\begin{aligned}&\left| (Q^{(0)}_{M,k}\varPi ^{(0)}_{k,\le k\nu _k})(x)\right| =\left| \sum ^{d_k}_{j=1}\langle (Q^{(0)}_{M,k}f_j)(x)|f_j(x)\rangle _{h^{L^k}}\right| \nonumber \\&\quad \le \Biggr (\sum ^{d_k}_{j=1}\left| (Q^{(0)}_{M,k}f_j)(x)\right| ^2_{h^{L^k}}\Biggr )^{\frac{1}{2}} \Biggr (\sum ^{d_k}_{j=1}\left| f_j(x)\right| ^2_{h^{L^k}}\Biggr )^{\frac{1}{2}}. \end{aligned}$$
(6.7)

From (6.7), (6.4), (6.3) and (6.1), (6.2) follows. \(\square \)

Fix a point \(p\in X\). Let \(x=(x_1,\ldots ,x_{2n-1})=(z,\theta )\) be canonical coordinates of \(X\) defined in some small neighborhood \(D\) of \(p\) and let \(s\) be a local section of \(L\) on \(D\), \(\left| s\right| ^2_{h^L}=\mathrm{e}^{-\phi }\). We take \(x\) and \(s\) so that (4.1) holds. Until further notice, we work with the local coordinates \(x\) and the local section \(s\) and we will use the same notations as Sect. 4. We identify \(D\) with some open set in \(\mathbb {C}^{n-1}\times \mathbb {R}\). Put

$$\begin{aligned} u(z,\theta )&= (2\pi )^{-\frac{n}{2}}\Biggr (\int _{\mathbb {R}_{p,0}}\det (M^\phi _p+2s{\mathcal L}_p)\mathrm{d}s\Biggr )^{-\frac{1}{2}}\nonumber \\&\quad \times \int \mathrm{e}^{i\theta \xi +\frac{\beta \theta }{2}+(-\xi +\frac{i}{2}\beta )\sum ^{n-1}_{j=1}\lambda _j\left| z_j\right| ^2}\det (M^\phi _p+2\xi {\mathcal L}_p){1\!\!1}_{\mathbb {R}_{p,0}}(\xi )\mathrm{d}\xi . \end{aligned}$$
(6.8)

\(u(z,\theta )\in C^\infty (\mathbb {C}^{n-1}\times \mathbb {R})\). We remind that \(\mathbb {R}_{p,0}\) is given by (2.15). Set

$$\begin{aligned} \alpha _k=k^{\frac{n}{2}}s^k\chi _1\left( \frac{\sqrt{k}}{\log k}z,\frac{\sqrt{k}}{\log k}\theta \right) u(\sqrt{k}z,k\theta )\in C^\infty _0(D,L^k), \end{aligned}$$
(6.9)

where \(\chi _1\in C^\infty \), \(0\le \chi _1\le 1\),

$$\begin{aligned} \mathrm{Supp\,}\chi _1\subset \left\{ (z,\theta )\in \mathbb {C}^{n-1}\times \mathbb {R};\, \left| z\right| \le 1,\left| \theta \right| \le 1\right\} , \end{aligned}$$

\(\chi _1(z,\theta )=1\) if \(\left| z\right| \le \frac{1}{2}\), \(\left| \theta \right| \le \frac{1}{2}\). We notice that

$$\begin{aligned} \mathrm{Supp\,}\alpha _k\subset \left\{ (z,\theta )\in \mathbb {C}^{n-1}\times \mathbb {R};\, \left| z\right| \le \frac{\log k}{\sqrt{k}}, \left| \theta \right| \le \frac{\log k}{\sqrt{k}}\right\} . \end{aligned}$$

Thus, for \(k\) large, \(\mathrm{Supp\,}\alpha _k\in D\) and \(\alpha _k\) is well defined. The following is well known (see [9, section 5]).

Proposition 6.2

With the notations used above\(,\) we have

$$\begin{aligned}&\lim _{k\rightarrow \infty }k^{-n}\left| \alpha _k(0)\right| ^2_{h^{L^k}}=(2\pi )^{-n}\int _{\mathbb {R}_{p,0}}\det (M^\phi _p+2s{\mathcal L}_p)\mathrm{d}s,\end{aligned}$$
(6.10)
$$\begin{aligned}&\lim _{k\rightarrow \infty }\left\| \alpha _k\right\| _{h^{L^k}}=1,\end{aligned}$$
(6.11)
$$\begin{aligned}&\lim _{k\rightarrow \infty }\left\| \left( \frac{1}{k}\Box ^{(0)}_{b,k}\alpha _k\right) ^m\right\| _{h^{L^k}}=0,\quad \forall m\in {\mathbb {N}}, \end{aligned}$$
(6.12)

and there is a sequence \(\gamma _k>0,\) independent of the point \(p\) and tending to zero as \(k\rightarrow \infty ,\) such that

$$\begin{aligned} \left( \frac{1}{k}\Box ^{(0)}_{b,k}\alpha _k\ |\ \alpha _k\right) _{h^{L^k}}\le \gamma _k,\quad \forall k>0. \end{aligned}$$
(6.13)

We have the following

Proposition 6.3

Let \(\nu _k>0\) be any sequence with \(\lim _{k\rightarrow \infty }\frac{\gamma _k}{\nu _k}=0\) and \(\nu _k\rightarrow 0\) as \(k\rightarrow \infty ,\) where \(\gamma _k\) is as in (6.13). Let \(\alpha _k\) be as in (6.9). Let

$$\begin{aligned}&\alpha _k=\alpha ^1_k+\alpha ^2_k,\nonumber \\&\alpha ^1_k\in \fancyscript{H}^0_{b,\le k\nu _k}(X,L^k),\quad \alpha ^2_k\in \fancyscript{H}^0_{b,>k\nu _k}(X,L^k). \end{aligned}$$
(6.14)

Then\(,\)

$$\begin{aligned} \lim _{k\rightarrow \infty }\left\| \alpha ^1_k\right\| _{h^{L^k}}=1 \end{aligned}$$
(6.15)

and

$$\begin{aligned} \lim _{k\rightarrow \infty }k^{-n}\left| \alpha ^1_k(0)\right| ^2_{h^{L^k}}=(2\pi )^{-n}\int _{\mathbb {R}_{p,0}}\det (M^\phi _p+2s{\mathcal L}_p)\mathrm{d}s. \end{aligned}$$
(6.16)

Moreover\(,\) on \(D,\) we put

$$\begin{aligned} \alpha ^2_k=k^{\frac{n}{2}}s^k\beta ^2_k,\quad \beta ^2_k\in C^\infty (D). \end{aligned}$$
(6.17)

Fix \(r>0\). Then\(,\) for every \(\varepsilon >0,\) there is a \(k_0>0\) such that for all \(k\ge k_0,\) we have \(F_k(D_{2r})\subset D\) and

$$\begin{aligned} \left| \beta ^2_k\left( \frac{z}{\sqrt{k}},\frac{\theta }{k}\right) \right| \le \varepsilon ,\quad \forall (z,\theta )\in D_r. \end{aligned}$$
(6.18)

In particular\(,\)

$$\begin{aligned} \lim _{k\rightarrow \infty }\left| \beta ^2_k\left( \frac{z}{\sqrt{k}},\frac{\theta }{k}\right) \right| =0,\quad \forall (z,\theta )\in D. \end{aligned}$$
(6.19)

Proof

From (2.9), we have

$$\begin{aligned} \left\| \alpha ^2_k\right\| ^2_{h^{L^k}}\leqslant \frac{1}{k\nu _k}\big (\Box ^{(0)}_{b,k}\alpha ^2_k\ \big |\ \alpha ^2_k\big )_{h^{L^k}}\leqslant \frac{1}{k\nu _k}\big (\Box ^{(0)}_{b,k}\alpha _k\ \big |\ \alpha _k\big )_{h^{L^k}}\leqslant \frac{\gamma _k}{\nu _k}\rightarrow 0, \end{aligned}$$

as \(k\rightarrow \infty \). Thus, \(\lim _{k\rightarrow \infty }\left\| \alpha ^2_k\right\| _{h^{L^k}}=0\). Since \(\left\| \alpha _k\right\| _{h^{L^k}}\rightarrow 1\) as \(k\rightarrow \infty \), (6.15) follows.

Now, we prove (6.18). As (6.17), on \(D\), we write \(\alpha ^2_k=s^kk^{\frac{n}{2}}\beta ^2_k\), \(\beta ^2_k\in C^\infty (D)\). From (4.22), we know that

$$\begin{aligned}&\sup _{(z,\theta )\in D_r}\left| F^*_k\beta ^2_k(z,\theta )\right| ^2=\sup _{(z,\theta )\in D_r}\left| \beta ^2_k\left( \frac{z}{\sqrt{k}},\frac{\theta }{k}\right) \right| ^2\nonumber \\&\quad \leqslant C_{r}\Biggr (\left\| F^*_k\beta ^2_k\right\| ^2_{kF^*_k\phi ,D_{2r}}+\sum ^n_{m=1}\left\| (\Box ^{(q)}_{k,k\phi ,(k)})^mF^*_k\beta ^2_k\right\| ^2_{kF^*_k\phi ,D_{2r}}\Biggr ), \end{aligned}$$
(6.20)

where \(C_r>0\) is independent of \(k\). Now, we have

$$\begin{aligned} \left\| F^*_k\beta ^2_k\right\| ^2_{kF^*_k\phi ,D_{2r}}\leqslant \left\| \alpha ^2_k\right\| ^2_{h^{L^k}}\rightarrow 0,\quad \hbox {as } k\rightarrow \infty . \end{aligned}$$
(6.21)

Moreover, from (4.17), it is easy to can check that for all \(m\in {\mathbb {N}}\),

$$\begin{aligned}&\Big \Vert \Bigr (\Box ^{(q)}_{b,k\phi ,(k)}\Bigr )^mF^*_k\beta ^2_k\Big \Vert ^2_{kF^*_k\phi ,D_{2r}}\leqslant \Big \Vert \Bigr (\tfrac{1}{k}\Box ^{(q)}_{b,k}\Bigr )^m\alpha ^2_k\Big \Vert ^2_{h^{L^k}}\nonumber \\&\quad \leqslant \Big \Vert \Bigr (\tfrac{1}{k}\Box ^{(q)}_{b,k}\Bigr )^m\alpha _k\Big \Vert ^2_{h^{L^k}}\rightarrow 0 \quad \text {as } k\rightarrow \infty . \end{aligned}$$
(6.22)

Here, we used (6.12). Combining (6.20) with (6.21) and (6.22), (6.18) follows.

