Izvestiya: Mathematics
RUS  ENG    JOURNALS   PEOPLE   ORGANISATIONS   CONFERENCES   SEMINARS   VIDEO LIBRARY   PACKAGE AMSBIB  
General information
Latest issue
Forthcoming papers
Archive
Impact factor
Guidelines for authors
Submit a manuscript

Search papers
Search references

RSS
Latest issue
Current issues
Archive issues
What is RSS



Izv. RAN. Ser. Mat.:
Year:
Volume:
Issue:
Page:
Find






Personal entry:
Login:
Password:
Save password
Enter
Forgotten password?
Register


Izvestiya: Mathematics, 2023, Volume 87, Issue 5, Pages 987–1010
DOI: https://doi.org/10.4213/im9336e
(Mi im9336)
 

On the positivity of direct image bundles

Zhi Lia, Xiangyu Zhoubc

a School of Science, Beijing University of Posts and Telecommunications, Beijing, China
b Institute of Mathematics, Academy of Mathematics and Systems Science, Chinese Academy of Sciences, Beijing, China
c Hua Loo-Keng Key Laboratory of Mathematics, Chinese Academy of Sciences, Beijing, China
References:
Abstract: In the present paper, we obtain an equivalent relation between the log-plurisubharmonicity of the relative Bergman kernel, the Griffiths and Nakano positivity for the direct image with the natural $L^2$ metric, by finding a converse of Berndtsson's theorem on the direct image. A converse of Berndtsson's generalization of Kiselman minimal principle is also obtained.
Keywords: $L^2$-methods, plurisubharmonic functions, direct images, positive hermitian holomorphic vector bundles, minimal principles, relative Bergman kernel.
Funding agency Grant number
National Key Research and Development Program of China 2021YFA1002600
2021YFA1003100
National Natural Science Foundation of China 12201060
12288201
This research is supported by National Key R&D Program of China (2021YFA1002600 and 2021YFA1003100). The first author is partially supported by the NSFC grant (no. 12201060), the second author is supported by the NSFC grant (no. 12288201).
Received: 22.03.2022
Revised: 17.04.2022
Russian version:
Izvestiya Rossiiskoi Akademii Nauk. Seriya Matematicheskaya, 2023, Volume 87, Issue 5, Pages 140–163
DOI: https://doi.org/10.4213/im9336
Bibliographic databases:
Document Type: Article
UDC: 517.550.7+517.553+517.554
Language: English
Original paper language: English

Dedicated to the centenary of V. S. Vladimirov

§ 1. Introduction

Let $H \times D \subset \mathbb{C}_{\tau}^m \times \mathbb{C}^n_z$ be a connected open subset and $\varphi$ be a function on $H \times D$ which is smooth up to the boundary. For each fixed ${\tau} \in H$, denote the weighted Bergman space by

$$ \begin{equation*} A^2 (D, \varphi_{\tau}) := \biggl\{ f \in \mathcal{O}(D) \colon \int_D |f|^2 e^{-\varphi_{\tau}} \, dV < \infty \biggr\}, \end{equation*} \notag $$
where $\varphi_{\tau} (z) = \varphi ({\tau}, z)$ and $dV$ is the Lebesgue measure on $\mathbb{C}^n$.

By the assumption that $\varphi$ is smooth up to the boundary, we have for all ${\tau} \in H$, $A^2 (D, \varphi_{\tau})$ are equal as vector spaces. Let $(\underline{\mathbb{C}},e^{-\varphi})\to H \times D$ denote the hermitian trivial holomorphic line bundle, and therefore, we can regard $\coprod_{{\tau} \in H} A^2 (D, \varphi_{\tau})$ as a trivial vector bundle over $H$ equipped with the natural $L^2$ metric, varies with ${\tau} \in H$, defined as for every $f \in A^2 (D, \varphi_{\tau})$,

$$ \begin{equation*} \|f\|^2_{\tau} = \int_D |f|^2 e^{- \varphi_{\tau}}\, dV. \end{equation*} \notag $$

It turns out that the bundle is a hermitian holomorphic vector bundle of infinite rank over $H$, which is called the direct image bundle over $H$ of the line bundle $(\underline{\mathbb{C}}, e^{-\varphi}) \to H \times D$. Throughout the present paper, a domain means a connected open set and the notation $\operatorname{PSH} (U)$ means the set of plurisubharmonic functions on a domain $U$.

A celebrated theorem of Berndtsson [1] asserts that if $\varphi$ is (strictly) plurisubharmonic and $D$ is pseudoconvex, then the direct image bundle is (positive) semi-positive in the sense of Nakano.

We consider the following setting. Let $D \subset \mathbb{C}^n$ be a bounded pseudoconvex domain and $\varphi \in \operatorname{PSH} (D) \cap C^{\infty} (\overline{D})$, for any domain $H \subset \mathbb{C}^m$ and any $\psi \in \operatorname{PSH} (H \times D) \cap C^{\infty} (\overline{H \times D})$. Take a trivial line bundle $(\underline{\mathbb{C}},e^{-k\pi^{\ast}\varphi-\psi})\to H\times D$, where $\pi \colon H \times D \to D$ is the projection to the second coordinate, then the direct image bundle is semi-positive in the sense of Nakano for any positive integer $k$, and thus, semi-positive in the sense of Griffiths.

In particular, the fiberwised Bergman kernel $B_{k,\tau} (z)$ weighted by $k \pi^{\ast} \varphi ({\tau}, {\cdot}\,)+\psi ({\tau}, {\cdot}\,)$ is log-plurisubharmonic in $(\tau, z)$. The latter result follows from Berndtsson’s log-plurisubharmonicity theorem (see [2]).

In general, Griffiths positivity could not imply Nakano positivity. A natural question arises from the theorem of Berndtsson as follows.

Question 1.1. In the above setting, for the direct image with the natural $L^2$ metric, are Griffiths (semi-)positivity and (semi-)Nakano positivity equivalent?

The first result in the present paper gives an answer to the above question by establishing a converse of the theorem of Berndtsson.

Theorem 1.2. Let $D \subset \mathbb{C}^n_z$ be a bounded pseudoconvex domain and $\varphi\colon D\to [-\infty,+\infty)$ be a function such that $\varphi \in L^1_{\mathrm{loc}}$, and, for any set $E$ of measure zero,

$$ \begin{equation*} \limsup_{z \to z_0,\, z \in D \setminus E} \varphi (z) = \varphi(z_0) \end{equation*} \notag $$
for any $z_0 \in D$. Assume further for any positive integer $k$, the Bergman kernel weighted by $k \varphi$ does not vanish on $D$. Then the following statements are equivalent:

(1) $\varphi \in \operatorname{PSH}(D)$;

(2) for any domain $H \subset \mathbb{C}^m_{\tau}$ and any locally bounded $\psi \in \operatorname{PSH} (H \times D)$, the Bergman kernel $B_{k, \tau} (z)$ of the direct image of the trivial line bundle $(\underline{\mathbb{C}}, e^{- k \pi^{\ast} \varphi - \psi})\to H \times D$ is log-plurisubharmonic;

(3) for any domain $H \subset \mathbb{C}^m_{\tau}$ and any locally bounded $\psi \in \operatorname{PSH} (H \times D)$, the direct image of the trivial line bundle $(\underline{\mathbb{C}}, e^{- k \pi^{\ast} \varphi - \psi}) \to H \times D$ is semi-positive in the sense of Griffiths;

(4) for any domain $H \subset \mathbb{C}^m_{\tau}$ and any locally bounded $\psi \in \operatorname{PSH} (H \times D)$, the direct image of the trivial line bundle $(\underline{\mathbb{C}}, e^{- k \pi^{\ast} \varphi - \psi}) \to H \times D$ is semi-positive in the sense of Nakano.

Remark 1.3. (1) In the above theorem, since $\psi$ is locally bounded, the direct image of the trivial line bundle $(\underline{\mathbb{C}}, e^{- k \pi^{\ast} \varphi - \psi}) \to H \times D$ is a trivial vector bundle and locally the fibers of $F$ are equal to $A^2(D,e^{-\varphi})$ as vector spaces.

(2) Usually the notion of positivity is defined for smooth hermitian metrics. In view of the recent works [1], [3]–[6], it is possible to define the positivity of Griffiths or Nakano for singular metrics and which will be explained in the next section.

(3) It is usually impossible to derive Nakano positivity from Griffiths positivity even in the situation of smooth hermitian metrics, see [7]. However, Theorem 1.2 indicates that in this case, one can get Nakano positivity from a certain condition involved with Griffiths positivity.

(4) The condition that $D$ is pseudoconvex can be omitted in the direction $(2)\Rightarrow (1)$.

Berndtsson’s theorem was obtained in [1] by Hörmander’s $L^2$-estimate for $\overline{\partial}$-operator. Recently Deng, Ning, Wang, and Zhou in [5] found a general criterion for the Nakano positivity for holomorphic vector bundles by the $L^p$-estimate for $\overline{\partial}$-operator. With this criterion, a generalized version of Berndtsson’s theorem with a lower-bounded estimate has been obtained.

As Nakano positivity implies Griffiths positivity, the direct image $F$ is also positive in the sense of Griffiths. In particular, the Bergman kernel $B_t (z)$ weighted by $\varphi_t$ is log-plurisubharmonic. This is also known as Berndtsson’s log-plurisubharmonicity theorem, see [2]. Guan–Zhou (see [8], [9]) proved Berndtsson’s log-plurisubharmonicity theorem and Griffiths positivity of the direct images by the $L^2$-extension theorem with optimal estimate. It should be noted that the $L^2$-extension theorem only with optimal estimate plays an essential role in Guan–Zhou’s argument. In [10], it is shown that a version of $L^2$-extension theorem with optimal estimate can also be derived by the theorem of Berndtsson. These results give unexpected relations between log-plurisubharmonicity, the optimal $L^2$-extension theorem, and positivity of the direct image.

Later, Deng, Wang, Zhang, and Zhou [4] found a new characterization for plurisubharmonic functions, and with this result they proved the Griffiths positivity of the direct image with the usual $L^2$-extension theorem. Thus, the Griffiths positivity and the $L^2$-extension theorem with optimal estimate is somewhat equivalent.

Another closely related topic is the minimal principle. Prékopa [11] found a version of the minimal principle in convex analysis. Later on, a celebrated theorem of Kiselman [12] generalizes Prékopa’s minimal principle to plurisubharmonic functions. In [13], Berndtsson found the following generalization of Kiselman’s minimal principle in an integral form.

Let $\varphi (\tau, z)$ be a plurisubharmonic function on $U_\tau \times V_z \subset \mathbb{C}^m_\tau \times \mathbb{C}^n_z$, where $V_z$ is pseudoconvex.

(1) Assume that $V$ is a Reinhardt domain and that $\varphi$ is independent of $\arg (z_j)$, $j = 1, 2, \dots, n$. Define $\widetilde{\varphi}$ by

$$ \begin{equation*} e^{- \widetilde{\varphi} (\tau)} = \int_V e^{- \varphi (\tau, z)}\, dV_z, \end{equation*} \notag $$
then $\widetilde{\varphi} \in \operatorname{PSH}(U)$. Moreover, the same conclusion holds if we only assume that $V$ and $\varphi (\tau, z)$ are invariant under the group action of $S^1$: $z \mapsto e^{i \theta} z$ for any $\theta \in \mathbb{R}$, provided $V$ contains the origin.

(2) Assume $V$ is a tube domain

$$ \begin{equation*} V = X+i\mathbb{R}^m \end{equation*} \notag $$
and that $\varphi$ is independent of $\operatorname{Im} (z_j)$, $j = 1, 2, \dots, n$. Define $\widetilde{\varphi}$ by
$$ \begin{equation*} e^{- \widetilde{\varphi} (\tau)} = \int_X e^{- \varphi (\tau, \mathrm{Rez})}\, dV_{\mathrm{Rez}}, \end{equation*} \notag $$
then $\widetilde{\varphi} \in \operatorname{PSH} (U)$.

