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

Invariant weighted Bergman metrics on domains

Sungmin Yoo Department of Mathematics, Incheon National University, 119 Academy-ro, Yeonsu-gu, Incheon, 22012, Republic of Korea
Abstract.

In this paper, we study the cases where the weighted Bergman metrics of a domain are invariant under biholomorphisms by introducing the concept of invariant weight assignments, focusing on two examples by Tian and Tsuji, respectively. Using Bergman’s minimum integral method and a domain version of the Tian-Yau-Zelditch expansion for the weighted Bergman kernels and metrics, we give an alternative proof of uniform convergence of Tian’s sequence of Bergman kernels and metrics on uniform squeezing domains. We also present a proof of the uniform convergence of Tsuji’s dynamical kernel sequence on uniform squeezing domains.

1. Introduction

Bergman [bergman1970] introduced the concept of the Bergman kernel and metric for bounded domains from the Hilbert space of L2L^{2}-holomorphic functions, known as the Bergman space. This seminal work laid the foundation for understanding complex manifolds from the perspective of function theory. Subsequently, Kobayashi [kobayashi1959geometry] extended these concepts to abstract complex manifolds by considering holomorphic (n,0)(n,0)-forms, sections of the canonical line bundle. The remarkable property of invariance under biholomorphisms renders the Bergman kernel form and metric canonical within the realm of Several Complex Variables and Complex Geometry.

These concepts found further generalization to polarized manifolds, which are compact complex manifolds endowed with ample line bundles [tian1990, donaldson2001scalar]. Their significance became particularly pronounced in the proof of the celebrated Tian-Yau-Donaldson conjecture [chen2015kahler1, chen2015kahler2, chen2015kahler3, tian2015k]. However, it is worth noting that the definition of the Bergman kernel function in this setting is contingent upon the choice of the hermitian metric and volume form in defining the inner product on the vector space of global sections.

Returning to the domain context, these results can be elucidated through the concept of weighted Bergman kernels. Let Ω\Omega be a domain in n\mathbb{C}^{n} and dλd\lambda be the standard Lebesgue measure. For a positive measurable function μ\mu on Ω\Omega, consider the space

𝒜2(Ω,μ):={u𝒪(Ω)|uΩ,μ2:=Ω|u|2μ𝑑λ<}.\mathcal{A}^{2}(\Omega,\mu):=\left\{u\in\mathcal{O}(\Omega)\ \Big{|}\ \left\|u\right\|^{2}_{\Omega,\mu}:=\int_{\Omega}|u|^{2}\mu d\lambda<\infty\right\}.

If the weight function μ\mu satisfies the admissible condition (e.g. continuous function), then the above space admits the reproducing kernel, called the weighted Bergman kernel or μ\mu-Bergman kernel. As in the classic case (μ=1Ω\mu=1_{\Omega}), this gives us a Kähler metric if Ω\Omega is bounded, which is called the weighted Bergman metric or μ\mu-Bergman metric. However, compared to the classic Bergman metric, there has been relatively little research on the geometry of weighted Bergman metrics, primarily due to their lack of biholomorphic invariance.

In this paper, we study the situation when the weighted Bergman metric is invariant under biholomorphisms, For this, we introduce the concept of (biholomorphically) invariant weight assignment. Roughly speaking, we consider an weight assignment \mathcal{M} for domains such that the weight function μΩ:=(Ω)\mu_{\Omega}:=\mathcal{M}(\Omega) only depends on the geometry of the domain Ω\Omega (for precise statements, see Definition 3.5).

Theorem 1 (Theorem 3.7).

If the assignment \mathcal{M} is invariant, the \mathcal{M}-weighted Bergman metric is invariant: gΩ,μΩ=FgF(Ω),μF(Ω),g_{\Omega,\mu_{\Omega}}=F^{*}g_{F(\Omega),\mu_{F(\Omega)}}, for any biholomorphism FF.

The above theorem yields that we can consider various types of biholomorphically invariant Kähler metrics on domains depending upon the choice of invariant assignments. As examples, we investigate two important invariant weighted Bergman metrics following the approaches of Tian [tian1990] and Tsuji [tsuji2010] in the setting of canonically polarized compact manifolds.

As the first example, consider the sequence of weight functions μΩ,mKE=det(gΩKE)(m1),\mu^{\rm KE}_{\Omega,m}={\det\left(g^{\rm KE}_{\Omega}\right)}^{-(m-1)}, where gΩKEg^{\rm KE}_{\Omega} is the Kähler-Einstein metric of the bounded pseudoconvex domain Ω\Omega. Then, the weighted Bergman kernels KΩ,μΩ,mKEK_{\Omega,\mu^{\rm KE}_{\Omega,m}} gives us a sequence of invariant weighted Bergman metrics gΩ,μΩ,mKEg_{\Omega,\mu^{\rm KE}_{\Omega,m}}. Denotes the corresponding holomorphic sectional curvature by HΩ,μΩ,mKE{H}_{\Omega,\mu^{\rm KE}_{\Omega,m}}. An appropriate normalization gives the following

Theorem 2 (Theorem 5.5).

If Ω\Omega is a bounded pseudoconvex domain (with the uniform squeezing property), then KΩ,μΩ,mKEm,1mgΩ,μΩ,mKE,mHΩ,μΩ,mKE\sqrt[m]{{K}_{\Omega,\mu^{\rm KE}_{\Omega,m}}},\frac{1}{m}{g}_{\Omega,\mu^{\rm KE}_{\Omega,m}},m{H}_{\Omega,\mu^{\rm KE}_{\Omega,m}} converges (uniformly) on Ω\Omega to det(gΩKE),gΩKE,HgΩKE\det\left(g^{\rm KE}_{\Omega}\right),g^{\rm KE}_{\Omega},H_{g^{\rm KE}_{\Omega}}, respectively, as mm\rightarrow\infty.

This theorem can be considered as a noncompact version of Tian’s theorem in [tian1990] for domains with the Kähler-Einstein metrics. For the proof, Tian [tian1990] constructed an asymptotic expansion of the Bergman metric of canonically polarized manifolds using the “peaked section” method, based on Hörmander’s theorem. For domains in n\mathbb{C}^{n}, this method can be understood in terms of the weighted version of the Bergman minimum integral method. In this paper, we present a straightforward proof of the following domain version of the Tian-Yau-Zelditch expansion for the weighted Bergman kernels, metrics, and curvatures, following the proofs in [tian1990, ruan1998canonical, lu2000lower, szekelyhidi2014introduction].

Theorem 3 (Theorem 4.10).

Let Ω\Omega be a pseudoconvex domain in n\mathbb{C}^{n} with a smooth strictly plurisubharmonic function φ\varphi. Suppose that (Ω,eφ)(\Omega,e^{-\varphi}) admits the positive definite weighted Bergman metric. Fix a point pΩp\in\Omega and a vector XnX\in\mathbb{C}^{n}. For sufficiently large mm, we have

KΩ,emφ(p)\displaystyle K_{\Omega,e^{-m\varphi}}(p) =det(φkl¯(p))emφ(p)(mπ)n(1Sφ(p)2m+O(1m2)),\displaystyle=\frac{\det{\left(\varphi_{k\overline{l}}(p)\right)}}{e^{-m\varphi(p)}}\left(\frac{m}{\pi}\right)^{n}\left(1-\frac{S_{\varphi}(p)}{2m}+O\Big{(}\frac{1}{m^{2}}\Big{)}\right),
gΩ,emφ(p;X)\displaystyle g_{\Omega,e^{-m\varphi}}(p;X) =mgφ(p;X)(1Rφ(p;X)m+O(1m2)),\displaystyle=mg_{\varphi}(p;X)\left(1-\frac{R_{\varphi}(p;X)}{m}+O\Big{(}\frac{1}{m^{2}}\Big{)}\right),
HΩ,emφ(p;X)\displaystyle H_{\Omega,e^{-m\varphi}}(p;X) =Hφ(p;X)m+O(1m2),\displaystyle=\frac{H_{\varphi}(p;X)}{m}+O\Big{(}\frac{1}{m^{2}}\Big{)},

where Sφ,Rφ,HφS_{\varphi},R_{\varphi},H_{\varphi} are the scalar, Ricci, holomorphic sectional curvature of the Kähler metric gφ:=i¯φg_{\varphi}:=i\partial\overline{\partial}\varphi, respectively.

Note that the above formula for KΩ,emφK_{\Omega,e^{-m\varphi}} differs slightly from the compact case, since the definition of the weighted Bergman kernel for domains depends on the standard Euclidean coordinates. For the proof, we use a classic version of Hörmander’s theorem for domains [hormander1965] as we use the standard Lebesgue measure dλd\lambda (the volume form of the incomplete Euclidean metric) instead of (i¯φ)n=det(φkl¯)dλ(i\partial\overline{\partial}\varphi)^{n}=\det{\left(\varphi_{k\overline{l}}\right)}d\lambda. Note that in our case, the metric i¯φi\partial\overline{\partial}\varphi not necessarily to be complete on Ω\Omega.

As the second example of invariant weighted Bergman metrics, we consider an iterative sequence of weighted functions, developed by Tsuji’s [tsuji2010]. Set μ~Ω,1B:=1Ω\widetilde{\mu}^{\rm B}_{\Omega,1}:=1_{\Omega}. For the given weight function μ~Ω,mB\widetilde{\mu}^{\rm B}_{\Omega,m}, consider the weighted Bergman kernel KΩ,μ~Ω,mBK_{\Omega,\widetilde{\mu}^{\rm B}_{\Omega,m}} of the space 𝒜μ~Ω,mB2(Ω)\mathcal{A}^{2}_{\widetilde{\mu}^{\rm B}_{\Omega,m}}(\Omega). Define the next weight function by

μ~Ω,m+1B:=1Cm1KΩ,μ~Ω,mB:=(mπ)n(1n2m)1KΩ,μ~Ω,mB.\widetilde{\mu}^{\rm B}_{\Omega,m+1}:=\frac{1}{C_{m}}\frac{1}{K_{\Omega,\widetilde{\mu}^{\rm B}_{\Omega,m}}}:=\left(\frac{m}{\pi}\right)^{n}\left(1-\frac{n}{2m}\right)\frac{1}{K_{\Omega,\widetilde{\mu}^{\rm B}_{\Omega,m}}}.

Then the corresponding weighted Bergman metrics are invariant under biholomorphisms (Proposition 5.6). Although we were unable to obtain the metric convergence result, we have the following convergence result on the potential level.

Theorem 4 (Theorem 5.12).

Let Ω\Omega be a uniform squeezing domain in n\mathbb{C}^{n}. There exists a uniform constant C>0C>0 satisfying

eCmdet(gΩKE)CmKΩ,μ~Ω,mBmeCmdet(gΩKE).e^{-\frac{C}{m}}\det\left(g^{\rm KE}_{\Omega}\right)\leq\sqrt[m]{C_{m}K_{\Omega,\widetilde{\mu}^{\rm B}_{\Omega,m}}}\leq e^{\frac{C}{m}}\det\left(g^{\rm KE}_{\Omega}\right).

Hence, 1mlog(CmKΩ,μ~Ω,mB)\frac{1}{m}\log\left(C_{m}K_{\Omega,\widetilde{\mu}^{\rm B}_{\Omega,m}}\right) uniformly converges to log(det(gΩKE))\log\left(\det\left(g^{\rm KE}_{\Omega}\right)\right) at the rate 1m\frac{1}{m}.

For the above uniform estimate, we leverage the Tian-Yau-Zelditch expansion, as in the proof for the case of compact canonically polarized manifolds [tsuji2010, song2010]. A difference is that we adjust the normalizing factor in the definition of the dynamical system of the Bergman kernel to achieve better convergence speed, following Berndtsson’s idea in [berndtsson2009] (compare with the convergence rate ‘logm/m\log m/m’ in [song2010]).

The results of this paper can be extended to non-compact manifolds with bounded geometry, but we focus solely on domains here. This is because the domain theory itself is important, especially in terms of providing numerous concrete examples.

Our paper is organized as follows: In Section 2, we provide a brief overview of known results concerning weighted Bergman kernels and metrics. In Section 3, we introduce the concept of invariant weighted Bergman metrics. Section 4 presents a proof of the Tian-Yau-Zelditch expansion for domains, and in Section 5, we establish a proof of the uniform convergence of Tian and Tsuji’s weighted Bergman sequence on uniform squeezing domains.

2. Weighted Bergman kernel and metric

In this section, we briefly review well-known results for the classic Bergman kernel,metric, and their generalization to the weighted setting including the minimum integral method.

2.1. Weighted Bergman space, kernel, and metric

Let Ω\Omega be a domain in n\mathbb{C}^{n} and dλd\lambda be the Lebesgue measure of n\mathbb{C}^{n}. Consider a measure dμd\mu on Ω\Omega, which is defined by

dμ(z):=μ(z)dλ(z),d\mu(z):=\mu(z)d\lambda(z),

where μ(z)\mu(z) is a Lebesgue measurable positive real-valued function on Ω\Omega. The function μ\mu is called an weight function. Denote by L2(Ω,μ)L^{2}(\Omega,\mu) the space of all Lebesgue measurable complex-valued μ\mu-square integrable functions on Ω\Omega, i.e.,

L2(Ω,μ):={u:Ω|uΩ,μ2:=Ω|u|2dμ<}.L^{2}(\Omega,\mu):=\left\{u:\Omega\rightarrow\mathbb{C}\ \Big{|}\ \left\|u\right\|^{2}_{\Omega,\mu}:=\int_{\Omega}|u|^{2}d\mu<\infty\right\}.

Then L2(Ω,μ)L^{2}(\Omega,\mu) is a separable Hilbert space with respect to the inner product:

u,vΩ,μ:=Ωuv¯𝑑μ=Ωuv¯μ𝑑λ.\langle u,v\rangle_{\Omega,\mu}:=\int_{\Omega}u\overline{v}d\mu=\int_{\Omega}u\overline{v}\mu d\lambda.

Consider the linear subspace of L2(Ω,μ)L^{2}(\Omega,\mu), which is defined by

𝒜2(Ω,μ):=𝒪(Ω)L2(Ω,μ),\mathcal{A}^{2}(\Omega,\mu):=\mathcal{O}(\Omega)\cap L^{2}(\Omega,\mu),

where 𝒪(Ω)\mathcal{O}(\Omega) is the set of all holomorphic functions on Ω\Omega. The space 𝒜2(Ω,μ)\mathcal{A}^{2}(\Omega,\mu) is called the μ\mu-Bergman space or weighted Bergman space. Note that if 𝒜2(Ω,μ)\mathcal{A}^{2}(\Omega,\mu) is a closed subspace of L2(Ω,μ)L^{2}(\Omega,\mu), then 𝒜2(Ω,μ)\mathcal{A}^{2}(\Omega,\mu) is also a Hilbert space.

Definition 2.1.

An weight μ\mu on Ω\Omega is called an admissible weight function if 𝒜2(Ω,μ)\mathcal{A}^{2}(\Omega,\mu) is a closed subspace of L2(Ω,μ)L^{2}(\Omega,\mu) and for each fixed point pΩp\in\Omega, the functional

Φp:uu(p)\Phi_{p}:u\rightarrow u(p)

is a continuous linear functional on 𝒜2(Ω,μ)\mathcal{A}^{2}(\Omega,\mu).

The following criterion is useful to check the condition for the admissibility.

Theorem 2.2 (Pasternak-Winiarski [pasternak1990dependence, pasternak1992weights]).

An weight function μ\mu on Ω\Omega is admissible if and only if for any compact subset AΩA\subset\Omega, there is a constant CA>0C_{A}>0 such that for all u𝒜2(Ω,μ)u\in\mathcal{A}^{2}(\Omega,\mu), it satisfies the following Cauchy type inequality:

supzA|u(z)|CAuΩ,μ.\sup_{z\in A}|u(z)|\leq C_{A}\left\|u\right\|_{\Omega,\mu}.
Remark 2.3.

It is well-known that the characteristic function μ:=1Ω\mu:=1_{\Omega} is admissible. In this case, 𝒜2(Ω):=𝒜2(Ω,1Ω)\mathcal{A}^{2}(\Omega):=\mathcal{A}^{2}(\Omega,1_{\Omega}) is called the (classic or unweighted) Bergman space. In general, if μ1\mu^{-1} is locally integrable on Ω\Omega, then μ\mu is admissible (cf. Pasternak-Winiarski [pasternak1990dependence]). Therefore, every positive continuous function on Ω\Omega is admissible.

Note that if μ\mu is admissible, we can guarantee the existence of the kernel function applying the Riesz representation theorem, i.e., there exists the unique function KμK_{\mu} on Ω×Ω\Omega\times\Omega such that for all u𝒜2(Ω,μ)u\in\mathcal{A}^{2}(\Omega,\mu), it satisfies the reproducing property:

u(z)=ΩKμ(z,w)u(w)𝑑μ(w).u(z)=\int_{\Omega}K_{\mu}(z,w)u(w)d\mu(w).

In this case, the function KμK_{\mu} is called the μ\mu-Bergman function or weighted Bergman kernel function. Similarly to the classic Bergman kernel function, we have the following

Theorem 2.4 (Pasternak-Winiarski [pasternak1990dependence]).

For an admissible weight μ\mu, the weighted Bergman kernel function KμK_{\mu} satisfies the following properties.

  • (1)

    For any complete orthonormal basis {uν}\{u^{\nu}\} and compact subset KΩK\subset\Omega, the series

    νuν(z)uν(w)¯\sum_{\nu}u^{\nu}(z)\overline{u^{\nu}(w)}

    converges uniformly on K×KK\times K to the weighted Bergman kernel KμK_{\mu}.

  • (2)

    KμK_{\mu} is conjugate symmetric: Kμ(z,w)=Kμ(w,z)¯.K_{\mu}(z,w)=\overline{K_{\mu}(w,z)}.

  • (3)

    KμK_{\mu} is real analytic.

  • (4)

    KμK_{\mu} is the integral kernel of the orthogonal projector Pμ:L2(Ω,μ)𝒜2(Ω,μ)P_{\mu}:L^{2}(\Omega,\mu)\rightarrow\mathcal{A}^{2}(\Omega,\mu), i.e., for all uL2(Ω,μ)u\in L^{2}(\Omega,\mu), we have

    Pμ[u](z)=ΩKμ(z,w)u(w)𝑑μ(w).P_{\mu}[u](z)=\int_{\Omega}K_{\mu}(z,w)u(w)d\mu(w).
Remark 2.5.

If μ\mu is admissible, the μ\mu-Bergman space 𝒜2(Ω,μ)\mathcal{A}^{2}(\Omega,\mu) inherits the separability from L2(Ω,μ)L^{2}(\Omega,\mu) hence a complete orthonormal basis always exists.