From (6.18), we deduce

$$\begin{aligned} \lim _{k\rightarrow \infty }\left| F^*_k\beta ^2_k(0)\right| ^2=\lim _{k\rightarrow \infty }\left| \beta ^2_k(0)\right| ^2= \lim _{k\rightarrow \infty }k^{-n}\left| \alpha ^2_k(0)\right| ^2_{h^{L^k}}=0. \end{aligned}$$
(6.23)

From this and (6.10), (6.16) follows. \(\square \)

Now, we can prove

Theorem 6.4

Let \(\delta _k=\min \left\{ \mu _k,\gamma _k\right\} ,\) where \(\mu _k\) is as in Theorem 2.1 and \(\gamma _k\) is as in (6.13). Let \(\nu _k>0\) be any sequence with \(\lim _{k\rightarrow \infty }\frac{\delta _k}{\nu _k}=0\) and \(\nu _k\rightarrow 0\) as \(k\rightarrow \infty \). Then\(,\)

$$\begin{aligned}&\lim _{k\rightarrow \infty }k^{-n}(Q^{(0)}_{M,k}\varPi ^{(0)}_{k,\,\leqslant k\nu _k})(x)\nonumber \\&\quad =(2\pi )^{-n}\int \mathrm{e}^{it\xi }\hat{\psi }(t)\chi _M(t)\det (M^\phi _x+2\xi {\mathcal L}_x){1\!\!1}_{\mathbb {R}_{x,0}}(\xi )\mathrm{d}td\xi \end{aligned}$$
(6.24)

for all \(x\in X,\) where \(\psi (\eta )\) is as in the discussion after (5.1) and \(\chi _M(t)\) is given by (5.6)\(,\) \(\hat{\psi }(t)=\int \mathrm{e}^{-it\eta }\psi (\eta )\mathrm{d}\eta .\) We remind that \(\mathrm{Supp\,}\psi \bigcap \mathbb {R}_{x,1}=\emptyset ,\) for every \(x\in X.\)

Proof

Let \(\nu _k>0\) be any sequence with \(\lim _{k\rightarrow \infty }\frac{\delta _k}{\nu _k}=0\) and \(\nu _k\rightarrow 0\) as \(k\rightarrow \infty \). Fix a point \(p\in X\). Let \(x=(x_1,\ldots ,x_{2n-1})=(z,\theta )\) be canonical coordinates of \(X\) defined in some small neighborhood \(D\) of \(p\) and let \(s\) be a local section of \(L\) on \(D\), \(\left| s\right| ^2_{h^L}=\mathrm{e}^{-\phi }\). As before we take \(x\) and \(s\) so that (4.1) hold and let \(\alpha ^1_k\in \fancyscript{H}^0_{b,\le k\nu _k}(X,L^k)\) be as in (6.14). We take

$$\begin{aligned} f^1_k:=\frac{\alpha ^1_k}{\left\| \alpha ^1_k\right\| _{h^{L^k}}},f^2_k,\ldots ,f^{d_k}_k \end{aligned}$$

to be an orthonormal frame for \(\fancyscript{H}^0_{b,\le k\nu _k}(X,L^k)\). From (6.15), (6.16) and (2.19), we conclude that

$$\begin{aligned} \lim _{k\rightarrow \infty }k^{-n}\left| f^1_k(0)\right| ^2_{h^{L^k}}=\lim _{k\rightarrow \infty }k^{-n}\varPi ^{(0)}_{k,\le k\nu _k}(0) =(2\pi )^{-n}\int _{\mathbb {R}_{p,0}}\det (M^\phi _p+2s{\mathcal L}_p)\mathrm{d}s.\nonumber \\ \end{aligned}$$
(6.25)

Thus,

$$\begin{aligned} \lim _{k\rightarrow \infty }k^{-n}\sum ^{d_k}_{j=2}\left| f^j_k(0)\right| ^2_{h^{L^k}}=0. \end{aligned}$$
(6.26)

Now,

$$\begin{aligned} (Q^{(0)}_{M,k}\varPi ^{(0)}_{k,\le k\nu _k})(0)\!=\!\langle (Q^{(0)}_{M,k}f^1_k)(0)|f^1_k(0)\rangle _{h^{L^k}}\!+\!\sum ^{d_k}_{j=2} \langle (Q^{(0)}_{M,k}f^j_k)(0)|f^j_k(0)\rangle _{h^{L^k}}.\qquad \end{aligned}$$
(6.27)

From (6.1) and (6.26), we have

$$\begin{aligned}&\lim _{k\rightarrow \infty }k^{-n}\left| \sum ^{d_k}_{j=2}\langle (Q^{(0)}_{M,k}f^j_k)(0)|f^j_k(0)\rangle _{h^{L^k}}\right| \nonumber \\&\quad \le \lim _{k\rightarrow \infty }k^{-n}\sqrt{\sum ^{d_k}_{j=2} \left| (Q^{(0)}_{M,k}f^j_k)(0)\right| ^2_{h^{L^k}}}\sqrt{\sum ^{d_k}_{j=2} \left| f^j_k(0)\right| ^2_{h^{L^k}}}\rightarrow 0,\quad \hbox {as }k\rightarrow \infty .\qquad \end{aligned}$$
(6.28)

Combining (6.28) with (6.27), we conclude that

$$\begin{aligned} \lim _{k\rightarrow \infty }k^{-n}(Q^{(0)}_{M,k}\varPi ^{(0)}_{k,\le \nu _k})(0)= \lim _{k\rightarrow \infty }k^{-n}\langle (Q^{(0)}_{M,k}f^1_k)(0)|f^1_k(0)\rangle _{h^{L^k}}. \end{aligned}$$
(6.29)

Let \(\alpha ^2_k\) be as in (6.14). From (6.18) and the definition of \(Q^{(0)}_{M,k}\) [see (5.7)], it is not difficult to see that

$$\begin{aligned} \lim _{k\rightarrow \infty }k^{-n}\langle (Q^{(0)}_{M,k}\alpha ^2_k)(0)| \alpha ^2_k(0)\rangle _{h^{L^k}}=0. \end{aligned}$$
(6.30)

Combining (6.30) with (6.29) and (6.15), we deduce

$$\begin{aligned} \lim _{k\rightarrow \infty }k^{-n}(Q^{(0)}_{M,k}\varPi ^{(0)}_{k,\le \nu _k})(0)=\lim _{k\rightarrow \infty } k^{-n}\langle (Q^{(0)}_{M,k}\alpha _k)(0)|\alpha _k(0)\rangle _{h^{L^k}}, \end{aligned}$$
(6.31)

where \(\alpha _k\) is as in (6.9). On \(D\), we put

$$\begin{aligned} Q^{(0)}_{M,k}\alpha _k=s^kq_k,\quad q_k\in C^\infty (D). \end{aligned}$$
(6.32)

By the definitions of \(Q^{(0)}_{M,k}\) and \(\alpha _k\) [see (5.7) and (6.9)], we can check that

$$\begin{aligned}&q_k(0) =k^{\frac{n}{2}}(2\pi )^{-\frac{n}{2}} \Biggr (\int _{\mathbb {R}_{p,0}}\det (M^\phi _p+2s{\mathcal L}_p) \mathrm{d}s\Biggr )^{-\frac{1}{2}}\nonumber \\&\quad \times \int \mathrm{e}^{-it\eta }\psi (\eta )\chi _M(t) \mathrm{e}^{-\frac{k}{2}\phi \left( 0,\frac{t}{k}\right) }\chi _1\left( 0,\frac{t}{\sqrt{k}\log k}\right) \mathrm{e}^{it\xi +\frac{\beta }{2}t}{1\!\!1}_{\mathbb {R}_{p,0}}(\xi )\det (M^\phi _p+2\xi {\mathcal L}_p)\mathrm{d}\xi \mathrm{d}t\mathrm{d}\eta .\nonumber \\ \end{aligned}$$
(6.33)

We notice that \(\frac{k}{2}\phi (0,\frac{t}{k})=\frac{\beta }{2}t+\epsilon _k(t)\), where \(\epsilon _k(t)\rightarrow 0\) as \(k\rightarrow \infty \), uniformly on \(\mathrm{Supp\,}\chi _M\) and \(\chi _1(0,\frac{t}{\sqrt{k}\log k})\rightarrow 1\) as \(k\rightarrow \infty \), uniformly on \(\mathrm{Supp\,}\chi _M\). Combining this observation with (6.33), (6.9) and (6.8), we can check that

$$\begin{aligned}&\lim _{k\rightarrow \infty }k^{-n}\langle (Q^{(0)}_{M,k}\alpha _k)(0)|\alpha _k(0)\rangle _{h^{L^k}}\nonumber \\&\quad =\lim _{k\rightarrow \infty }k^{-\frac{n}{2}}q_k(0)\overline{u(0,0)}\mathrm{e}^{-k\phi (0)}\nonumber \\&\quad =(2\pi )^{-n}\int \mathrm{e}^{-it\eta +it\xi }\psi (\eta )\chi _M(t){1\!\!1}_{\mathbb {R}_{p,0}}(\xi )\det (M^\phi _p+2\xi {\mathcal L}_p)\mathrm{d}\xi \mathrm{d}t\mathrm{d}\eta \nonumber \\&\quad =(2\pi )^{-n}\int \mathrm{e}^{it\xi }\hat{\psi }(t)\chi _M(t){1\!\!1}_{\mathbb {R}_{p,0}}(\xi )\det (M^\phi _p+2\xi {\mathcal L}_p)\mathrm{d}\xi \mathrm{d}t, \end{aligned}$$
(6.34)

where \(\hat{\psi }(t):=\int \mathrm{e}^{-it\eta }\psi (\eta )\mathrm{d}\eta \), \(u\) is as in (6.8). From (6.34) and (6.31), (6.24) follows. We get Theorem 6.4. \(\square \)

We need

Theorem 6.5

Let \(\delta _k>0,\) \(\delta _k\rightarrow 0,\) as \(k\rightarrow \infty ,\) be as in Theorem 6.4 and let \(\nu _k>0\) be any sequence with \(\lim _{k\rightarrow \infty }\frac{\delta _k}{\nu _k}=0\) and \(\nu _k\rightarrow 0\) as \(k\rightarrow \infty .\) Then\(,\) there is a \(k_0>0\) such that for all \(k\ge k_0,\)

$$\begin{aligned}&\left| \int _X(Q^{(0)}_{M,k}\varPi ^{(0)}_{k,\le k\nu _k})(x)\mathrm{d}v_X(x)\right| \nonumber \\&\quad \ge \frac{k^n}{2}(2\pi )^{1-n}\int _X\Biggr (\int \psi (\xi )\det (M^\phi _x+2\xi {\mathcal L}_x) {1\!\!1}_{\mathbb {R}_{x,0}}(\xi )\mathrm{d}\xi \Biggr )\mathrm{d}v_X(x). \end{aligned}$$
(6.35)

Proof

For each \(x\in X\), put

$$\begin{aligned} C(x):=(2\pi )^{-n}\int \mathrm{e}^{it\xi }\hat{\psi }(t)\chi _M(t)\det (M^\phi _x+2\xi {\mathcal L}_x){1\!\!1}_{\mathbb {R}_{x,0}}(\xi )\mathrm{d}\xi \mathrm{d}t. \end{aligned}$$
(6.36)

From (6.2), (6.24) and the Lebesgue dominated Theorem, we conclude that

$$\begin{aligned} \int _X(Q^{(0)}_{M,k}\varPi ^{(0)}_{k,\le k\nu _k})(x)\mathrm{d}v_X(x)=k^n\int _XC(x)\mathrm{d}v_X(x)+o(k^n) \end{aligned}$$

and hence

$$\begin{aligned} \left| \int _X(Q^{(0)}_{M,k}\varPi ^{(0)}_{k,\le k\nu _k})(x)\mathrm{d}v_X(x)\right| \ge k^n\left| \int _XC(x)\mathrm{d}v_X(x)\right| +o(k^n). \end{aligned}$$
(6.37)

We first claim that for each \(x\in X\), \(C(x)\) is real. We notice that \(\overline{\hat{\psi }(t)}=\hat{\psi }(-t)\) and \(\chi _M(t)=\chi _M(-t)\). From this observation, we can check that

$$\begin{aligned} \overline{C(x)}&= (2\pi )^{-n}\int \mathrm{e}^{-it\xi }\overline{\hat{\psi }(t)}\chi _M(t)\det (M^\phi _x+2\xi {\mathcal L}_x) {1\!\!1}_{\mathbb {R}_{x,0}}(\xi )\mathrm{d}\xi \mathrm{d}t\\&= (2\pi )^{-n}\int \mathrm{e}^{-it\xi }\hat{\psi }(-t)\chi _M(-t)\det (M^\phi _x+2\xi {\mathcal L}_x) {1\!\!1}_{\mathbb {R}_{x,0}}(\xi )\mathrm{d}\xi \mathrm{d}t\\&= (2\pi )^{-n}\int \mathrm{e}^{it\xi }\hat{\psi }(t)\chi _M(t)\det (M^\phi _x+2\xi {\mathcal L}_x) {1\!\!1}_{\mathbb {R}_{x,0}}(\xi )\mathrm{d}\xi \mathrm{d}t=C(x). \end{aligned}$$

Thus, \(C(x)\) is real.