Let us consider a case of the above minimal principle. Let $D \subset \mathbb{C}^n$ be a pseudoconvex Reinhardt domain containing the origin and $\varphi \in \operatorname{PSH} (D)$. For any domain $H \subset \mathbb{C}^m_{\tau}$ and any $\psi \in \operatorname{PSH} (H \times D)$, the function $\widetilde{\psi}$, as defined by

$$ \begin{equation*} e^{- \widetilde{\psi} (\tau)} = \int_D e^{- k \pi^{\ast} \varphi (\tau, z) - \psi(\tau, z)} \, dV_z, \end{equation*} \notag $$
is plurisubharmonic for any $k > 0$, provided $\psi$ and $\varphi$ are invariant with respect to the group action of $S^1$: the maps $z \mapsto e^{i \theta} z$ for $\theta \in \mathbb{R}$.

Another natural question asks whether there exists a converse of the above minimal principle.

Question 1.4. Let $D \subset \mathbb{C}^n$ be a domain and $\varphi$ be a function on $D$. Can one derive the plurisubharmonicity of $\varphi$ from the plurisubharmonicity of $\widetilde{\psi}$ defined in the form of

$$ \begin{equation*} e^{- \widetilde{\psi} (\tau)} = \int_D e^{- \pi^{\ast} \varphi (\tau, z) - \psi(\tau, z)} \,dV_z? \end{equation*} \notag $$

The second result of this paper provides an answer to this question by showing that a necessary condition is also sufficient for a function to be plurisubharmonic. Actually, we could have a more general version for the converse.

Theorem 1.5. Let $D \subset \mathbb{C}^n$ be a bounded domain and $\varphi$ be a function on $D$ such that $\varphi \in L^1_{\mathrm{loc}}$, and for any set $E$ of measure zero,

$$ \begin{equation*} \limsup_{z \to z_0,\, z \in D \setminus E} \varphi (z) = \varphi(z_0) \end{equation*} \notag $$
for any $z_0 \in D$. If, for any domain $H \subset \mathbb{C}^m_{\tau}$ and any plurisubharmonic function $\psi (\tau, z)$ on $H_{\tau} \times D_z$, the function $\widetilde{\psi} (\tau)$ defined by
$$ \begin{equation*} e^{- \widetilde{\psi} (\tau)} = \int_D e^{- k \pi^{\ast} \varphi (\tau, z) - \psi (\tau, z)}\, dV_z \end{equation*} \notag $$
is plurisubharmonic for any $k > 0$, then $\varphi$ is plurisubharmonic on $D$, provided
$$ \begin{equation*} \int_D e^{- k \varphi}\, dV_z \end{equation*} \notag $$
is finite for every $k > 0$.

The Berndtsson’s generalization of Kiselman’s minimal principle was proved in [13] by a converse version of Hörmander’s $L^2$-estimates in dimension one and also can be easily derived from [5]. Inspired by Ball, Barthe, and Naor [14], Cordero–Erausquin [15] proved the Kiselman–Berndtsson’s minimal principle by computing the complex Hessian directly by the solution to the $L^2$-existence of $\overline{\partial}$-operator due to Hörmander [16] and Kohn (see [17]). Also Cordero–Erausquin found that the condition of invariance for $V$ and $\varphi$ under the maps $z \mapsto e^{i \theta}$, for any $\theta \in \mathbb{R}$, can be relaxed. And that is why in Theorem 1.5 the condition on the invariance for $D$, $\varphi$ or $\psi$ is omitted.

Remark 1.6. The Kiselman–Berndtsson’s minimal principle also follows from the theorem of Berndtsson [1]. By invariance of $\varphi$ and $V$, the Bergman kernel $B_\tau (\zeta, 0)$ with respect to $\varphi (t, z)$ for a fixed $\tau$ is just a constant, which is equal to

$$ \begin{equation*} \biggl( \int_V e^{- \varphi (\tau, z)}\, dV_z \biggr)^{- 1}, \end{equation*} \notag $$
and it follows from the theorem of Berndtsson that
$$ \begin{equation*} \widetilde{\varphi} (\tau) = - \log \int_V e^{- \varphi (\tau, z)}\, dV_z \end{equation*} \notag $$
is plurisubharmonic.

§ 2. Preliminaries

In this section we will recall some notions and results which are needed in proving Theorem 1.2 and Theorem 1.5.

2.1. $L^2$-estimates for $\overline{\partial}$-operator

The following $L^2$-existence theorem is a fundamental result in the $L^2$-methods.

Theorem 2.1 (Hörmander’s $L^2$-existence theorem, see [7], [16]). Let $D \subset \mathbb{C}^n$ be a pseudoconvex domain and $\varphi$ be a plurisubharmonic function on $D$ such that

$$ \begin{equation*} i\, \partial\, \overline{\partial} \varphi \geqslant \theta \end{equation*} \notag $$
in the sense of distributions for some continuous positive $(1, 1)$-form $\theta$ on $D$. Then, for any $(n, 1)$-form $f$ on $D$ such that $\overline{\partial} f = 0$ with
$$ \begin{equation*} \int_D | f |^2_{\theta} e^{- \varphi} < \infty, \end{equation*} \notag $$
there exists an $(n, 0)$-form $u \in L^2 (D, e^{- \varphi})$ satisfying $\overline{\partial} u = f$ and
$$ \begin{equation*} \int_D | u |^2 e^{- \varphi} \leqslant \int_D | f |^2_{\theta} e^{-\varphi}. \end{equation*} \notag $$

If we write $\theta = i \sum_{j, k} \theta_{j \overline{k}}\,\, d z_j \wedge d \overline{z}_k$ and $f = \sum_j f_{\overline{j}}\,\, d z \wedge d \overline{z}_j$, $d z = d z_1 \wedge d z_2 \wedge \cdots \wedge d z_n$. Since $(\theta_{j \overline{k}})$ is positive definite, the inverse of $(\theta_{j \overline{k}})$ is well-defined and will be denoted as $(\theta^{\overline{k} j})$. Therefore, the norm of $f$ with respect to $\theta$ is defined to be

$$ \begin{equation*} \int_D | f |^2_{\theta} e^{- \varphi} = \int_D \sum_{j, k} \theta^{\overline{k} j} f_j \overline{f}_k e^{- \varphi}. \end{equation*} \notag $$

Hörmander’s $L^2$-existence theorem (for a pseudoeffective trivial line bundle $(\underline{\mathbb{C}}, e^{- \varphi}))$ was generalized by Demailly [18] (see also [19]) to a Nakano positive hermitian holomorphic vector bundle on a complete Kähler manifold.

Recently in [5], the authors have introduced a notion of a hermitian holomorphic vector bundles satisfying the $L^2$-estimates for $\overline{\partial}$-operator, and established a converse version of Theorem 2.1. For the sake of completeness, we state Deng, Ning, Wang, Zhou’s results in [5] as follows.

Definition 2.2 (see [5]). Let $(X, \omega)$ be a Kähler manifold of dimension $n$, which admits a positive hermitian holomorphic line bundle and $p > 0$. We say a hermitian holomorphic vector bundle $(E, h)$ (may be of infinite rank) satisfies the optimal $L^p$-estimate if for any positive holomorphic hermitian line bundle $(A, h_A)$ over $X$, for any $f \in C_{\mathrm{c}}^{\infty} \bigl(X, \bigwedge^{n, 1} T^{\ast}_X \otimes E \otimes A\bigr)$ with $\overline{\partial} f = 0$, there exists $u \in L^p \bigl(X, \bigwedge^{n, 0} T^{\ast}_X \otimes E \otimes A\bigr)$ such that

$$ \begin{equation*} \overline{\partial} u = f \end{equation*} \notag $$
and
$$ \begin{equation} \int_X | u |^p_{h \otimes h_A} \, d V_{\omega} \leqslant \int_X \langle B^{-1}_{A, h_A} f, f \rangle^{p/2}\, d V_{\omega}, \end{equation} \tag{2.1} $$
provided that the right-hand side of (2.1) is finite, where $B_{A, h_A} = [i \Theta_{A, h_A} \otimes \mathrm{Id}_E, \Lambda_{\omega}]$.

By Hörmander and Demailly’s $L^2$-existence theorem, a Nakano semi-positive hermitian holomorphic vector bundle on a complete Kähler manifold satisfies the optimal $L^2$-estimate.

Remark 2.3. Although Definition 2.2 defines the notion via $L^p$-estimate condition, we only need the $L^2$-estimate condition in the present paper.

Theorem 2.4 (see [5]). Let $(X, \omega)$ be a Kähler manifold of dimension $n$ which admits a positive holomorphic hermitian line bundle, $(E, h)$ be a smooth hermitian holomorphic vector bundle over $X$. If $(E, h)$ satisfies the optimal $L^2$-estimate, then $(E, h)$ is semi-positive in the sense of Nakano.

2.2. A new characterization for plurisubharmonic functions

The extension of holomorphic functions or sections of a holomorphic vector bundle with suitable integrable conditions plays an important role in several complex variables and complex geometry. The reader is referred to [20], [21] for surveys on this topic.

Theorem 2.5 (Ohsawa–Takegoshi, see [22]). Let $D\subset\mathbb{C}^n$ be a pseudoconvex domain and $\varphi$ be a plurisubharmonic function on $D$. Let $H\subset D$ be a closed submanifold. Then there exists a constant $C$ which depends only on the diameter of $D$ such that for any holomorphic function $f$ on $H$ with

$$ \begin{equation*} \int_{H}|f|^2e^{-\varphi}\, dV_H<\infty, \end{equation*} \notag $$
there exists a holomorphic function $F$ on $D$ such that $F|_H=f$ and
$$ \begin{equation*} \int_{D}|F|^2e^{-\varphi}\, d V_D\leqslant C\int_{H}|f|^2e^{-\varphi}\, dV_H<\infty. \end{equation*} \notag $$
where $d V_D$ and $d V_H$ denote the Lebesgue measure on $D$ and $H$, respectively.

In [4], Deng, Wang, Zhang, and Zhou found a new characterization involving an $L^p$-extension property for holomorphic sections which can be regarded as a converse to Theorem 2.5.

Definition 2.6 (Multiple $L^p$-extension property, see [4]). Let $(E, h)$ be a holomorphic vector bundle over a bounded domain $D \subset \mathbb{C}^n$ equipped with a singular Finsler metric $h$. Let $p > 0$ be a fixed constant. Assume that, for any $z \in D$, any nonzero element $a\in E_z$ with finite norm $| a|$, and any $m \geqslant 1$, there is a holomorphic section $f_m$ of $E^{\otimes m}$ on $D$ such that $f_m (z) = a^{\otimes m} $ and satisfies the following estimate:

$$ \begin{equation*} \int_D | f_m |^p \leqslant C_m | a^{\otimes m} |^p = C_m | a|^{m p}, \end{equation*} \notag $$
where $C_m$ are constants independent of $z$ and satisfying $1 / m \log C_m \to 0$ as $m \to \infty$. Then $(E, h)$ is said to have multiple $L^p$-extension property.

With the multiple $L^p$-extension property, one can expect the Griffiths positivity.

Theorem 2.7 (see [4]). Let $(E,h)$ be a holomorphic vector bundle over a bounded domain $D\subset\mathbb{C}^n$ equipped with a singular Finsler metric $h$ such that the norm of any local holomorphic section of $E^*$ is upper semi-continuous. If $(E,h)$ has multiple $L^p$-extension property for some $p>0$, then $(E,h)$ is semi-positive in the sense of Griffiths, namely, $\log|u|^2$ is plurisubharmonic for any local holomorphic section $u$ of $E^*$.

Remark 2.8. In [4], in the case that $(\underline{\mathbb{C}},h)$ is a trivial line bundle with $h=e^{-\varphi}$, it is actually showed that if $h=e^{-\varphi}$ satisfies multiple $L^p$-extension property, there exists a sequence of functions $\{\varphi_m\}\subset \operatorname{PSH}(D)$ converging to $\varphi$ decreasingly. In particular, $\varphi_m\to\varphi$ in $L^1_{\mathrm{loc}}$ as $m\to \infty$.