From now on, we always assume that our weight function μ\mu is admissible on Ω\Omega. Then we can define a symmetric tensor

gΩ,μ:=j,k=1ngjk¯(z)dzjdzk¯=j,k=1n2logKΩ,μ(z,z)zjzk¯dzjdzk¯g_{\Omega,\mu}:=\sum^{n}_{j,k=1}g_{j\overline{k}}(z)dz_{j}\otimes d\overline{z_{k}}=\sum^{n}_{j,k=1}\frac{\partial^{2}\log K_{\Omega,\mu}(z,z)}{\partial z_{j}\partial\overline{z_{k}}}dz_{j}\otimes d\overline{z_{k}}

for all zΩz\in\Omega satisfying KΩ,μ(z,z)>0K_{\Omega,\mu}(z,z)>0. We will call this the μ\mu-Bergman metric or weighted Bergman (psuedo)-metric of Ω\Omega. If gΩ,μ(z)g_{\Omega,\mu}(z) is positive-definite for all zΩz\in\Omega, we say that (Ω,μ)(\Omega,\mu) admits the weighted Bergman metric. In this case, the weighted Bergman metric is a (real-analytic) Kähler metric. Denote the corresponding Bergman length square of X=j=1nXjzj|z=pTpΩnX=\sum_{j=1}^{n}X_{j}\frac{\partial}{\partial z_{j}}\big{|}_{z=p}\in T_{p}\Omega\cong\mathbb{C}^{n} at pΩp\in\Omega by

gΩ,μ(p;X):=j,k=1ngjk¯(p)XjXk¯.g_{\Omega,\mu}(p;X):=\sum\limits_{j,k=1}^{n}g_{j\overline{k}}(p)X_{j}\overline{X_{k}}.

The holomorphic sectional curvature of gΩ,μg_{\Omega,\mu} at pp along XnX\in\mathbb{C}^{n} is

HΩ,μ(p;X):=(i,j,k,l=1nRij¯kl¯(p)XiXj¯XkXl¯)(j,k=1ngjk¯(p)XjXk¯)2,H_{\Omega,\mu}(p;X):=\left(\sum\limits_{i,j,k,l=1}^{n}R_{i\overline{j}k\overline{l}}(p)X_{i}\overline{X_{j}}X_{k}\overline{X_{l}}\right)\left(\sum\limits_{j,k=1}^{n}g_{j\overline{k}}(p)X_{j}\overline{X_{k}}\right)^{-2},

where Rij¯kl¯=2gij¯zkzl¯+gαβ¯giβ¯zkgαj¯zl¯R_{i\overline{j}k\overline{l}}=-\frac{\partial^{2}g_{i\overline{j}}}{\partial z_{k}\partial\overline{z_{l}}}+g^{\alpha\overline{\beta}}\frac{\partial g_{i\overline{\beta}}}{\partial z_{k}}\frac{\partial g_{\alpha\overline{j}}}{\partial\overline{z_{l}}} are coefficients of the curvature tensor of gΩ,μg_{\Omega,\mu}.

2.2. Bergman’s special basis and minimum integrals

Now we discuss a special way to construct a complete orthonormal basis of 𝒜2(Ω,μ)\mathcal{A}^{2}(\Omega,\mu), which were developed by Bergman (in the unweighted case). Suppose that μ\mu is an admissible weight on Ω\Omega. Let (z1,,zn)(z_{1},\ldots,z_{n}) be the standard Euclidean coordinates for n\mathbb{C}^{n}.

For the multi-indices α=(α1,,αn)n\alpha=(\alpha_{1},\ldots,\alpha_{n})\in\mathbb{N}^{n} with |α|:=i=1nαi|\alpha|:=\sum_{i=1}^{n}\alpha_{i} , we will denote the holomorphic derivatives of a function uu by

Dαu:=Dzαu=|α|z1α1znαnu.D^{\alpha}u:=D^{\alpha}_{z}u=\frac{\partial^{|\alpha|}}{\partial z_{1}^{\alpha_{1}}\cdots\partial z_{n}^{\alpha_{n}}}u.
Definition 2.6 (Bergman’s special basis).

Fix a point pΩp\in\Omega. We say that a complete orthonormal basis {uα}\{u^{\alpha}\} for 𝒜2(Ω,μ)\mathcal{A}^{2}(\Omega,\mu) is special at pp if it satisfies that Dαuα(p)0D^{\alpha}u^{\alpha}(p)\neq 0 and Dβuα(p)=0D^{\beta}u^{\alpha}(p)=0 if β<α\beta<\alpha, where the order for multi-indices is given by the lexicographic order.

For each multi-index α\alpha, define the subsets of 𝒜μ2(Ω):=𝒜2(Ω,μ)\mathcal{A}^{2}_{\mu}(\Omega):=\mathcal{A}^{2}(\Omega,\mu) by

μα(Ω):={u𝒜μ2(Ω):Dαu(p)=1,Dβu(p)=0ifβ<α}\mathcal{E}^{\alpha}_{\mu}(\Omega):=\left\{u\in\mathcal{A}^{2}_{\mu}(\Omega):D^{\alpha}u(p)=1,\ D^{\beta}u(p)=0\ \ {\rm if\ }\beta<\alpha\right\}

If the set μα(Ω)\mathcal{E}^{\alpha}_{\mu}(\Omega) is non-empty, one can check that there exists a L2L^{2}-minimal element vαv^{\alpha} in μα(Ω)\mathcal{E}^{\alpha}_{\mu}(\Omega). Then {uα}\{u^{\alpha}\} with uα:=vα/vαΩ,μu^{\alpha}:=v^{\alpha}/\left\|v^{\alpha}\right\|_{\Omega,\mu} is a complete orthonormal basis for 𝒜μ2(Ω)\mathcal{A}^{2}_{\mu}(\Omega), which is special at the point pp.

Proposition 2.7.

If Ω\Omega is a bounded domain with an admissible weight μL1(Ω)\mu\in L^{1}(\Omega), then the diagonal part of weighted Bergman kernel is positive everywhere, and the weighted Bergman metric is positive-definite everywhere.

Proof.

For the multi-indices α=(α1,,αn)n\alpha=(\alpha_{1},\ldots,\alpha_{n})\in\mathbb{N}^{n} , denote the monomials (including the constant function 1=z𝟎1=z^{\bf 0} for 𝟎:=(0,,0){\bf 0}:=(0,\ldots,0)) by

zα:=z1α1znαn.z^{\alpha}:=z_{1}^{\alpha_{1}}\cdots z_{n}^{\alpha_{n}}.

Since Ω\Omega is bounded and μL1(Ω)\mu\in L^{1}(\Omega), zαμα(Ω)z^{\alpha}\in\mathcal{E}^{\alpha}_{\mu}(\Omega) for all α\alpha. Then, the subsets μα(Ω)\mathcal{E}^{\alpha}_{\mu}(\Omega) are non-empty for any point pΩp\in\Omega so that there exists a special basis {uα}\{u^{\alpha}\} at pp. The conclusion follows from the following equalities.

KΩ,μ(p):=KΩ,μ(p,p)=|u𝟎(p)|2=|v𝟎(p)|2/v𝟎Ω,μ2,K_{\Omega,\mu}(p):=K_{\Omega,\mu}(p,p)=\left|u^{\bf 0}(p)\right|^{2}=\left|v^{\bf 0}(p)\right|^{2}/\left\|v^{\bf 0}\right\|^{2}_{\Omega,\mu},
2zjzk¯logKΩ,μ(z)|p=|u𝟎(p)|2|α|=1uαzj(p)uαzk(p)¯.\left.\frac{\partial^{2}}{\partial z_{j}\partial\overline{z_{k}}}\log K_{\Omega,\mu}(z)\right|_{p}=\left|u^{\bf 0}(p)\right|^{-2}\sum_{|\alpha|=1}\frac{\partial u^{\alpha}}{\partial z_{j}}(p)\overline{\frac{\partial u^{\alpha}}{\partial z_{k}}(p)}.

Definition 2.8.

Fix a point pΩp\in\Omega and a nonzero vector XnX\in\mathbb{C}^{n}. Define subsets:

μ0(Ω)\displaystyle\mathcal{E}^{0}_{\mu}(\Omega) :={u𝒜μ2(Ω):u(p)=1}=μ𝟎(Ω),\displaystyle:=\left\{u\in\mathcal{A}^{2}_{\mu}(\Omega):u(p)=1\right\}=\mathcal{E}^{\bf 0}_{\mu}(\Omega),
μ1(Ω)\displaystyle\mathcal{E}^{1}_{\mu}(\Omega) :={u𝒜μ2(Ω):u(p)=0,DXu(p)=1},\displaystyle:=\left\{u\in\mathcal{A}^{2}_{\mu}(\Omega):u(p)=0,\ D_{X}u(p)=1\right\},
μ2(Ω)\displaystyle\mathcal{E}^{2}_{\mu}(\Omega) :={u𝒜μ2(Ω):u(p)=0,du(p)=0,DXDXu(p)=1},\displaystyle:=\left\{u\in\mathcal{A}^{2}_{\mu}(\Omega):u(p)=0,\ du(p)=0,\ D_{X}D_{X}u(p)=1\right\},

where DXD_{X} is the directional derivative along XX.

Definition 2.9.

Minimum integrals of the weighted Bergman kernels are defined by

IΩ,μ0(p):=infuμ0(Ω)uΩ,μ2,IΩ,μ1(p;X):=infuμ1(Ω)uΩ,μ2,IΩ,μ2(p;X):=infuμ2(Ω)uΩ,μ2.I^{0}_{\Omega,\mu}(p):=\inf_{u\in\mathcal{E}^{0}_{\mu}(\Omega)}\|u\|_{\Omega,\mu}^{2},\ \ \ \ I^{1}_{\Omega,\mu}(p;X):=\inf_{u\in\mathcal{E}^{1}_{\mu}(\Omega)}\|u\|_{\Omega,\mu}^{2},\ \ \ \ I^{2}_{\Omega,\mu}(p;X):=\inf_{u\in\mathcal{E}^{2}_{\mu}(\Omega)}\|u\|_{\Omega,\mu}^{2}.

For simplicity, we will use the following notations from now on.

Iμ0:=IΩ,μ0(p),Iμ1:=IΩ,μ1(p;X),Iμ2:=IΩ,μ2(p;X).I^{0}_{\mu}:=I^{0}_{\Omega,\mu}(p),\ \ \ \ I^{1}_{\mu}:=I^{1}_{\Omega,\mu}(p;X),\ \ \ \ I^{2}_{\mu}:=I^{2}_{\Omega,\mu}(p;X).
DX0u:=u,DX1u:=DXu,DX2u:=DXDXu.D_{X}^{0}u:=u,\ \ \ \ D_{X}^{1}u:=D_{X}u,\ \ \ \ D_{X}^{2}u:=D_{X}D_{X}u.

We can generalize the Bergman-Fuks formula to the weighted cases, using the same proof. For reader’s convenience, we briefly sketch the proof here.

Theorem 2.10 (Bergman-Fuks formula).
KΩ,μ(p)=1Iμ0,gΩ,μ(p;X)=Iμ0Iμ1,HΩ,μ(p;X)=2(Iμ1)2Iμ2Iμ0.K_{\Omega,\mu}(p)=\frac{1}{I^{0}_{\mu}},\ \ \ \ g_{\Omega,\mu}(p;X)=\frac{I^{0}_{\mu}}{I^{1}_{\mu}},\ \ \ \ H_{\Omega,\mu}(p;X)=2-\frac{(I^{1}_{\mu})^{2}}{I^{2}_{\mu}I^{0}_{\mu}}.
Proof.

For j=0,1,2j=0,1,2, denote the minimizer of the set μj(Ω)\mathcal{E}^{j}_{\mu}(\Omega) by vjv^{j}. Consider an orthonormal basis of 𝒜μ2(Ω)\mathcal{A}^{2}_{\mu}(\Omega) including uj:=vj/vjΩ,μu^{j}:=v^{j}/\|v^{j}\|_{\Omega,\mu}. Note that uk(p)=0u^{k}(p)=0 if k>0k>0, and duk(p)=0du^{k}(p)=0 if k>1k>1. Let K:=KΩ,μK:=K_{\Omega,\mu}, KX:=DXKΩ,μK_{X}:=D_{X}K_{\Omega,\mu}, and KX¯:=DX¯KΩ,μK_{\overline{X}}:=\overline{D_{X}}K_{\Omega,\mu}. Similarly, for the directional derivatives of any function, use sub-indices. Then

K(p)=|u0(p)|2,KX(p)=uX0(p)u0(p)¯,KXX¯(p)=|uX0(p)|2+|uX1(p)|2.K(p)=|u^{0}(p)|^{2},\ \ K_{X}(p)=u_{X}^{0}(p)\cdot\overline{u^{0}(p)},\ \ K_{X\overline{X}}(p)=|u^{0}_{X}(p)|^{2}+|u^{1}_{X}(p)|^{2}.

Therefore, we have

g(p;X):=gΩ,μ(p;X)=KXX¯(p)K(p)KX(p)KX¯(p)K2(p)=|uX1(p)|2|u0(p)|2.g(p;X):=g_{\Omega,\mu}(p;X)=\frac{K_{X\overline{X}}(p)K(p)-K_{X}(p)K_{\overline{X}}(p)}{K^{2}(p)}=\frac{|u^{1}_{X}(p)|^{2}}{|u^{0}(p)|^{2}}.

Similarly, a long and tedious computation shows that HΩ,μ(p;X)=2|u0(p)|2|uXX2(p)|2|uX1(p)|4.H_{\Omega,\mu}(p;X)=2-\frac{|u^{0}(p)|^{2}|u^{2}_{XX}(p)|^{2}}{|u^{1}_{X}(p)|^{4}}. Then, the conclusion follows from |DXjuj(p)|2=1/vjΩ,μ2=1/Iμj|D_{X}^{j}u^{j}(p)|^{2}=1/\|v^{j}\|_{\Omega,\mu}^{2}=1/I^{j}_{\mu}. ∎

2.3. The weighted Bergman kernel and metric of the ball

In [forelli1974projections], Forelli and Rudin computed the Bergman kernel function of the unit ball 𝔹n(1)\mathbb{B}^{n}(1) in n\mathbb{C}^{n} with the admissible weight (1|z|2)m(1-|z|^{2})^{m}. Using the transformation formula (3.1), one can obtain the weighted Bergman kernel and metric of a dilated ball. Let 𝔹n(r)\mathbb{B}^{n}(r) be the ball in n\mathbb{C}^{n} with the radius rr centered at the origin. Then, for the admissible weight μ(w):=(r2|w|2r2)m\mu(w):=\left(\frac{r^{2}-|w|^{2}}{r^{2}}\right)^{m},

K𝔹n(r),μ(w)=1cm(r)(r2r2|w|2)n+m+1,K_{\mathbb{B}^{n}(r),\mu}(w)=\frac{1}{c_{m}(r)}\left(\frac{r^{2}}{r^{2}-|w|^{2}}\right)^{n+m+1},

where cm(r)c_{m}(r) denotes the weighted volume of the ball 𝔹n(r)\mathbb{B}^{n}(r) for integers m0m\geq 0:

cm(r):=𝔹n(r)(r2|w|2r2)m𝑑λ(w)=(πr2)nΓ(m+1)Γ(n+m+1)=(πr2)nm!(n+m)!.c_{m}(r):=\int_{\mathbb{B}^{n}(r)}\Big{(}\frac{r^{2}-|w|^{2}}{r^{2}}\Big{)}^{m}d\lambda(w)=(\pi r^{2})^{n}\frac{\Gamma(m+1)}{\Gamma(n+m+1)}=(\pi r^{2})^{n}\frac{m!}{(n+m)!}.

The corresponding weighted Bergman metric is given by

g𝔹n(r),μ=j,k=1n(n+m+1)(r2|w|2)δjk¯+wj¯wk(r2|w|2)2dwjdwk¯.g_{\mathbb{B}^{n}(r),\mu}=\sum\limits_{j,k=1}^{n}(n+m+1)\frac{(r^{2}-|w|^{2})\delta_{j\overline{k}}+\overline{w_{j}}w_{k}}{(r^{2}-|w|^{2})^{2}}dw_{j}\otimes d\overline{w_{k}}.

The holomorphic sectional curvatures of g𝔹n(r),μg_{\mathbb{B}^{n}(r),\mu} are constant H𝔹n(r),μ(w)=2n+m+1.H_{\mathbb{B}^{n}(r),\mu}(w)=-\frac{2}{n+m+1}.

3. Invariance of the weighted Bergman kernel and metric

3.1. Transformation formula for the weighted Bergman kernels

It is well-known that the classic Bergman kernel function satisfies the transformation formula for biholomorphisms. In the weighted case, we can generalize it as follows.

Proposition 3.1.

Let Ω\Omega and Ω\Omega^{\prime} be domains in n\mathbb{C}^{n} and F:ΩΩF:\Omega\rightarrow\Omega^{\prime} be a biholomorphism. Let μ\mu be an admissible weight function of Ω\Omega.

  1. (1)

    Let hh be a non-vanishing holomorphic function on Ω\Omega. Then the function

    μ(F(z)):=|h(z)|2μ(z)\mu^{\prime}(F(z)):=|h(z)|^{2}\mu(z)

    is an admissible weight function on Ω\Omega^{\prime}.

  2. (2)

    In the above case, we have the following transformation formula:

    (3.1) KΩ,μ(z,w)=𝒥(F(z))h(z)KΩ,μ(F(z),F(w))𝒥(F(w))h(w)¯,K_{\Omega,\mu}(z,w)=\mathcal{J}(F(z))h(z)\cdot K_{\Omega^{\prime},\mu^{\prime}}(F(z),F(w))\cdot\overline{\mathcal{J}(F(w))h(w)},

    where 𝒥(F):=detJF\mathcal{J}(F):=\det J_{\mathbb{C}}F is the determinant of complex Jacobian of FF.

Proof.
  1. (1)

    Let uu be a function in 𝒜2(Ω,μ)\mathcal{A}^{2}(\Omega^{\prime},\mu^{\prime}). Set ζ=F(z)\zeta=F(z). Then we have

    uΩ,μ2:=Ω|u(ζ)|2μ(ζ)𝑑λ(ζ)=Ω|u(F(z))h(z)𝒥(F(z))|2μ(z)𝑑λ(z).\left\|u\right\|^{2}_{\Omega^{\prime},\mu^{\prime}}:=\int_{\Omega^{\prime}}|u(\zeta)|^{2}\mu^{\prime}(\zeta)d\lambda(\zeta)=\int_{\Omega}\left|u(F(z))h(z)\mathcal{J}(F(z))\right|^{2}\mu(z)d\lambda(z).

    This implies that

    v(z):=u(F(z))h(z)𝒥(F(z))𝒜2(Ω,μ).v(z):=u(F(z))h(z)\mathcal{J}(F(z))\in\mathcal{A}^{2}(\Omega,\mu).

    Let AA^{\prime} be a compact subset of Ω\Omega^{\prime} so that A:=F1(A)A:=F^{-1}(A^{\prime}) is a compact subset of Ω\Omega. Since μ\mu is admissible, there is a constant CA>0C_{A}>0 such that

    supzA|v(z)|CAvΩ,μ.\sup_{z\in A}|v(z)|\leq C_{A}\left\|v\right\|_{\Omega,\mu}.