Now, we claim that \(\int _XC(x)\mathrm{d}v_X(x)\) is positive and

$$\begin{aligned} \int _XC(x)\mathrm{d}v_X(x)\!>\!\frac{1}{2}(2\pi )^{1-n}\int _X\Bigr (\int \psi (\xi )\det (M^\phi _x+2\xi {\mathcal L}_x) {1\!\!1}_{\mathbb {R}_{x,0}}(\xi )\mathrm{d}\xi \Bigr )\mathrm{d}v_X(x).\qquad \end{aligned}$$
(6.38)

We have

$$\begin{aligned}&C(x)=(2\pi )^{-n}\int \mathrm{e}^{it\xi }\hat{\psi }(t)\det (M^\phi _x+2\xi {\mathcal L}_x) {1\!\!1}_{\mathbb {R}_{x,0}}(\xi )\mathrm{d}\xi \mathrm{d}t\nonumber \\&\qquad \qquad +(2\pi )^{-n}\int \mathrm{e}^{it\xi }\hat{\psi }(t)(\chi _M(t)-1)\det (M^\phi _x+2\xi {\mathcal L}_x) {1\!\!1}_{\mathbb {R}_{x,0}}(\xi )\mathrm{d}\xi \mathrm{d}t\nonumber \\&\quad =(2\pi )^{1-n}\int \psi (\xi )\det (M^\phi _x+2\xi {\mathcal L}_x) {1\!\!1}_{\mathbb {R}_{x,0}}(\xi )\mathrm{d}\xi \nonumber \\&\qquad +(2\pi )^{-n}\int \mathrm{e}^{it\xi }\hat{\psi }(t)(\chi _M(t)-1)\det (M^\phi _x+2\xi {\mathcal L}_x) {1\!\!1}_{\mathbb {R}_{x,0}}(\xi )\mathrm{d}\xi \mathrm{d}t. \end{aligned}$$
(6.39)

Here, we used Fourier’s inversion formula. Since \(0\le \chi _M\le 1\) and \(\chi _M=1\) if \(-M\le t\le M\), we have

$$\begin{aligned}&\left| \int \mathrm{e}^{it\xi }\hat{\psi }(t)(\chi _M(t)-1)\det (M^\phi _x+2\xi {\mathcal L}_x) {1\!\!1}_{\mathbb {R}_{x,0}}(\xi )\mathrm{d}\xi \mathrm{d}t\right| \nonumber \\&\quad =\left| \int _{\left| t\right| \ge M}\mathrm{e}^{it\xi }\hat{\psi }(t)(\chi _M(t)-1)\det (M^\phi _x+2\xi {\mathcal L}_x){1\!\!1}_{\mathbb {R}_{x,0}}(\xi )\mathrm{d}\xi \mathrm{d}t\right| \nonumber \\&\quad \le \int _{\left| t\right| \ge M}\left| \hat{\psi }(t)\right| \mathrm{d}t\left| \int \det (M^\phi _x+2\xi {\mathcal L}_x) {1\!\!1}_{\mathbb {R}_{x,0}}(\xi )\mathrm{d}\xi \right| \nonumber \\&\quad \le \frac{2C_0}{M}\int \det (M^\phi _x+2\xi {\mathcal L}_x) {1\!\!1}_{\mathbb {R}_{x,0}}(\xi )\mathrm{d}\xi , \end{aligned}$$
(6.40)

where \(C_0=\sup _{t\in \mathbb {R}}t^2\left| \hat{\psi }(t)\right| \). Combining (6.40) with (6.39), we get

$$\begin{aligned} C(x)&\ge (2\pi )^{1-n}\int \psi (\xi )\det (M^\phi _x+2\xi {\mathcal L}_x) {1\!\!1}_{\mathbb {R}_{x,0}}(\xi )\mathrm{d}\xi \nonumber \\&-\frac{2C_0}{M}(2\pi )^{-n}\int \det (M^\phi _x+2\xi {\mathcal L}_x) {1\!\!1}_{\mathbb {R}_{x,0}}(\xi )\mathrm{d}\xi . \end{aligned}$$

Combining this with (5.4), (6.38) follows.

From (6.38) and (6.37), we obtain (6.35). \(\square \)

7 The asymptotic behavior of \(\Big (Q^{(1)}_{M,k}\varPi ^{(1)}_{k,\le k\nu _k}\overline{Q^{(1)}_{M,k}}\Big )(x)\)

We will use the same notations as before. Fix \(p\in X\). Let \(x=(x_1,\ldots ,x_{2n-1})=(z,\theta )\) be canonical coordinates of \(X\) defined in some small neighborhood \(D\) of \(p\) and let \(s\) be a local section of \(L\) on \(D\), \(\left| s\right| ^2_{h^L}=\mathrm{e}^{-\phi }\). We take \(x\) and \(s\) so that (4.1) hold. Until further notice, we work with the local coordinates \(x\) and the local section \(s\). We also write \(t\) to denote the coordinate \(\theta \). We identify \(D\) with some open set in \(H_n=\mathbb {C}^{n-1}\times \mathbb {R}\). Let \(\nu _k>0\) be any sequence with \(\nu _k\rightarrow 0\) as \(k\rightarrow \infty \). We are going to estimate \(\limsup _{k\rightarrow \infty }k^{-n}(Q^{(1)}_{M,k}\varPi ^{(1)}_{k,\le k\nu _k}\overline{Q^{(1)}_{M,k}})(p)\). For the convenience of the reader we recall some notations we used before. Let \(e_j(z,\theta )\), \(j=1,\ldots ,n-1\), denote the basis of \(T^{*(0,1)}X\), dual to \(\overline{U}_j(z,\theta )\), \(j=1,\ldots ,n-1\), where \(U_j\), \(j=1,\ldots ,n-1\), are as in (4.3). For \(f\in \Omega ^{0,1}(X,L^k)\), we write \(f=\sum ^{n-1}_{j=1}f_je_j\), \(f_j\in C^\infty (X,L^k)\), \(j=1,\ldots ,n-1\). We call \(f_j\) the component of \(f\) along \(e_j\). As (2.21), for \(j=1,\ldots ,n-1\), we define

$$\begin{aligned} (Q^{(1)}_{M,k}S^{(1)}_{k,\,\leqslant k\nu _k,j}\overline{Q^{(1)}_{M,k}})(y):=\sup _{\alpha \in \,\fancyscript{H}_{b,\le k\nu _k}^1(X, L^k),\,\left\| \alpha \right\| _{h^{L^k}}=1}\left| (Q^{(1)}_{M,k}\alpha )_j(y)\right| ^2_{h^{L^k}}, \end{aligned}$$
(7.1)

where \((Q^{(1)}_{M,k}\alpha )_j\) denotes the component of \(Q^{(1)}_{M,k}\alpha \) along \(e_j\). From (2.23), we know that

$$\begin{aligned} (Q^{(1)}_{M,k}\varPi ^{(1)}_{k,\,\leqslant k\nu _k}\overline{Q^{(1)}_{M,k}})(y)=\sum ^{n-1}_{j=1}(Q^{(1)}_{M,k}S^{(1)}_{k,\,\leqslant k\nu _k,j}\overline{Q^{(1)}_{M,k}})(y),\quad \forall y\in D. \end{aligned}$$
(7.2)

We consider \(H_n\). Let \(\psi (\eta )\) be as in the discussion after (5.1) and let \(\chi _M(t)\) be as in (5.6). The operator \(Q^{(1)}_{M,H_n}\) is a continuous operator \(\Omega ^{0,1}(H_n)\rightarrow \Omega ^{0,1}(H_n)\) defined as follows. Let \(u\in \Omega ^{0,1}(H_n)\). We write \(u=\sum ^{n-1}_{j=1}u_j\mathrm{d}\overline{z}_j\), \(u_j\in C^\infty (H_n)\), \(j=1,\ldots ,n-1\). Then,

$$\begin{aligned} (Q^{(1)}_{M,H_n}u)(z,\theta )&= \sum ^{n-1}_{j=1}\left( \int \mathrm{e}^{-it\eta }\psi (\eta )\chi _M(t)\mathrm{e}^{-\frac{\beta }{2} (t+\theta )}u_j(z,t+\theta )\mathrm{d}t\mathrm{d}\eta \right) \mathrm{d}\overline{z}_j\nonumber \\&:= \int \mathrm{e}^{-it\eta }\psi (\eta )\chi _M(t)\mathrm{e}^{-\frac{\beta }{2}(t+\theta )}u(z,t+\theta )\mathrm{d}t\mathrm{d}\eta . \end{aligned}$$
(7.3)

We remind that \(\beta \) is as in (4.1). For \(j=1,\ldots ,n-1\), put [compare (7.1)]

$$\begin{aligned}&(Q^{(1)}_{M,H_n}S^{(1)}_{j,H_n}\overline{Q^{(1)}_{M,H_n}})(0)\nonumber \\&\quad =\sup \left\{ \left| (Q^{(1)}_{M,H_n}\alpha )_j(0)\right| ^2;\, \alpha \in \Omega ^{0,1}(H_n),\Box ^{(1)}_{b,H_n}\alpha =0,\left\| \alpha \right\| _{\phi _0}=1\right\} , \end{aligned}$$
(7.4)

where

$$\begin{aligned} (Q^{(1)}_{M,H_n}\alpha )(x)=\sum ^{n-1}_{j=1}(Q^{(1)}_{M,H_n}\alpha )_j(x)\mathrm{d}\overline{z}_j,\quad (Q^{(1)}_{M,H_n}\alpha )_j\in C^\infty (H_n),\qquad j=1,\ldots ,n-1, \end{aligned}$$

and

$$\begin{aligned} \left\| \alpha \right\| ^2_{\phi _0}=\int \left| \alpha (z,\theta )\right| ^2\mathrm{e}^{-\phi _0(z,\theta )}\mathrm{d}v(z)\mathrm{d}\theta , \quad \mathrm{d}v(z)=2^{n-1}\mathrm{d}x_1\mathrm{d}x_2\ldots \mathrm{d}x_{2n-1}. \end{aligned}$$

We recall that \(\phi _0\) is as in the discussion after (4.23). We first need

Theorem 7.1

We have

$$\begin{aligned} \limsup _{k\rightarrow \infty }k^{-n}\left( Q^{(1)}_{M,k}\varPi ^{(1)}_{k,\le k\nu _k}\overline{Q^{(1)}_{M,k}}\right) (0)\leqslant \sum ^{n-1}_{j=1} \left( Q^{(1)}_{M,H_n}S^{(1)}_{j,H_n}\overline{Q^{(1)}_{M,H_n}}\right) (0). \end{aligned}$$

Proof

Fix \(j\in \left\{ 1,2,\ldots ,n-1\right\} \). We claim that

$$\begin{aligned} \limsup _{k\rightarrow \infty }k^{-n}\left( Q^{(1)}_{M,k}S^{(1)}_{k,\le k\nu _k,j}\overline{Q^{(1)}_{M,k}}\right) (0)\leqslant \left( Q^{(1)}_{M,H_n}S^{(1)}_{j,H_n}\overline{Q^{(1)}_{M,H_n}}\right) (0). \end{aligned}$$
(7.5)

The definition (7.1) of \((Q^{(1)}_{M,k}S^{(1)}_{k,\le k\nu _k,j}\overline{Q^{(1)}_{M,k}})(0)\) yields a sequence

$$\begin{aligned} \alpha _{k_s}\in \fancyscript{H}^1_{b,\le k_s\nu _{k_s}}(X, L^{k_s}),\quad k_1<k_2<\cdots , \end{aligned}$$

such that \(\left\| \alpha _{k_s}\right\| _{h^{L^{k_s}}}=1\) and

$$\begin{aligned} \lim _{s\rightarrow \infty }k_s^{-n}\left| (Q^{(1)}_{M,k_s}\alpha _{k_s})_j(0)\right| ^2_{h^{L^{k_s}}}=\limsup _{k\rightarrow \infty }k^{-n}(Q^{(1)}_{M,k}S^{(1)}_{k,\le k\nu _k,j}\overline{Q^{(1)}_{M,k}})(0), \end{aligned}$$
(7.6)

where \((Q^{(1)}_{M,k_s}\alpha _{k_s})_j\) is the component of \(Q^{(1)}_{M,k_s}\alpha _{k_s}\) along \(e_j\). On \(D\), we write

$$\begin{aligned} \alpha _{k_s}=s^{k_s}\widetilde{\alpha }_{k_s},\quad \widetilde{\alpha }_{k_s}\in \Omega ^{0,1}(D), \end{aligned}$$

and on \(D_{\log k_s}\), put

$$\begin{aligned} \gamma _{k_s}=k_s^{-\frac{n}{2}}F^*_{k_s}\widetilde{\alpha }_{k_s}\in F^*_{k_s}\Omega ^{0,1}(D_{\log k_s}). \end{aligned}$$