2.3. Positivity of hermitian holomorphic vector bundles

We first recall in this subsection some basic notions on positivity of smooth hermitian holomorphic vector bundles.

Definition 2.9. Let $M$ be a complex manifold of dimension $n$ and $(E, h)\,{\to}\, M$ be a hermitian holomorphic vector bundle of finite or infinite rank. Let $\Theta =\sum_{j,k} \Theta_{j \overline{k}} \,\, d z_j \wedge d \overline{z}_k$ be the Chern curvature tensor of $E$.

(1) $(E, h)$ is positive in the sense of Griffiths if, for any section $u$ of $E$ and any vector $v \in \mathbb{C}^n$, one has

$$ \begin{equation*} \sum (\Theta_{j \overline{k}} u, u)_h v_j \overline{v}_k \geqslant \delta \| u \|^2_h | v |^2, \end{equation*} \notag $$
for some $\delta > 0$.

(2) $(E, h)$ is positive in the sense of Nakano if, for any $n$-tuple $(u_1, \dots, u_n)$ of sections of $E$, one has

$$ \begin{equation*} \sum (\Theta_{j \overline{k}} u_j, u_k)_h \geqslant \delta \sum \| u_j \|^2_h, \end{equation*} \notag $$
for some $\delta>0$.

Semi-positivity concepts in two senses could be defined accordingly.

It is easy to see that Nakano (semi-)positivity implies Griffiths (semi-)positivity. However, the converse does not hold in general, see [7]. In a more analytical way, Griffiths semi-positivity is equivalent to the fact that $\log | u |^2_h$ is plurisubharmonic for any local holomorphic section $u$ of the dual bundle $E^{\ast}$. However, such duality property fails for Nakano positivity, see [7].

For a holomorphic line bundle with a hermitian metric which can be locally written as $e^{-\varphi}$, the notions of Griffiths semi-positivity and Nakano semi-positivity coincide and being semi-positive means $\varphi \in \operatorname{PSH}$ locally.

As we have mentioned, it is possible to define the positivity of Griffiths and Nakano for a holomorphic vector bundle with a (possibly) singular hermitian metric.

By a singular hermitian metric we mean a measurable map from the base space to the space of non-negative hermitian forms which are finite almost everywhere on each fiber. By DNWZ’s results stated in the above Definition 2.2 and Theorem 2.4 (see [5]), it is reasonable to introduce the following notion.

Definition 2.10 (see [6]). Let $M$ be a complex manifold of dimension $n$ and let $(E, h) \to M$ be a hermitian holomorphic vector bundle of finite or infinite rank with a (possibly) singular hermitian metric $h$.

(1) $(E,h)$ is semi-positive in the sense of Griffiths if, for any holomorphic section $u$ of the dual bundle, $\log ||u||^2_{h^{\ast}}$ is plurisubharmonic.

(2) $(E,h)$ is semi-positive in the sense of Nakano if $(E,h)$ is semi-positive in the sense of Griffiths and satisfies the optimal $L^2$-estimate as in Definition 2.2 locally.

The following lemma and remark interpret the behaviour of $[i \Theta_{E, h}, \Lambda_{\omega}]$ acting on the bundle valued $(n,1)$-forms.

Lemma 2.11 (see [7]). Let $(X, \omega)$ be a Kähler manifold, $(E, h) \to X$ be a hermitian vector bundle of rank $r$ which is positive (respectively, semi-positive) in the sense of Nakano. Then $[i\Theta_{E, h}, \Lambda_{\omega}]$ is positive definite (respectively, semi-positive definite) on $\bigwedge^{n, 1} T^{\ast}_X \otimes E$.

Remark 2.12. Actually, if we fix a point $z_0 \in X$ and choose a local coordinate neighbourhood centered at $z_0$ such that $(\partial / \partial z_1, \partial / \partial z_2, \dots, \partial / \partial z_n)$ is an orthonormal basis of $T X$ at $z_0$. We also choose an orthonormal basis of $E_{{z_0} }$, denoted by $(e_1, e_2, \dots, e_r)$. We can write

$$ \begin{equation*} \omega = i \sum_{j} d z_j \wedge d \overline{z}_j+O(\|z\|^2) \end{equation*} \notag $$
and
$$ \begin{equation*} i \Theta_{E, h} |_{z_0} = i \sum_{j, k, \lambda, \mu} c_{j \overline{k} \lambda \mu}\, d z_j \wedge d \overline{z}_k \otimes e^{\ast}_{\lambda} \otimes e_{\mu}. \end{equation*} \notag $$
One may calculate the following identity:
$$ \begin{equation*} \langle [i \Theta_{E, h}, \Lambda_{\omega}] u, u \rangle = \sum_{j, k, \lambda, \mu} c_{j \overline{k} \lambda \mu} u_{j, \lambda} \overline{u}_{k, \mu}, \end{equation*} \notag $$
where $u = \sum_{k,\lambda} u_{k, \lambda}\,\, d z \wedge d \overline{z}_k \otimes e_{\lambda}$.

2.4. The weighted Bergman kernel

Recall the definition of the weighted Bergman kernel. Let $D \subset\mathbb{C}^n$ be a domain and $\varphi$ be an upper semi-continuous function on $D$. As usual, we denote

$$ \begin{equation*} A^2 (D, \varphi) = \biggl\{ f \in \mathcal{O} (D) \colon \| f \|^2_{\varphi} := \int_D | f |^2 e^{- \varphi} < \infty \biggr\}. \end{equation*} \notag $$
For $z \in D$, the weighted Bergman kernel is defined to be
$$ \begin{equation*} B_{\varphi} (z) = \bigl(\inf \{ \| f \|^2_{\varphi} \colon f \in A^2 (D, \varphi),\, f(z) = 1 \}\bigr)^{- 1}, \end{equation*} \notag $$
if there exists $f \in A^2 (D, \varphi)$ with $f (z) \neq 0$ and otherwise $B_{\varphi}(z)$ is defined to be $0$.

It is clearly that

$$ \begin{equation*} B_{\varphi} (z) = \sup \{ | f (z) |^2 \colon f \in A^2 (D, \varphi),\, \| f \|^2_{\varphi} = 1 \}. \end{equation*} \notag $$
If there is no confusion, we will call $B_{\varphi} (z)$ the Bergman kernel for short and omit the weight function $\varphi$.

Proposition 2.13. Let $D\,{\subset}\, \mathbb{C}^n$ be a bounded domain, $\varphi$ be an upper semi-continuous function on $D$ and $B_{\varphi}(z)$ be the (diagonal) Bergman kernel weighted by $\varphi$. Let $z \in D$ be an arbitrary point, then $B_{\varphi}(z)>0$ if and only if there exists a holomorphic function on $D$ which is square integrable with respect to $\varphi$ such that $f (z) = 1$ and

$$ \begin{equation*} \frac{1}{B_{\varphi} (z)} = \int_D | f |^2 e^{- \varphi}\, d V. \end{equation*} \notag $$
In particular, $e^{- \varphi} \in L^1_{\mathrm{loc}}$ near $z$.

Proof. We assume that $B_{\varphi}(z)>0$ for some $z \in D$. By the definition of the Bergman kernel,
$$ \begin{equation*} \infty > \frac{1}{B_{\varphi}(z)} = \inf \biggl\{ \int_D | f |^2 e^{- \varphi} \, d V \colon f \in \mathcal{O} (D), \, f (z) = 1 \biggr\}. \end{equation*} \notag $$
Thus, there exists a sequence of holomorphic functions $\{ f_n \}$ such that, for every $n = 1, 2, \dots$, $f_n (z) = 1$ and
$$ \begin{equation} \lim_{n \to \infty} \int_D | f_n |^2 e^{- \varphi}\, d V = \frac{1}{B(z)}, \end{equation} \tag{2.2} $$
and therefore, there exists a constant $C$ such that, for any $n = 1, 2, \dots$,
$$ \begin{equation*} \int_D | f_n |^2 e^{- \varphi} < C. \end{equation*} \notag $$
We fixed an arbitrary $n$, since $f_n (z) = 1$, there exists a neighbourhood $U$ of $z$ such that
$$ \begin{equation*} \frac{1}{2} < | f_n(z) |^2 < \frac{3}{2} \end{equation*} \notag $$
on $U$ and
$$ \begin{equation*} \frac{1}{2} \int_U e^{- \varphi}\, d V < \int_D | f_n |^2 e^{- \varphi} \, d V < C, \end{equation*} \notag $$
and thus $e^{- \varphi}$ is locally integrable near $z$.

Again by (2.2), for any compact subset $K \subset D$, since $\varphi$ is an upper semi-continuous, there exists a constant $M_K$ such that

$$ \begin{equation*} e^{- \varphi} \geqslant e^{- M_K} \end{equation*} \notag $$
on $K$. Therefore, for $n = 1, 2, \dots$,
$$ \begin{equation*} e^{- M_K} \int_K | f_n |^2\, d V \leqslant \int_K | f_n |^2 e^{- \varphi}\, d V \leqslant \int_D | f_n |^2 e^{\varphi}\, d V < C, \end{equation*} \notag $$
that is to say, $\{ f_n \}$ is locally $L^2$-bounded uniformly. Thus there exists a subsequence of $\{ f_n \}$, which is sill denoted by itself, converging to some holomorphic function $f$ uniformly on any compact subset of $D$. In particular,
$$ \begin{equation*} f (z) = \lim_{n \to \infty} f_n (z) = 1 \end{equation*} \notag $$
and
$$ \begin{equation*} \int_D | f |^2 e^{- \varphi} \, d V \leqslant \liminf_{n \to \infty} \int_D | f_n |^2 e^{-\varphi} \, d V = \frac{1}{B_{\varphi}(z)}. \end{equation*} \notag $$
By the above inequality, $f$ is square integrable with respect to $\varphi$ on $D$. On the other hand, one has by definition,
$$ \begin{equation*} \frac{1}{B_{\varphi}(z)} \leqslant \int_D | f |^2 e^{- \varphi}\, d V, \end{equation*} \notag $$
and so we have
$$ \begin{equation*} \frac{1}{B_{\varphi}(z)} = \int_D | f |^2 e^{- \varphi}\, d V. \end{equation*} \notag $$

The converse is obvious. $\Box$

Lemma 2.14. Let $\varphi \colon D \to [- \infty, \infty)$ be an upper semi-continuous function on a domain $D \subset \mathbb{C}^n$. Then the weighted Bergman kernel $B_{\varphi}(z)$ is continuous on $D$ and $\log B_{\varphi}(z)$ is plurisubharmonic on $D$.

Proof. We first show that $B_{\varphi} (z)$ is upper semi-continuous. Let $z \in D$ be an arbitrary point and $z_j \in D$ that is convergent to $z$ as $j \to \infty$. Then, for any $\varepsilon > 0$, there exists $f_j \in A^2 (D, \varphi)$ such that $\| f_j \|^2_{\varphi} = 1$ and
$$ \begin{equation*} B_{\varphi} (z_j) - \varepsilon < | f_j (z_j) |^2. \end{equation*} \notag $$
Since $\varphi$ is locally upper-bounded, it follows that $\{ f_j \}$ is a normal family on $D$ and thus there is a subsequence, which is still denoted by itself, converging to some $f \in \mathcal{O} (D)$ uniformly on any compact subset of $D$.

By Fatou’s lemma, one has

$$ \begin{equation*} \| f \|^2_{\varphi} \leqslant \liminf_{j \to \infty} \| f_j \|^2_{\varphi} = 1 \end{equation*} \notag $$
and thus
$$ \begin{equation*} B_{\varphi} (z) \geqslant \frac{| f (z) |^2}{\| f \|^2_{\varphi}} \geqslant | f (z) |^2 = \lim_{j \to \infty} | f_j (z_j) |^2 \geqslant \limsup_{j \to \infty} B_{\varphi} (z_j) - \varepsilon, \end{equation*} \notag $$
and we conclude that $B_{\varphi} (z)$ is upper semi-continuous by letting $\varepsilon \to 0$.