    Choose CA:=CAsupzA|1h(z)𝒥(F(z))|>0C_{A^{\prime}}:=C_{A}\sup\limits_{z\in A}\left|\frac{1}{h(z)\mathcal{J}(F(z))}\right|>0. Then we have

    supζA|u(ζ)|=supzA|v(z)1h(z)𝒥(F(z))|CAuΩ,μ.\sup_{\zeta\in A^{\prime}}|u(\zeta)|=\sup_{z\in A}\left|v(z)\frac{1}{h(z)\mathcal{J}(F(z))}\right|\leq C_{A^{\prime}}\left\|u\right\|_{\Omega^{\prime},\mu^{\prime}}.

    By Theorem 2.2, μ\mu^{\prime} is an admissible weight function of Ω\Omega^{\prime}.

  2. (2)

    Let ξ=F(w)\xi=F(w). For any function u𝒜2(Ω,μ)u\in\mathcal{A}^{2}(\Omega,\mu), we have

    Ω𝒥(F(z))h(z)KΩ,μ(F(z),F(w))𝒥(F(w))h(w)¯u(w)μ(w)𝑑λ(w)\displaystyle\int_{\Omega}\mathcal{J}(F(z))h(z)K_{\Omega^{\prime},\mu^{\prime}}(F(z),F(w))\overline{\mathcal{J}(F(w))h(w)}u(w)\mu(w)d\lambda(w)
    =\displaystyle= Ω𝒥(F(z))h(z)KΩ,μ(F(z),ξ)𝒥(F(w))h(F1(ξ))¯u(F1(ξ))μ(ξ)dλ(ξ)|h(F1(ξ))𝒥(F(w))|2\displaystyle\int_{\Omega^{\prime}}\mathcal{J}(F(z))h(z)K_{\Omega^{\prime},\mu^{\prime}}(F(z),\xi)\overline{\mathcal{J}(F(w))h(F^{-1}(\xi))}u(F^{-1}(\xi))\frac{\mu^{\prime}(\xi)d\lambda(\xi)}{\left|h(F^{-1}(\xi))\mathcal{J}(F(w))\right|^{2}}
    =\displaystyle= Ω𝒥(F(z))h(z)KΩ,μ(F(z),ξ)(𝒥(F(w))h(F1(ξ)))1u(F1(ξ))𝑑μ(ξ)\displaystyle\int_{\Omega^{\prime}}\mathcal{J}(F(z))h(z)K_{\Omega^{\prime},\mu^{\prime}}(F(z),\xi)\left(\mathcal{J}\left(F\left(w\right)\right)h\left(F^{-1}\left(\xi\right)\right)\right)^{-1}u(F^{-1}(\xi))d\mu^{\prime}(\xi)
    =\displaystyle= 𝒥(F(z))h(z)(𝒥(F(z))h(z))1u(F1(F(z)))=u(z)\displaystyle\mathcal{J}(F(z))h(z)\left(\mathcal{J}(F(z))h(z)\right)^{-1}u(F^{-1}(F(z)))=u(z)

    The uniqueness of the kernel function implies the equation (3.1).

Remark 3.2.

The above proposition is a generalization of Lemma 1 in [dragomir1994weighted] when h1h\equiv 1.

3.2. Invariant weighted Bergman metrics

Recall that in the unweighted case (μΩ=1Ω\mu_{\Omega}=1_{\Omega}), the definitions of the classic Bergman kernel and metric depend only on the geometry of domain Ω\Omega. One of the most important properties of the classic Bergman metric is that it is invariant under biholomorphisms, in the sense that for any biholomorphism FF, we have

gΩ,1Ω=FgF(Ω),1F(Ω).g_{\Omega,1_{\Omega}}=F^{*}g_{F(\Omega),1_{F(\Omega)}}.

Since this invariance comes from the transformation formula of Bergman kernels under biholomorphisms, we need to check the following weighted version, which is a generalization of Theorem 1 in [dragomir1994weighted] when h1h\equiv 1.

Proposition 3.3.

Suppose that (Ω,μ)(\Omega,\mu) and (Ω,μ)(\Omega^{\prime},\mu^{\prime}) both admit the weighted Bergman metrics gΩ,μg_{\Omega,\mu} and gΩ,μg_{\Omega^{\prime},\mu^{\prime}} respectively. Let F:ΩΩF:\Omega\rightarrow\Omega^{\prime} be a biholomorphism. If μ\mu and μ\mu^{\prime} satisfy the relation:

μ(F(z))=|h(z)|2μ(z)\mu^{\prime}(F(z))=|h(z)|^{2}\mu(z)

for some non-vanishing holomorphic function hh of Ω\Omega, then FF is an isometry with respect to the weighted Bergman metrics, i.e.,

gΩ,μ=FgΩ,μ.g_{\Omega,\mu}=F^{*}g_{\Omega^{\prime},\mu^{\prime}}.
Proof.

The transformation formula (3.1) for the weighted case implies that

KΩ,μ(z)=KΩ,μ(F(z))|𝒥(F(z))h(z)|2.K_{\Omega,\mu}(z)=K_{\Omega^{\prime},\mu^{\prime}}(F(z))|\mathcal{J}(F(z))h(z)|^{2}.

By taking logarithm and mixed derivatives to both sides, we can show that

2zjzk¯logKΩ,μ(z)=l,m=1n2ζlζm¯logKΩ,μ(F(z))Fl(z)zj(Fm(z)zk)¯.\frac{\partial^{2}}{\partial z_{j}\partial\overline{z_{k}}}\log K_{\Omega,\mu}(z)=\sum^{n}_{l,m=1}\frac{\partial^{2}}{\partial\zeta_{l}\partial\overline{\zeta_{m}}}\log K_{\Omega^{\prime},\mu^{\prime}}(F(z))\frac{\partial F_{l}(z)}{\partial z_{j}}\overline{\left(\frac{\partial F_{m}(z)}{\partial z_{k}}\right)}.

This implies that for any non-zero vector XnX\in\mathbb{C}^{n}, we have

gΩ,μ(z,X)=gΩ,μ(F(z),dF(X)),g_{\Omega,\mu}(z,X)=g_{\Omega^{\prime},\mu^{\prime}}(F(z),dF(X)),

as we required. ∎

Definition 3.4.

Let 𝒟\mathcal{D} be a collection of domains in n\mathbb{C}^{n} and \mathcal{M} be an assignment of an admissible weight function μΩ:=(Ω)\mu_{\Omega}:=\mathcal{M}(\Omega) to each domain Ω𝒟\Omega\in\mathcal{D}. We will call the weighted Bergman kernel KΩ,μΩK_{\Omega,\mu_{\Omega}} \mathcal{M}-Bergman kernel” of Ω\Omega.
Suppose that for any domain Ω𝒟\Omega\in\mathcal{D}, gμΩg_{\mu_{\Omega}} is positive-definite. In this case, we will call the weighted Bergman metric gΩ,μΩg_{\Omega,\mu_{\Omega}} \mathcal{M}-Bergman metric” of Ω\Omega.

Definition 3.5.

We say that the assignment \mathcal{M} is invariant (under biholomorphisms) if for any biholomorphism FF of Ω𝒟\Omega\in\mathcal{D}, it satisfies that

μF(Ω)F=|hF|2μΩ\mu_{F(\Omega)}\circ F=|h_{F}|^{2}\mu_{\Omega}

for some non-vanishing holomorphic function hFh_{F} (depending on FF) of Ω\Omega. We say that an invariant assignment \mathcal{M} is canonical (of level m+m\in\mathbb{N}^{+}) if it satisfies that

μF(Ω)F=|𝒥(F)|2(m1)μΩ.\mu_{F(\Omega)}\circ F=|\mathcal{J}(F)|^{2(m-1)}\mu_{\Omega}.
Remark 3.6.

\mathcal{M} and hFh_{F} can be considered as a hermitian metric and the transition function of a trivial line bundle LL over Ω\Omega, respectively. In the case that LL is the canonical line bundle, the transition function is hF=𝒥(F)h_{F}=\mathcal{J}(F).

As a corollary of Proposition 3.1 and Proposition 3.3, we obtain the following

Theorem 3.7.

If an assignment \mathcal{M} is invariant, the \mathcal{M}-Bergman metric is invariant under biholomorphisms, i.e.,

gΩ,μΩ=FgF(Ω),μF(Ω).g_{\Omega,\mu_{\Omega}}=F^{*}g_{F(\Omega),\mu_{F(\Omega)}}.

If an invariant assignment \mathcal{M} is canonical of level mm, the \mathcal{M}-Bergman kernel satisfies the following transformation formula:

KΩ,μΩ=KF(Ω),μF(Ω)|𝒥(F)|2mK_{\Omega,\mu_{\Omega}}=K_{F(\Omega),\mu_{F(\Omega)}}|\mathcal{J}(F)|^{2m}

so that the corresponding volume form is biholomorphically invariant:

KΩ,μΩ1m(z)λ(z)=KF(Ω),μF(Ω)1m(w)λ(w),K_{\Omega,\mu_{\Omega}}^{\frac{1}{m}}(z)\lambda(z)=K_{F(\Omega),\mu_{F(\Omega)}}^{\frac{1}{m}}(w)\lambda(w),

where w=F(z)w=F(z).

Remark 3.8.

If \mathcal{M} is a canonical invariant assignment of level mm, we will call the normalized function KΩ,μΩ1mK_{\Omega,\mu_{\Omega}}^{\frac{1}{m}}\mathcal{M}-normalized Bergman kernel” of Ω\Omega. The level mm means the tensor power of the canonical line bundle.

Example 3.9.

Let 𝒟bp\mathcal{D}^{\rm bp} be a collection of bounded pseudoconvex domains in n\mathbb{C}^{n}. By the famous work of Cheng-Yau[cheng1980existence] and Mok-Yau[mok1983completeness], every Ω𝒟bp\Omega\in\mathcal{D}^{\rm bp} admits unique complete Kähler-Einstein metric gΩKEg^{\rm KE}_{\Omega}. Define an admissible assignment KE\mathcal{M}^{\rm KE} by

KE(Ω):=eφΩKE=1det(gΩKE).\mathcal{M}^{\rm KE}(\Omega):=e^{-\varphi^{\rm KE}_{\Omega}}=\frac{1}{\det\left(g^{\rm KE}_{\Omega}\right)}.

By the uniqueness of the Kähler-Einstein metric and the volume form:

det(gΩKE(z))=det(gF(Ω)KE(F(z)))|𝒥(F(z))|2,\det\left(g^{\rm KE}_{\Omega}(z)\right)=\det\left(g^{\rm KE}_{F(\Omega)}(F(z))\right)|\mathcal{J}(F(z))|^{2},

Proposition 3.3 and Theorem 3.7 imply that KE\mathcal{M}^{\rm KE}-Bergman metric is invariant under biholomorphisms.

Example 3.10.

Let Ω\Omega be a bounded domain in n\mathbb{C}^{n}. Then the diagonal part of the classic Bergman kernel function KΩK_{\Omega} is a positive smooth strictly plurisubharmonic function. Let 𝒟b\mathcal{D}^{\rm b} be a collection of bounded domains in n\mathbb{C}^{n}. Define an admissible assignment B\mathcal{M}^{\rm B} by

B(Ω):=1KΩ.\mathcal{M}^{\rm B}(\Omega):=\frac{1}{K_{\Omega}}.

By the transformation formula for Bergman kernels:

KΩ(z,z)=KF(Ω)(F(z),F(z))|𝒥(F(z))|2,K_{\Omega}(z,z)=K_{F(\Omega)}(F(z),F(z))|\mathcal{J}(F(z))|^{2},

Proposition 3.3 and Theorem 3.7 imply that B\mathcal{M}^{\rm B}-Bergman metric is invariant under biholomorphisms.

4. Estimates of the weighted Bergman kernels and metrics

In [tian1990], Tian constructed an asymptotic expansion of the sequence of the weighted Bergman kernels and metrics for canonically polarized manifolds. Later, this result is improved by Ruan, Zelditch, Lu, and so on [ruan1998canonical, zelditch1998szego, lu2000lower]. This is called the Tian-Yau-Zelditch expansion. For the proof, Tian used the “peaked section” method, based on the standard ¯\overline{\partial}-estimates for complete manifolds.

For domains in n\mathbb{C}^{n}, this method can be understood in terms of the weighted version of the Bergman minimum integral method and Hörmander’s classic L2¯L^{2}-\overline{\partial} theorem for domains. In this section, we present a straightforward proof of a domain version of the Tian-Yau-Zelditch expansion for the weighted Bergman kernels, metrics, and curvatures.

4.1. Estimates of the minimum integrals

Let Ω\Omega be a pseudoconvex domain in n\mathbb{C}^{n} and φ\varphi be a smooth strictly plurisubharmonic function on Ω\Omega. Choose an weight function μ:=eφ\mu:=e^{-\varphi}. Fix a point pΩp\in\Omega and a vector XnX\in\mathbb{C}^{n}. Consider the minimum integrals of the weighted Bergman kernels and denote the minimizers of μj(Ω)𝒜μ2(Ω)\mathcal{E}^{j}_{\mu}(\Omega)\subset\mathcal{A}^{2}_{\mu}(\Omega) by vjv^{j}, i.e.,

Iμj:=infuμj(Ω)uΩ,μ2=vjΩ,μ2.I^{j}_{\mu}:=\inf_{u\in\mathcal{E}^{j}_{\mu}(\Omega)}\|u\|_{\Omega,\mu}^{2}=\|v^{j}\|^{2}_{\Omega,\mu}.

Since it is hard to obtain the explicit minimizer vjμj(Ω)v^{j}\in\mathcal{E}^{j}_{\mu}(\Omega) in general, we will approximate it by another function via the following slight variant of Hörmander’s theorem to construct a holomorphic function in μj(Ω)\mathcal{E}^{j}_{\mu}(\Omega).

Theorem 4.1 (Theorem 5 in [gallagher2017dimension]).

Let Ω\Omega be a pseudoconvex domain in n\mathbb{C}^{n}, and let ψ\psi be a plurisubharmonic function in Ω\Omega. Suppose that gg is a ¯\overline{\partial}-closed CC^{\infty}-smooth (0,1)(0,1)-form which has a compact support on some open subset UΩU\subset\Omega. If there exists a positive constant CC such that

ψ(z)C|z|2\psi(z)-C|z|^{2}

is plurisubharmonic on UU, then there exists a CC^{\infty}-smooth function uL2(Ω,eψ)u\in L^{2}(\Omega,e^{-\psi}) solving ¯u=g\overline{\partial}u=g and satisfying

Ω|u|2eψ𝑑λ1CU|g|2eψ𝑑λ.\int_{\Omega}|u|^{2}e^{-\psi}d\lambda\leq\frac{1}{C}\int_{U}|g|^{2}e^{-\psi}d\lambda.

Using Theorem 4.1 with a similar technique used in the localization lemma of minimum integrals (cf. [greene2006function, kim1996boundary]), we can construct an explicit holomorphic function in μj(Ω)\mathcal{E}^{j}_{\mu}(\Omega) as follows:

Let U,VU,V be bounded neighborhoods of pp satisfying VUΩV\subset\subset U\subset\subset\Omega. Choose a cut-off function χUCc(U)\chi_{U}\in C^{\infty}_{c}(U) such that χU=1\chi_{U}=1 on VV and 0χU10\leq\chi_{U}\leq 1 on UU. Let jj^{\prime} be any integer j>jj^{\prime}>j. Consider an upper-bounded function ξ\xi on Ω\Omega satisfying eξCΩe^{\xi}\leq C_{\Omega} and

(4.1) eξ(z):=O(|zp|2(n+j))nearp.e^{\xi(z)}:=O\left(|z-p|^{2(n+j^{\prime})}\right)\ \ \ {\rm near\ }p.

By the boundedness of UU and VV, we may assume that on U\VU\backslash V,

|¯χU|2eξCU\V|\overline{\partial}\chi_{U}|^{2}e^{-\xi}\leq C_{U\backslash V}

for some positive constant CU\V>0C_{U\backslash V}>0.

Proposition 4.2.

Suppose that ψ:=φ+ξ\psi:=\varphi+\xi is plurisubharmonic on Ω\Omega, and

ψ(z)CU|z|2\psi(z)-C_{U}|z|^{2}

is plurisubharmonic on UU for some positive constant CUC_{U}. Let ηjμj(W)\eta^{j}\in\mathcal{E}^{j}_{\mu}(W) be a function on some neighborhood WW of pp satisfying UWΩU\subset W\subset\Omega. Then, there exists a holomorphic function u^jμj(Ω)\widehat{u}^{j}\in\mathcal{E}^{j}_{\mu}(\Omega) satisfying

u^jχUηjΩ,μ2CηjU\V,μ2,\|\widehat{u}^{j}-\chi_{U}\cdot\eta^{j}\|^{2}_{\Omega,\mu}\leq C\|\eta^{j}\|^{2}_{U\backslash V,\mu},

where CC depends only on CΩ,CU\V,CUC_{\Omega},C_{U\backslash V},C_{U}.

Proof.

Apply Theorem 4.1 with a ¯\overline{\partial}-closed CC^{\infty}-smooth (0,1)(0,1)-form gg, defined by

g:=¯(χUηj).g:=\overline{\partial}\left(\chi_{U}\cdot\eta^{j}\right).

Then there exists a CC^{\infty}-smooth function ujL2(Ω,eψ)u^{j}\in L^{2}(\Omega,e^{-\psi}) solving ¯u=g\overline{\partial}u=g and satisfying

(4.2) Ω|uj|2eψ𝑑λ1CUU\V|¯χU|2|ηj|2eψ𝑑λCU\VCUU\V|ηj|2eφ𝑑λ.\int_{\Omega}|u^{j}|^{2}e^{-\psi}d\lambda\leq\frac{1}{C_{U}}\int_{U\backslash V}|\overline{\partial}\chi_{U}|^{2}|\eta^{j}|^{2}e^{-\psi}d\lambda\leq\frac{C_{U\backslash V}}{C_{U}}\int_{U\backslash V}|\eta^{j}|^{2}e^{-\varphi}d\lambda.