We recall that \(F^*_{k_s}\) is the scaling map given by (4.9). It is not difficult to see that

$$\begin{aligned} \left\| \gamma _{k_s}\right\| _{k_sF^*_{k_s}\phi ,D_{\log k_s}}\leqslant 1. \end{aligned}$$

Moreover, from (4.17) and (4.2), it is straightforward to see that

$$\begin{aligned} \left\| (\Box ^{(1)}_{b,k_s\phi ,(k_s)})^m\gamma _{k_s}\right\| _{k_sF^*_{k_s}\phi ,D_{\log k_s}}\le \frac{1}{k^m_s}\left\| (\Box ^{(1)}_{b,k_s})^m\alpha _{k_s}\right\| _{h^{L^{k_s}}}\le \nu ^m_{k_s}, \quad \forall m\in {\mathbb {N}}. \end{aligned}$$

Proposition 4.3 yields a subsequence \(\left\{ \gamma _{k_{s_u}}\right\} \) of \(\left\{ \gamma _{k_s}\right\} \) such that for each \(t\) in the set \(\left\{ 1,2,\ldots ,n-1\right\} \), \(\gamma _{k_{s_u},t}\) converges uniformly with all derivatives on any compact subset of \(H_n\) to a smooth function \(\gamma _t\), where \(\gamma _{k_{s_u},t}\) denotes the component of \(\gamma _{k_{s_u}}\) along \(e_t(\frac{z}{\sqrt{k}},\frac{\theta }{k})\). Set \(\gamma =\sum ^{n-1}_{t=1}\gamma _t\mathrm{d}\overline{z}_t\). Then, we have \(\Box ^{(1)}_{b,H_n}\gamma =0\) and, by (4.24), \(\left\| \gamma \right\| _{\phi _0}\leqslant 1\). Thus,

$$\begin{aligned} \left| (Q^{(1)}_{M,H_n}\gamma )_j(0)\right| ^2\leqslant \frac{\left| (Q^{(1)}_{M,H_n}\gamma )_j(0)\right| ^2}{\left\| \gamma \right\| ^2_{\phi _0}}\leqslant \left( Q^{(1)}_{M,H_n}S^{(1)}_{j,H_n}\overline{Q^{(1)}_{M,H_n}}\right) (0), \end{aligned}$$
(7.7)

where

$$\begin{aligned} Q^{(1)}_{M,H_n}\gamma =\sum ^{n-1}_{t=1}(Q^{(1)}_{M,H_n}\gamma )_t\mathrm{d}\overline{z}_t,\qquad (Q^{(1)}_{M,H_n}\gamma )_t\in C^\infty (H_n),\quad t=1,\ldots ,n-1. \end{aligned}$$

We claim that

$$\begin{aligned} \lim _{u\rightarrow \infty }k_{s_u}^{-n}\left| \left( Q^{(1)}_{M,k_{s_u}}\alpha _{k_{s_u}}\right) _j(0)\right| ^2 =\left| \left( Q^{(1)}_{M,H_n}\gamma \right) _j(0)\right| ^2. \end{aligned}$$
(7.8)

We write

$$\begin{aligned} \widetilde{\alpha }_{k_{s_u}}=\sum ^{n-1}_{j=1}\widetilde{\alpha }_{k_{s_u},j}e_j=\sum ^{n-1}_{j=1}\hat{\alpha }_{k_{s_u},j}\mathrm{d}\overline{z}_j. \end{aligned}$$

Since \(e_t=\mathrm{d}\overline{z}_t+O(\left| (z,\theta )\right| \), \(t=1,\ldots ,n-1\), we conclude that for all \(t=1,\ldots ,n-1\),

$$\begin{aligned} \lim _{u\rightarrow \infty }k_{s_u}^{-\frac{n}{2}}F^*_{k_{s_u}}\widetilde{\alpha }_{k_{s_u},t}= \lim _{u\rightarrow \infty }k_{s_u}^{-\frac{n}{2}}F^*_{k_{s_u}}\hat{\alpha }_{k_{s_u},t}=\gamma _t. \end{aligned}$$
(7.9)

Moreover, from the definition of \(Q^{(1)}_{M,k_{s_u}}\) [see (5.13)], it is easy to see that

$$\begin{aligned} \left| (Q^{(1)}_{M,k_{s_u}}\alpha _{k_{s_u}})_j(0)\right| _{h^{L^k}}=\left| \int \mathrm{e}^{-it\eta }\psi (\eta )\chi _M(t)\mathrm{e}^{-\frac{k_{s_u}}{2}(F^*_{k_{s_u}}\phi )(0,t)}F^*_{k_{s_u}}\hat{\alpha }_{k_{s_u},j}(0,t)\mathrm{d}t\right| . \end{aligned}$$
(7.10)

Combining (7.10) with (7.9), (7.3) and notice that \(-\frac{k}{2}(F^*_k\phi )(0,t)\rightarrow -\frac{\beta }{2}t\), as \(k\rightarrow \infty \), uniformly on \(\mathrm{Supp\,}\chi _M\), (7.8) follows. The claim (7.5) follows from (7.6), (7.7) and (7.8). Finally, (7.5) and (7.2) imply the conclusion of the theorem. \(\square \)

To estimate \(\sum ^{n-1}_{j=1} (Q^{(1)}_{M,H_n}S^{(1)}_{j,H_n}\overline{Q^{(1)}_{M,H_n}})(0)\), we need the some preparation. Put

$$\begin{aligned} \Phi _0=\sum ^{n-1}_{j,t=1}\mu _{j,\,t}\overline{z}_jz_t, \end{aligned}$$
(7.11)

where \(\mu _{j,t}\), \(j,t=1,\ldots ,n-1\), are as in (4.1). Note that

$$\begin{aligned} \phi _0(z,\theta )=\Phi _0(z)+\beta \theta . \end{aligned}$$

For \(q=0,1,\ldots ,n-1\), we denote by \(L^2_{(0,q)}(H_n, \Phi _0)\) the completion of \(\Omega _0^{0,q}(H_n)\) with respect to the norm \(\Vert \cdot \Vert _{\Phi _0}\), where

$$\begin{aligned} \Vert u\Vert ^2_{\Phi _0}=\int _{H_n}\left| u\right| ^2\mathrm{e}^{-\Phi _0}\mathrm{d}v(z)\mathrm{d}\theta ,\quad u\in \Omega _0^{(0,q)}(H_n). \end{aligned}$$

Let \(u(z, \theta )\in \Omega ^{0,1}(H_n)\) with \(\left\| u\right\| _{\phi _0}=1\), \(\Box ^{(1)}_{b,H_n}u=0\). Put \(v(z, \theta )=u(z, \theta )\mathrm{e}^{-\frac{\beta }{2}\theta }\). We have

$$\begin{aligned} \int _{H_n}\!\left| v(z,\theta )\right| ^2\mathrm{e}^{-\Phi _0(z)}\mathrm{d}v(z)\mathrm{d}\theta =1. \end{aligned}$$

Choose \(\chi (\theta )\in C^\infty _0(\mathbb {R})\) so that \(\chi (\theta )=1\) when \(\left| \theta \right| <1\) and \(\chi (\theta )=0\) when \(\left| \theta \right| >2\) and set \(\chi _j(\theta )=\chi (\theta /j)\), \(j\in {\mathbb {N}}\). Let

$$\begin{aligned} \hat{v}_j(z, \eta )=\int _{\mathbb {R}}\!v(z,\theta )\chi _j(\theta )\mathrm{e}^{-i\theta \eta }\mathrm{d}\theta \in \Omega ^{0,1}(H_n),\quad j=1,2,\ldots . \end{aligned}$$
(7.12)

From Parseval’s formula, we have

$$\begin{aligned}&\int _{H_n}\!\left| \hat{v}_j(z,\eta )-\hat{v}_t(z,\eta )\right| ^2\mathrm{e}^{-\Phi _0(z)}\mathrm{d}\eta \mathrm{d}v(z)\\&\quad =2\pi \int _{H_n}\!\left| v(z,\theta )\right| ^2\left| \chi _j(\theta )-\chi _t(\theta )\right| ^2\mathrm{e}^{-\Phi _0(z)}\mathrm{d}\theta \mathrm{d}v(z)\rightarrow 0,\quad j,t\rightarrow \infty . \end{aligned}$$

Thus, there is \(\hat{v}(z, \eta )\in L^2_{(0,1)}(H_n, \Phi _0)\) such that \(\hat{v}_j(z,\eta )\rightarrow \hat{v}(z, \eta )\) in \(L^2_{(0,1)}(H_n, \Phi _0)\). We have

$$\begin{aligned} \int \left| \hat{v}(z,\eta )\right| ^2\mathrm{e}^{-\Phi _0(z)}\mathrm{d}v(z)\mathrm{d}\eta =2\pi . \end{aligned}$$
(7.13)

We call \(\hat{v}(z, \eta )\) the Fourier transform of \(v(z, \theta )\) with respect to \(\theta \). Formally,

$$\begin{aligned} \hat{v}(z, \eta )=\int _{\mathbb {R}}\! \mathrm{e}^{-i\theta \eta }v(z,\theta )\mathrm{d}\theta . \end{aligned}$$
(7.14)

The following theorem is one of the main technical results in [9] (see [9, section 3], for the proof).

Theorem 7.2

With the notations used above. Let \(u(z, \theta )\in \Omega ^{0,1}(H_n)\) with \(\left\| u\right\| _{\phi _0}=1,\) \(\Box ^{(1)}_{b,H_n}u=0\) and let \(\hat{v}(z, \eta )\in L^2_{(0,1)}(H_n, \Phi _0)\) be the Fourier transform of the function \(u(z,\theta )\mathrm{e}^{-\frac{\beta }{2}\theta }\) with respect to \(\theta \) \((\)see the discussion before (7.14)\().\) Then, for almost all \(\eta \in \mathbb {R},\) we have \(\hat{v}(z,\eta )\) is smooth with respect to \(z\) and

$$\begin{aligned} \int _{\mathbb {C}^{n-1}}\left| \hat{v}(z,\eta )\right| ^2\mathrm{e}^{-\Phi _0(z)}\mathrm{d}v(z)<\infty \end{aligned}$$

and

$$\begin{aligned} \left| \hat{v}(z,\eta )\right| ^2 \le (2\pi )^{-n+1}\mathrm{e}^{\Phi _0(z)}{1\!\!1}_{\mathbb {R}_{p,1}}(\eta )\left| \det (M^\phi _p+2\eta {\mathcal L}_p)\right| \int _{\mathbb {C}^{n-1}}\left| \hat{v}(w,\eta )\right| ^2\mathrm{e}^{-\Phi _0(w)}\mathrm{d}v(w)\nonumber \\ \end{aligned}$$
(7.15)

for all \(z\in \mathbb {C}^{n-1}\).

Now, we can prove

Proposition 7.3

Let \(u(z, \theta )\in \Omega ^{0,1}(H_n)\) with \(\left\| u\right\| _{\phi _0}=1,\) \(\Box ^{(1)}_{b,H_n}u=0\). We have

$$\begin{aligned} \left| (Q^{(1)}_{M,H_n}u)(0)\right| ^2\le \frac{E^2}{n-1}, \end{aligned}$$
(7.16)

where \(E\) is as in (5.3).