Let $w \in D$ be a fixed point, then, for any $\varepsilon > 0$, there exists $f \in A^2 (D, \varphi)$ such that

$$ \begin{equation*} \frac{| f (w) |^2}{\| f \|^2_{\varphi}} > B_{\varphi} (w) - \varepsilon. \end{equation*} \notag $$
Thus
$$ \begin{equation*} B_{\varphi} (z) \geqslant \frac{| f (z) |^2}{\| f \|^2_{\varphi}} = \lim_{w \to z} \frac{| f (w) |^2}{\| f \|^2_{\varphi}} \geqslant \liminf_{w \to z} B (w) - \varepsilon, \end{equation*} \notag $$
and the lower semi-continuity follows by letting $\varepsilon \to 0$.

For the log-plurisubharmonicity, note that $\log B_{\varphi} (z)$ is upper semi-continuous and

$$ \begin{equation*} \log B_{\varphi} (z) = \sup \{ \log | f (z) |^2 \colon f \in A^2 (D, \varphi),\, \| f \|^2_{\varphi} = 1 \}. \end{equation*} \notag $$
$\Box$

Remark 2.15. It follows from Lemma 2.14 that the set

$$ \begin{equation*} \{ z \in D \colon B_{\varphi} (z) = 0 \} \end{equation*} \notag $$
is of measure zero since $\log B_{\varphi} (z)$ is plurisubharmonic.

2.5. Miscellaneous

We will also use the following classical results from real and complex analysis.

Proposition 2.16 (see [23]). If $f \in L^1_{\mathrm{loc}}(\mathbb{R}^n)$, then almost every $x \in \mathbb{R}^n$ is a Lebesgue point of $f$. By a Lebesgue point of $f$ we mean a point $x \in \mathbb{R}^n$ which satisfies

$$ \begin{equation*} \lim_{r \to 0} \frac{1}{\sigma_n r^{2 n}} \int_{B (x, r)} | f (y) - f (x) | \, d V (y) = 0, \end{equation*} \notag $$
where $\sigma_n$ is the volume of $n$-dimensional unit ball.

Lemma 2.17. Let $D \subset \mathbb{R}^n$ be a domain and $f$ be a measurable function on $D$ such that $f \in L^1_{\mathrm{loc}}$ near almost every $x \in D$. Then almost every $x \in D$ is a Lebesgue point of $f$.

Proof. Denote by $E = \{ x \in D \colon f \notin L^1_{\mathrm{loc}} \text{ near } x\}$, then $E$ is a set of measure zero. Assume there exists $\{ x_k \} \subset E$ such that $x_k \to x$ as $k \to \infty$. If $f \in L^1_{\mathrm{loc}}$ near $x$, then there is a neighbourhood $U$ of $x$ such that $f \in L^1 (U)$. However, when $k$ is large enough, $x_k \in U$ and then $f \in L^1_{\mathrm{loc}}$ near $x_k$. Thus $E$ is a closed subset of $D$.

Since for every $x \in D \setminus E$, there exists a neighbourhood $V$ of $x$ such that $\overline{V} \subset D$ and $f \in L^1 (V)$, therefore, $f \in L^1_{\mathrm{loc}} (D\setminus E)$ and by the above proposition, almost every $x \in D \setminus E$ is a Lebesgue point of $f$. Since $E$ is a set of measure zero, almost every $x \in D$ is a Lebesgue point of $f$. $\Box$

Proposition 2.18 (see [24]). Let $D\subset\mathbb{C}^n$ be a domain and $\{u_j\}$ be a sequence of functions in $\operatorname{PSH}(D)$ which is locally uniformly bounded from above on $D$. If $\{u_j\}$ does not converge to $-\infty$ locally uniformly on $D$, then it admits a subsequence which converges to some $u\in \operatorname{PSH}(D)$ in $L^1_{\mathrm{loc}}(D)$.

Proposition 2.19 (see [25]). Let $D \subset \mathbb{R}^n$ be a domain and $u \colon D \to [- \infty, \infty)$ be an upper semi-continuous function. Then $u$ is subharmonic if and only if

$$ \begin{equation*} M (x_0, r) = \frac{1}{c_n} \int_{| y | = 1} u (x_0+r y)\, d V_y \end{equation*} \notag $$
is an increasing function of $r \in [0, \delta]$, where $c_n$ is the volume of the unit sphere in $\mathbb{R}^n$, $d V$ is the Lebesgue measure and $x_0 \in \{ x \in D\colon y \in D \textit{ if } | y - x | < \delta\}$, $\delta > 0$. In particular, if $u$ is subharmonic on $D$,
$$ \begin{equation*} u (x_0) \leqslant \frac{1}{\sigma_n r^{2n}} \int_{B (x_0, r)} u (y)\, d V_y, \end{equation*} \notag $$
where $\sigma_n$ is the volume of unit ball in $\mathbb{R}^n$.

The following corollary is straightforward.

Corollary 2.20. Let $D \subset \mathbb{R}^n$ be a domain and $u$ be a subharmonic function on $D$. Then, for any set $E$ of measure zero,

$$ \begin{equation*} \limsup_{x \to x_0,\, x \in D \setminus E} u (x) = u (x_0) \end{equation*} \notag $$
for any $x_0\in D$.

§ 3. Proof of Theorem 1.2

In this section, we prove Theorem 1.2. The proof is divided into several parts.

Remark 3.1. Before we prove the above theorem, note that if we choose $\psi$ to be lower-bounded, then $\log B_{k\varphi}(z)\not\equiv -\infty$ implies $\log B_{k\varphi+\psi}(z) \not\equiv-\infty$. Indeed, if $\log B_{k\varphi}(z)> -\infty$ for some $z\in D$, there exists a holomorphic function $f\in A^2(D,e^{-\varphi})$ on $D$ such that $f(z)\neq 0$. Since $\psi$ is lower-bounded, there exists some $C>0$ such that

$$ \begin{equation*} \int_{D}|f|^2e^{-k\varphi-\psi}\leqslant C \int_{D}|f|^2e^{-k\varphi}<\infty, \end{equation*} \notag $$
and it follows that $\log B_{k\varphi+\psi}(z)>-\infty$.

The following result establishes implication $(2)\,{\Rightarrow}\,(1)$ in Theorem 1.2.

Proposition 3.2. With the same assumptions and notation in Theorem 1.2, if, for any domain $H \subset \mathbb{C}^m_{\tau}$ and any locally bounded $\psi \in \operatorname{PSH} (H \,{\times}\, D)$, the Bergman kernel $B_{k, \tau} (z)$ of the direct image of the trivial line bundle $(\underline{\mathbb{C}}, e^{- k \pi^{\ast} \varphi - \psi})\to H \times D$ is log-plurisubharmonic, then $\varphi$ is plurisubharmonic.

Proof. Since $D$ is bounded, we may assume that the diameter of $D$ is $R$, then, for any $z, z_0 \in D$, $\log | z - z_0 |^2 < \log R^2$. Choose $\psi (\tau, z) = \max \{ \log | z - z_0 |^2 - \operatorname{Re} \tau,\, 0 \}$ and $H = \{\tau \in \mathbb{C}\colon \operatorname{Re} \tau < \log R^2 \}$, the half plane of $\mathbb{C}$. Hence $\psi$ is a plurisubharmonic function defined on $H \times D$ and depends only on $t = \operatorname{Re} \tau$.

By our hypothesis,

$$ \begin{equation*} \log B_{k, \tau, p}(z) \end{equation*} \notag $$
is plurisubharmonic for any $p > 0$ and positive integer $k$, where
$$ \begin{equation*} B_{k, \tau, p} (z) = \sup \biggl\{ \frac{1}{\int_D |f|^2 e^{- k \varphi - p \psi_{\tau}}} \colon f \in A^2 (D, k \varphi+p \psi_{\tau}),\, f (z) = 1 \biggr\}, \end{equation*} \notag $$
is the Bergman kernel weighted by $k \varphi+p \psi_{\tau}$.

By our assumption and Remark 3.1, it follows from $\psi\geqslant 0$ that, for every positive integer $k$, $\log B_{k, \tau, p} (z) \,{\not\equiv}\, {-}\infty$. It follows from Lemma 2.14 that $\{ z \,{\in}\, D \colon B_{k, \tau} (z) \,{=}\, 0 \}$ is a set of measure zero for every positive integer $k$ and

$$ \begin{equation*} \bigcup_k \{ z \in D \colon B_{k, \tau} (z) = 0 \} \end{equation*} \notag $$
is a set of measure zero. An appeal to Proposition 2.13 and Remark 2.15 shows that $e^{- k \varphi - p \psi_{\tau}} \in L^1_{\mathrm{loc}}$ near almost every $z \in D$ and $k \in \mathbb{N}^+$.

Since $e^{- p \psi_{\tau}}$ is bounded for any fixed $\tau$, $e^{- k \varphi} \in L^1_{\mathrm{loc}}$ near almost every $z \in D$ and $k \in \mathbb{N}^+$. It then follows from Lemma 2.17 that almost every $z \in D$ is a Lebesgue point of $e^{- k \varphi}$, $k = 1, 2, \dots$ . Note that with this choice of $H$ and $\psi$, one has for any $k,p\in\mathbb{N}^+$ and $\tau\in H$,

$$ \begin{equation*} \{ z \in D \colon B_{k, \tau,p} (z) = 0 \} \subset \{z\in D\colon B_{k,\log R^2}(z)=0\}. \end{equation*} \notag $$

Let $z_0 \in D$ be an arbitrary point such that $z_0$ is a Lebesgue point of $e^{- k \varphi}$ for any $k \in \mathbb{N}^+$ and

$$ \begin{equation*} z_0 \in \bigcap_k \{ z \in D \colon B_{k, \tau, p} (z) \neq 0 \}, \end{equation*} \notag $$
we may also assume that $\varphi (z_0) > - \infty$.

For any $f \in A^2 (D, k \varphi+p \psi_{\tau})$ with $f (z_0) = 1$, we have

$$ \begin{equation*} \int_D |f|^2 e^{- k \varphi - p \psi_{\tau}}=\int_{D (\log |z - z_0 |^2 < t)} |f|^2 e^{- k \varphi}+\int_{D (\log |z - z_0 |^2 \geqslant t)} |f|^2 e^{- k \varphi - p \psi_{\tau}}, \end{equation*} \notag $$
where
$$ \begin{equation*} D (\log |z - z_0 |^2 < t) = \{z \in D \colon \log |z - z_0 |^2 < t\} \end{equation*} \notag $$
and similarly, for $D (\log |z - z_0 |^2 \geqslant t)$.