Since the right hand side integral is finite by assumption, the left hand side integral

Ω|uj|2eψ𝑑λ=Ω|uj|2eξeφ𝑑λ\int_{\Omega}|u^{j}|^{2}e^{-\psi}d\lambda=\int_{\Omega}\frac{|u^{j}|^{2}}{e^{\xi}}e^{-\varphi}d\lambda

is also finite. This with the property (4.1) implies that Dαuj(p)=0,D^{\alpha}u^{j}(p)=0, for all α=(α1,,αn)\alpha=(\alpha_{1},\ldots,\alpha_{n}) with |α|j|\alpha|\leq j. Then the function u^j\widehat{u}^{j}, defined by

u^j:=χUηjuj\widehat{u}^{j}:=\chi_{U}\cdot\eta^{j}-u^{j}

satisfies that

Dαu^j(p)=Dαηj(p),D^{\alpha}\widehat{u}^{j}(p)=D^{\alpha}\eta^{j}(p),

for all α=(α1,,αn)\alpha=(\alpha_{1},\ldots,\alpha_{n}) with |α|j|\alpha|\leq j. Since eξCΩe^{\xi}\leq C_{\Omega}, the L2L^{2}-estimate (4.2) implies that

(4.3) u^jχUηjΩ,μ2=ujΩ,μ2CΩCU\VCUηjU\V,μ2,\|\widehat{u}^{j}-\chi_{U}\cdot\eta^{j}\|^{2}_{\Omega,\mu}=\|u^{j}\|_{\Omega,\mu}^{2}\leq\frac{C_{\Omega}C_{U\backslash V}}{C_{U}}\|\eta^{j}\|_{U\backslash V,\mu}^{2},

as we required. ∎

Remark 4.3.

In the localization lemma of minimum integrals for domains (cf. [greene2006function, kim1996boundary]), the plurisubharmonic function ξ(z):=log|zp|2(n+j)\xi(z):=\log|z-p|^{2(n+j^{\prime})} is usually applied so that the constant CC depends on the distance from pp to boundary of Ω\Omega. However, we want to prove results for not only the pointwise convergence but also the uniform convergence in this paper. Therefore, we will use a modified function ξ\xi by an appropriate cut-off function, following Tian’s idea in [tian1990].

Note that since vjv^{j} is the minimizer of μj(Ω)𝒜μ2(Ω)\mathcal{E}^{j}_{\mu}(\Omega)\subset\mathcal{A}^{2}_{\mu}(\Omega), we have an upper bound estimate for the minimum integral:

Iμj:=vjΩ,μ2u^jΩ,μ2ηjU,μ2+CηjU\V,μ2.I^{j}_{\mu}:=\|v^{j}\|^{2}_{\Omega,\mu}\leq\|\widehat{u}^{j}\|^{2}_{\Omega,\mu}\leq\|\eta^{j}\|^{2}_{U,\mu}+C\|\eta^{j}\|_{U\backslash V,\mu}^{2}.

For a sharp estimate of ηjμ2\|\eta^{j}\|_{\mu}^{2}, we will change the coordinates locally so that μ=eφ\mu=e^{-\varphi} has good representation, using the following finite order approximation of the Bochner coordinates (or K-coordinates). For later uses, we present a proof with details here.

Lemma 4.4.

There exist a neighborhood WW of pp and an injective holomorphic map f:WBn(rp)f:W\rightarrow B^{n}(r_{p}) with f(p)=0f(p)=0 such that for a holomorphic function h:Wh:W\rightarrow\mathbb{C}, the Taylor expansion of the function

Φ:=(φ2Re(h))f1{\Phi}:=(\varphi-2{\rm Re}(h))\circ f^{-1}

at the origin in the new coordinate system w=(w1,.wn)=f(z)w=(w_{1},\ldots.w_{n})=f(z) satisfies that

Φ(w)=|w|2+14i,j,k,l=1nΦij¯kl¯(0)wiwj¯wkwl¯+O(|w|5),\Phi(w)=|w|^{2}+\frac{1}{4}\sum_{i,j,k,l=1}^{n}\Phi_{i\overline{j}k\overline{l}}(0)w_{i}\overline{w_{j}}w_{k}\overline{w_{l}}+O(|w|^{5}),

where O(|w|5)O(|w|^{5}) denotes terms which are at least quintic in w,w¯w,\overline{w}-variables. Moreover, the radius rpr_{p} only depends on values

{Dzαφjk¯(p)},\{D_{z}^{\alpha}\varphi_{j\overline{k}}(p)\},

for all α=(α1,,αn)\alpha=(\alpha_{1},\ldots,\alpha_{n}) with 0|α|20\leq|\alpha|\leq 2 and j,k=1,,nj,k=1,\ldots,n.

Proof.

Let ζ:=zp\zeta:=z-p. The Taylor expansion of φ\varphi at pp is

φ(z)=\displaystyle\varphi(z)=\ h(z)+h(z)¯+i,j=1nφij¯(p)ζiζj¯+k=1n(ζkhk(z)¯+ζk¯hk(z))\displaystyle h(z)+\overline{h(z)}+\sum_{i,j=1}^{n}\varphi_{i\overline{j}}(p)\zeta_{i}\overline{\zeta_{j}}+\sum_{k=1}^{n}\left(\zeta_{k}\overline{h_{k}(z)}+\overline{\zeta_{k}}h_{k}(z)\right)
+i,j,k,l=1n14φij¯kl¯(p)ζiζj¯ζkζl¯+O(|zp|5),\displaystyle+\sum_{i,j,k,l=1}^{n}\frac{1}{4}\varphi_{i\overline{j}k\overline{l}}(p)\zeta_{i}\overline{\zeta_{j}}\zeta_{k}\overline{\zeta_{l}}+O(|z-p|^{5}),

where

h(z):=12φ(p)+1|α|4Dzαφ(p)|α|!(zp)αandhk(z):=2|α|3Dzαφk¯(p)|α|!(zp)α.h(z):=\frac{1}{2}\varphi(p)+\sum_{1\leq|\alpha|\leq 4}\frac{D^{\alpha}_{z}\varphi(p)}{|\alpha|!}(z-p)^{\alpha}{\rm\ \ and\ \ }h_{k}(z):=\sum_{2\leq|\alpha|\leq 3}\frac{D^{\alpha}_{z}\varphi_{\overline{k}}(p)}{|\alpha|!}(z-p)^{\alpha}.

Then the function

ϕ(z):=φ(z)h(z)h(z)¯\phi(z):=\varphi(z)-h(z)-\overline{h(z)}

satisfies that for all α,β\alpha,\beta with 0|α|4, 1|β|30\leq|\alpha|\leq 4,\ 1\leq|\beta|\leq 3, we have

(4.4) Dzαϕ(p)=0,andDzβϕk¯(p)=Dzβφk¯(p).D^{\alpha}_{z}\phi(p)=0,{\rm\ \ and\ \ }D^{\beta}_{z}\phi_{\overline{k}}(p)=D^{\beta}_{z}\varphi_{\overline{k}}(p).

Define w=f(z)=(f1(z),,fn(z))w=f(z)=(f_{1}(z),\ldots,f_{n}(z)) by

(4.5) fj(z):=k=1nφk¯j(p){1|α|3Dzαφk¯(p)|α|!(zp)α},f_{j}(z):=\sum_{k=1}^{n}\sqrt{\varphi}^{\overline{k}j}(p)\left\{\sum_{1\leq|\alpha|\leq 3}\frac{D^{\alpha}_{z}\varphi_{\overline{k}}(p)}{|\alpha|!}(z-p)^{\alpha}\right\},

where (φk¯j(p))\left(\sqrt{\varphi}^{\overline{k}j}(p)\right) is the matrix whose square is the inverse matrix of (φjk¯(p))\left({\varphi}_{j\overline{k}}(p)\right). Consider the Taylor coefficients of the function

Φ(w):=(φ2Re(h))f1(w)=ϕf1(w)=ϕ(z).{\Phi}(w):=(\varphi-2{\rm Re}(h))\circ f^{-1}(w)=\phi\circ f^{-1}(w)=\phi(z).

Then a direct computation with (4.4) and (4.5) implies that

Φij¯(0)=δij,DwαΦ(p)=0,andDwβΦk¯(p)=Dw¯βΦk(p)=0,\Phi_{i\overline{j}}(0)=\delta_{ij},\ \ D^{\alpha}_{w}\Phi(p)=0,{\rm\ \ and\ \ }D^{\beta}_{w}\Phi_{\overline{k}}(p)=D^{\beta}_{\overline{w}}\Phi_{k}(p)=0,

for all α,β\alpha,\beta with 0|α|4, 1|β|30\leq|\alpha|\leq 4,\ 1\leq|\beta|\leq 3, as we required. ∎

Remark 4.5.

One can check that the coordinates w=(w1,.wn)=f(z)w=(w_{1},\ldots.w_{n})=f(z) are Kähler normal coordinates of the Kähler metric i¯Φ=i¯φi\partial\overline{\partial}\Phi=i\partial\overline{\partial}\varphi. If the given strictly plurisubharmonic function φ\varphi is real-analytic, we can construct a better coordinate system in the sense that DwαΦjk¯(0)=0D_{w}^{\alpha}\Phi_{j\overline{k}}(0)=0 for all α\alpha with |α|1|\alpha|\geq 1. This is called the Bochner normal coordinates (or K-coordinates), unique up to unitary linear transformations. In fact, we can express the coordinate transform explicitly in terms of the polarization φ(z,ζ¯)\varphi(z,\overline{\zeta}) of the given analytic function φ(z)=φ(z,z¯)\varphi(z)=\varphi(z,\overline{z}):

wj(z)=k=1nφk¯j(p){ζk¯|ζ=pφ(z,ζ¯)ζk¯|ζ=pφ(ζ,ζ¯)}.w_{j}(z)=\sum_{k=1}^{n}\sqrt{\varphi}^{\overline{k}j}(p)\left\{\frac{\partial}{\partial\overline{\zeta_{k}}}\Big{|}_{\zeta=p}\varphi(z,\overline{\zeta})-\frac{\partial}{\partial\overline{\zeta_{k}}}\Big{|}_{\zeta=p}\varphi(\zeta,\overline{\zeta})\right\}.

In the case that the potential function φ\varphi is the logarithm of the classic Bergman kernel function, the Bochner normal coordinates are called the Bergman representative coordinates (See [yoo2017]).

Since (w1,,wn)(w_{1},\ldots,w_{n}) is a holomorphic normal coordinate system of the Kähler metric gφ:=i¯φ=i¯Φg_{\varphi}:=i\partial\overline{\partial}\varphi=i\partial\overline{\partial}\Phi and Φ\Phi is a potential function, one can check that

Φij¯kl¯(0)=α,β,γ,σ=1n(φαβ¯γσ¯+φμν¯φαν¯γφμβ¯σ¯)zαwizβ¯wj¯zγwkzσ¯wl¯|z=p,-\Phi_{i\overline{j}k\overline{l}}(0)=\sum^{n}_{\alpha,\beta,\gamma,\sigma=1}\left(-\varphi_{\alpha\overline{\beta}\gamma\overline{\sigma}}+\varphi^{\mu\overline{\nu}}\varphi_{\alpha\overline{\nu}\gamma}\varphi_{\mu\overline{\beta}\overline{\sigma}}\right)\frac{\partial z_{\alpha}}{\partial w_{i}}\frac{\partial\overline{z_{\beta}}}{\partial\overline{w_{j}}}\frac{\partial z_{\gamma}}{\partial w_{k}}\frac{\partial\overline{z_{\sigma}}}{\partial\overline{w_{l}}}\Big{|}_{z=p},

are the local expressions of the curvature 44-tensor, i.e.,

Rij¯kl¯(w)dwidwj¯dwkdwl¯=Rαβ¯γσ¯(z)dzαdzβ¯dzγdzσ¯,R_{i\overline{j}k\overline{l}}(w)dw_{i}\otimes d\overline{w_{j}}\otimes dw_{k}\otimes d\overline{w_{l}}=R_{\alpha\overline{\beta}\gamma\overline{\sigma}}(z)dz_{\alpha}\otimes d\overline{z_{\beta}}\otimes dz_{\gamma}\otimes d\overline{z_{\sigma}},

where Rij¯kl¯(w):=Φij¯kl¯(w)R_{i\overline{j}k\overline{l}}(w):=-\Phi_{i\overline{j}k\overline{l}}(w). Denote the Ricci and holomorphic sectional curvature at pp along XX by Sφ(p),Rφ(p;X)S_{\varphi}(p),R_{\varphi}(p;X) and Hφ(p;X)H_{\varphi}(p;X), respectively. For instance, if w1|0=df(X|p)\frac{\partial}{\partial w_{1}}\big{|}_{0}=df(X|_{p}), then

Hφ(p;X):=R11¯11¯(p)=Φ11¯11¯(0),H_{\varphi}(p;X):=R_{1\overline{1}1\overline{1}}(p)=-\Phi_{1\overline{1}1\overline{1}}(0),
Rφ(p;X):=i=1nR11¯ii¯(p)=i=1nΦ11¯ii¯(0).R_{\varphi}(p;X):=\sum_{i=1}^{n}R_{1\overline{1}i\overline{i}}(p)=-\sum_{i=1}^{n}\Phi_{1\overline{1}i\overline{i}}(0).

The scalar curvature is given by

Sφ(p):=i,j=1nRii¯jj¯(0)=i,j=1nΦii¯jj¯(0).S_{\varphi}(p):=\sum_{i,j=1}^{n}R_{i\overline{i}j\overline{j}}(0)=-\sum_{i,j=1}^{n}\Phi_{i\overline{i}j\overline{j}}(0).

4.2. Tian’s sequence of Bergman kernels and metrics

Let Ω\Omega be a pseudoconvex domain in n\mathbb{C}^{n} and φ\varphi be a smooth strictly plurisubharmonic function on Ω\Omega. Consider a sequence of weight functions μm+1:=emφ\mu_{m+1}:=e^{-m\varphi} for non-negative integers m0m\geq 0, and the corresponding weighted Bergman spaces

𝒜μm+12(Ω)=𝒜2(Ω,emφ).\mathcal{A}^{2}_{\mu_{m+1}}(\Omega)=\mathcal{A}^{2}(\Omega,e^{-m\varphi}).

Fix a point pΩp\in\Omega and a vector XnX\in\mathbb{C}^{n}. For j=0,1,2j=0,1,2, consider the minimum integrals of the weighted Bergman kernels and denote the minimizers of μm+1j(Ω)\mathcal{E}^{j}_{\mu_{m+1}}(\Omega) by vm+1jv^{j}_{m+1}, i.e.,

Iμm+1j:=infuμm+1juμm+12=vm+1jμm+12.I^{j}_{\mu_{m+1}}:=\inf_{u\in\mathcal{E}^{j}_{\mu_{m+1}}}\|u\|_{\mu_{m+1}}^{2}=\|v^{j}_{m+1}\|^{2}_{\mu_{m+1}}.

For an upper estimate of the minimum integral, we will apply Propostion 4.2 with some suitable function ηj=ηm+1jμm+1j(W)\eta^{j}=\eta^{j}_{m+1}\in\mathcal{E}^{j}_{\mu_{m+1}}(W), defined on some neighborhood WΩW\subset\Omega of pp.

Proposition 4.6.

There exists m0>0m_{0}>0 such that for all mm0m\geq m_{0}, there is a holomorphic function u^m+1jm+1j(Ω)\widehat{u}^{j}_{m+1}\in\mathcal{E}^{j}_{m+1}(\Omega) satisfying

(4.6) u^m+1jμm+12=emφ(p)det(φkl¯(p))1(gφ(p;X))j1j!mj(πm)n(1+cjm+O(1m2)),\|\widehat{u}^{j}_{m+1}\|^{2}_{\mu_{m+1}}=\frac{e^{-m\varphi(p)}}{\det{\left(\varphi_{k\overline{l}}(p)\right)}}\frac{1}{\left(g_{\varphi}(p;X)\right)^{j}}\cdot\frac{1}{j!m^{j}}\left(\frac{\pi}{m}\right)^{n}\left(1+\frac{c_{j}}{m}+O\Big{(}\frac{1}{m^{2}}\Big{)}\right),

and the constants cjc_{j} are given by

c0=Sφ(p)2,c1=Sφ(p)+2Rφ(p;X)2,c2=Sφ(p)+4Rφ(p;X)+Hφ(p;X)2,c_{0}=\frac{S_{\varphi}(p)}{2},\ c_{1}=\frac{S_{\varphi}(p)+2R_{\varphi}(p;X)}{2},\ c_{2}=\frac{S_{\varphi}(p)+4R_{\varphi}(p;X)+H_{\varphi}(p;X)}{2},

where gφ(p;X)g_{\varphi}(p;X) is the length square of XX with respect to the Kähler metric gφ:=i¯φg_{\varphi}:=i\partial\overline{\partial}\varphi.

Proof.

Note that by definition, uμm+1j(Ω)u\in\mathcal{E}^{j}_{\mu_{m+1}}(\Omega) implies that DcXj(1cju)=1D^{j}_{cX}\left(\frac{1}{c^{j}}u\right)=1 for any constant c>0c>0. Therefore, we may assume that by normalizing XX,

gφ(p;X)=|X|i¯φ2=1.g_{\varphi}(p;X)=|X|^{2}_{i\partial\overline{\partial}\varphi}=1.

Let ff be the holomorphic mapping, and let WW be the neighborhood of pp in Lemma 4.4. Consider neighborhoods VmUmWV_{m}\subset U_{m}\subset W of pp such that

f(Vm)=Bn(rm/2)f(Um)=Bn(rm)f(W)=Bn(rp).f(V_{m})=B^{n}(r_{m}/2)\subset f(U_{m})=B^{n}(r_{m})\subset f(W)=B^{n}(r_{p}).

The radius rmr_{m} will be chosen later so that rm0r_{m}\searrow 0 as mm\rightarrow\infty. Define a holomorphic function ηm+1j\eta^{j}_{m+1} on WW by

ηm+1j(z):=1j!emh(p)𝒥(f(p))(f1(z))jemh(z)𝒥(f(z)),\eta^{j}_{m+1}(z):=\frac{1}{j!e^{mh(p)}\mathcal{J}(f(p))}(f_{1}(z))^{j}e^{mh(z)}\mathcal{J}(f(z)),

where hh and f(z)=(f1(z),,fn(z))f(z)=(f_{1}(z),\ldots,f_{n}(z)) are the holomorphisms in Lemma 4.4. We may further assume that dfp(X)df_{p}(X) and w1\frac{\partial}{\partial w_{1}} are parallel by modifying ff via an unitary linear transform. Then

ηm+1jμm+1j(W).\eta^{j}_{m+1}\in\mathcal{E}^{j}_{\mu_{m+1}}(W).

Let χ:[0,)[0,1]\chi:[0,\infty)\rightarrow[0,1] be a cut-off function satisfying χ(t)=1\chi(t)=1 for t12t\leq\frac{1}{2}, χ(t)=0\chi(t)=0 for t1t\geq 1, 0χ(t)40\leq-\chi^{\prime}(t)\leq 4, and |χ′′(t)|8|\chi^{\prime\prime}(t)|\leq 8. Then the function

χUm(z):=χ(|f(z)|/rm)Cc(Um)\chi_{U_{m}}(z):=\chi(|f(z)|/r_{m})\in C^{\infty}_{c}({U_{m}})

satisfies that χUm=1\chi_{U_{m}}=1 on Vm{V_{m}} and 0χUm10\leq\chi_{U_{m}}\leq 1 on Um{U_{m}}. Define a non-positive function ξm\xi_{m} on Ω\Omega by

ξm:=(n+j)χUmlog(|f|2/rm2).\xi_{m}:=(n+j^{\prime})\chi_{U_{m}}\cdot\log(|f|^{2}/r^{2}_{m}).