Proof

Let \(\varphi \in C^\infty _0(\mathbb {C}^{n-1},\mathbb {R})\) such that \(\int _{\mathbb {C}^{n-1}}\!\varphi (z)\mathrm{d}v(z)=1\), \(\varphi \geqslant 0\), \(\varphi (z)=0\) if \(\left| z\right| >1\). Put \(g_m(z)=m^{2n-2}\varphi (mz)\mathrm{e}^{\Phi _0(z)}\), \(m=1,2,\ldots \). Then, \(\int _{\mathbb {C}^{n-1}}\!g_m(z)\mathrm{e}^{-\Phi _0(z)}\mathrm{d}v(z)=1\) and

$$\begin{aligned} (Q^{(1)}_{M,H_n}u)(0)&= \lim _{m\rightarrow \infty }\int \mathrm{e}^{-it\eta }\psi (\eta )\chi _M(t)\mathrm{e}^{-\frac{\beta }{2}t}\mathrm{e}^{-\Phi _0(z)}g_m(z)u(z,t)\mathrm{d}t\mathrm{d}v(z)\nonumber \\&= \lim _{m\rightarrow \infty }\int \hat{\psi }(t)\chi _M(t)\mathrm{e}^{-\frac{\beta }{2}t}\mathrm{e}^{-\Phi _0(z)}g_m(z)u(z,t)\mathrm{d}t\mathrm{d}v(z). \end{aligned}$$
(7.17)

Choose \(\chi (t)\in C^\infty _0(\mathbb {R})\) so that \(\chi (t)=1\) when \(\left| t\right| <1\) and \(\chi (t)=0\) when \(\left| t\right| >2\) and set \(\chi _j(t)=\chi (t/j)\), \(j\in {\mathbb {N}}\). For each \(m\), we have

$$\begin{aligned}&\int \hat{\psi }(t)\chi _M(t)\mathrm{e}^{-\frac{\beta }{2}t}\mathrm{e}^{-\Phi _0(z)}g_m(z)u(z,t)\mathrm{d}t\mathrm{d}v(z)\nonumber \\&\quad =\lim _{j\rightarrow \infty }\int \hat{\psi }(t)\chi _M(t)\mathrm{e}^{-\frac{\beta }{2}t}\mathrm{e}^{-\Phi _0(z)}g_m(z)u(z,t)\chi _j(t)\mathrm{d}t\mathrm{d}v(z). \end{aligned}$$
(7.18)

From Parseval’s formula, we can check that for each \(j\),

$$\begin{aligned}&\int \hat{\psi }(t)\chi _M(t)\mathrm{e}^{-\frac{\beta }{2}t}\mathrm{e}^{-\Phi _0(z)}g_m(z)u(z,t)\chi _j(t)\mathrm{d}t\mathrm{d}v(z)\nonumber \\&\quad =\frac{1}{2\pi }\int \alpha (\eta )\hat{v}_j(z,\eta )g_m(z)\mathrm{e}^{-\Phi _0(z)}\mathrm{d}\eta \mathrm{d}v(z), \end{aligned}$$
(7.19)

where \(\hat{v}_j(z,\eta )\) is as in (7.12) and

$$\begin{aligned} \alpha (\eta )=\int \mathrm{e}^{-it\eta }\hat{\psi }(\eta )\chi _M(t)\mathrm{d}t. \end{aligned}$$
(7.20)

From (7.19) and (7.18), we obtain for each \(m\),

$$\begin{aligned}&\int \hat{\psi }(\eta )\chi _M(t)\mathrm{e}^{-\frac{\beta }{2}t}\mathrm{e}^{-\Phi _0(z)}g_m(z)u(z,t)\mathrm{d}v(z)\mathrm{d}t\nonumber \\&\quad =\frac{1}{2\pi }\int \hat{v}(z,\eta )g_m(z)\alpha (\eta )\mathrm{e}^{-\Phi _0(z)}\mathrm{d}v(z)\mathrm{d}\eta , \end{aligned}$$
(7.21)

where \(\hat{v}(z,\eta )\) is as in (7.14). Now,

$$\begin{aligned} \alpha (\eta )&= \int \mathrm{e}^{-it\eta }\hat{\psi }(t)\chi _M(t)\mathrm{d}t\\&= \int \mathrm{e}^{-it\eta }\hat{\psi }(t)\mathrm{d}t+\int \mathrm{e}^{-it\eta }\hat{\psi }(t)(\chi _M(t)-1)\mathrm{d}t\\&= (2\pi )\psi (\eta )+\alpha _1(\eta ), \end{aligned}$$

where

$$\begin{aligned} \alpha _1(\eta )=\int \mathrm{e}^{-it\eta }\hat{\psi }(t)(\chi _M(t)-1)\mathrm{d}t. \end{aligned}$$

Combining this with (7.21), we have

$$\begin{aligned}&\int \hat{\psi }(t)\chi _M(t)\mathrm{e}^{-\frac{\beta }{2}t}g_m(z)u(z,t)\mathrm{e}^{-\Phi _0(z)}\mathrm{d}v(z)\mathrm{d}t\nonumber \\&\quad =\int \hat{v}(z,\eta )g_m(z)\psi (\eta )\mathrm{e}^{-\Phi _0(z)}\mathrm{d}v(z)\mathrm{d}\eta +\frac{1}{2\pi }\int \hat{v}(z,\eta )g_m(z)\alpha _1(\eta )\mathrm{e}^{-\Phi _0(z)}\mathrm{d}v(z)\mathrm{d}\eta .\nonumber \\ \end{aligned}$$
(7.22)

Since \(\hat{v}(z,\eta )\in L^2_{(0,1)}(H_n,\Phi _0)\), it is easy to see that

$$\begin{aligned} \int \left| \psi (\eta )\right| \left| \hat{v}(z,\eta )\right| \left| g_m(z)\right| \mathrm{e}^{-\Phi _0(z)}\mathrm{d}\eta \mathrm{d}v(z)<\infty ,\quad \forall m>0. \end{aligned}$$
(7.23)

From (7.15), we see that \(\hat{v}(z,\eta )=0\) almost everywhere on \(\mathbb {R}\setminus \mathbb {R}_{p,1}\), for every \(z\in \mathbb {C}^{n-1}\). Since \(\mathrm{Supp\,}\psi \bigcap \mathbb {R}_{p,1}=\emptyset \) [see the discussion after (5.1)], we conclude that for each \(m>0\),

$$\begin{aligned} z\rightarrow \int \psi (\eta )\hat{v}(z,\eta )g_m(z)\mathrm{e}^{-\Phi _0(z)}\mathrm{d}\eta =0. \end{aligned}$$
(7.24)

From (7.23), (7.24) and Fubini’s theorem, we obtain

$$\begin{aligned} \int \hat{v}(z, \eta )g_m(z)\psi (\eta )\mathrm{e}^{-\Phi _0(z)}\mathrm{d}\eta \mathrm{d}v(z)=0 \end{aligned}$$
(7.25)

for every \(m>0\). From (7.25) and (7.22), we get for each \(m\),

$$\begin{aligned}&\int \hat{\psi }(t)\chi _M(t)\mathrm{e}^{-\frac{\beta }{2}t}g_m(z)u(z,t)\mathrm{e}^{-\Phi _0(z)}\mathrm{d}v(z)\mathrm{d}t\nonumber \\&\quad =\frac{1}{2\pi }\int \hat{v}(z,\eta )g_m(z)\alpha _1(\eta )\mathrm{e}^{-\Phi _0(z)}\mathrm{d}v(z)\mathrm{d}\eta .\nonumber \\ \end{aligned}$$
(7.26)

Since \(0\le \chi _M\le 1\) and \(\chi _M=1\) if \(-M\le t\le M\), we have

$$\begin{aligned} \left| \alpha _1(\eta )\right| =\left| \int _{\left| t\right| \ge M}\mathrm{e}^{-it\eta }\hat{\psi }(t)(\chi _M(t)-1)\mathrm{d}t\right| \le \int _{\left| t\right| \ge M}\left| \hat{\psi }(t)\right| \mathrm{d}t\le \frac{2C_0}{M},\quad \forall \eta \in \mathbb {R},\nonumber \\ \end{aligned}$$
(7.27)

where \(C_0=\sup _{t\in \mathbb {R}}t^2\left| \hat{\psi }(t)\right| \). Put

$$\begin{aligned} f(\eta ):=\int _{\mathbb {C}^{n-1}}\left| \hat{v}(z,\eta )\right| ^2\mathrm{e}^{-\Phi _0(z)}\mathrm{d}v(z). \end{aligned}$$

From (7.27) and (7.26), we have for each \(m\),

$$\begin{aligned}&\left| \int \hat{\psi }(\eta )\chi _M(t)\mathrm{e}^{- \frac{\beta }{2}t}g_m(z)u(z,t)\mathrm{e}^{-\Phi _0(z)}\mathrm{d}v(z)\mathrm{d}t\right| \nonumber \\&\quad \le \frac{2C_0}{M}\frac{1}{2\pi }\int \left| \hat{v}(z,\eta )\right| g_m(z)\mathrm{e}^{-\Phi _0(z)}\mathrm{d}v(z)\mathrm{d}\eta =\frac{2C_0}{M}\frac{1}{2\pi }\int _{\left| z\right| \le 1}\left| \hat{v} \left( \frac{z}{m},\eta \right) \right| \varphi (z)\mathrm{d}v(z)\mathrm{d}\eta \nonumber \\&\quad \mathop {\le }\limits ^{\mathrm{by}\, (7.15)}\frac{2C_0}{M}(2\pi )^{-\frac{n+1}{2}}\int _{\left| z\right| \le 1} \mathrm{e}^{\Phi (\frac{z}{m})}{1\!\!1}_{\mathbb {R}_{p,1}}(\eta ) \left| \det (M^{\phi }_p+2\eta {\mathcal L}_p)\right| ^{\frac{1}{2}}\sqrt{f(\eta )}\varphi (z)\mathrm{d}v(z)\mathrm{d}\eta \nonumber \\&\quad \le \frac{2C_0}{M}(2\pi )^{-\frac{n+1}{2}}\sup \{\mathrm{e}^{\Phi _0(\frac{z}{m})};\, \left| z\right| \le 1\}\Biggr (\int _{\mathbb {R}_{p,1}} \left| \det (M^{\phi }_p+2\eta {\mathcal L}_p)\right| \mathrm{d}\eta \Biggr )^{\frac{1}{2}}\Biggr (\int f(\eta )\mathrm{d}\eta \Biggr )^{\frac{1}{2}}\nonumber \\&\quad \mathop {=}\limits ^{\mathrm{by}\, (7.13)}\frac{2C_0}{M}(2\pi )^{-\frac{n}{2}}\sup \{\mathrm{e}^{\Phi _0(\frac{z}{m})};\, \left| z\right| \le 1\}\Biggr (\int _{\mathbb {R}_{p,1}}\left| \det (M^{\phi }_p+2\eta {\mathcal L}_p)\right| \mathrm{d}\eta \Biggr )^{\frac{1}{2}}. \end{aligned}$$
(7.28)

Combining (7.28) with (7.17) and (5.5), we get

$$\begin{aligned} \left| (Q^{(1)}_{M,H_n}u)(0)\right| \le \frac{2C_0}{M}\Bigr ((2\pi )^{-n}\int _{\mathbb {R}_{p,1}}\left| \det (M^{\phi }_p+2\eta {\mathcal L}_p)\right| \mathrm{d}\eta \Bigr )^{\frac{1}{2}}<\frac{E}{\sqrt{n-1}}, \end{aligned}$$

where \(E\) is as in (5.3). (7.16) follows. \(\square \)

In view of Proposition 7.3, we have proved that for all \(u(z, \theta )\in \Omega ^{0,1}(H_n)\) with \(\left\| u\right\| _{\phi _0}=1\), \(\Box ^{(1)}_{b,H_n}u=0\), we have

$$\begin{aligned} \left| (Q^{(1)}_{M,H_n}u)(0)_j\right| ^2\le \left| (Q^{(1)}_{M,H_n}u)(0)\right| ^2< \frac{E^2}{n-1}, \end{aligned}$$

for all \(j=1,\ldots ,n-1\), where \((Q^{(1)}_{M,H_n}u)(0)=\sum ^{n-1}_{j=1}(Q^{(1)}_{M,H_n}u)_j(0)\mathrm{d}\overline{z}_j\) and \(E\) is as in (5.3). Thus, for every \(j=1,\ldots ,n-1\), we have

$$\begin{aligned} \left( Q^{(1)}_{M,H_n}S^{(1)}_{j,H_n}\overline{Q^{(1)}_{M,H_n}}\right) (0) <\frac{E^2}{n-1} \end{aligned}$$

and

$$\begin{aligned} \sum ^{n-1}_{j=1}(Q^{(1)}_{M,H_n}S^{(1)}_{j,H_n}\overline{Q^{(1)}_{M,H_n}})(0) <E^2. \end{aligned}$$
(7.29)

From (7.29) and Theorem 7.1, we obtain the main result of this section

Theorem 7.4

Let \(\nu _k>0\) be any sequence with \(\nu _k\rightarrow 0\) as \(k\rightarrow \infty \). For each \(x\in X,\) we have

$$\begin{aligned} \limsup _{k\rightarrow \infty }k^{-n}(Q^{(1)}_{M,k}\varPi ^{(1)}_{k,\le k\nu _k}\overline{Q^{(1)}_{M,k}})(x)< E^2, \end{aligned}$$
(7.30)

where \(E\) is as in (5.3).