Since

$$ \begin{equation*} \limsup_{z \to z_0,\, z \in D \setminus E} \varphi (z) = \varphi (z_0), \end{equation*} \notag $$
for any $E \subset D$ of measure zero, it follows that
$$ \begin{equation*} \liminf_{z \to z_0,\, z \in D} |f (z) |^2 e^{- \varphi (z)} = e^{- \varphi(z_0)}. \end{equation*} \notag $$
That is to say, for any $\varepsilon > 0$, there exists $\delta \in\mathbb{R}$ such that on $D (\log |z - z_0 |^2 < \delta)$,
$$ \begin{equation*} |f (z) |^2 e^{- k \varphi (z)} > e^{- k \varphi (z_0)} - \varepsilon, \end{equation*} \notag $$
and thus if $t < \delta$,
$$ \begin{equation*} \int_{D (\log |z - z_0 |^2 < t)} |f|^2 e^{- k \varphi} > (e^{- k \varphi(z_0)} - \varepsilon) \sigma_{2 n} e^{nt}, \end{equation*} \notag $$
that is,
$$ \begin{equation*} \frac{1}{\sigma_{2 n} e^{nt}} \int_{D (\log |z - z_0 |^2 < t)} |f|^2 e^{- k \varphi} > e^{- k \varphi (z_0)} - \varepsilon, \end{equation*} \notag $$
where $\sigma_{2 n}$ is the volume of $2 n$-dimensional unit ball. It follows from $B_{k, \tau, p} (z_0) > 0$ and Proposition 2.13 that there exists $f_{\tau, p} \in A^2 (D, k\varphi+p \psi_{\tau})$ with $f_{\tau, p} (z_0) = 1$ such that
$$ \begin{equation*} \frac{1}{B_{k, \tau, p} (z_0)} = \int_D |f_{\tau, p} |^2 e^{- k \varphi - p \psi_{\tau}}. \end{equation*} \notag $$
Now we have
$$ \begin{equation} \begin{aligned} \, &\log B_{k, \tau, p} (z_0)+\log \sigma_{2 n}+nt = -\log \int_D |f_{\tau, p} |^2 e^{- k \varphi - p \psi_{\tau}}+\log \sigma_{2 n}+nt \nonumber \\ &\quad =- \log \biggl( \int_{D (\log |z - z_0 |^2 < t)} |f_{\tau, p} |^2 e^{-k \varphi } +\int_{D (\log |z - z_0 |^2 \geqslant t)} |f_{\tau, p} |^2 e^{- k \varphi - p \psi_{\tau}} \biggr) \nonumber \\ &\quad\qquad +\log \sigma_{2 n}+nt \nonumber \\ &\quad \leqslant-\log \biggl( \int_{D (\log |z - z_0 |^2 < t)} |f_{\tau, p}|^2 e^{- k \varphi } \biggr)+\log \sigma_{2 n}+nt \nonumber \\ &\quad= \log \biggl( \frac{1}{\sigma_{2 n} e^{nt}} \int_{D (\log |z - z_0|^2 < t)} |f_{\tau, p} |^2 e^{- k \varphi } \biggr)^{-1} \leqslant \log (e^{- k \varphi (z_0)} - \varepsilon)^{- 1}, \end{aligned} \end{equation} \tag{3.1} $$
for any $p$ and $t < 0$ small enough.

On the other hand, since $B_{k, \tau, p}$ depends only on $t = \operatorname{Re} \tau$,

$$ \begin{equation*} \log B_{k, \tau, p} (z_0)+\log \sigma_{2 n}+nt \end{equation*} \notag $$
is a convex function. It follows from (3.1) that $\log B_{k, \tau, p} (z_0)+\log \sigma_{2 n}+nt$ is upper-bounded as $t\to-\infty$ and hence an increasing function. Hence, for any $t < \log R^2$, we have
$$ \begin{equation} \sigma_{2 n} e^{n \log R^2} B_{k, \log R^2} (z_0) = \sigma_{2 n} e^{n \log R^2} B_{k, \log R^2, p} (z_0) \geqslant \sigma_{2 n} e^{nt} B_{k, \tau, p} (z_0), \end{equation} \tag{3.2} $$
for any $p > 0$, the first equality holds because $\log |z - z_0 | < \log R^2$ for any $z, z_0 \in D$.

Now, for any $f \in \mathcal{O} (D)$ with $f (z_0) = 1$ satisfying

$$ \begin{equation*} \int_D |f|^2 e^{- k \varphi} < \infty, \end{equation*} \notag $$
since $\psi \geqslant 0$, we have
$$ \begin{equation*} \int_D |f|^2 e^{- k \varphi - p \psi_{\tau}} \leqslant \int_D |f|^2 e^{- k \varphi} < \infty. \end{equation*} \notag $$
It then follows that
$$ \begin{equation*} \sigma^{- 1}_{2 n} e^{-nt} \frac{1}{B_{k, \tau, p}} \leqslant \sigma^{-1}_{2 n} e^{-nt} \int_D |f|^2 e^{- k \varphi - p \psi_{\tau}}. \end{equation*} \notag $$
Therefore, by (3.2) and the above inequality,
$$ \begin{equation*} \begin{aligned} \, \sigma^{- 1}_{2 n} R^{- 2 n} \frac{1}{B_{k, \log R^2}}&\leqslant \sigma^{- 1}_{2 n} e^{- nt} \int_{D (\log |z - z_0 |^2 < t)} |f|^2 e^{- k \varphi} \\ &\qquad+\sigma^{- 1}_{2 n} e^{- nt} \int_{D (\log |z - z_0 |^2 \geqslant t)} |f|^2 e^{- k \varphi - p \psi}. \end{aligned} \end{equation*} \notag $$
Now making $p \to \infty$, we have
$$ \begin{equation*} \sigma^{- 1}_{2 n} R^{- 2 n} \frac{1}{B_{\log R^2}} \leqslant \sigma^{- 1}_{2 n} e^{- nt} \int_{D (\log |z - z_0|^2 < t)} |f|^2 e^{- k \varphi}, \end{equation*} \notag $$
and further, by letting $t \to - \infty$, since $z_0$ is a Lebesgue point of $e^{- k\varphi}$ and since $\varphi (z_0) > - \infty$, we have
$$ \begin{equation} \sigma^{- 1}_{2 n} R^{- 2 n} \frac{1}{B_{k, \log R^2}} \leqslant e^{- k \varphi (z_0)}. \end{equation} \tag{3.3} $$
By our assumption that $B_{k, \log R^2} (z_0) = B_{k, \log R^2, p} (z_0) > 0$ and by Proposition 2.13, there exists $F \in \mathcal{O} (D)$ with $F (z_0) = 1$ satisfying
$$ \begin{equation*} \sigma^{- 1}_{2 n} R^{- 2 n} \int_D |F|^2 e^{- k \varphi} = \sigma^{-1}_{2 n} R^{- 2 n} \frac{1}{B_{k, \log R^2}}, \end{equation*} \notag $$
it then follows from (3.3) that
$$ \begin{equation*} \sigma^{- 1}_{2 n} R^{- 2 n} \int_D |F|^2 e^{- k \varphi} \leqslant e^{-k \varphi (z_0)}. \end{equation*} \notag $$

By Theorem 2.7 and Remark 2.8, there exists a sequence of plurisubharmonic functions $\{\varphi_m \} \subset \operatorname{PSH} (D)$ converging to $\varphi$ decreasingly on any Lebesgue point $z$ of $e^{- k\varphi}$ with $B_{k,{\tau},p}(z)>0$, $\varphi (z) > - \infty$ for any $k\in\mathbb{N}^+, {\tau}\in H$ and $p>0$. Since $\{\varphi_m \} \subset \operatorname{PSH} (D)$ does not converge to $- \infty$ uniformly on $D$, there exists a plurisubharmonic function $\widetilde{\varphi}$ on $D$ such that

$$ \begin{equation*} \lim_{m \to \infty} \varphi_m (z) = \widetilde{\varphi} (z). \end{equation*} \notag $$

By Lemma 2.17, almost every $z \in D$ is a Lebesgue point of $e^{- k \varphi}$ for any $k \in \mathbb{N}^+$. Since $\bigcup_k \{z \in D \colon B_{k,{\tau},p} (z) = 0 \}$ and $\{ z \in D \colon \varphi (z) = - \infty \}$ are subsets of measure zero, it follows that, for almost every $z \in D$,

$$ \begin{equation*} \varphi (z) = \widetilde{\varphi}(z). \end{equation*} \notag $$

On the other hand, on $D$ we have

$$ \begin{equation} \limsup_{w \to z,\, w \in D \setminus E} \varphi (w) = \varphi (z) \end{equation} \tag{3.4} $$
for any subset $E$ of measure zero. For any $z \in D$, we may choose $E$ in (3.4) to be the union of the set of non Lebesgue points of $e^{- k \varphi}$, $k = 1, 2, \dots $, $\bigcup_k \{ z \in D\colon B_{k, \log R^2} (z) = 0 \}$ and $\{ z \in D \colon \varphi (z) = - \infty \}$, then
$$ \begin{equation*} \widetilde{\varphi} (z) = \limsup_{w \to z,\, w \in D \setminus E} \widetilde{\varphi} (w) = \limsup_{w \to z,\, w \in D \setminus E} \varphi (w) = \varphi (z), \end{equation*} \notag $$
therefore, $\varphi$ is a plurisubharmonic function on $D$. $\Box$

The next result establishes implication $(1)\,{\Rightarrow}\,(4)$ in Theorem 1.2.

Proposition 3.3. With the same assumptions and notation in Theorem 1.2, if $\varphi$ is plurisubharmonic, then, for any domain $H \subset \mathbb{C}^m_{\tau}$ and any locally bounded $\psi \in \operatorname{PSH} (H \times D)$, the direct image bundle $F$ of the trivial line bundle $(\underline{\mathbb{C}}, e^{- k \pi^{\ast} \varphi - \psi}) \to H \times D$ is semi-positive in the sense of Nakano.

Proof. By Theorem 1.3 in [5], $F$ is semi-positive in the sense of Griffiths. We may assume that $H$ is pseudoconvex. By Definition 2.10, it suffices to show that, for any $f \in C^{\infty}_{\mathrm{c}} (H, \bigwedge^{m, 1} T^{\ast}_H \otimes F)$ with $\overline{\partial}_{\tau} f = 0$ and for any plurisubharmonic function $\rho$ on $H$, there exists $u \in L^2 (H, \bigwedge^{m, 0} T^{\ast}_H \otimes F)$ satisfying $\overline{\partial} u = f$ and
$$ \begin{equation*} \int_H | u |^2_{\tau} e^{- \rho} \, d V_{\tau} \leqslant \int_H \langle [i\,\partial\,\overline{\partial}\rho\otimes \mathrm{Id}_{F},\Lambda_{\omega}] ^{-1} f, f \rangle^2_{\tau} e^{- \rho}\, d V_{\tau}. \end{equation*} \notag $$

Let $f \in C^{\infty}_{\mathrm{c}} (H, \bigwedge^{m, 1} T^{\ast}_H \otimes F)$ with $\overline{\partial}_{\tau} f = 0$. We can regard $f$ as $f (\tau, z)$ which is smooth and compact supported in $\tau$ and holomorphic in $z$, therefore, $\overline{\partial} f = \overline{\partial}_{\tau} f+\overline{\partial}_z f = 0$. We write $f = \sum_{j = 1}^m f_j (\tau, z)\, d \tau \wedge d \overline{\tau}_j$ with $f_j(\tau, z) \in F_{\tau}$ for $\tau \in H$ and $d \tau$ denotes $d \tau_1\, \wedge d \tau_2 \wedge \dots \wedge d \tau_m$.

Since $\overline{\partial} f = 0$, it follows from Hörmander’s $L^2$-existence theorem (Theorem 2.1) that there exists $u \in L^2 (H \times D)$ such that $\overline{\partial}_{\tau}u=f$ and

$$ \begin{equation} \int_{H \times D} | u |^2 e^{- k \varphi - \psi - \rho} \leqslant \int_{H \times D} | f |^2_{i\, \partial\, \overline{\partial} \rho} e^{- k \varphi - \psi - \rho}. \end{equation} \tag{3.5} $$
It follows from $\overline{\partial} u = f$ that $\partial u/\partial \overline{z}_j = 0$ for $j = 1, 2, \dots, n$, that is $\overline{\partial}_{\tau}u=f$. Therefore, we can regard $u$ as a section of $F$ which satisfies the integral condition.

By the Fubini theorem, we may rewrite (3.5) as

$$ \begin{equation*} \begin{aligned} \, \int_H | u |^2_{\tau} e^{- \rho} &=\int_{H \times D} | u |^2 e^{- k \varphi - \psi - \rho} \leqslant \int_{H \times D} | f |^2_{i \,\partial\, \overline{\partial} \rho} e^{- k \varphi - \psi - \rho} \\ &= \int_H \langle [i\, \partial\,\overline{\partial}\rho\otimes \operatorname{Id}_{F},\Lambda_{\omega}] ^{-1} f, f \rangle^2_{\tau} e^{- \rho}, \end{aligned} \end{equation*} \notag $$
the last equality follows from Lemma 2.11 and Remark 2.12. Thus, by Definition 2.10, $F$ is semi-positive in the sense of Nakano. $\Box$

Remark 3.4. Proposition 3.3 is closely related to the question raised by Lempert in [26] which asked whether a $C^2$ hermitian metric whose curvature dominates $0$ is positive in the sense of Nakano. Actually, results in [5] lead to an affirmative answer to this question, see [27], [28].