Then the definition (4.5) implies that there exists a positive constant Cφ(p)C_{\varphi}(p), depending only derivatives of φjk¯(p)\varphi_{j\overline{k}}(p), satisfying

|f(z)|2=|w|2Cφ(p)|zp|2.|f(z)|^{2}=|w|^{2}\leq C_{\varphi}(p)|z-p|^{2}.

A direct computation shows that for a constant C=Cφ(p)>0C^{\prime}=C^{\prime}_{\varphi}(p)>0, independent of mm,

ξm+C(n+j)rm2|z|2\xi_{m}+\frac{C^{\prime}(n+j^{\prime})}{r_{m}^{2}}|z|^{2}

is plurisubharmonic. Consider the function

ψm:=mφ+ξm=(mφC(n+j)rm2|z|2)+(ξm+C(n+j)rm2|z|2).\psi_{m}:=m\varphi+\xi_{m}=\left(m\varphi-\frac{C^{\prime}(n+j^{\prime})}{r_{m}^{2}}|z|^{2}\right)+\left(\xi_{m}+\frac{C^{\prime}(n+j^{\prime})}{r_{m}^{2}}|z|^{2}\right).

Note that since φ\varphi is strictly plurisubharmonic on Ω\Omega, there exists a positive constant C′′=C′′(φ,W)C^{\prime\prime}=C^{\prime\prime}(\varphi,W) such that φC′′|z|2\varphi-C^{\prime\prime}|z|^{2} is plurisubharmonic on WΩW\subset\subset\Omega. Then,

ψm(C′′mC(n+j)rm2)|z|2\psi_{m}-\left(C^{\prime\prime}m-\frac{C^{\prime}(n+j^{\prime})}{r_{m}^{2}}\right)|z|^{2}

is plurisubharmonic on UmWU_{m}\subset W. Now take rm2:=(logm)2mr_{m}^{2}:=\frac{(\log m)^{2}}{m}. For sufficiently large mm, there exists a positive constant CC, independent of mm such that

CUm:=C′′mC(n+j)rm2=m(C′′C(n+j)(logm)2)>mCC_{U_{m}}:=C^{\prime\prime}m-\frac{C^{\prime}(n+j^{\prime})}{r_{m}^{2}}=m\left(C^{\prime\prime}-\frac{C^{\prime}(n+j^{\prime})}{(\log m)^{2}}\right)>\frac{m}{C}

is positive, and ψm\psi_{m} is plurisubharmonic on Ω\Omega. Then, we can apply Proposition 4.2 with ηmj\eta^{j}_{m} and ψm\psi_{m}. Note that

|¯χUm|2eξmCrm|\overline{\partial}\chi_{U_{m}}|^{2}e^{-\xi_{m}}\leq Cr_{m}

on Um\VmU_{m}\backslash V_{m}. Proposition 4.2 implies that there exists a holomorphic function

u^m+1j:=χUmηm+1jum+1jμm+1j(Ω)\widehat{u}^{j}_{m+1}:=\chi_{U_{m}}\cdot\eta^{j}_{m+1}-u^{j}_{m+1}\in\mathcal{E}^{j}_{\mu_{m+1}}(\Omega)

satisfying

u^m+1jΩ,μm+12=χUmηm+1jΩ,μm+122ReχUmηm+1j,um+1jΩ,μm+1+um+1jΩ,μm+12,\|\widehat{u}^{j}_{m+1}\|^{2}_{\Omega,\mu_{m+1}}=\|\chi_{U_{m}}\cdot\eta^{j}_{m+1}\|^{2}_{\Omega,\mu_{m+1}}-2{\rm Re}\langle\chi_{U_{m}}\cdot\eta^{j}_{m+1},u^{j}_{m+1}\rangle_{\Omega,\mu_{m+1}}+\|u^{j}_{m+1}\|^{2}_{\Omega,\mu_{m+1}},

and

um+1jΩ,μm+12C(logm)2ηm+1jUm\Vm,μm+12.\|u^{j}_{m+1}\|^{2}_{\Omega,\mu_{m+1}}\leq\frac{C}{(\log m)^{2}}\|\eta^{j}_{m+1}\|^{2}_{U_{m}\backslash V_{m},\mu_{m+1}}.

Note that for any neighborhood UWU\subset W,

ηm+1jU,μm+12\displaystyle\|\eta^{j}_{m+1}\|^{2}_{U,\mu_{m+1}} =1|j!emh(p)𝒥(f(p))|2U|(f1(z))j|2em(φ2Re(h))|𝒥(f(z))|2𝑑λ(z)\displaystyle=\frac{1}{|j!e^{mh(p)}\mathcal{J}(f(p))|^{2}}\int_{U}\left|(f_{1}(z))^{j}\right|^{2}e^{-m(\varphi-2{\rm Re}(h))}\left|\mathcal{J}(f(z))\right|^{2}d\lambda(z)
=emφ(p)(j!)2det(φkl¯(p))f(U)|(w1)j|2emΦ(w)𝑑λ(w),\displaystyle=\frac{e^{-m\varphi(p)}}{(j!)^{2}\det{\left(\varphi_{k\overline{l}}(p)\right)}}\int_{f(U)}\left|(w_{1})^{j}\right|^{2}e^{-m{\Phi}(w)}d\lambda(w),

since 2Re(h(p))=φ(p)2{\rm Re}(h(p))=\varphi(p) and |𝒥(f(p))|2=det(φkl¯(p))|\mathcal{J}(f(p))|^{2}=\det{\left(\varphi_{k\overline{l}}(p)\right)} (see the proof of Lemma 4.4).

On the other hand, the Taylor expansion of Φ{\Phi} implies that

(4.7) emΦ(w)=em|w|2(1m4i,j,k,l=1nΦij¯kl¯(0)wiwj¯wkwl¯O(m|w|5)).e^{-m\Phi(w)}=e^{-m|w|^{2}}\left(1-\frac{m}{4}\sum_{i,j,k,l=1}^{n}\Phi_{i\overline{j}k\overline{l}}(0)w_{i}\overline{w_{j}}w_{k}\overline{w_{l}}-O(m|w|^{5})\right).

so that we have

Bn(rm)\Bn(rm/2)|(w1)j|2emΦ(w)𝑑λ(w)=O(emrm2)=O(mlogm)=O(mr),\int_{B^{n}(r_{m})\backslash B^{n}(r_{m}/2)}\left|(w_{1})^{j}\right|^{2}e^{-m{\Phi}(w)}d\lambda(w)=O\left(e^{-mr_{m}^{2}}\right)=O\left(m^{-\log m}\right)=O\left(m^{-r}\right),

for any r>0r>0. This implies that

u^m+1jΩ,μm+12=emφ(p)(j!)2det(φkl¯(p))(Bn(rm)|χ(|w|rm)(w1)j|2emΦ(w)𝑑λ(w)+O(mr)).\|\widehat{u}^{j}_{m+1}\|^{2}_{\Omega,\mu_{m+1}}=\frac{e^{-m\varphi(p)}}{(j!)^{2}\det{\left(\varphi_{k\overline{l}}(p)\right)}}\left(\int_{B^{n}(r_{m})}\left|\chi\left(\frac{|w|}{r_{m}}\right)(w_{1})^{j}\right|^{2}e^{-m{\Phi}(w)}d\lambda(w)+O\left(m^{-r}\right)\right).

Note that

Bn(rm2)|(w1)j|2emΦ𝑑λBn(rm)|χ(|w|rm)(w1)j|2emΦ𝑑λBn(rm)|(w1)j|2emΦ𝑑λ,\int_{B^{n}(\frac{r_{m}}{2})}\left|(w_{1})^{j}\right|^{2}e^{-m{\Phi}}d\lambda\leq\int_{B^{n}(r_{m})}\left|\chi\left(\frac{|w|}{r_{m}}\right)(w_{1})^{j}\right|^{2}e^{-m{\Phi}}d\lambda\leq\int_{B^{n}(r_{m})}\left|(w_{1})^{j}\right|^{2}e^{-m{\Phi}}d\lambda,

and

𝔹(0,rm)|wα|2em|w|2𝑑λ=n|wα|2em|w|2𝑑λ+O(mr).\int_{\mathbb{B}(0,r_{m})}|w^{\alpha}|^{2}e^{-m|w|^{2}}d\lambda=\int_{\mathbb{C}^{n}}|w^{\alpha}|^{2}e^{-m|w|^{2}}d\lambda+O(m^{-r}).

Hence, we can use the following Laplace type integral:

(mπ)nn|wα|2em|w|2dλ=α1!αn!mα1++αn=:α!m|α|.\left(\frac{m}{\pi}\right)^{n}\int_{\mathbb{C}^{n}}|w^{\alpha}|^{2}e^{-m|w|^{2}}d\lambda=\frac{\alpha_{1}!\cdots\alpha_{n}!}{m^{\alpha_{1}+\cdots+\alpha_{n}}}=:\frac{\alpha!}{m^{|\alpha|}}.

Therefore, with the condition gφ(p;X)=1g_{\varphi}(p;X)=1, we have

u^m+1jμm+12=emφ(p)det(φkl¯(p))1j!mj(πm)n(1+cjm+O(1m2)),\|\widehat{u}^{j}_{m+1}\|^{2}_{\mu_{m+1}}=\frac{e^{-m\varphi(p)}}{\det{\left(\varphi_{k\overline{l}}(p)\right)}}\frac{1}{j!m^{j}}\left(\frac{\pi}{m}\right)^{n}\left(1+\frac{c_{j}}{m}+O\Big{(}\frac{1}{m^{2}}\Big{)}\right),

Finally, the conclusion follows from the below lemma. ∎

Lemma 4.7.

The constants are given by

c0=Sφ(p)2,c1=Sφ(p)+2Rφ(p;X)2,c2=Sφ(p)+4Rφ(p;X)+Hφ(p;X)2.c_{0}=\frac{S_{\varphi}(p)}{2},\ c_{1}=\frac{S_{\varphi}(p)+2R_{\varphi}(p;X)}{2},\ c_{2}=\frac{S_{\varphi}(p)+4R_{\varphi}(p;X)+H_{\varphi}(p;X)}{2}.
Proof.

Note that

cjm=j!mjm4(mπ)ni,t,k,l=1n{Φit¯kl¯(0)n|w1j|2wiwt¯wkwl¯em|w|2𝑑λ}.\frac{c_{j}}{m}=-j!m^{j}\frac{m}{4}\left(\frac{m}{\pi}\right)^{n}\sum_{i,t,k,l=1}^{n}\left\{\Phi_{i\overline{t}k\overline{l}}(0)\int_{\mathbb{C}^{n}}|w_{1}^{j}|^{2}w_{i}\overline{w_{t}}w_{k}\overline{w_{l}}e^{-m|w|^{2}}\ d\lambda\right\}.

For any multi-indices αβ\alpha\neq\beta, we have nwαwβ¯em|w|2𝑑λ=0.\int_{\mathbb{C}^{n}}w^{\alpha}\overline{w^{\beta}}e^{-m|w|^{2}}d\lambda=0. This implies that

i,j,k,l=1nnΦij¯kl¯(0)wiwj¯wkwl¯em|w|2𝑑λ\displaystyle\sum_{i,j,k,l=1}^{n}\int_{\mathbb{C}^{n}}\Phi_{i\overline{j}k\overline{l}}(0)w_{i}\overline{w_{j}}w_{k}\overline{w_{l}}\ e^{-m|w|^{2}}d\lambda
=\displaystyle= ni=1nΦii¯ii¯(0)|wi|4em|w|2dλ+nij(Φii¯jj¯(0)+Φij¯ji¯(0))|wi|2|wj|2em|w|2dλ\displaystyle\int_{\mathbb{C}^{n}}\sum_{i=1}^{n}\Phi_{i\overline{i}i\overline{i}}(0)|w_{i}|^{4}e^{-m|w|^{2}}d\lambda+\int_{\mathbb{C}^{n}}\sum_{i\neq j}\left(\Phi_{i\overline{i}j\overline{j}}(0)+\Phi_{i\overline{j}j\overline{i}}(0)\right)|w_{i}|^{2}|w_{j}|^{2}e^{-m|w|^{2}}d\lambda
=\displaystyle= (πm)n1m2{2i=1nΦii¯ii¯(0)+ij(Φii¯jj¯(0)+Φij¯ji¯(0))}\displaystyle\left(\frac{\pi}{m}\right)^{n}\frac{1}{m^{2}}\left\{2\sum_{i=1}^{n}\Phi_{i\overline{i}i\overline{i}}(0)+\sum_{i\neq j}\left(\Phi_{i\overline{i}j\overline{j}}(0)+\Phi_{i\overline{j}j\overline{i}}(0)\right)\right\}
=\displaystyle= (πm)n2m2i,jΦii¯jj¯(0)=(πm)n2m2Sφ(p).\displaystyle\left(\frac{\pi}{m}\right)^{n}\frac{2}{m^{2}}\sum_{i,j}\Phi_{i\overline{i}j\overline{j}}(0)=-\left(\frac{\pi}{m}\right)^{n}\frac{2}{m^{2}}S_{\varphi}(p).

Therefore, c0=Sφ(p)/2c_{0}=S_{\varphi}(p)/2. The constant c1c_{1} can be computed by

i,j,k,l=1nnΦij¯kl¯(0)|w1|2wiwj¯wkwl¯em|w|2𝑑λ\displaystyle\sum_{i,j,k,l=1}^{n}\int_{\mathbb{C}^{n}}\Phi_{i\overline{j}k\overline{l}}(0)|w_{1}|^{2}w_{i}\overline{w_{j}}w_{k}\overline{w_{l}}\ e^{-m|w|^{2}}d\lambda
=\displaystyle= nΦ11¯11¯(0)|w1|6em|w|2𝑑λ+ni2Φii¯ii¯(0)|w1|2|wi|4em|w|2dλ\displaystyle\int_{\mathbb{C}^{n}}\Phi_{1\overline{1}1\overline{1}}(0)|w_{1}|^{6}e^{-m|w|^{2}}d\lambda+\int_{\mathbb{C}^{n}}\sum_{i\geq 2}\Phi_{i\overline{i}i\overline{i}}(0)|w_{1}|^{2}|w_{i}|^{4}e^{-m|w|^{2}}d\lambda
+ni2(Φ11¯ii¯(0)+Φii¯11¯(0)+Φ1i¯i1¯(0)+Φi1¯1i¯(0))|w1|4|wi|2em|w|2dλ\displaystyle+\int_{\mathbb{C}^{n}}\sum_{i\geq 2}\left(\Phi_{1\overline{1}i\overline{i}}(0)+\Phi_{i\overline{i}1\overline{1}}(0)+\Phi_{1\overline{i}i\overline{1}}(0)+\Phi_{i\overline{1}1\overline{i}}(0)\right)|w_{1}|^{4}|w_{i}|^{2}e^{-m|w|^{2}}d\lambda
+nij2(Φii¯jj¯(0)+Φij¯ji¯(0))|w1|2|wi|2|wj|2em|w|2dλ\displaystyle+\int_{\mathbb{C}^{n}}\sum_{i\neq j\geq 2}\left(\Phi_{i\overline{i}j\overline{j}}(0)+\Phi_{i\overline{j}j\overline{i}}(0)\right)|w_{1}|^{2}|w_{i}|^{2}|w_{j}|^{2}e^{-m|w|^{2}}d\lambda
=\displaystyle= (πm)n1m3{6Φ11¯11¯(0)+2i2Φii¯ii¯(0)+8i2Φ11¯ii¯(0)+2ij2Φii¯jj¯(0)}\displaystyle\left(\frac{\pi}{m}\right)^{n}\frac{1}{m^{3}}\left\{6\Phi_{1\overline{1}1\overline{1}}(0)+2\sum_{i\geq 2}\Phi_{i\overline{i}i\overline{i}}(0)+8\sum_{i\geq 2}\Phi_{1\overline{1}i\overline{i}}(0)+2\sum_{i\neq j\geq 2}\Phi_{i\overline{i}j\overline{j}}(0)\right\}
=\displaystyle= (πm)n1m3{4Φ11¯11¯(0)+2i=1nΦii¯ii¯(0)+4i2Φ11¯ii¯(0)+2ijΦii¯jj¯(0)}\displaystyle\left(\frac{\pi}{m}\right)^{n}\frac{1}{m^{3}}\left\{4\Phi_{1\overline{1}1\overline{1}}(0)+2\sum_{i=1}^{n}\Phi_{i\overline{i}i\overline{i}}(0)+4\sum_{i\geq 2}\Phi_{1\overline{1}i\overline{i}}(0)+2\sum_{i\neq j}\Phi_{i\overline{i}j\overline{j}}(0)\right\}
=\displaystyle= (πm)n1m3{4i=1nΦ11¯ii¯(0)+2i,jΦii¯jj¯(0)}=(πm)n1m3(4Rφ(p;X)+2Sφ(p)),\displaystyle\left(\frac{\pi}{m}\right)^{n}\frac{1}{m^{3}}\left\{4\sum_{i=1}^{n}\Phi_{1\overline{1}i\overline{i}}(0)+2\sum_{i,j}\Phi_{i\overline{i}j\overline{j}}(0)\right\}=-\left(\frac{\pi}{m}\right)^{n}\frac{1}{m^{3}}\left(4R_{\varphi}(p;X)+2S_{\varphi}(p)\right),

Similarly, one can show that

i,j,k,l=1nnΦij¯kl¯(0)|w1|4wiwj¯wkwl¯em|w|2𝑑λ\displaystyle\sum_{i,j,k,l=1}^{n}\int_{\mathbb{C}^{n}}\Phi_{i\overline{j}k\overline{l}}(0)|w_{1}|^{4}w_{i}\overline{w_{j}}w_{k}\overline{w_{l}}\ e^{-m|w|^{2}}d\lambda
=\displaystyle= (πm)n1m4(4Hφ(p;X)+16Rφ(p;X)+4Sφ(p)).\displaystyle-\left(\frac{\pi}{m}\right)^{n}\frac{1}{m^{4}}\left(4H_{\varphi}(p;X)+16R_{\varphi}(p;X)+4S_{\varphi}(p)\right).

Define sub-spaces of the weighted Bergman space m:=𝒜μm2(Ω)\mathcal{H}_{m}:=\mathcal{A}^{2}_{\mu_{m}}(\Omega) of co-dimension 11:

m0\displaystyle\mathcal{H}^{0}_{m} :={um:u(p)=0},\displaystyle:=\left\{u\in\mathcal{H}_{m}:u(p)=0\right\},
m1\displaystyle\mathcal{H}^{1}_{m} :={um0:DXu(p)=0},\displaystyle:=\left\{u\in\mathcal{H}^{0}_{m}:D_{X}u(p)=0\right\},
m2\displaystyle\mathcal{H}^{2}_{m} :={um1:DXDXu(p)=0}.\displaystyle:=\left\{u\in\mathcal{H}^{1}_{m}:D_{X}D_{X}u(p)=0\right\}.