The proof of the following theorem is essentially the same as the proof of (6.1). We omit the proof.

Theorem 7.5

For any sequence \(\nu _k>0\) with \(\nu _k\rightarrow 0\) as \(k\rightarrow \infty ,\) there is a constant \(C>0\) independent of \(k,\) such that

$$\begin{aligned} \left| (Q^{(1)}_{M,k}\varPi ^{(1)}_{k,\,\leqslant k\nu _k}\overline{Q^{(1)}_{M,k}})(x)\right| \leqslant Ck^n,\quad \forall x\in X. \end{aligned}$$
(7.31)

Now, we can prove

Theorem 7.6

Let \(\nu _k>0\) be any sequence with \(\nu _k\rightarrow 0\) as \(k\rightarrow \infty \). Then\(,\) there is a \(k_0>0\) such that for all \(k\ge k_0,\)

$$\begin{aligned} \int _X(Q^{(1)}_{M,k}\varPi ^{(1)}_{k,\le k\nu _k}\overline{Q^{(1)}_{M,k}})(x)\mathrm{d}v_X(x)\le k^n\int _XE^2\mathrm{d}v_X(x), \end{aligned}$$
(7.32)

where \(E\) is as in (5.3).

Proof

In view of Theorem 7.5, \(\sup _k k^{-n}Q^{(1)}_{M,k}\varPi ^{(1)}_{k,\le k\nu _k}\overline{Q^{(1)}_{M,k}})(\cdot )\) is integrable on \(X\). Thus, we can apply Fatou’s lemma and we get using Theorem 7.4:

$$\begin{aligned}&\limsup _{k\rightarrow \infty }k^{-n}\int _X(Q^{(1)}_{M,k}\varPi ^{(1)}_{k,\le k\nu _k}\overline{Q^{(1)}_{M,k}})(x)\mathrm{d}v_X(x)\\&\quad \le \int _X\limsup _{k\rightarrow \infty }k^{-n}(Q^{(1)}_{M,k}\varPi ^{(1)}_{k,\le k\nu _k}\overline{Q^{(1)}_{M,k}})(x)\mathrm{d}v_X(x)\\&\quad <\int _XE^2\mathrm{d}v_X(x). \end{aligned}$$

The theorem follows. \(\square \)

8 The proof of Theorem 1.15

Let \(\delta _k>0\), \(\delta _k\rightarrow \infty \) as \(k\rightarrow \infty \), be as in Theorem 6.4 and let \(\nu _k>0\) be any sequence with \(\lim _{k\rightarrow \infty }\frac{\delta _k}{\nu _k}=0\) and \(\nu _k\rightarrow 0\) as \(k\rightarrow \infty \). Let \(\gamma _{1,k}<\gamma _{2,k}<\cdots <\gamma _{m_k,k}\) be all the distinct non-zero eigenvalues of \(\Box ^{(0)}_{b,k}\) between \(0\) and \(k\nu _k\). Thus, \(\gamma _{1,k}>0\) and \(\gamma _{m_k,k}\le k\nu _k\). We notice that \(\gamma _{j,k}\), \(j=1,\ldots ,m_k\), are also eigenvalues of \(\Box ^{(1)}_{b,k}\). For \(\mu \in \mathbb {R}\), let \(\fancyscript{H}^q_{b,\mu }(X,L^k)\) denote the space spanned by the eigenforms of \(\Box ^{(q)}_{b,k}\) whose eigenvalues are \(\lambda \). For each \(j\in \left\{ 1,\ldots ,m_k\right\} \), let \(f^1_{j,k},f^2_{j,k},\ldots ,f^{d_{j,k}}_{j,k}\) be an orthonormal basis for \(\fancyscript{H}^0_{b,\gamma _{j,k}}(X,L^k)\), where \(d_{j,k}=\mathrm{dim\,}\fancyscript{H}^0_{b,\gamma _{j,k}}(X,L^k)\). Let \(f^1_{0,k},f^2_{0,k},\ldots ,f^{d_{0,k}}_{0,k}\) be an orthonormal basis for \(\fancyscript{H}^0_b(X,L^k)\), where \(d_{0,k}=\mathrm{dim\,}\fancyscript{H}^0_b(X,L^k)\).

Let \(Q^{(0)}_{M,k}\) and \(Q^{(1)}_{M,k}\) be as in (5.7) and (5.13), respectively. By the definition of \((Q^{(0)}_{M,k}\varPi ^{(0)}_{k,\le k\nu _k})(x)\) [see (2.13)], we have

$$\begin{aligned}&(Q^{(0)}_{M,k}\varPi ^{(0)}_{k,\le k\nu _k})(x)\nonumber \\&\quad =\sum ^{d_{0,k}}_{t=1}\langle (Q^{(0)}_{M,k}f^t_{0,k})(x)|f^t_{0,k}(x)\rangle _{h^{L^k}}+\sum ^{m_k}_{j=1}\sum ^{d_{j,k}}_{t=1}\langle (Q^{(0)}_{M,k}f^t_{j,k})(x)|f^t_{j,k}(x)\rangle _{h^{L^k}}. \end{aligned}$$
(8.1)

From (8.1) and (6.35), we conclude that

$$\begin{aligned}&\sum ^{d_{0,k}}_{t=1}\left| (Q^{(0)}_{M,k}f^t_{0,k}\ |\ f^t_{0,k})_{h^{L^k}}\right| +\sum ^{m_k}_{j=1}\sum ^{d_{j,k}}_{t=1}\left| (Q^{(0)}_{M,k}f^t_{j,k}\ |\ f^t_{j,k})_{h^{L^k}}\right| \nonumber \\&\quad \ge \frac{k^n}{2}(2\pi )^{1-n}\int _X\Bigr (\int \psi (\xi )\det (M^\phi _x+2\xi {\mathcal L}_x) {1\!\!1}_{\mathbb {R}_{x,0}}(\xi )\mathrm{d}\xi \Bigr )\mathrm{d}v_X(x), \end{aligned}$$
(8.2)

for \(k\) large. For \(j=1,\ldots ,m_k\), we put

$$\begin{aligned} g^t_{j,k}=\frac{1}{\left\| \overline{\partial }_{b,k}f^t_{j,k}\right\| _{h^{L^k}}}\overline{\partial }_{b,k}f^t_{j,k}=\frac{1}{\sqrt{\gamma _{j,k}}}\overline{\partial }_{b,k}f^t_{j,k} \in \fancyscript{H}^1_{b,\gamma _{j,k}}(X,L^k),\quad t=1,\ldots ,d_{j,k}. \end{aligned}$$

For each \(j=1,\ldots ,m_k\),

$$\begin{aligned} (Q^{(1)}_{M,k}g^t_{j,k}\ |\ g^t_{j,k})_{h^{L^k}}&= \frac{1}{\gamma _{j,k}}(Q^{(1)}_{M,k}\overline{\partial }_{b,k}f^t_{j,k}\ |\ \overline{\partial }_{b,k}f^t_{j,k})_{h^{L^k}}\nonumber \\&= \frac{1}{\gamma _{j,k}}(\overline{\partial }_{b,k}Q^{(0)}_{M,k}f^t_{j,k}\ |\ \overline{\partial }_{b,k}f^t_{j,k})_{h^{L^k}}\ \hbox {here we used (5.18)}\nonumber \\&= \frac{1}{\gamma _{j,k}}(Q^{(0)}_{M,k}f^t_{j,k}\ |\ \Box ^{(0)}_{b,k}f^t_{j,k})_{h^{L^k}}\nonumber \\&= (Q^{(0)}_{M,k}f^t_{j,k}\ |\ f^t_{j,k})_{h^{L^k}},\quad t=1,\ldots ,d_{j,k}. \end{aligned}$$
(8.3)

Hence,

$$\begin{aligned}&\sum ^{m_k}_{j=1}\sum ^{d_{j,k}}_{t=1}\left| (Q^{(0)}_{M,k}f^t_{j,k}\ |\ f^t_{j,k})_{h^{L^k}}\right| \nonumber \\&\quad =\sum ^{m_k}_{j=1}\sum ^{d_{j,k}}_{t=1}\left| (Q^{(1)}_{M,k}g^t_{j,k}\ |\ g^t_{j,k})_{h^{L^k}}\right| \nonumber \\&\quad \le \sqrt{\sum ^{m_k}_{j=1}\sum ^{d_{j,k}}_{t=1}\left\| Q^{(1)}_{M,k}g^t_{j,k}\right\| ^2_{h^{L^k}}} \sqrt{\sum ^{m_k}_{j=1}\sum ^{d_{j,k}}_{t=1}\left\| g^t_{j,k}\right\| ^2_{h^{L^k}}}. \end{aligned}$$
(8.4)

Since \(\left\| g^t_{j,k}\right\| ^2_{h^{L^k}}=\left\| f^t_{j,k}\right\| ^2_{h^{L^k}}=1\), for every \(j\) and \(t\), it is obviously that

$$\begin{aligned} \sqrt{\sum ^{m_k}_{j=1}\sum ^{d_{j,k}}_{t=1}\left\| g^t_{j,k}\right\| ^2_{h^{L^k}}} =\sqrt{\sum ^{m_k}_{j=1}\sum ^{d_{j,k}}_{t=1}\left\| f^t_{j,k}\right\| ^2_{h^{L^k}}}. \end{aligned}$$

Combining this with (8.4) and (8.2), we get

$$\begin{aligned}&\sum ^{d_{0,k}}_{t=1}\left| (Q^{(0)}_{M,k}f^t_{0,k}\ |\ f^t_{0,k})_{h^{L^k}}\right| +\sqrt{\sum ^{m_k}_{j=1}\sum ^{d_{j,k}}_{t=1}\left\| Q^{(1)}_{M,k}g^t_{j,k}\right\| ^2_{h^{L^k}}} \sqrt{\sum ^{m_k}_{j=1}\sum ^{d_{j,k}}_{t=1}\left\| f^t_{j,k}\right\| ^2_{h^{L^k}}}\nonumber \\&\quad \ge \frac{k^n}{2}(2\pi )^{1-n}\int _X\Bigr (\int \psi (\xi )\det (M^\phi _x+2\xi {\mathcal L}_x) {1\!\!1}_{\mathbb {R}_{x,0}}(\xi )\mathrm{d}\xi \Bigr )\mathrm{d}v_X(x), \end{aligned}$$
(8.5)

for \(k\) large.