With Propositions 3.2 and 3.3, one reaches Theorem 1.2 immediately.

We continue the proof of Theorem 1.2. $(4)\Rightarrow(3)$ follows from Definition 2.10. It follows from the assumptions and Remark 3.1 that fiberwised Bergman kernel $B_{k,\tau} $ does not vanish for any $k\in\mathbb{N}^+$ and $\tau\in H$. $(3)\Rightarrow (2)$ follows from that $B_{k,{\tau}}$ can be regarded as a non-trivial holomorphic section of the dual bundle. Indeed, for any $f \in \mathcal{O} (H \times D)$ such that $f (\,{\cdot}\,, z)$ is square integrable with respect to $k \varphi+\psi$,

$$ \begin{equation*} f (\tau, w) = \int_D f (\tau,{\cdot}\,) B_{k,{\tau}} (\,{\cdot}\,, w) e^{- k \varphi - \psi}, \end{equation*} \notag $$
for any $w \in D$. It again follows from Definition 2.10 that $B_{k,{\tau}}(z)$ is log-plurisubharmonic. $\Box$

Remark 3.5. With the same assumptions, it can be seen from the proof of Theorem 1.2 that it only needs $ B_{k,\tau} (z)$, a specific family of elements in the dual bundle, to satisfy the log-plurisubharmonic property and non-triviality to obtain Nakano positivity. Moreover, it follows that if one has just $\log B_{k,\tau} (z)\not\equiv -\infty$ (for any $k\in\mathbb{N}^+$ and $\tau\in H$) to be plurisubharmonic, then $\log | \xi |^2_\tau$ is plurisubharmonic for every $\xi$ in the dual bundle.

The method adopted here with a suitable modification can also be used to obtain an analogue of Theorem 1.2 for the direct image of a hermitian holomorphic vector bundle. More precisely, we have the following result.

Theorem 3.6. Let $D \subset \mathbb{C}^n_z$ be a bounded domain and $(E, h) \to D$ be a holomorphic vector bundle with a hermitian metric $h$ whose minimal eigenvalue is locally lower bounded. We assume that, for any open set $U\subset D$ and any local holomorphic section $\xi \in H^0(U, E^{\ast})$,

$$ \begin{equation*} \limsup_{z \to z_0,\, z \in U \setminus E} \log | \xi (z) |_h = \log | \xi (z_0) |_h, \end{equation*} \notag $$
for every $z_0 \in U$, where $E$ is any set of measure zero.

Assume that for any domain $H \subset \mathbb{C}^m_{\tau}$ and any locally bounded $\psi \in \operatorname{PSH} (H \times D)$, the direct image $F$ (over $H$) of $(\pi^{\ast} E^{\otimes k}, \pi^{\ast} h^k e^{- \psi}) \to H \times D$ is semi-positive in the sense of Griffiths for any positive integer $k$. Assume further that

(1) for any constant local section $a$ of $E$, $| a^{\otimes k}|^2_{h^k}$ is locally integrable for every $k \in \mathbb{N}^+$;

(2) for almost every $z \in D$, $k \in \mathbb{N}^+$, $\tau \in H$ and $a \in E_z$ with $|a^{\otimes k} |^2_{h^k e^{- \psi_{\tau}}} < \infty$ at $z$, there exists a holomorphic section $v_{k, \tau}$ of $E^{\otimes k}$ such that $v_{k, \tau} (z) = a^{\otimes k}$ and

$$ \begin{equation*} \| v_{k, \tau} \|^2_{h^k e^{- \psi_{\tau}}}<\infty. \end{equation*} \notag $$

Then $(E, h)$ is semi-positive in the sense of Griffiths.

Proof. Since $D \subset \mathbb{C}^n$ is a bounded domain, we assume that the diameter of $D$ is $R$, therefore, for any $z, z_0 \in D$, one has $\log |z - z_0 |^2 \leqslant \log R^2$. Again, when there is not any confusion, we will write $h$ instead of $\pi^{\ast} h$. Denote by $\psi (\tau, z) = \max \{\log |z - z_0 |^2 - \operatorname{Re} \tau,\, 0\}$ and $H = \{\tau \in \mathbb{C}\colon \operatorname{Re} \tau < \log R^2 \}$, the half plane of $\mathbb{C}$. Then $\psi$ is a plurisubharmonic function defined on $H \times D$ and depends only on $t = \operatorname{Re} \tau$. For a given positive integer $k$ and for any $p > 0$,
$$ \begin{equation*} F_{k, \tau, p} = \biggl\{ u \in H^0 (D, E^{\otimes k}) \colon \|f\|^2_{k, \tau, p} = \int_D |u|^2_{h^k} e^{- p \psi_\tau} < \infty \biggr\}. \end{equation*} \notag $$
As vector spaces, all $F_{k, \tau, p}$ are locally equal for any $p$. Note that
$$ \begin{equation*} \lim_{p \to \infty} \|u\|^2_{k, \tau, p} = \int_{D (\log|z-z_0|^2 < t)} |u|^2_{h^k}, \end{equation*} \notag $$
and $\|u\|^2_{k, \log R^2} = \|u\|^2_{k, \log R^2, p}$ for any $p > 0$, where $D (\log |z - z_0 |^2 < t) = \{z \in D\colon \log |z - z_0 |^2 < t\}$ and similarly $\{z \in D \colon \log |z - z_0 |^2 \geqslant t\}$ will be denoted by $D (\log |z - z_0 |^2 \geqslant t)$, etc.

Without loss of generality, we assume that $E$ is trivial on $D$ since it only needs the local property of $E$ in the last step of our proof. By our assumption, any constant section $a\in H^0 (D, E)$, $| a^{\otimes k}|^2_{h^k}$ is locally integrable on $D$, and by Lemma 2.17, almost every $z\in D$ is a Lebesgue point for $|a^{\otimes k}|^2_{h^k}$ for every $k\in\mathbb{N}^+$. Without loss of generality, we may assume $|a^{\otimes k}|^2_{h^k e^{-\psi_{\tau}"}}<\infty$ at $z$.

By our second condition, there exists a section $v_{k,\tau, p}\in L^2(D, E^{\otimes k}, h^{\otimes k} e^{- p \psi_\tau})$ such that $v_{k, \tau, p} (z_0) = a^{\otimes k}$ and $v_{k, \tau, p}$ has the minimal $L^2$-norm, in particular, $v_{k, \log R^2}$ is the minimal $L^2$-extension on $D$ such that $v_{k,\log R^2} (z_0) = a^{\otimes k}$ and $\|v_{k, \log R^2}\|^2_{h^k }<\infty$. Up to a set of measure zero, we may also assume that $z_0$ is a Lebesgue point of $\| v_{k, \log R^2} \|^2_{h^k }$.

It follows that

$$ \begin{equation} \|v_{k,\tau, p} \|^2_{k,\tau, p} = \int_D |v_{k, \tau, p} |^2_{h^{ k}} e^{- p\psi_\tau} = \sup \frac{| \langle \xi, v_{k, \tau, p} \rangle |^2}{\| \xi \|^2_{k, \tau, p}}, \end{equation} \tag{3.6} $$
where the supreme is taken over elements in $F^{\ast}_{k,\tau, p}$ that vanish on $I = \{s \in F_{k,\tau, p}\colon s (z_0) = 0\}$. We can take $\xi = \xi_g$ of the form
$$ \begin{equation*} \langle \xi_g, u \rangle = (\xi_g, u) |_{z_0}, \end{equation*} \notag $$
where $g$ is a smooth section of $E^{\otimes k}$ with compact support and $(\xi_g, u)$ represents the standard (global) Euclidean inner product, that is, $g \cdot \overline{u}^{\,T}$. It is obvious that such $\xi_g$ forms a dense subspace of the subspace formed by all elements in $F^{\ast}_{k,\tau, p}$ that vanish on $I = \{s \in F_{k,\tau, p} \colon s (z_0) = 0\}$.

By our hypothesis, $(F, \|\,{\cdot}\,\|_{k, \tau, p})$ is semi-positive in the sense of Griffiths, it follows that $\log \| \xi_g \|^2_{k, \tau, p}$ is a plurisubharmonic function. Since $\log \| \xi_g \|^2_{k, \tau, p}$ depends only on $t = \operatorname{Re} \tau$, it follows that this is a convex function.

Since the minimal eigenvalue of $h$ is locally lower bounded, there exist $C_1 > 0$ such that, for any $u \in F_{k, \tau, p}$, one has

$$ \begin{equation*} |u|^2 \leqslant C_1 |u|^2_{h^k}, \end{equation*} \notag $$
on $D (\log |z - z_0 |^2 \leqslant t)$, where again $|u|^2$ denotes the standard Euclidean norm of $u$ on $E$. It then follows form the sub-mean inequality that
$$ \begin{equation*} |u (z_0) |^2 \leqslant \sigma_{2 n}^{- 1} e^{- nt} \int_{D (\log |z - z_0 |^2 < t)} |u|^2 \leqslant C_1 \sigma_{2 n}^{- 1} e^{- nt} \int_{D (\log |z - z_0 |^2 < t)} |u|^2_{h^k}, \end{equation*} \notag $$
which means that $\| \xi_g \|^2_{k, \tau, p} \sigma_{2 n} e^{nt}$ is bounded above as $t$ small enough and $p$ large enough. Since $\log \| \xi_g \|^2_{k, \tau, p}$ is convex, we have
$$ \begin{equation*} \log \| \xi_g \|^2_{k, \tau, p}+\log \sigma_{2 n}+nt \end{equation*} \notag $$
is an increasing function.

By (3.6), we have, for any $t < \log R^2$ and any $p > 0$,

$$ \begin{equation*} \sigma_{2 n}^{- 1} R^{- 2 n} \|v_{k, \log R^2} \|^2_{k, \log R^2} = \sup \frac{| \langle \xi_g, v_{k, \log R^2} \rangle |^2}{\sigma_{2 n} R^{2 n} \| \xi_g \|^2_{k, \log R^2}} \leqslant \sup \frac{| \langle \xi_g, v_{k, \tau, p} \rangle|^2}{\sigma_{2 n} e^{nt} \| \xi_g \|^2_{k, \tau, p}}. \end{equation*} \notag $$
Another appeal to (3.6) shows that
$$ \begin{equation*} \begin{aligned} \, \sup \frac{| \langle \xi_g, v_{k, \tau, p} \rangle |^2}{\sigma_{2 n} e^{nt}\| \xi_g \|^2_{k, \tau, p}} &= \sigma_{2 n}^{- 1} e^{- nt} \int_D |v_{k,\tau, p} |^2_{h^k} e^{- p \psi_\tau} \leqslant \sigma_{2 n}^{- 1} e^{- nt} \int_D |v_{k, \log R^2}|^2_{h^k} e^{- p \psi_\tau} \\ &=\sigma_{2 n}^{- 1} e^{- nt} \int_{D (\log |z - z_0 |^2 < t)} |v_{k, \log R^2} |^2_{h^k} \\ &\qquad +\sigma_{2 n}^{- 1} e^{- nt} \int_{D (\log |z - z_0 |^2 \geqslant t)} |v_{k, \log R^2} |^2_{h^k} e^{- p \psi_\tau} \end{aligned} \end{equation*} \notag $$
and
$$ \begin{equation} \begin{aligned} \, \sigma_{2 n}^{- 1} R^{- 2 n} \|v_{k, \log R^2} \|^2_{k, \log R^2} &\leqslant \sigma_{2 n}^{- 1} e^{- nt} \int_{D (\log |z - z_0 |^2 < t)} |v_{k, \log R^2}|^2_{h^k} \nonumber \\ &\qquad+\sigma_{2 n}^{- 1} e^{- nt} \int_{D (\log |z - z_0 |^2 \geqslant t)} |v_{k, \log R^2} |^2_{h^k} e^{- p \psi_\tau}. \end{aligned} \end{equation} \tag{3.7} $$