The following proposition shows that the function u^mj\widehat{u}^{j}_{m} is asymptotically orthogonal to the subspace mj\mathcal{H}^{j}_{m} as mm\rightarrow\infty.

Proposition 4.8.

For all vm+1jv\in\mathcal{H}^{j}_{m+1}, we have

(4.8) |u^m+1j,vμm+1|=O(1m)u^m+1jμm+1vμm+1.|\langle\widehat{u}^{j}_{m+1},v\rangle_{\mu_{m+1}}|=O\Big{(}\frac{1}{m}\Big{)}\|\widehat{u}^{j}_{m+1}\|_{\mu_{m+1}}\|v\|_{\mu_{m+1}}.
Proof.

Let vv be a function in m+1j\mathcal{H}^{j}_{m+1}. On WW, we can represent vv as

v(z)=v(z)emh(z)𝒥(f(z))emh(z)𝒥(f(z))=:v~(f(z))emh(z)𝒥(f(z)).v(z)=\frac{v(z)}{e^{mh(z)}\mathcal{J}(f(z))}e^{mh(z)}\mathcal{J}(f(z))=:\widetilde{v}(f(z))\ e^{mh(z)}\mathcal{J}(f(z)).

As in Proposition 4.6, we may assume that dfp(X)=w1df_{p}(X)=\frac{\partial}{\partial w_{1}}. Since vm+1jv\in\mathcal{H}^{j}_{m+1},

(4.9) (w1)kv~(0)=0,forall 0kj.\left(\frac{\partial}{\partial w_{1}}\right)^{k}\widetilde{v}(0)=0,\ \ \ \ {\rm for\ all\ \ }0\leq k\leq j.

Recall that

u^m+1j:=χUmηm+1jum+1jμm+1j(Ω)\widehat{u}^{j}_{m+1}:=\chi_{U_{m}}\cdot\eta^{j}_{m+1}-u^{j}_{m+1}\in\mathcal{E}^{j}_{\mu_{m+1}}(\Omega)

and

ηm+1j(z)=1j!emh(p)𝒥(f(p))(f1(z))jemh(z)𝒥(f(z))=:c(p,m)(f1(z))jemh(z)𝒥(f(z)).\eta^{j}_{m+1}(z)=\frac{1}{j!e^{mh(p)}\mathcal{J}(f(p))}(f_{1}(z))^{j}e^{mh(z)}\mathcal{J}(f(z))=:c(p,m)(f_{1}(z))^{j}e^{mh(z)}\mathcal{J}(f(z)).

Using the Taylor expansion (4.7), we can show that

u^m+1j,vμm+1\displaystyle\langle\widehat{u}^{j}_{m+1},v\rangle_{\mu_{m+1}} =χUmηm+1jum+1j,vΩ,μm+1\displaystyle=\langle\chi_{U_{m}}\cdot\eta^{j}_{m+1}-u^{j}_{m+1},v\rangle_{\Omega,\mu_{m+1}}
=χUmηm+1j,vUm,μm+1um+1j,vΩ,μm+1\displaystyle=\langle\chi_{U_{m}}\cdot\eta^{j}_{m+1},v\rangle_{U_{m},\mu_{m+1}}-\langle u^{j}_{m+1},v\rangle_{\Omega,\mu_{m+1}}
=c(p,m)𝔹(0,rm)w1jv~(w)¯em|w|2(1+O(m|w|4))𝑑λ+O(mr)vμm+1.\displaystyle=c(p,m)\int_{\mathbb{B}(0,r_{m})}w_{1}^{j}\ \overline{\widetilde{v}(w)}e^{-m|w|^{2}}\left(1+O(m|w|^{4})\right)d\lambda+O(m^{-r})\|v\|_{\mu_{m+1}}.

Note that the condition (4.9) implies that

𝔹(0,rm)w1jv~(w)¯em|w|2𝑑λ=0.\int_{\mathbb{B}(0,r_{m})}w_{1}^{j}\ \overline{\widetilde{v}(w)}e^{-m|w|^{2}}d\lambda=0.

Applying the Cauchy-Schwarz inequality to the integral, we obtain

|𝔹(0,rm)c(p,m)w1jv~(w)¯em|w|2O(m|w|4)𝑑λ|\displaystyle\Big{|}\int_{\mathbb{B}(0,r_{m})}c(p,m)w_{1}^{j}\ \overline{\widetilde{v}(w)}e^{-m|w|^{2}}O(m|w|^{4})d\lambda\Big{|}
C(m2𝔹(0,rm)|c(p,m)|2|w1j|2|w|8emΦ(w)𝑑λ)12vμm+1\displaystyle\leq C\left(m^{2}\int_{\mathbb{B}(0,r_{m})}|c(p,m)|^{2}|w_{1}^{j}|^{2}|w|^{8}e^{-m{\Phi}(w)}d\lambda\right)^{\frac{1}{2}}\|v\|_{\mu_{m+1}}
Cmu^mjμm+1vμm+1,\displaystyle\leq\frac{C}{m}\|\widehat{u}^{j}_{m}\|_{\mu_{m+1}}\|v\|_{\mu_{m+1}},

as we required. ∎

Lemma 4.9.

Let vm+1jv^{j}_{m+1} be the minimizers of μm+1j\mathcal{E}^{j}_{\mu_{m+1}}, i.e., Iμm+1j=vm+1jμm+12I^{j}_{\mu_{m+1}}=\|v^{j}_{m+1}\|^{2}_{\mu_{m+1}} for j=0,1,2j=0,1,2. Then, we have

(1O(1m2))u^m+1jμm+12vm+1jμm+12u^m+1jμm+12.\left(1-O\left(\frac{1}{m^{2}}\right)\right)\|\widehat{u}^{j}_{m+1}\|^{2}_{\mu_{m+1}}\leq\|v^{j}_{m+1}\|^{2}_{\mu_{m+1}}\leq\|\widehat{u}^{j}_{m+1}\|^{2}_{\mu_{m+1}}.
Proof.

Choose an orthogonal basis {vm+1k}k=0\{v^{k}_{m+1}\}_{k=0}^{\infty} of the weighted Bergman space 𝒜μm+12(Ω)\mathcal{A}^{2}_{\mu_{m+1}}(\Omega) including the minimizers vm+10,vm+11,vm+12v^{0}_{m+1},v^{1}_{m+1},v^{2}_{m+1}. Consider an orthogonal decomposition:

u^m+1j=k=0jakvm+1k+v~m+1j+1,\widehat{u}^{j}_{m+1}=\sum_{k=0}^{j}a_{k}v^{k}_{m+1}+\widetilde{v}_{m+1}^{j+1},

where v~m+1j+1m+1j\widetilde{v}_{m+1}^{j+1}\in\mathcal{H}^{j}_{m+1}. Since vmjmjv^{j}_{m}\in\mathcal{E}^{j}_{m}, we have aj=1a_{j}=1, and ak=0a_{k}=0 if 0k<j0\leq k<j so that

u^m+1j=vm+1j+v~m+1j+1,\widehat{u}^{j}_{m+1}=v^{j}_{m+1}+\widetilde{v}_{m+1}^{j+1},

and vm+1jv^{j}_{m+1} is orthogonal to v~m+1j+1\widetilde{v}_{m+1}^{j+1}. Proposition 4.8 implies that

v~m+1j+12=u^m+1jvm+1j2=u^m+1j,u^m+1jvm+1jCmu^m+1jv~m+1j+1\|\widetilde{v}_{m+1}^{j+1}\|^{2}=\|\widehat{u}^{j}_{m+1}-v^{j}_{m+1}\|^{2}=\langle\widehat{u}^{j}_{m+1},\widehat{u}^{j}_{m+1}-v^{j}_{m+1}\rangle\leq\frac{C}{m}\|\widehat{u}_{m+1}^{j}\|\|\widetilde{v}_{m+1}^{j+1}\|

so that

v~m+1j+12Cm2u^m+1j2.\|\widetilde{v}_{m+1}^{j+1}\|^{2}\leq\frac{C}{m^{2}}\|\widehat{u}_{m+1}^{j}\|^{2}.

The conclusion follows from the following:

vm+1j2=u^m+1j2v~m+1j+12(1Cm2)u^m+1j2.\|v_{m+1}^{j}\|^{2}=\|\widehat{u}_{m+1}^{j}\|^{2}-\|\widetilde{v}_{m+1}^{j+1}\|^{2}\geq\left(1-\frac{C}{m^{2}}\right)\|\widehat{u}_{m+1}^{j}\|^{2}.

We have the following asymptotic expansion for the sequence of the weighted Bergman kernels, metrics, and curvatures with respect to the weight μm+1:=emφ\mu_{m+1}:=e^{-m\varphi}.

Theorem 4.10.

Let Ω\Omega be a pseudoconvex domain in n\mathbb{C}^{n} with a smooth strictly plurisubharmonic function φ\varphi. Suppose that (Ω,eφ)(\Omega,e^{-\varphi}) admits the positive definite weighted Bergman metric. Fix a point pΩp\in\Omega and a vector XnX\in\mathbb{C}^{n}. There exists m0>0m_{0}>0 such that for all mm0m\geq m_{0}, we have

KΩ,μm+1(p)\displaystyle K_{\Omega,\mu_{m+1}}(p) =det(φkl¯(p))emφ(p)(mπ)n(1Sφ(p)2m+O(1m2)),\displaystyle=\frac{\det{\left(\varphi_{k\overline{l}}(p)\right)}}{e^{-m\varphi(p)}}\left(\frac{m}{\pi}\right)^{n}\left(1-\frac{S_{\varphi}(p)}{2m}+O\Big{(}\frac{1}{m^{2}}\Big{)}\right),
gΩ,μm+1(p;X)\displaystyle g_{\Omega,\mu_{m+1}}(p;X) =mgφ(p;X)(1Rφ(p;X)m+O(1m2)),\displaystyle=mg_{\varphi}(p;X)\left(1-\frac{R_{\varphi}(p;X)}{m}+O\Big{(}\frac{1}{m^{2}}\Big{)}\right),
HΩ,μm+1(p;X)\displaystyle H_{\Omega,\mu_{m+1}}(p;X) =Hφ(p;X)m+O(1m2).\displaystyle=\frac{H_{\varphi}(p;X)}{m}+O\Big{(}\frac{1}{m^{2}}\Big{)}.
Proof.

Let vm+1jv^{j}_{m+1} be minimizers of μm+1j\mathcal{E}^{j}_{\mu_{m+1}}, i.e., Iμm+1j=vm+1jμm+12I^{j}_{\mu_{m+1}}=\|v^{j}_{m+1}\|^{2}_{\mu_{m+1}}. Proposition 4.6 and Lemma 4.9 imply that

Iμm+1j=emφ(p)det(φkl¯(p))1gφ(p;X)j1j!mj(πm)n(1+cjm+O(1m2)).I^{j}_{\mu_{m+1}}=\frac{e^{-m\varphi(p)}}{\det{\left(\varphi_{k\overline{l}}(p)\right)}}\frac{1}{g_{\varphi}(p;X)^{j}}\frac{1}{j!m^{j}}\left(\frac{\pi}{m}\right)^{n}\left(1+\frac{c_{j}}{m}+O\Big{(}\frac{1}{m^{2}}\Big{)}\right).

Then, the conclusion follows from the Bergman-Fuks formula in Theorem 2.10:

KΩ,μm+1(p)=1Iμm+10,gΩ,μm+1(p;X)=Iμm+10Iμm+11,HΩ,μm+1(p;X)=2(Iμm+11)2Iμm+12Iμm+10.K_{\Omega,\mu_{m+1}}(p)=\frac{1}{I^{0}_{\mu_{m+1}}},\ \ \ \ g_{\Omega,\mu_{m+1}}(p;X)=\frac{I^{0}_{\mu_{m+1}}}{I^{1}_{\mu_{m+1}}},\ \ \ \ H_{\Omega,\mu_{m+1}}(p;X)=2-\frac{(I^{1}_{\mu_{m+1}})^{2}}{I^{2}_{\mu_{m+1}}I^{0}_{\mu_{m+1}}}.

Corollary 4.11.

For μm:=e(m1)φ\mu_{m}:=e^{-(m-1)\varphi}, we have the following

KΩ,μm(p)\displaystyle K_{\Omega,\mu_{m}}(p) =emφ(p)det(φkl¯(p))eφ(p)(mπ)n(1Sφ(p)+2n2m+O(1m2)),\displaystyle=e^{m\varphi(p)}\frac{\det{\left(\varphi_{k\overline{l}}(p)\right)}}{e^{\varphi(p)}}\left(\frac{m}{\pi}\right)^{n}\left(1-\frac{S_{\varphi}(p)+2n}{2m}+O\Big{(}\frac{1}{m^{2}}\Big{)}\right),
gΩ,μm(p;X)\displaystyle g_{\Omega,\mu_{m}}(p;X) =mgφ(p;X)(1Rφ(p;X)+1m+O(1m2)),\displaystyle=mg_{\varphi}(p;X)\left(1-\frac{R_{\varphi}(p;X)+1}{m}+O\Big{(}\frac{1}{m^{2}}\Big{)}\right),
HΩ,μm(p;X)\displaystyle H_{\Omega,\mu_{m}}(p;X) =Hφ(p;X)m+O(1m2).\displaystyle=\frac{H_{\varphi}(p;X)}{m}+O\Big{(}\frac{1}{m^{2}}\Big{)}.
Proof.

Use the following modification:

Iμmj\displaystyle I^{j}_{\mu_{m}} =e(m1)φ(p)det(φkl¯(p))1gφ(p;X)j1j!(m1)j(πm1)n(1+cjm1+O(1(m1)2))\displaystyle=\frac{e^{-(m-1)\varphi(p)}}{\det{\left(\varphi_{k\overline{l}}(p)\right)}}\frac{1}{g_{\varphi}(p;X)^{j}}\frac{1}{j!(m-1)^{j}}\left(\frac{\pi}{m-1}\right)^{n}\left(1+\frac{c_{j}}{m-1}+O\Big{(}\frac{1}{(m-1)^{2}}\Big{)}\right)
=e(m1)φ(p)det(φkl¯(p))1(gφ(p;X))j1j!mj(πm)n(1+cj+n+jm+O(1m2)).\displaystyle=\frac{e^{-(m-1)\varphi(p)}}{\det{\left(\varphi_{k\overline{l}}(p)\right)}}\frac{1}{\left(g_{\varphi}(p;X)\right)^{j}}\frac{1}{j!m^{j}}\left(\frac{\pi}{m}\right)^{n}\left(1+\frac{c_{j}+n+j}{m}+O\Big{(}\frac{1}{m^{2}}\Big{)}\right).

Corollary 4.12.

As mm\rightarrow\infty, we have the following pointwise convergences:

KΩ,μmm(p)eφ(p),1mgΩ,μm(p;X)gφ(p;X),mHΩ,μm(p;X)Hφ(p;X).\sqrt[m]{K_{\Omega,\mu_{m}}}(p)\rightarrow e^{\varphi(p)},\ \ \ \ \frac{1}{m}g_{\Omega,\mu_{m}}(p;X)\rightarrow g_{\varphi}(p;X),\ \ \ \ mH_{\Omega,\mu_{m}}(p;X)\rightarrow H_{\varphi}(p;X).
Remark 4.13.

The convergence speed at pp depends on the big OO-terms in the previous Theorem, which are determined by the derivatives {Dzαφjk¯(p)}\{D_{z}^{\alpha}\varphi_{j\overline{k}}(p)\} of the metric gφ=i¯φg_{\varphi}=i\partial\overline{\partial}\varphi, and the constant C′′C^{\prime\prime}, the lower bound of eigenvalues of the levi form i¯φi\partial\overline{\partial}\varphi on WW.

In [tian1990], Tian proved the C2C^{2}-convergence of the sequence of weighted Bergman metrics (C4C^{4}-convergence of the sequence of the weighted Bergman kernels) for compact polarized manifolds. Later, Ruan[ruan1998canonical] and Zelditch[zelditch1998szego], respectively improved this result up to to CC^{\infty} level. Since this is a local statement (pointwise convergence), the same result also holds for the non-compact case. Therefore, the convergence in Corollary 4.12 can be improved up to all CkC^{k}-derivatives of the weighted Bergman kernels.

5. Convergence of invariant weighted Bergman sequences

For compact manifolds, the pointwise convergence implies the uniform convergence. More generally, it is known that the convergence of Tian’s sequence is uniform if the given (possibly non-compact) manifold satisfies the conditions of the bounded geometry (cf. [ma2015exponential]). Bounded domains in n\mathbb{C}^{n} with the properties of uniformly squeezing [yeung2009] (also called the holomorphic homogeneous regular [liu2004canonical]) belong to this category. The proof of the uniform convergence in this case is much simpler thanks to the existence of the global coordinates and the transformation formula for weighted Bergman kernels.

In this section, we will focus two sequences of the weighted Bergman kernels, developed by Tian and Tsuji respectively. They are important examples in the sense that these sequences converge to the volume form of the unique Kähler-Einstein metric, and the corresponding weighted Bergman metrics are biholomorphically invariant. Since we will prove the uniform convergence of these sequences on uniform squeezing domains, we first briefly review known related results.

5.1. Uniform squeezing domains

Definition 5.1.

A (bounded) domain Ωn\Omega\subset\mathbb{C}^{n} is called an uniform squeezing domain if for any point pΩp\in\Omega, there exist r(0,1]r\in(0,1] (independent of pp) and a biholomorphism FpF_{p} on Ω\Omega satisfying Fp(p)=0F_{p}(p)=0 and

𝔹n(r)Ωp𝔹n(1),\mathbb{B}^{n}(r)\subset\Omega_{p}\subset\mathbb{B}^{n}(1),

where Ωp:=Fp(Ω)\Omega_{p}:=F_{p}(\Omega). The supremum of such r(0,1]r\in(0,1] is called the uniform squeezing number.

Remark 5.2.

There are many important examples of bounded domains satisfying the uniform squeezing property such as homogeneous domains, convex domains, strongly pseudoconvex domains (see [yeung2009, deng2016properties, kim2016uniform]).