We can check that for each \(j=1,\ldots ,m_k\), \(g^t_{j,k}\), \(t=1,\ldots ,d_{j,k}\) is an orthonormal basis of the space \(\overline{\partial }_{b,k}\fancyscript{H}^0_{b,\gamma _{j,k}}(X,L^k)\subset \fancyscript{H}^1_{b,\gamma _{j,k}}(X,L^k)\). From this observation and the definition of \((Q^{(1)}_{M,k}\varPi ^{(1)}_{k,\le k\nu _k}\overline{Q^{(1)}_{M,k}})(x)\) [see (2.14)], we conclude that

$$\begin{aligned} \sum ^{m_k}_{j=1}\sum ^{d_{j,k}}_{t=1}\left| Q^{(1)}_{M,k}g^t_{j,k}\right| ^2_{h^{L^k}}(x)\le (Q^{(1)}_{M,k}\varPi ^{(1)}_{k,\le k\nu _k}\overline{Q^{(1)}_{M,k}})(x). \end{aligned}$$
(8.6)

Thus,

$$\begin{aligned} \sum ^{m_k}_{j=1}\sum ^{d_{j,k}}_{t=1}\left\| Q^{(1)}_{M,k}g^t_{j,k}\right\| ^2_{h^{L^k}}\le \int _X(Q^{(1)}_{M,k}\varPi ^{(1)}_{k,\le k\nu _k}\overline{Q^{(1)}_{M,k}})(x)\mathrm{d}v(x). \end{aligned}$$
(8.7)

Combining (8.7) with (7.32), we get

$$\begin{aligned} \sum ^{m_k}_{j=1}\sum ^{d_{j,k}}_{t=1}\left\| Q^{(1)}_{M,k}g^t_{j,k}\right\| ^2_{h^{L^k}} \le k^n\int _XE^2\mathrm{d}v_X(x) \end{aligned}$$
(8.8)

for \(k\) large, where \(E\) is as in (5.3).

From (2.19) and (2.18), we conclude that

$$\begin{aligned} \int _X\varPi ^{(0)}_{k,\,\leqslant k\nu _k}(x)\mathrm{d}v_X(x)=k^n(2\pi )^{-n}\int _X\Bigr (\int _{\mathbb {R}_{x,0}}\left| \det (M^\phi _x+2s{\mathcal L}_x)\right| \mathrm{d}s\Bigr )\mathrm{d}v_X(x)+o(k^n),\nonumber \\ \end{aligned}$$
(8.9)

for \(k\) large. It is obviously the case that

$$\begin{aligned} \sum ^{m_k}_{j=1}\sum ^{d_{j,k}}_{t=1}\left\| f^t_{j,k}\right\| ^2_{h^{L^k}}\le \int _X\varPi ^{(0)}_{k,\,\leqslant k\nu _k}(x)\mathrm{d}v_X(x). \end{aligned}$$

Combining this with (8.9), we get

$$\begin{aligned}&\sum ^{m_k}_{j=1}\sum ^{d_{j,k}}_{t=1}\left\| f^t_{j,k}\right\| ^2_{h^{L^k}}\nonumber \\&\quad \le k^n2(2\pi )^{-n}\int _X\Bigr (\int \det (M^\phi _x+2\xi {\mathcal L}_x) {1\!\!1}_{\mathbb {R}_{x,0}}(\xi )\mathrm{d}\xi \Bigr )\mathrm{d}v_X(x) \end{aligned}$$
(8.10)

for \(k\) large. From (8.8), (8.10), (8.5) and (5.3), we obtain

Theorem 8.1

Let \(f^1_{0,k},f^2_{0,k},\ldots ,f^{d_{0,k}}_{0,k}\) be an orthonormal basis for \(\fancyscript{H}^0_b(X,L^k),\) where \(d_{0,k}=\mathrm{dim\,}\fancyscript{H}^0_b(X,L^k).\) Then\(,\) for \(k\) large\(,\) we have

$$\begin{aligned}&\sum ^{d_{0,k}}_{t=1}\left| (Q^{(0)}_{M,k}f^t_{0,k}\ |\ f^t_{0,k})_{h^{L^k}}\right| \nonumber \\&\quad \ge \frac{k^n}{4}(2\pi )^{1-n}\int _X\Bigr (\int \psi (\xi )\det (M^\phi _x+2\xi {\mathcal L}_x) {1\!\!1}_{\mathbb {R}_{x,0}}(\xi )\mathrm{d}\xi \Bigr )\mathrm{d}v_X(x). \end{aligned}$$
(8.11)

The following is straightforward

Lemma 8.2

For \(k\) large\(,\) there is a constant \(C>0\) independent of \(k,\) such that

$$\begin{aligned} \left\| Q^{(0)}_{M,k}u\right\| ^2_{h^{L^k}}\le C\left\| u\right\| ^2_{h^{L^k}},\quad \forall u\in C^\infty (X,L^k). \end{aligned}$$

Proof

Let \(D\Subset D'\Subset D''\Subset X\) be open sets of \(X\) and let \(s\) be a local section of \(L\) on \(D''\). We assume that there exist canonical coordinates \(x=(x_1,\ldots ,x_{2n-1})=(z,\theta )\) on \(D''\). Let \(\chi _M\) be as in (5.6). For \(k\) large, we have

$$\begin{aligned} \left\{ \Phi ^{\frac{t}{k}}(x)\in D';\, \forall x\in D, t\in \mathrm{Supp\,}\chi _M\right\} \end{aligned}$$

and \(\mathrm{Supp\,}f(\Phi ^{\frac{t}{k}}x)\subset D'\), \(\forall t\in \mathrm{Supp\,}\chi _M\), \(\forall f\in C^\infty _0(D,L^k)\). In canonical coordinates \(x=(z,\theta )\), we have \(\Phi ^{\frac{t}{k}}(x)=(z,\frac{t}{k}+\theta )\). Let \(m(z,\theta )\mathrm{d}v(z)\mathrm{d}\theta \) be the volume form on \(D''\), where \(\mathrm{d}v(z)=2^{n-1}\mathrm{d}x_1\mathrm{d}x_2\ldots \mathrm{d}x_{2n-2}\). Since \(m(z,\theta )\) is strictly positive, for \(k\) large, there is a constant \(C_1>0\) independent of \(k\), such that

$$\begin{aligned} m(z,\theta )\le C_1m\left( z,\theta +\frac{t}{k}\right) ,\quad \forall (z,\theta )\in D',\ t\in \mathrm{Supp\,}\chi _M. \end{aligned}$$
(8.12)

Let \(u\in C^\infty _0(D,L^k)\). On \(D''\), we write \(u=s^k\widetilde{u}\), \(\widetilde{u}\in C^\infty _0(D)\). From the definition of \(Q^{(0)}_{M,k}\) [see (5.7)], we can check that for \(k\) large,

$$\begin{aligned}&\int \left| (Q^{(0)}_{M,k}u)(x)\right| ^2_{h^{L^k}}\mathrm{d}v_X(x)\nonumber \\&\quad =\int _{D'}\left| \int \mathrm{e}^{-it\eta }\psi (\eta )\chi _M(t) \mathrm{e}^{-\frac{k}{2}\phi \left( z,\theta +\frac{t}{k }\right) }\widetilde{u}\left( z,\theta +\frac{t}{k}\right) \mathrm{d}t\mathrm{d}\eta \right| ^2m(z,\theta )\mathrm{d}v(z)\mathrm{d}\theta \nonumber \\&\quad \le \widetilde{C}\int _{(z,\theta )\in D'}\chi _M(t)\mathrm{e}^{-k\phi \left( z,\theta +\frac{t}{k}\right) }\left| \widetilde{u}\left( z,\theta +\frac{t}{k}\right) \right| ^2m(z,\theta )\mathrm{d}td\theta \mathrm{d}v(z)\nonumber \\&\quad \le \widetilde{C}C_1\int _{(z,\theta )\in D'}\chi _M(t)\mathrm{e}^{-k\phi \left( z,\theta +\frac{t}{k}\right) }\left| \widetilde{u}\left( z,\theta +\frac{t}{k}\right) \right| ^2m\left( z,\theta +\frac{t}{k}\right) \mathrm{d}td\theta \mathrm{d}v(z)\nonumber \\&\quad =\widetilde{C}C_1\int _{(z,\lambda -\frac{t}{k})\in D'}\chi _M(t)\mathrm{e}^{-k\phi (z,\lambda )}\left| \widetilde{u}(z,\lambda )\right| ^2m(z,\lambda )\mathrm{d}td\lambda \mathrm{d}v(z)\nonumber \\&\quad \le C\int \mathrm{e}^{-k\phi (z,\theta )}\left| \widetilde{u}(z,\theta )\right| ^2m(z,\theta )\mathrm{d}v(z)\mathrm{d}\theta =C\left\| u\right\| ^2_{h^{L^k}}. \end{aligned}$$
(8.13)

where \(\widetilde{C}>0\), \(C>0\) are independent of \(k\) and \(u\) and \(C_1\) is as in (8.12). From (8.13) and using partition of unity, the lemma follows. \(\square \)

Proof of Theorem 1.15

From Lemma 8.2 and (8.11), we see that for \(k\) large,

$$\begin{aligned} \sqrt{C}d_{0,k}&= \sqrt{C}\sum ^{d_{0,k}}_{t=1}\left\| f^t_{0,k}\right\| ^2_{h^{L^k}} \ge \sum ^{d_{0,k}}_{t=1}\left| (Q^{(0)}_{M,k}f^t_{0,k}\ |\ f^t_{0,k})_{h^{L^k}}\right| \\&\ge \frac{k^n}{4}(2\pi )^{1-n}\int _X\Bigr (\int \psi (\xi )\det (M^\phi _x+2\xi {\mathcal L}_x) {1\!\!1}_{\mathbb {R}_{x,0}}(\xi )\mathrm{d}\xi \Bigr )\mathrm{d}v_X(x), \end{aligned}$$

where \(C>0\) is the constant as in Lemma 8.2 and \(d_{0,k}=\mathrm{dim\,}\fancyscript{H}^0_b(X,L^k)\). Theorem 1.15 follows. \(\square \)

9 Examples

In this section, some examples are collected. The aim is to illustrate the main results in some simple situations.

9.1 Compact Heisenberg groups

Let \(\lambda _1,\ldots ,\lambda _{n-1}\) be given non-zero integers. Let \({\fancyscript{C}}H_n=(\mathbb {C}^{n-1}\times \mathbb {R})/_\sim \), where \((z, \theta )\sim (\widetilde{z}, \widetilde{\theta })\) if

$$\begin{aligned}&\widetilde{z}-z=(\alpha _1,\ldots ,\alpha _{n-1})\in \sqrt{2\pi }{\mathbb {Z}}^{n-1}+i\sqrt{2\pi }{\mathbb {Z}}^{n-1},\\&\widetilde{\theta }-\theta -i\sum ^{n-1}_{j=1}\lambda _j(z_j\overline{\alpha }_j-\overline{z}_j\alpha _j)\in \pi {\mathbb {Z}}. \end{aligned}$$

We can check that \(\sim \) is an equivalence relation and \({\fancyscript{C}}H_n\) is a compact manifold of dimension \(2n-1\). The equivalence class of \((z, \theta )\in \mathbb {C}^{n-1}\times \mathbb {R}\) is denoted by \([(z, \theta )]\). For a given point \(p=[(z, \theta )]\), we define \(T^{1, 0}_p{\fancyscript{C}}H_n\) to be the space spanned by

$$\begin{aligned} \textstyle \big \{\frac{\partial }{\partial z_j}+i\lambda _j\overline{z}_j\frac{\partial }{\partial \theta },\ j=1,\ldots ,n-1\big \}. \end{aligned}$$

It is easy to see that the definition above is independent of the choice of a representative \((z, \theta )\) for \([(z, \theta )]\). Moreover, we can check that \(T^{1, 0}{\fancyscript{C}}H_n\) is a CR structure and \(T:=\frac{\partial }{\partial \theta }\) is a rigid global real vector field. Thus, \(({\fancyscript{C}}H_n, T^{1, 0}{\fancyscript{C}}H_n)\) is a compact generalized Sasakian CR manifold of dimension \(2n-1\). Let \(J\) denote the canonical complex structure on \(X\times \mathbb {R}\) given by \(J\frac{\partial }{\partial t}=T\), where \(t\) denotes the coordinate of \(\mathbb {R}\). We take a Hermitian metric \(\langle \,\cdot \,|\,\cdot \,\rangle \) on the complexified tangent bundle \(\mathbb {C}T{\fancyscript{C}}H_n\) such that

$$\begin{aligned} \Big \lbrace \tfrac{\partial }{\partial z_j}+i\lambda _j\overline{z}_j\tfrac{\partial }{\partial \theta }, \tfrac{\partial }{\partial \overline{z}_j}-i\lambda _jz_j\tfrac{\partial }{\partial \theta }, -\tfrac{\partial }{\partial \theta };\, j=1,\ldots ,n-1\Big \rbrace \end{aligned}$$

is an orthonormal basis. The dual basis of the complexified cotangent bundle is

$$\begin{aligned} \Big \lbrace \mathrm{d}z_j,\, \mathrm{d}\overline{z}_j,\, \omega _0:=-\mathrm{d}\theta +\textstyle \sum ^{n-1}_{j=1}(i\lambda _j\overline{z}_j\mathrm{d}z_j-i\lambda _jz_j\mathrm{d}\overline{z}_j); j=1,\ldots ,n-1 \Big \rbrace . \end{aligned}$$

The Levi form \({\mathcal L}_p\) of \({\fancyscript{C}}H_n\) at \(p\in {\fancyscript{C}}H_n\) is given by \({\mathcal L}_p=\sum ^{n-1}_{j=1}\lambda _j\mathrm{d}z_j\wedge \mathrm{d}\overline{z}_j\).