Letting $p \to \infty$, we find that the second term of the right-hand side of (3.7) tends to zero. Since $z_0$ is a Lebesgue point of $|v_{k, \log R^2} |^2_{h^k}$, we have

$$ \begin{equation*} \lim_{t \to - \infty} \sigma_{2 n}^{- 1} e^{- nt} \int_{D (\log |z - z_0 |^2 < t)} |v_{k, \log R^2} |^2_{h^k} = |a^{\otimes k} |^2_{h^k}|_{z_0}. \end{equation*} \notag $$
Thus, by letting $t \to - \infty$ and using (3.7), we get
$$ \begin{equation*} \sigma_{2 n}^{- 1} R^{- 2 n} \|v_{\log R^2} \|^2_{\log R^2} \leqslant \liminf_{t \to - \infty} \sigma_{2 n}^{- 1} e^{-nt} \int_{D (\log |z - z_0 |^2 < t)} |v_{k, \log R^2} |^2_{h^k} = |a^{\otimes k} |^2_{h^k} |_{z_0}. \end{equation*} \notag $$

Now let $\widetilde{\xi}$ be a local holomorphic section of $E^{\ast}$ on a Stein neighbourhood $U$ of $z_0$. We may assume that at $z_0$, $|a|_h = 1$ and $\langle \widetilde{\xi} (z_0), a\rangle_h = | \widetilde{\xi} (z_0) |_h$. We regard $\widetilde{\xi}^{\otimes k}$ as a local holomorphic section of $(E^{\ast})^{\otimes k}$. It then follows that

$$ \begin{equation*} | \widetilde{\xi}^{\otimes k} |_{h^k} = | \xi |^k_h,\qquad | \widetilde{\xi}^{\otimes k} (z_0) |_{h^k} = \langle \widetilde{\xi}^{\otimes k} (z_0), a^{\otimes k} \rangle \end{equation*} \notag $$
and by definition,
$$ \begin{equation} | \widetilde{\xi} (z) |^k_h \geqslant \frac{| \langle \widetilde{\xi}^{\otimes k} (z), v_{k, \log R^2} \rangle |}{|v_{k, \log R^2} |_{h^k}}, \end{equation} \tag{3.8} $$
where $z \in U$, $v_{k, \log R^2}$ is the minimal $L^2$ holomorphic section of $E^{\otimes k}$ such that $v_{k, \log R^2} (z_0) = a^{\otimes k}$. We rewrite (3.8) as
$$ \begin{equation} e^{- k \log | \widetilde{\xi} (z) |_h} \leqslant e^{- \log | \langle \widetilde{\xi}^{\otimes k} (z), v_{k, \log R^2} \rangle |} |v_{k, \log R^2} |_{h^k}. \end{equation} \tag{3.9} $$
Since $\widetilde{\xi}^{\otimes k}$ and $v_{k, \log R^2}$ are local holomorphic sections of $(E^{\ast})^{\otimes k}$ and $E^{\otimes k}$, respectively, $\langle\widetilde{\xi}^{\otimes k} (z), v_{k, \log R^2} \rangle$ is a holomorphic function on $U$. By the Ohsawa–Takegoshi extension theorem, there is a holomorphic section $H$ on $U$ such that $H (z_0) = 1$ and
$$ \begin{equation*} \int_U |H|^2 e^{- 2 \log | \langle \widetilde{\xi}^{\otimes k} (z), v_{k, \log R^2} \rangle |} \leqslant Ce^{- 2 \log | \langle \widetilde{\xi}^{\otimes k} (z_0), a^{\otimes k} \rangle |} = Ce^{- 2 \log | \widetilde{\xi}^k (z_0) |_{h^k}}. \end{equation*} \notag $$
It then follows from (3.9) that
$$ \begin{equation*} \begin{aligned} \, \int_U |H| e^{- k \log | \widetilde{\xi} (z) |_h} &\leqslant \int_U |H| e^{- \log |\langle \widetilde{\xi}^{\otimes k} (z), v_{k, \log R^2} \rangle |} |v_{k, \log R^2}|_{h^k} \\ &\leqslant \biggl[ \int_U |H|^2 e^{- 2 \log | \langle \widetilde{\xi}^{\otimes k}(z), v_{k, \log R^2} \rangle |} \int_U |v_{k, \log R^2} |^2_{h^k} \biggr]^{1/2} \\ &\leqslant \bigl[ Ce^{- 2 \log | \widetilde{\xi}^k (z_0) |_{h^k}} \sigma_{2 n}^{-1} R^{- 2 n} \bigr]^{1/2} = \sqrt{C \sigma_{2 n}^{- 1} R^{- 2 n}}\, e^{- k \log |\widetilde{\xi} (z_0)|}. \end{aligned} \end{equation*} \notag $$
By Theorem 2.7 and Remark 2.8, there exists a sequence of plurisubharmonic functions $\{\varphi_m \} \subset \operatorname{PSH} (U)$ converging to $\log | \widetilde{\xi} |_h$ decreasingly on the set of Lebesgue point of $| a^{\otimes k} |_{h^k}$ up to a set of measure zero, $k = 1, 2, 3, \dots$, with $| a^{\otimes k}|_{h^k} < \infty$ and the requirement of our second condition. Since $\{\varphi_m \} \subset \operatorname{PSH} (U)$ does not converge to $- \infty$ uniformly on $U$, there exists a plurisubharmonic function $\widetilde{\varphi}$ on $U$ such that
$$ \begin{equation*} \lim_{m \to \infty} \varphi_m (z) = \widetilde{\varphi} (z). \end{equation*} \notag $$
By Lemma 2.17, almost every $z \in U$ is a Lebesgue point of $| a^{\otimes k}|_{h^k}$ (up to a set of measure zero) for any $k \in \mathbb{N}^+$. Since the set of points which do not satisfy our second condition and the set of points at which $|a|_h = \infty$ are of measure zero, it follows that, for almost every $z \in U$,
$$ \begin{equation} \log | \widetilde{\xi} |_h = \widetilde{\varphi} (z). \end{equation} \tag{3.10} $$
On the other hand, we have on $U$,
$$ \begin{equation} \limsup_{z \to z_0,\, z \in U \setminus E} \log | \widetilde{\xi} (z) |_h = \log |\widetilde{\xi} (z_0)|_h \end{equation} \tag{3.11} $$
for any subset $E$ of measure zero. For any $z \in U$, we may choose $E$ in (3.11) to be the union of the set of non Lebesgue points of $|a^{\otimes k}|_{h^k}, k = 1, 2, \dots$,
$$ \begin{equation*} \bigcup_k \{ z\in U\colon \text{there is not minimal } L^2 \text{-extension which equals } a^{\otimes k} \text{ at } z\} \end{equation*} \notag $$
and $\{ z \in U \colon |a|_{h} = - \infty \}$, then
$$ \begin{equation*} \widetilde{\varphi} (z) = \limsup_{w \to z,\, w \in U \setminus E} \widetilde{\varphi} (w) = \limsup_{w \to z,\, w \in U \setminus E} \log | \widetilde{\xi} (w) |_h = \log |\widetilde{\xi} (z) |_h, \end{equation*} \notag $$
therefore, $\log | \widetilde{\xi} |_h$ is a plurisubharmonic function on $U$. $\Box$

§ 4. Proof of Theorem 1.5

In this section, we prove Theorem 1.5 with the method which is similar to that of Theorem 1.2.

Proof of Theorem 1.5. We suppose the diameter of $D$ is $R$, then, for any $z, w \in D$, we have $\log | z - w |^2 < \log R^2$. By our hypothesis,
$$ \begin{equation*} \widetilde{\psi} (\tau) = -\log \int_D e^{- k \pi^{\ast} \varphi (\tau, z) - \psi (\tau, z)}\, d V_z = -\log \int_D e^{- k \varphi (z) - \psi (\tau, z)}\, d V_z \end{equation*} \notag $$
is plurisubharmonic for any $k > 0$ and any plurisubharmonic function $\psi(\tau, z)$ on $H_{\tau} \times D_z$. All the integrals are taken with respect to $d V_z$ except written explicitly. Since $e^{-k \varphi}$ is integrable for $k\in\mathbb{N}^+$, almost every $z\in D$ is a Lebesgue point of $e^{-k \varphi}$, $k=1,2,\dots$ . Since $\varphi\in L^1_{\mathrm{loc}}(D)$, for almost every $z\in D$, $\varphi(z)>-\infty$.

In particular, we choose that $H = \{ \tau \in \mathbb{C} \colon \operatorname{Re} \tau < \log R^2 \}$ is the half plane of $\mathbb{C}$ and $\psi (\tau, z) = \max \{ \log | z - z_0 |^2 - \operatorname{Re} \tau,\, 0 \}$, where $z_0$ is a Lebesgue point of $e^{-k\varphi}$, $k=1,2,\dots$, with $\varphi(z)>-\infty$. Then $\widetilde{\psi}$ is a plurisubharmonic function defined on $H \times D$ and depends only on $t = \operatorname{Re} \tau$. Note that

$$ \begin{equation*} \widetilde{\psi} (\tau) = - \log \int_D e^{- k \varphi (z) - p \psi(\tau, z)} \end{equation*} \notag $$
and
$$ \begin{equation*} \begin{aligned} \, &\log \int_D e^{- k \varphi (z) - p \psi (\tau, z)} \\ &\qquad= \log \biggl(\int_{D (\log | z - z_0 |^2 < t)} e^{- k \varphi(z)} + \int_{D (\log | z - z_0 |^2 \geqslant t)} e^{- k\varphi (z) - p \psi (\tau, z)}\biggr) \\ &\qquad >\log \int_{D (\log | z - z_0 |^2 < t)} e^{- k \varphi(z)}, \end{aligned} \end{equation*} \notag $$
where again $D (\log | z - z_0 |^2 < t)$ denotes $\{ z \in D \colon \log | z - z_0 |^2 < t \}$ and similarly, for $D (\log | z - z_0 |^2 \geqslant t)$. Since
$$ \begin{equation*} \limsup_{z\to z_0,\, z\in D\setminus E}\varphi(z)=\varphi(z_0), \end{equation*} \notag $$
for any set $E$ of measure zero, it follows that
$$ \begin{equation*} \liminf_{z\to z_0,\, z\in D}e^{-\varphi(z)}=e^{-\varphi(z_0)}. \end{equation*} \notag $$
That is to say, for any $\varepsilon > 0$, there exists $\delta \in\mathbb{R}$ such that on $D (\log | z - z_0 |^2 < \delta)$,
$$ \begin{equation*} e^{- k \varphi (z)}>e^{- k \varphi (z_0)} - \varepsilon, \end{equation*} \notag $$
and thus if $t<\delta$,
$$ \begin{equation*} \int_{D (\log | z - z_0 |^2 < t)} e^{- k \varphi}>(e^{- k \varphi (z_0)} - \varepsilon) \sigma_{_{2 n}} e^{nt}, \end{equation*} \notag $$
that is,
$$ \begin{equation*} \frac{1}{\sigma_{_{2 n}} e^{nt}} \int_{D (\log | z - z_0 |^2 < t)} e^{- k \varphi}> e^{- k \varphi (z_0)} - \varepsilon, \end{equation*} \notag $$
where $\sigma_{2 n}$ is the volume of $2 n$-dimensional unit ball. Thus for any $p$, when $t < 0$ small enough, we have
$$ \begin{equation} \begin{aligned} \, &\widetilde{\psi} (\tau)+\log \sigma_{2 n}+n t = - \log \frac{1}{\sigma_{2 n} e^{n t}} \int_D e^{- k \varphi (z) - p\psi (\tau, z)} \nonumber \\ &\qquad\leqslant -\log \frac{1}{\sigma_{2 n} e^{n t}} \int_{D (\log | z - z_0 |^2 < t)} e^{- k \varphi (z)} \leqslant -\log ( e^{- k \varphi (z_0)} - \varepsilon). \end{aligned} \end{equation} \tag{4.1} $$
On the other hand, since $\widetilde{\psi} (\tau)+\log \sigma_{2 n}+n t$ depends only on $t = \operatorname{Re} \tau$, $\widetilde{\psi} (\tau)+\log \sigma_{2 n}+n t$ is a convex function. It then follows from (4.1) that
$$ \begin{equation*} \widetilde{\psi} (\tau)+\log \sigma_{2 n}+n t \end{equation*} \notag $$
is an increasing function. Now, for any $t < \log R^2$,
$$ \begin{equation} \widetilde{\psi} (\log R^2)+\log \sigma_{2 n}+n \log R^2 \geqslant \widetilde{\psi} (t)+\log \sigma_{2 n}+n t. \end{equation} \tag{4.2} $$
Since $\psi \geqslant 0$, it follows from (4.2) that
$$ \begin{equation*} \begin{aligned} \, &\frac{1}{\sigma_{2 n} R^{2 n}} \int_D e^{- k \varphi(z)} \leqslant \frac{1}{\sigma_{2 n} e^{n t}} \int_D e^{- k \varphi(z) - p \psi (\tau, z)} \\ &\qquad =\frac{1}{\sigma_{2 n} e^{n t}} \int_{D (\log | z - z_0 |^2 < t)} e^{- k \varphi (z)}+\frac{1}{\sigma_{2 n} e^{n t}} \int_{D (\log | z - z_0 |^2 \geqslant t)} e^{- k \varphi (z) - p \psi (\tau, z)}. \end{aligned} \end{equation*} \notag $$
Therefore, let $p \to \infty$, we have
$$ \begin{equation*} \frac{1}{\sigma_{2 n} R^{2 n}} \int_D e^{- k \varphi (z)} \leqslant \frac{1}{\sigma_{2 n} e^{n t}} \int_{D (\log | z - z_0 |^2 < t)} e^{- k \varphi (z)}, \end{equation*} \notag $$
and now, making $t \to - \infty$, since $z_0$ is a Lebesgue point of $e^{-\varphi}$, we have
$$ \begin{equation*} \frac{1}{\sigma_{2 n} R^{2 n}} \int_D e^{- k \varphi (z)} \leqslant e^{- k \varphi (z_0)}. \end{equation*} \notag $$