Recall that by the famous theorem by Cheng-Yau [cheng1980existence] and Mok-Yau [mok1983completeness], every bounded pseudoconvex domain Ω\Omega admits the unique complete Kähler-Einstein metric

gΩKE=gΩ,αβ¯KE(z)dzαdzβ¯g^{\rm KE}_{\Omega}=g^{\rm KE}_{\Omega,\alpha\overline{\beta}}(z)dz_{\alpha}\otimes d\overline{z_{\beta}}

satisfying

Rαβ¯=gΩ,αβ¯KE,R_{\alpha\overline{\beta}}=-g^{\rm KE}_{\Omega,\alpha\overline{\beta}},

where Rαβ¯:=2zαzβ¯logdet(gΩ,γδ¯KE)R_{\alpha\overline{\beta}}:=-\frac{\partial^{2}}{\partial z_{\alpha}\partial\overline{z_{\beta}}}\log\det(g^{\rm KE}_{\Omega,\gamma\overline{\delta}}) is the Ricci tensor of the metric gΩKEg^{\rm KE}_{\Omega}. Then a potential function of the Kähler-Einstein metric, defined by

φΩKE:=logdet(gΩKE),\varphi^{\rm KE}_{\Omega}:=\log\det(g^{\rm KE}_{\Omega}),

is smooth strictly plurisubharmonic on Ω\Omega, where (gΩKE)(g^{\rm KE}_{\Omega}) denotes the matrix representation of the Kähler-Eintstein metric with respect to the standard Euclidean coordinates for nΩ\mathbb{C}^{n}\supset\Omega.

Theorem 5.3 (Yeung [yeung2009]).

Let Ω\Omega be a uniform squeezing domain in n\mathbb{C}^{n}. Then Ω\Omega is pseudoconvex so that admits the unique complete Kähler-Einstein metric gΩKEg^{\rm KE}_{\Omega}. Moreover, (Ω,gΩKE)(\Omega,g^{\rm KE}_{\Omega}) has bounded geometry of infinite order in the sense that for every positive integer kk, there exists a positive constant CkC_{k} (independent of pp) satisfying

φΩpKECk(𝔹n(r/2))Ck,\|\varphi^{\rm KE}_{\Omega_{p}}\|_{C^{k}(\mathbb{B}^{n}(r/2))}\leq C_{k},

where

φΩpKE(Fp(z)):=logdet(gΩpKE(Fp(z)))=log(det(gΩKE(z))/|𝒥(Fp(z))|2).{\varphi^{\rm KE}_{\Omega_{p}}(F_{p}(z))}:=\log\det(g^{\rm KE}_{\Omega_{p}}(F_{p}(z)))=\log(\det(g^{\rm KE}_{\Omega}(z))/|\mathcal{J}(F_{p}(z))|^{2}).

5.2. Tian’s Bergman sequence related to KE metric

Let 𝒟bp\mathcal{D}^{\rm bp} be a collection of bounded pseudoconvex domains in n\mathbb{C}^{n}. Define a sequence of admissible assignments mKE\mathcal{M}^{\rm KE}_{m} on 𝒟bp\mathcal{D}^{\rm bp} for m+m\in\mathbb{N}_{+} by

mKE(Ω):=μΩ,mKE:=e(m1)φΩKE=1det(gΩKE)(m1),\mathcal{M}^{\rm KE}_{m}(\Omega):=\mu^{\rm KE}_{\Omega,m}:=e^{-(m-1)\varphi^{\rm KE}_{\Omega}}=\frac{1}{{\det\left(g^{\rm KE}_{\Omega}\right)}^{(m-1)}},

for Ω𝒟bp\Omega\in\mathcal{D}^{\rm bp}. Denote the mKE\mathcal{M}^{\rm KE}_{m}-Bergman kernel by KΩ,mKE:=KΩ,μΩ,mKEK^{\rm KE}_{\Omega,m}:=K_{\Omega,\mu^{\rm KE}_{\Omega,m}} and the mKE\mathcal{M}^{\rm KE}_{m}-Bergman metric by gΩ,mKE:=gΩ,μΩ,mKEg^{\rm KE}_{\Omega,m}:=g_{\Omega,\mu^{\rm KE}_{\Omega,m}}, and the curvature by HΩ,mKE:=HΩ,μΩ,mKEH^{\rm KE}_{\Omega,m}:=H_{\Omega,\mu^{\rm KE}_{\Omega,m}}.

Proposition 5.4.

The assignment mKE\mathcal{M}^{\rm KE}_{m} is invariant and canonical of level mm so that the mKE\mathcal{M}^{\rm KE}_{m}-Bergman metrics are invariant under biholomorphisms.

Proof.

Let Ω𝒟bp\Omega\in\mathcal{D}^{\rm bp} be a bounded pseudoconvex domain. By the uniqueness of the Kähler-Einstein metric and the volume form, for any biholomorphism FF, we have

det(gΩKE(z))m=det(gF(Ω)KE(F(z)))m|𝒥(F(z))|2m.\det\left(g^{\rm KE}_{\Omega}(z)\right)^{m}=\det\left(g^{\rm KE}_{F(\Omega)}(F(z))\right)^{m}|\mathcal{J}(F(z))|^{2m}.

This implies that

μF(Ω),m+1KE(F(z))=1det(gF(Ω)KE(F(z)))m=|𝒥(F(z))|2mdet(gΩKE(z))m=|𝒥(F(z))|2mμΩ,m+1KE(z).\mu^{\rm KE}_{F(\Omega),m+1}(F(z))=\frac{1}{{\det(g^{\rm KE}_{F(\Omega)}(F(z)))}^{m}}=\frac{|\mathcal{J}(F(z))|^{2m}}{\det\left(g^{\rm KE}_{\Omega}(z)\right)^{m}}=|\mathcal{J}(F(z))|^{2m}\mu^{\rm KE}_{\Omega,m+1}(z).

Therefore, the assignment mKE\mathcal{M}^{\rm KE}_{m} is invariant and canonical of level mm. Theorem 3.7 imply that mKE\mathcal{M}^{\rm KE}_{m}-Bergman metrics are invariant under biholomorphisms. ∎

Consider the sequence of mKE\mathcal{M}^{\rm KE}_{m}-normalized Bergman kernels, metrics, and curvatures:

K~Ω,mKE:=KΩ,mKEm,g~Ω,mKE:=1mgΩ,mKE,andH~Ω,mKE:=mHΩ,mKE.\widetilde{K}^{\rm KE}_{\Omega,m}:=\sqrt[m]{K^{\rm KE}_{\Omega,m}},\ \ \ \ \widetilde{g}^{\rm KE}_{\Omega,m}:=\frac{1}{m}g^{\rm KE}_{\Omega,m},\ \ \ {\rm\ and\ }\ \ \ \widetilde{H}^{\rm KE}_{\Omega,m}:=mH^{\rm KE}_{\Omega,m}.

Let φ:=φΩKE=logdet(gΩKE)\varphi:=\varphi^{\rm KE}_{\Omega}=\log\det(g^{\rm KE}_{\Omega}). Then gφ=i¯φ=gΩKEg_{\varphi}=i\partial\overline{\partial}\varphi=g^{\rm KE}_{\Omega} and Hφ=HΩKE:=HgΩKEH_{\varphi}=H^{\rm KE}_{\Omega}:=H_{g^{\rm KE}_{\Omega}}. Therefore, Corollary 4.11 and Corollary 4.12 imply the pointwise convergences for the above sequences. For uniform squeezing domains, in fact, the convergences are uniform.

Theorem 5.5.

If Ω\Omega has the uniform squeezing property, we have the following uniform convergences:

K~Ω,mKEdet(gΩKE),g~Ω,mKEgΩKE,H~Ω,mKEHΩKE:=HgΩKE,\widetilde{K}^{\rm KE}_{\Omega,m}\rightarrow\det\left(g^{\rm KE}_{\Omega}\right),\ \ \ \widetilde{g}^{\rm KE}_{\Omega,m}\rightarrow g^{\rm KE}_{\Omega},\ \ \ \widetilde{H}^{\rm KE}_{\Omega,m}\rightarrow H^{\rm KE}_{\Omega}:=H_{g^{\rm KE}_{\Omega}},

as mm\rightarrow\infty.

Proof.

The transformation formula for the uniform squeezing map FpF_{p} implies that

KΩ,mKE(p)=KΩp,mKE(0)|𝒥(Fp(p))|2m.K^{\rm KE}_{\Omega,m}(p)=K^{\rm KE}_{\Omega_{p},m}(0)|\mathcal{J}(F_{p}(p))|^{2m}.

Therefore, for arbitrary point pΩp\in\Omega, we have

KΩ,mKE(p)det(gΩKE(p))m=KΩp,mKE(0)det(gΩpKE(0))m.\frac{K^{\rm KE}_{\Omega,m}(p)}{\det\left(g^{\rm KE}_{\Omega}(p)\right)^{m}}=\frac{K^{\rm KE}_{\Omega_{p},m}(0)}{{\det(g^{\rm KE}_{\Omega_{p}}(0))}^{m}}.

Apply Corollary 4.11 to Ωp\Omega_{p} with φΩpKE=logdet(gΩpKE)\varphi^{\rm KE}_{\Omega_{p}}=\log\det(g^{\rm KE}_{\Omega_{p}}) at the origin, we obtain

KΩp,mKE(0)det(gΩpKE(0))m=KΩp,mKE(0)emφΩpKE(0)=(mπ)n(1n2m+O(1m2)),\frac{K^{\rm KE}_{\Omega_{p},m}(0)}{{\det(g^{\rm KE}_{\Omega_{p}}(0))}^{m}}=\frac{K^{\rm KE}_{\Omega_{p},m}(0)}{e^{m\varphi^{\rm KE}_{\Omega_{p}}(0)}}=\left(\frac{m}{\pi}\right)^{n}\left(1-\frac{n}{2m}+O\Big{(}\frac{1}{m^{2}}\Big{)}\right),

since det(φΩpKE(0))=eφΩpKE(0)\det{(\varphi^{\rm KE}_{\Omega_{p}}(0))}=e^{\varphi^{\rm KE}_{\Omega_{p}}(0)} and SφΩpKE(0)=nS_{\varphi^{\rm KE}_{\Omega_{p}}}(0)=-n by the Kähler-Einstein condition. Taking the mm-th root to the above equation, we have

K~Ω,mKEdet(gΩKE)(p)=(mπ)nm(1n2m+O(1m2))1m.\frac{\widetilde{K}^{\rm KE}_{\Omega,m}}{\det(g^{\rm KE}_{\Omega})}(p)=\left(\frac{m}{\pi}\right)^{\frac{n}{m}}\left(1-\frac{n}{2m}+O\Big{(}\frac{1}{m^{2}}\Big{)}\right)^{\frac{1}{m}}.

Similarly, Corollary 4.11 implies that

g~Ω,mKE=gΩKE(1+O(1m2)),H~Ω,mKE=HΩKE(1+O(1m))\widetilde{g}^{\rm KE}_{\Omega,m}=g^{\rm KE}_{\Omega}\left(1+O\left(\frac{1}{m^{2}}\right)\right),\ \ \ \ \widetilde{H}^{\rm KE}_{\Omega,m}=H^{\rm KE}_{\Omega}\left(1+O\left(\frac{1}{m}\right)\right)

Finally, the conclusion follows from Remark 4.13 and Theorem 5.3. ∎

5.3. Tsuji’s iterative Bergman sequence

Let Ω\Omega be a bounded domain in n\mathbb{C}^{n}. Define a sequence of weight functions μΩ,mB\mu^{\rm B}_{\Omega,m} for m1m\geq 1 inductively as follows. Set μΩ,1B:=1Ω\mu^{\rm B}_{\Omega,1}:=1_{\Omega}. For the given admissible weight μΩ,mB\mu^{\rm B}_{\Omega,m}, consider the weighted Bergman space

𝒜Bm2(Ω):=𝒜μΩ,mB2(Ω)=𝒜2(Ω,μΩ,mB).\mathcal{A}^{2}_{{\rm B}_{m}}(\Omega):=\mathcal{A}^{2}_{\mu^{\rm B}_{\Omega,m}}(\Omega)=\mathcal{A}^{2}(\Omega,\mu^{\rm B}_{\Omega,m}).

Denotes the corresponding weighted Bergman kernel of 𝒜Bm2(Ω)\mathcal{A}^{2}_{{\rm B}_{m}}(\Omega) by

KΩ,mB:=KΩ,μΩ,mB.K^{\rm B}_{\Omega,m}:=K_{\Omega,\mu^{\rm B}_{\Omega,m}}.

Define the next admissible weight function by

μΩ,m+1B:=1KΩ,mB=1KΩ,μΩ,mB.\mu^{\rm B}_{\Omega,m+1}:=\frac{1}{K^{\rm B}_{\Omega,m}}=\frac{1}{K_{\Omega,\mu^{\rm B}_{\Omega,m}}}.

Inductively, consider the weighted Bergman space with respect to the above weight:

𝒜Bm+12(Ω):=𝒜μΩ,m+1B2(Ω)=𝒜2(Ω,1KΩ,mB).\mathcal{A}^{2}_{{\rm B}_{m+1}}(\Omega):=\mathcal{A}^{2}_{\mu^{\rm B}_{\Omega,m+1}}(\Omega)=\mathcal{A}^{2}(\Omega,\frac{1}{K^{\rm B}_{\Omega,m}}).

Denotes the weighted Bergman metric of the kernel KΩ,mBK^{\rm B}_{\Omega,m} by

gΩ,mB:=gΩ,μΩ,mB.g^{\rm B}_{\Omega,m}:=g_{\Omega,\mu^{\rm B}_{\Omega,m}}.

Let 𝒟b\mathcal{D}^{\rm b} be a collection of bounded domains in n\mathbb{C}^{n}. Define a sequence of admissible assignments mB\mathcal{M}^{\rm B}_{m} on 𝒟b\mathcal{D}^{\rm b} by

mB(Ω):=μΩ,mB.\mathcal{M}^{\rm B}_{m}(\Omega):=\mu^{\rm B}_{\Omega,m}.
Proposition 5.6.

The assignment mB\mathcal{M}^{\rm B}_{m} is invariant and canonical of level mm so that the mB\mathcal{M}^{\rm B}_{m}-Bergman metrics are invariant under biholomorphisms.

Proof.

Let Ω𝒟b\Omega\in\mathcal{D}^{\rm b} be a bounded domain, and let FF be a biholomorphism on Ω\Omega. By the transformation formula for the classic Bergman kernels, we have

KΩ,1B(z)=KF(Ω),1B(F(z))|𝒥(F(z))|2.K^{\rm B}_{\Omega,1}(z)=K^{\rm B}_{F(\Omega),1}(F(z)){|\mathcal{J}(F(z))|}^{2}.

This implies that

μF(Ω),2B(F(z))=1KF(Ω),1B(F(z))=|𝒥(F(z))|2KΩ,1B(z)=|𝒥(F(z))|2μΩ,2B(z).\mu^{\rm B}_{F(\Omega),2}(F(z))=\frac{1}{K^{\rm B}_{F(\Omega),1}(F(z))}=\frac{{|\mathcal{J}(F(z))|}^{2}}{K^{\rm B}_{\Omega,1}(z)}={|\mathcal{J}(F(z))|}^{2}\mu^{\rm B}_{\Omega,2}(z).

Inductively, if we have

μF(Ω),mB(F(z))=|𝒥(F(z))|2(m1)μΩ,mB(z),\mu^{\rm B}_{F(\Omega),m}(F(z))={|\mathcal{J}(F(z))|}^{2(m-1)}\mu^{\rm B}_{\Omega,m}(z),

then the transformation formula in Proposition 3.1 implies that

μF(Ω),m+1B(F(z))=1KF(Ω),mB(F(z))=|𝒥(F(z))|2mKΩ,mB(z)=|𝒥(F(z))|2mμΩ,m+1B(z).\mu^{\rm B}_{F(\Omega),m+1}(F(z))=\frac{1}{K^{\rm B}_{F(\Omega),m}(F(z))}=\frac{{|\mathcal{J}(F(z))|}^{2m}}{K^{\rm B}_{\Omega,m}(z)}={|\mathcal{J}(F(z))|}^{2m}\mu^{\rm B}_{\Omega,m+1}(z).

Therefore, Theorem 3.7 imply that mB\mathcal{M}^{\rm B}_{m}-Bergman metrics are invariant under biholomorphisms. ∎

Recall that for a bounded pseudoconvex domain Ω\Omega, we defined a sequence of admissible weight function

μΩ,mKE:=e(m1)φΩKE=1det(gΩKE)(m1).\mu^{\rm KE}_{\Omega,m}:=e^{-(m-1)\varphi^{\rm KE}_{\Omega}}=\frac{1}{{\det(g^{\rm KE}_{\Omega})}^{(m-1)}}.

Denotes the corresponding weighted Bergman space by

𝒜KEm2(Ω):=𝒜μΩ,mKE2(Ω)=𝒜2(Ω,e(m1)φΩKE).\mathcal{A}^{2}_{{\rm KE}_{m}}(\Omega):=\mathcal{A}^{2}_{\mu^{\rm KE}_{\Omega,m}}(\Omega)=\mathcal{A}^{2}(\Omega,e^{-(m-1)\varphi^{\rm KE}_{\Omega}}).

Fix a point pΩp\in\Omega and a nonzero vector XnX\in\mathbb{C}^{n}. Denotes the subsets of 𝒜Bm2(Ω)\mathcal{A}^{2}_{{\rm B}_{m}}(\Omega) and 𝒜KEm2(Ω)\mathcal{A}^{2}_{{\rm KE}_{m}}(\Omega) by

Bmj(Ω):=μΩ,mBj(Ω),KEmj(Ω):=μΩ,mKEj(Ω).\mathcal{E}^{j}_{{\rm B}_{m}}(\Omega):=\mathcal{E}^{j}_{\mu^{\rm B}_{\Omega,m}}(\Omega),\ \ \ \ \ \mathcal{E}^{j}_{{\rm KE}_{m}}(\Omega):=\mathcal{E}^{j}_{\mu^{\rm KE}_{\Omega,m}}(\Omega).

Denotes the corresponding minimizers by vBmjv^{j}_{{\rm B}_{m}} and vKEmjv^{j}_{{\rm KE}_{m}}, i.e.,

IBmj:=infBmj(Ω)uμΩ,mB2=vBmjμΩ,mB2,IKEmj:=infBmj(Ω)uμΩ,mKE2=vKEmjμΩ,mKE2.I^{j}_{{\rm B}_{m}}:=\inf_{\mathcal{E}^{j}_{{\rm B}_{m}}(\Omega)}\|u\|_{\mu^{\rm B}_{\Omega,m}}^{2}=\|v^{j}_{{\rm B}_{m}}\|^{2}_{\mu^{\rm B}_{\Omega,m}},\ \ \ \ \ I^{j}_{{\rm KE}_{m}}:=\inf_{\mathcal{E}^{j}_{{\rm B}_{m}}(\Omega)}\|u\|_{\mu^{\rm KE}_{\Omega,m}}^{2}=\|v^{j}_{{\rm KE}_{m}}\|^{2}_{\mu^{\rm KE}_{\Omega,m}}.
Lemma 5.7.