Now, we construct a generalized Sasakian CR line bundle \((L,J)\) over \({\fancyscript{C}}H_n\). Let \(L=(\mathbb {C}^{n-1}\times \mathbb {R}\times \mathbb {C})/_\equiv \) where \((z,\theta ,\eta )\equiv (\widetilde{z}, \widetilde{\theta }, \widetilde{\eta })\) if

$$\begin{aligned}&(z,\theta )\sim (\widetilde{z}, \widetilde{\theta }),\\&\widetilde{\eta }\exp \left( \widetilde{\theta }+i\sum ^{n-1}_{j=1}\lambda _j\left| \widetilde{z}_j\right| ^2\right) =\eta \exp \left( \theta +i\sum ^{n-1}_{j=1}\lambda _j\left| z_j\right| ^2\right) \exp \left( \sum ^{n-1}_{j,t=1}\mu _{j,t} \left( z_j\overline{\alpha }_t+\frac{1}{2}\alpha _j\overline{\alpha }_t\right) \right) , \end{aligned}$$

where \(\alpha =(\alpha _1,\ldots ,\alpha _{n-1})=\widetilde{z}-z\), \(\mu _{j,t}=\mu _{t,j}\), \(j, t=1,\ldots ,n-1\), are given integers. We can check that \(\equiv \) is an equivalence relation and \((L,J)\) is a generalized Sasakian CR line bundle over \({\fancyscript{C}}H_n\). For \((z, \theta , \eta )\in \mathbb {C}^{n-1}\times \mathbb {R}\times \mathbb {C}\), we denote \([(z, \theta , \eta )]\) its equivalence class. It is straightforward to see that the pointwise norm

$$\begin{aligned} \big |[(z, \theta , \eta )]\big |^2_{h^L}:=\left| \eta \right| ^2\exp \big (2\theta -\textstyle \sum ^{n-1}_{j,t=1} \mu _{j,t}z_j\overline{z}_t\big ) \end{aligned}$$

is well defined. In local coordinates \((z, \theta , \eta )\), the weight function of this metric is

$$\begin{aligned} \phi =-2\theta +\sum ^{n-1}_{j,t=1}\mu _{j,t}z_j\overline{z}_t. \end{aligned}$$

We can check that \(T\phi =-2\). Thus, \((L,J,h^L)\) is a rigid generalized Sasakian CR line bundle over \({\fancyscript{C}}H_n\). Note that

$$\begin{aligned} \textstyle \overline{\partial }_b=\sum ^{n-1}_{j=1}\mathrm{d}\overline{z}_j\wedge (\frac{\partial }{\partial \overline{z}_j}-i\lambda _jz_j\frac{\partial }{\partial \theta }),\quad \partial _b=\sum ^{n-1}_{j=1}\mathrm{d}z_j\wedge (\frac{\partial }{\partial z_j}+i\lambda _j\overline{z}_j\frac{\partial }{\partial \theta }). \end{aligned}$$

Thus, \(d(\overline{\partial }_b\phi -\partial _b\phi )=2\sum ^{n-1}_{j,t=1}\mu _{j,t}\mathrm{d}z_j\wedge \mathrm{d}\overline{z}_t\) and for any \(p\in {\fancyscript{C}}H_n\),

$$\begin{aligned} M^\phi _p=\sum ^{n-1}_{j,t=1}\mu _{j,t}\mathrm{d}z_j\wedge \mathrm{d}\overline{z}_t. \end{aligned}$$

From this and Theorem 1.15, we obtain

Theorem 9.1

If the matrix \(\left( \mu _{j,t}\right) ^{n-1}_{j,t=1}\) is positive definite and \(Y(0),\) \(Y(1)\) hold on \({\fancyscript{C}}H_n,\) then for \(k\) large\(,\) there is a constant \(c>0\) independent of \(k,\) such that

$$\begin{aligned} \mathrm{dim\,}H^0_b({\fancyscript{C}}H_n, L^k)\ge ck^n. \end{aligned}$$

9.2 Holomorphic line bundles over a complex torus

Let

$$\begin{aligned} T_n:=\mathbb {C}^n/(\sqrt{2\pi }{\mathbb {Z}}^n+i\sqrt{2\pi }{\mathbb {Z}}^n) \end{aligned}$$

be the flat torus. Let \(\lambda =\left( \lambda _{j,t}\right) ^{n}_{j,t=1}\), where \(\lambda _{j,t}=\lambda _{t,j}\), \(j, t=1,\ldots ,n\), are given integers. Let \(L_\lambda \) be the holomorphic line bundle over \(T_n\) with curvature the \((1,1)\)-form \(\Theta _\lambda =\sum ^n_{j,t=1}\lambda _{j,t}\mathrm{d}z_j\wedge \mathrm{d}\overline{z}_t\). More precisely, \(L_\lambda :=(\mathbb {C}^n\times \mathbb {C})/_\sim \), where \((z, \theta )\sim (\widetilde{z}, \widetilde{\theta })\) if

$$\begin{aligned} \widetilde{z}-z=(\alpha _1,\ldots ,\alpha _n)\in \sqrt{2\pi }{\mathbb {Z}}^n+i\sqrt{2\pi }{\mathbb {Z}}^n,\quad \widetilde{\theta }=\textstyle \exp \big (\sum ^n_{j,t=1}\lambda _{j,t}(z_j\overline{\alpha }_t+\tfrac{1}{2}\alpha _j\overline{\alpha }_t\,)\big )\theta . \end{aligned}$$

We can check that \(\sim \) is an equivalence relation and \(L_\lambda \) is a holomorphic line bundle over \(T_n\). For \([(z, \theta )]\in L_\lambda \), we define the Hermitian metric by

$$\begin{aligned} \big \vert [(z, \theta )]\big \vert ^2:=\left| \theta \right| ^2\textstyle \exp (-\sum ^n_{j,t=1}\lambda _{j,t}z_j\overline{z}_t) \end{aligned}$$

and it is easy to see that this definition is independent of the choice of a representative \((z, \theta )\) of \([(z, \theta )]\). We denote by \(\phi _\lambda (z)\) the weight of this Hermitian fiber metric. Note that \(\partial \overline{\partial }\phi _\lambda =\Theta _\lambda \).

Let \(L^*_\lambda \) be the dual bundle of \(L_\lambda \) and let \(\left\| \,\cdot \,\right\| _{L^*_\lambda }\) be the norm of \(L^*_\lambda \) induced by the Hermitian fiber metric on \(L_\lambda \). Consider the compact CR manifold of dimension \(2n+1\): \(X=\{v\in L^*_\lambda ;\, \left\| v\right\| _{L^*_\lambda }=1\}\); this is the boundary of the Grauert tube associated to \(L^*_\lambda \). The manifold \(X\) is equipped with a natural \(S^1\)-action. Locally, \(X\) can be represented in local holomorphic coordinates \((z,\eta )\), where \(\eta \) is the fiber coordinate, as the set of all \((z,\eta )\) such that \(\left| \eta \right| ^2\mathrm{e}^{\phi _\lambda (z)}=1\). The \(S^1\)-action on \(X\) is given by \(\mathrm{e}^{i\theta }\circ (z,\eta )=(z,\mathrm{e}^{i\theta }\eta )\), \(\mathrm{e}^{i\theta }\in S^1\), \((z,\eta )\in X\). Let \(T\) be the global real vector field on \(X\) determined by \(Tu(x)=\frac{\partial }{\partial \theta }u(\mathrm{e}^{i\theta }\circ x)\big |_{\theta =0}\), for all \(u\in C^\infty (X)\). We can check that \(T\) is a rigid global real vector field on \(X\). Thus, \(X\) is a compact generalized Sasakian CR manifold of dimension \(2n+1\). Let \(J\) denote the canonical complex structure on \(X\times \mathbb {R}\) given by \(J\frac{\partial }{\partial t}=T\), where \(t\) denotes the coordinate of \(\mathbb {R}\).

Let \(\pi :L^*_\lambda \rightarrow T_n\) be the natural projection from \(L^*_\lambda \) onto \(T_n\). Let \(\mu =\left( \mu _{j,t}\right) ^{n}_{j,t=1}\), where \(\mu _{j,t}=\mu _{t,j}\), \(j, t=1,\ldots ,n\), are given integers. Let \(L_\mu \) be another holomorphic line bundle over \(T_n\) determined by the constant curvature form \(\Theta _\mu =\sum ^n_{j,t=1}\mu _{j,t}\mathrm{d}z_j\wedge \mathrm{d}\overline{z}_t\) as above. The pullback line bundle \(\pi ^*L_\mu \) is a holomorphic line bundle over \(L^*_\lambda \). If we restrict \(\pi ^*L_\mu \) on \(X\), then we can check that \((\pi ^*L_\mu ,J)\) is a generalized Sasakian CR line bundle over \(X\).

The Hermitian fiber metric on \(L_\mu \) induced by \(\phi _\mu \) induces a Hermitian fiber metric on \(\pi ^*L_\mu \) that we shall denote by \(h^{\pi ^*L_\mu }\). We let \(\psi \) to denote the weight of \(h^{\pi ^*L_\mu }\). The part of \(X\) that lies over a fundamental domain of \(T_n\) can be represented in local holomorphic coordinates \((z, \xi )\), where \(\xi \) is the fiber coordinate, as the set of all \((z, \xi )\) such that \(r(z, \xi ):=\left| \xi \right| ^2\exp (\sum ^n_{j,t=1}\lambda _{j,t}z_j\overline{z}_t)-1=0\) and the weight \(\psi \) may be written as \(\psi (z, \xi )=\sum ^n_{j,t=1}\mu _{j,t}z_j\overline{z}_t\). From this we see that \((\pi ^*L_\mu ,J,h^{\pi ^*L_\mu })\) is a rigid generalized Sasakian CR line bundle over \(X\). It is straightforward to check that for any \(p\in X\), we have \(M^\psi _p=\frac{1}{2}d(\overline{\partial }_b\psi -\partial _b\psi )(p)|_{T^{1, 0}X}=\sum ^n_{j,t=1}\mu _{j,t}\mathrm{d}z_j\wedge \mathrm{d}\overline{z}_t\). From this observation and Theorem 1.15, we obtain

Theorem 9.2

If the matrix \(\left( \mu _{j,t}\right) ^{n-1}_{j,t=1}\) is positive definite and \(Y(0),\) \(Y(1)\) hold on \(X,\) then for \(k\) large\(,\) there is a constant \(c>0\) independent of \(k,\) such that

$$\begin{aligned} \dim H^0_b(X, (\pi ^*L_\mu )^k)\ge ck^{n+1}. \end{aligned}$$