By Theorem 2.7 and Remark 2.8, there exists a sequence of plurisubharmonic functions $\{\varphi_m\}\subset \operatorname{PSH}(D)$ converging to $\varphi$ decreasingly on any Lebesgue point of $e^{-k\varphi}$, $k=1,2,\dots$, with $\varphi(z)>-\infty$. Since $\{\varphi_m\}$ does not converge to $-\infty$ uniformly on $D$, there exists a plurisubharmonic function $\widetilde{\varphi}$ on $D$ such that

$$ \begin{equation*} \lim_{m\to \infty}\varphi_m(z)=\widetilde{\varphi}(z). \end{equation*} \notag $$
By Proposition 2.16, almost every $z\in D$ is a Lebesgue point of $e^{-k\varphi}$, $k=1,2,\dots$, and moreover, almost every point of $D$ satisfies $\varphi(z)>-\infty$. It then follows that, for almost every $z\in D$,
$$ \begin{equation*} \varphi(z)=\widetilde{\varphi}(z). \end{equation*} \notag $$
On the other hand, we have on $D$,
$$ \begin{equation} \limsup_{w\to z,\, w\in D\setminus E}\varphi(w)=\varphi(z) \end{equation} \tag{4.3} $$
for any subset $E$ of measure zero. For any $z\in D$, we may choose $E$ in (4.3) to be the union of the non Lebesgue points of $e^{-k\varphi}$, $k=1,2,\dots$, and $\{z\in D\colon \varphi(z)=-\infty\}$, then
$$ \begin{equation*} \widetilde{\varphi}(z)=\limsup_{w\to z,\, w\in D\setminus E}\widetilde{\varphi}(w) =\limsup_{w\to z,\, w\in D\setminus E}\varphi(w)=\varphi(z), \end{equation*} \notag $$
therefore, $\varphi$ is a plurisubharmonic function on $D$.

Remark 4.1. As we have mentioned, the conditions on group action invariance for $D$ and $\varphi$ can be omitted.


Bibliography

1. B. Berndtsson, “Curvature of vector bundles associated to holomorphic fibrations”, Ann. of Math. (2), 169:2 (2009), 531–560  crossref  mathscinet  zmath
2. B. Berndtsson, “Subharmonicity properties of the Bergman kernel and some other functions associated to pseudoconvex domains”, Ann. Inst. Fourier (Grenoble), 56:6 (2006), 1633–1662  crossref  mathscinet  zmath
3. B. Berndtsson and M. Păun, “Bergman kernels and the pseudoeffectivity of relative canonical bundles”, Duke Math. J., 145:2 (2008), 341–378  crossref  mathscinet  zmath
4. Fusheng Deng, Zhiwei Wang, Liyou Zhang, and Xiangyu Zhou, New characterizations of plurisubharmonic functions and positivity of direct image sheaves, arXiv: 1809.10371
5. Fusheng Deng, Jiafu Ning, Zhiwei Wang, and Xiangyu Zhou, “Positivity of holomorphic vector bundles in terms of $L^p$-conditions for $\overline{\partial}$”, Math. Ann., 385:1-2 (2023), 575–607  crossref  mathscinet  zmath; arXiv: 2001.01762
6. T. Inayama, “Nakano positivity of singular Hermitian metrics and vanishing theorems of Demailly–Nadel–Nakano type”, Algebr. Geom., 9:1 (2022), 69–92  crossref  mathscinet  zmath
7. J.-P. Demailly, Complex analytic and differential geometry, 2012 http://www-fourier.ujf-grenoble.fr/~demailly/manuscripts/agbook.pdf
8. Qi'an Guan and Xiangyu Zhou, “A solution of an $L^2$ extension problem with an optimal estimate and applications”, Ann. of Math. (2), 181:3 (2015), 1139–1208  crossref  mathscinet  zmath
9. Qi'an Guan and XiangYu Zhou, “Strong openness of multiplier ideal sheaves and optimal $L^2$ extension”, Sci. China Math., 60:6 (2017), 967–976  crossref  mathscinet  zmath
10. B. Berndtsson and L. Lempert, “A proof of the Ohsawa–Takegoshi theorem with sharp estimates”, J. Math. Soc. Japan, 68:4 (2016), 1461–1472  crossref  mathscinet  zmath
11. A. Prékopa, “On logarithmic concave measures and functions”, Acta Sci. Math. (Szeged), 34 (1973), 335–343  mathscinet  zmath
12. C. O. Kiselman, “The partial Legendre transformation for plurisubharmonic functions”, Invent. Math., 49:2 (1978), 137–148  crossref  mathscinet  zmath  adsnasa
13. B. Berndtsson, “Prekopa's theorem and Kiselman's minimum principle for plurisubharmonic functions”, Math. Ann., 312:4 (1998), 785–792  crossref  mathscinet  zmath
14. K. Ball, F. Barthe, and A. Naor, “Entropy jumps in the presence of a spectral gap”, Duke Math. J., 119:1 (2003), 41–63  crossref  mathscinet  zmath
15. D. Cordero-Erausquin, “On Berndtsson's generalization of Prékopa's theorem”, Math. Z., 249:2 (2005), 401–410  crossref  mathscinet  zmath
16. L. Hörmander, “$L^2$ estimates and existence theorems for the $\overline{\partial}$ operator”, Acta. Math., 113:1 (1965), 89–152  crossref  mathscinet  zmath
17. So-Chin Chen and Mei-Chi Shaw, Partial differential equations in several complex variables, AMS/IP Stud. Adv. Math., 19, Amer. Math. Soc., Providence, RI; International Press, Boston, MA, 2001  crossref  mathscinet  zmath
18. J.-P. Demailly, “Estimations $L^2$ pour l'opérateur $\overline{\partial}$ d'un fibré vectoriel holomorphe semi-positif au-dessus d'une variété Kählerienne complète”, Ann. Sci. École Norm. Sup. (4), 15:3 (1982), 457–511  crossref  mathscinet  zmath
19. J.-P. Demailly, Analytic methods in algebraic geometry, Surv. Mod. Math., 1, International Press, Somerville, MA; Higher Education Press, Beijing, 2010/2012  mathscinet  zmath
20. T. Ohsawa, $L^2$ approaches in several complex variables. Development of Oka–Cartan theory by $L^2$ estimates for the $\overline\partial$ operator, Springer Monogr. Math., Springer, Tokyo, 2015  crossref  mathscinet  zmath
21. Xiangyu Zhou, “A survey on $L^2$ extension problem”, Complex geometry and dynamics, Abel Symp., 10, Springer, Cham, 2015, 291–309  crossref  mathscinet  zmath
22. T. Ohsawa and K. Takegoshi, “On the extension of $L^2$ holomorphic functions”, Math. Z., 195:2 (1987), 197–204  crossref  mathscinet  zmath
23. W. Rudin, Real and complex analysis, 3rd ed., McGraw-Hill Book Co., New York, 1987  mathscinet  zmath
24. V. Guedj and A. Zeriahi, Degenerate complex Monge–Ampère equations, EMS Tracts Math., 26, Eur. Math. Soc. (EMS), Zürich, 2017  crossref  mathscinet  zmath
25. L. Hörmander, Notions of convexity, Progr. Math., 127, Birkhäuser Boston, Inc., Boston, MA, 1994  crossref  mathscinet  zmath
26. L. Lempert, “Modules of square integrable holomorphic germs”, Analysis meets geometry, Trends Math., Birkhäuser/Springer, Cham, 2017, 311–333  crossref  mathscinet  zmath
27. Zhuo Liu, Hui Yang, and Xiangyu Zhou, “New properties of multiplier submodule sheaves”, C. R. Math. Acad. Sci. Paris, 360 (2022), 1205–1212  crossref  mathscinet  zmath
28. Zhuo Liu, Hui Yang, and Xiangyu Zhou, On the multiplier submodule sheaves associated to singular Nakano semi-positive metrics, arXiv: 2111.13452
29. H. Raufi, Log concavity for matrix-valued functions and a matrix-valued Prékopa theorem, arXiv: 1311.7343
30. F. Maitani and H. Yamaguchi, “Variation of Bergman metrics on Riemann surfaces”, Math. Ann., 330:3 (2004), 477–489  crossref  mathscinet  zmath

Citation: Zhi Li, Xiangyu Zhou, “On the positivity of direct image bundles”, Izv. RAN. Ser. Mat., 87:5 (2023), 140–163; Izv. Math., 87:5 (2023), 987–1010
Citation in format AMSBIB
\Bibitem{LiZho23}
\by Zhi~Li, Xiangyu~Zhou
\paper On the positivity of direct image bundles
\jour Izv. RAN. Ser. Mat.
\yr 2023
\vol 87
\issue 5
\pages 140--163
\mathnet{http://mi.mathnet.ru/im9336}
\crossref{https://doi.org/10.4213/im9336}
\mathscinet{http://mathscinet.ams.org/mathscinet-getitem?mr=4666684}
\adsnasa{https://adsabs.harvard.edu/cgi-bin/bib_query?2023IzMat..87..987L}
\transl
\jour Izv. Math.
\yr 2023
\vol 87
\issue 5
\pages 987--1010
\crossref{https://doi.org/10.4213/im9336e}
\isi{https://gateway.webofknowledge.com/gateway/Gateway.cgi?GWVersion=2&SrcApp=Publons&SrcAuth=Publons_CEL&DestLinkType=FullRecord&DestApp=WOS_CPL&KeyUT=001101882800008}
\scopus{https://www.scopus.com/record/display.url?origin=inward&eid=2-s2.0-85177196280}
Linking options:
  • https://www.mathnet.ru/eng/im9336
  • https://doi.org/10.4213/im9336e
  • https://www.mathnet.ru/eng/im/v87/i5/p140
  • Citing articles in Google Scholar: Russian citations, English citations
    Related articles in Google Scholar: Russian articles, English articles
    Известия Российской академии наук. Серия математическая Izvestiya: Mathematics
    Statistics & downloads:
    Abstract page:380
    Russian version PDF:1
    English version PDF:74
    Russian version HTML:39
    English version HTML:153
    References:107
    First page:12
     
      Contact us:
     Terms of Use  Registration to the website  Logotypes © Steklov Mathematical Institute RAS, 2024