Let Ω\Omega be a bounded pseudoconvex domain in n\mathbb{C}^{n}. Suppose that μΩ,mBμΩ,mKE\frac{\mu^{\rm B}_{\Omega,m}}{\mu^{\rm KE}_{\Omega,m}} is bounded from below and above. Fix a point pΩp\in\Omega and a nonzero vector XnX\in\mathbb{C}^{n}. Then, for j=0,1,2j=0,1,2, we have

infΩ(μΩ,mBμΩ,mKE)IKEmjIBmjsupΩ(μΩ,mBμΩ,mKE)IKEmj.\inf_{\Omega}\left(\frac{\mu^{\rm B}_{\Omega,m}}{\mu^{\rm KE}_{\Omega,m}}\right)I^{j}_{{\rm KE}_{m}}\leq I^{j}_{{\rm B}_{m}}\leq\sup_{\Omega}\left(\frac{\mu^{\rm B}_{\Omega,m}}{\mu^{\rm KE}_{\Omega,m}}\right)I^{j}_{{\rm KE}_{m}}.
Proof.

By the boundedness assumption,

Bmj(Ω)=KEmj(Ω).\mathcal{E}^{j}_{{\rm B}_{m}}(\Omega)=\mathcal{E}^{j}_{{\rm KE}_{m}}(\Omega).

The definition of the minimum integrals shows that

IBmj=vBmjμΩ,mB2vKEmjμΩ,mB2supΩ(μΩ,mBμΩ,mKE)vKEmjμΩ,mKE2=supΩ(μΩ,mBμΩ,mKE)IKEmj.I^{j}_{{\rm B}_{m}}=\|v^{j}_{{\rm B}_{m}}\|^{2}_{\mu^{\rm B}_{\Omega,m}}\leq\|v^{j}_{{\rm KE}_{m}}\|^{2}_{\mu^{\rm B}_{\Omega,m}}\leq\sup_{\Omega}\left(\frac{\mu^{\rm B}_{\Omega,m}}{\mu^{\rm KE}_{\Omega,m}}\right)\|v^{j}_{{\rm KE}_{m}}\|^{2}_{\mu^{\rm KE}_{\Omega,m}}=\sup_{\Omega}\left(\frac{\mu^{\rm B}_{\Omega,m}}{\mu^{\rm KE}_{\Omega,m}}\right)I^{j}_{{\rm KE}_{m}}.

Similarly, we have

IBmj=vBmjμΩ,mB2infΩ(μΩ,mBμΩ,mKE)vBmjμΩ,mKE2infΩ(μΩ,mBμΩ,mKE)vKEmjμΩ,mKE2=infΩ(μΩ,mBμΩ,mKE)IKEmj.I^{j}_{{\rm B}_{m}}=\|v^{j}_{{\rm B}_{m}}\|^{2}_{\mu^{\rm B}_{\Omega,m}}\geq\inf_{\Omega}\left(\frac{\mu^{\rm B}_{\Omega,m}}{\mu^{\rm KE}_{\Omega,m}}\right)\|v^{j}_{{\rm B}_{m}}\|^{2}_{\mu^{\rm KE}_{\Omega,m}}\geq\inf_{\Omega}\left(\frac{\mu^{\rm B}_{\Omega,m}}{\mu^{\rm KE}_{\Omega,m}}\right)\|v^{j}_{{\rm KE}_{m}}\|^{2}_{\mu^{\rm KE}_{\Omega,m}}=\inf_{\Omega}\left(\frac{\mu^{\rm B}_{\Omega,m}}{\mu^{\rm KE}_{\Omega,m}}\right)I^{j}_{{\rm KE}_{m}}.

Proposition 5.8.

Set Lm:=infΩ(μΩ,mBμΩ,mKE)L_{m}:=\inf\limits_{\Omega}\left(\frac{\mu^{\rm B}_{\Omega,m}}{\mu^{\rm KE}_{\Omega,m}}\right) and Um:=supΩ(μΩ,mBμΩ,mKE)U_{m}:=\sup\limits_{\Omega}\left(\frac{\mu^{\rm B}_{\Omega,m}}{\mu^{\rm KE}_{\Omega,m}}\right). Then we have

1UmKΩ,mKEKΩ,mB1LmKΩ,mKE,LmUmgΩ,mKEgΩ,mBUmLmgΩ,mKE,\frac{1}{U_{m}}K^{\rm KE}_{\Omega,m}\leq K^{\rm B}_{\Omega,m}\leq\frac{1}{L_{m}}K^{\rm KE}_{\Omega,m},\ \ \ \ \ \frac{L_{m}}{U_{m}}g^{\rm KE}_{\Omega,m}\leq g^{\rm B}_{\Omega,m}\leq\frac{U_{m}}{L_{m}}g^{\rm KE}_{\Omega,m},
(LmUm)2(2HΩ,mKE)2HΩ,mB(UmLm)2(2HΩ,mKE).\left(\frac{L_{m}}{U_{m}}\right)^{2}\left(2-H^{\rm KE}_{\Omega,m}\right)\leq 2-H^{\rm B}_{\Omega,m}\leq\left(\frac{U_{m}}{L_{m}}\right)^{2}\left(2-H^{\rm KE}_{\Omega,m}\right).
Proof.

Apply Lemma 5.7 and Theorem 2.10. ∎

Remark 5.9.

Note that if limmLmm=limmUmm=1\lim\limits_{m\rightarrow\infty}\sqrt[m]{L_{m}}=\lim\limits_{m\rightarrow\infty}\sqrt[m]{U_{m}}=1, the convergence of Tian’s sequence of normalized Bergman kernels K~Ω,mKE\widetilde{K}^{\rm KE}_{\Omega,m} implies the convergence of KΩ,mBm\sqrt[m]{K^{\rm B}_{\Omega,m}}. Moreover, if limmUmLm=1\lim\limits_{m\rightarrow\infty}\frac{U_{m}}{L_{m}}=1, then the convergence of Tian’s sequence of normalized Bergman metrics g~Ω,mKE\widetilde{g}^{\rm KE}_{\Omega,m} implies the convergence of 1mgΩ,mB\frac{1}{m}g^{\rm B}_{\Omega,m}.

Lemma 5.10.

Let Ω\Omega be a uniform squeezing domain in n\mathbb{C}^{n} with the squeezing number r(0,1]r\in(0,1]. For any integer m1m\geq 1, there exist uniform constants C,Dm>0C,D_{m}>0 satisfying

(πr2)nmCmDmμΩ,m+1BμΩ,m+1KECmπnmDm.\frac{(\pi r^{2})^{nm}}{C^{m}}D_{m}\leq\frac{\mu^{\rm B}_{\Omega,m+1}}{\mu^{\rm KE}_{\Omega,m+1}}\leq C^{m}\pi^{nm}D_{m}.
Proof.

For any point pΩp\in\Omega,

μΩ,m+1BμΩ,m+1KE(p)=det(gΩKE)mKΩ,mB(p)=det(gΩpKE)mKΩp,mB(0).\frac{\mu^{\rm B}_{\Omega,m+1}}{\mu^{\rm KE}_{\Omega,m+1}}(p)=\frac{{\det(g^{\rm KE}_{\Omega})}^{m}}{K^{\rm B}_{\Omega,m}}(p)=\frac{{\det(g^{\rm KE}_{\Omega_{p}})}^{m}}{K^{\rm B}_{\Omega_{p},m}}(0).

Note that the uniform squeezing property implies that there exists a positive constant C>0C>0 satisfying

1Cdet(gΩpKE)(0)C.\frac{1}{C}\leq\det(g^{\rm KE}_{\Omega_{p}})(0)\leq C.

Moreover, the squeezing property yields that

K𝔹n(1),mB(0)KΩp,mB(0)K𝔹n(r),mB(0).K^{\rm B}_{\mathbb{B}^{n}(1),m}(0)\leq K^{\rm B}_{\Omega_{p},m}(0)\leq K^{\rm B}_{\mathbb{B}^{n}(r),m}(0).

The computation in Section 2.3 shows that

K𝔹n(r),mB(0)\displaystyle K^{\rm B}_{\mathbb{B}^{n}(r),m}(0) =k=0m11ck(r)=k=0m1(1(πr2)n(k(n+1)+n)!(k(n+1))!)\displaystyle=\prod^{m-1}_{k=0}\frac{1}{c_{k}(r)}=\prod^{m-1}_{k=0}\left(\frac{1}{(\pi r^{2})^{n}}\frac{(k(n+1)+n)!}{(k(n+1))!}\right)
=1(πr2)nm(mn+m1)!(m1)!(n+1)m1=:1(πr2)nm1Dm.\displaystyle=\frac{1}{(\pi r^{2})^{nm}}\frac{(mn+m-1)!}{(m-1)!(n+1)^{m-1}}=:\frac{1}{(\pi r^{2})^{nm}}\frac{1}{D_{m}}.

Remark 5.11.

Unfortunately, limmDm=0\lim\limits_{m\rightarrow\infty}D_{m}=0 so that limmLmm=limmUmm=0\lim\limits_{m\rightarrow\infty}\sqrt[m]{L_{m}}=\lim\limits_{m\rightarrow\infty}\sqrt[m]{U_{m}}=0. However, since limmmnDm1/m=(en+1)n\lim\limits_{m\rightarrow\infty}m^{n}D_{m}^{1/m}=\left(\frac{e}{n+1}\right)^{n}, a normalization of the convergence speed implies that

r2nC(en+1)nlimm(mπ)nmμΩ,m+1BμΩ,m+1KEmC(en+1)n.\frac{r^{2n}}{C}\left(\frac{e}{n+1}\right)^{n}\leq\lim_{m\rightarrow\infty}\sqrt[m]{\left(\frac{m}{\pi}\right)^{nm}\frac{\mu^{\rm B}_{\Omega,m+1}}{\mu^{\rm KE}_{\Omega,m+1}}}\leq C\left(\frac{e}{n+1}\right)^{n}.

As in the case of compact canonically polarized manifolds (cf. [tsuji2010, song2010]), for convergences, we consider a modified sequence of weight functions μ~Ω,mB\widetilde{\mu}^{\rm B}_{\Omega,m} for m1m\geq 1 inductively as follows. Set μ~Ω,1B:=1Ω\widetilde{\mu}^{\rm B}_{\Omega,1}:=1_{\Omega}. For the given admissible weight μ~Ω,mB\widetilde{\mu}^{\rm B}_{\Omega,m}, consider the weighted Bergman space and the corresponding Bergman kernel:

𝒜μ~Ω,mB2(Ω)=𝒜2(Ω,μ~Ω,mB),KΩ,mB~:=KΩ,μ~Ω,mB.\mathcal{A}^{2}_{\widetilde{\mu}^{\rm B}_{\Omega,m}}(\Omega)=\mathcal{A}^{2}(\Omega,\widetilde{\mu}^{\rm B}_{\Omega,m}),\ \ \ \ K^{\rm\widetilde{B}}_{\Omega,m}:=K_{\Omega,\widetilde{\mu}^{\rm B}_{\Omega,m}}.

Define the next admissible weight function by

μ~Ω,m+1B:=(mπ)n(1n2m)1KΩ,mB~=(mπ)n(1n2m)1KΩ,μ~Ω,mB.\widetilde{\mu}^{\rm B}_{\Omega,m+1}:=\left(\frac{m}{\pi}\right)^{n}\left(1-\frac{n}{2m}\right)\frac{1}{K^{\rm\widetilde{B}}_{\Omega,m}}=\left(\frac{m}{\pi}\right)^{n}\left(1-\frac{n}{2m}\right)\frac{1}{K_{\Omega,\widetilde{\mu}^{\rm B}_{\Omega,m}}}.

Following Berndtsson’s idea in [berndtsson2009], we used the normalization factor (mπ)n(1n2m)\left(\frac{m}{\pi}\right)^{n}\left(1-\frac{n}{2m}\right) rather than (mπ)n\left(\frac{m}{\pi}\right)^{n} for a better convergence rate (cf. [tsuji2010, song2010]). The below theorem generalize Theorem 1.2 in [tsuji2013dynamical].

Theorem 5.12.

Let Ω\Omega be a uniform squeezing domain in n\mathbb{C}^{n}. There exists a uniform constant C>0C>0 satisfying

eCmμ~Ω,m+1BμΩ,m+1KEmeCm.e^{-\frac{C}{m}}\leq\sqrt[m]{\frac{\widetilde{\mu}^{\rm B}_{\Omega,m+1}}{\mu^{\rm KE}_{\Omega,m+1}}}\leq e^{\frac{C}{m}}.

In particular, we have the following uniform convergence:

limmKΩ,mB~m=det(gΩKE).\lim_{m\rightarrow\infty}\sqrt[m]{K^{\rm\widetilde{B}}_{\Omega,m}}=\det\left(g^{\rm KE}_{\Omega}\right).
Proof.

Let m0m_{0} be the constant in Theorem 4.10 and Corollary 4.11. Lemma 5.10 implies that there exist uniform constants Lm0,Um0>0L_{m_{0}},U_{m_{0}}>0 satisfying

Lm0μ~Ω,m0BμΩ,m0KEUm0.L_{m_{0}}\leq\frac{\widetilde{\mu}^{\rm B}_{\Omega,m_{0}}}{\mu^{\rm KE}_{\Omega,m_{0}}}\leq U_{m_{0}}.

For the proof, we use mathematical induction on mm. Let mm0m\geq m_{0}. Assume that there exist uniform constants L~m,U~m>0\widetilde{L}_{m},\widetilde{U}_{m}>0 satisfying

(5.1) L~mμ~Ω,mBμΩ,mKEU~m.\widetilde{L}_{m}\leq\frac{\widetilde{\mu}^{\rm B}_{\Omega,m}}{\mu^{\rm KE}_{\Omega,m}}\leq\widetilde{U}_{m}.

From the above assumption, we want to show that there exists a constant C>0C>0 satisfying

L~m(1Cm2)μ~Ω,m+1BμΩ,m+1KEU~m(1+Cm2).\widetilde{L}_{m}\left(1-\frac{C}{m^{2}}\right)\leq\frac{\widetilde{\mu}^{\rm B}_{\Omega,m+1}}{\mu^{\rm KE}_{\Omega,m+1}}\leq\widetilde{U}_{m}\left(1+\frac{C}{m^{2}}\right).

Apply the same proof of Lemma 5.7 to μ~Ω,mBμΩ,mKE\frac{\widetilde{\mu}^{\rm B}_{\Omega,m}}{\mu^{\rm KE}_{\Omega,m}}. Then, the induction hypothesis (5.1) imply that for any pΩp\in\Omega,

(5.2) L~mIKEm0(p)IB~m0(p)U~mIKEm0(p).\widetilde{L}_{m}I^{0}_{{\rm KE}_{m}}(p)\leq I^{0}_{{\rm\widetilde{B}}_{m}}(p)\leq\widetilde{U}_{m}I^{0}_{{\rm KE}_{m}}(p).

Recall that for all mm0m\geq m_{0}, we have

IKEm0(p)=1KΩ,mKE(p)\displaystyle I^{0}_{{\rm KE}_{m}}(p)=\frac{1}{K^{\rm KE}_{\Omega,m}(p)} =1det(gΩKE(p))m(πm)n(1+n2m+O(1m2))\displaystyle=\frac{1}{{\det(g^{\rm KE}_{\Omega}(p))}^{m}}\left(\frac{\pi}{m}\right)^{n}\left(1+\frac{n}{2m}+O\Big{(}\frac{1}{m^{2}}\Big{)}\right)
=μΩ,m+1KE(p)(πm)n(1+n2m+O(1m2)).\displaystyle=\mu^{\rm KE}_{\Omega,m+1}(p)\left(\frac{\pi}{m}\right)^{n}\left(1+\frac{n}{2m}+O\Big{(}\frac{1}{m^{2}}\Big{)}\right).

This implies that

IKEm0(p)(mπ)n(1n2m)=μΩ,m+1KE(p)(1+O(1m2)).I^{0}_{{\rm KE}_{m}}(p)\left(\frac{m}{\pi}\right)^{n}\left(1-\frac{n}{2m}\right)=\mu^{\rm KE}_{\Omega,m+1}(p)\left(1+O\Big{(}\frac{1}{m^{2}}\Big{)}\right).

In other words, there exists a constant C>0C>0 satisfying

(5.3) μΩ,m+1KE(p)(1Cm2)IKEm0(p)(mπ)n(1n2m)μΩ,m+1KE(p)(1+Cm2).\mu^{\rm KE}_{\Omega,m+1}(p)\left(1-\frac{C}{m^{2}}\right)\leq I^{0}_{{\rm KE}_{m}}(p)\left(\frac{m}{\pi}\right)^{n}\left(1-\frac{n}{2m}\right)\leq\mu^{\rm KE}_{\Omega,m+1}(p)\left(1+\frac{C}{m^{2}}\right).

On the other hand, by the definition, we have

IB~m0(p)=1KΩ,mB~(p)=1(mπ)n(1n2m)μ~Ω,m+1B(p).I^{0}_{{\rm\widetilde{B}}_{m}}(p)=\frac{1}{K^{\rm\widetilde{B}}_{\Omega,m}(p)}=\frac{1}{\left(\frac{m}{\pi}\right)^{n}\left(1-\frac{n}{2m}\right)}\widetilde{\mu}^{\rm B}_{\Omega,m+1}(p).

From the inequalities (5.2) and (5.3), we obtain that

L~m(1Cm2)μ~Ω,m+1BμΩ,m+1KEU~m(1+Cm2)\widetilde{L}_{m}\left(1-\frac{C}{m^{2}}\right)\leq\frac{\widetilde{\mu}^{\rm B}_{\Omega,m+1}}{\mu^{\rm KE}_{\Omega,m+1}}\leq\widetilde{U}_{m}\left(1+\frac{C}{m^{2}}\right)

as we required, which completes the induction step. Therefore, it follows that

Lm0k=m0m(1Ck2)μ~Ω,m+1BμΩ,m+1KEUm0k=m0m(1+Ck2).L_{m_{0}}\prod_{k=m_{0}}^{m}\left(1-\frac{C}{k^{2}}\right)\leq\frac{\widetilde{\mu}^{\rm B}_{\Omega,m+1}}{\mu^{\rm KE}_{\Omega,m+1}}\leq U_{m_{0}}\prod_{k=m_{0}}^{m}\left(1+\frac{C}{k^{2}}\right).

Finally, the conclusion follows from the following upper bound estimates:

log(k=m0m(1+Ck2))k=1m(log(1+Ck2))Ck=1m1k2C′′.\log\left(\prod_{k=m_{0}}^{m}\left(1+\frac{C}{k^{2}}\right)\right)\leq\sum_{k=1}^{m}\left(\log\left(1+\frac{C}{k^{2}}\right)\right)\leq C^{\prime}\sum_{k=1}^{m}\frac{1}{k^{2}}\leq C^{\prime\prime}.

Changing the signs of the above inequalities yields the lower bound estimates. ∎

Acknowledgements. The author would like to thank Professor Kang-Tae Kim for his suggestion of this work and valuable comments. He also would like to thank Professor Jun-Muk Hwang for his encouragement to write this paper, and Professor Bo Berndtsson for sharing the note [berndtsson2009]. This work was supported by Incheon National University Research Grant in 2022.

References