Hostname: page-component-7bb8b95d7b-lvwk9 Total loading time: 0 Render date: 2024-09-27T22:15:52.612Z Has data issue: false hasContentIssue false

Second-order estimates for collapsed limits of Ricci-flat Kähler metrics

Published online by Cambridge University Press:  05 January 2023

Kyle Broder*
Affiliation:
School of Mathematics and Physics, The University of Queensland, St. Lucia, QLD 4067, Australia
Rights & Permissions [Opens in a new window]

Abstract

We show that the singularities of the twisted Kähler–Einstein metric arising as the longtime solution of the Kähler–Ricci flow or in the collapsed limit of Ricci-flat Kähler metrics are intimately related to the holomorphic sectional curvature of reference conical geometry. This provides an alternative proof of the second-order estimate obtained by Gross, Tosatti, and Zhang (2020, Preprint, arXiv:1911.07315) with explicit constants appearing in the divisorial pole.

Type
Article
Copyright
© The Author(s), 2023. Published by Cambridge University Press on behalf of The Canadian Mathematical Society

1 Introduction

Declare a compact Kähler manifold $(X, \omega )$ to be Calabi–Yau if the canonical bundle $K_X$ is holomorphically torsion, i.e., $K_X^{\otimes \ell } \simeq \mathcal {O}_X$ for some $\ell \in \mathbb {N}$ . A fiber space is understood to mean a surjective holomorphic map $f : X \longrightarrow Y$ from a Kähler manifold X onto a normal irreducible reduced projective variety Y, with connected, positive-dimensional fibers. A fiber space is said to be Calabi–Yau if the smooth fibers are Calabi–Yau in the above sense. Such Calabi–Yau fiber spaces arise in the study of the Kähler–Ricci flow on compact Kähler manifolds with semi-ample canonical bundle (where f is given by the Iitaka map $\Phi _{| K_X^{\otimes \ell } |}$ ), and collapsed limits of Ricci-flat Kähler metrics [Reference Broder1, Reference Gross, Tosatti and Zhang12Reference Gross, Tosatti and Zhang14, Reference Li and Tosatti21, Reference Song and Tian28Reference Song, Tian and Zhang30, Reference Tosatti34Reference Tosatti and Zhang41]. These families of Kähler metrics are known to converge to a twisted Kähler–Einstein metric $\omega _{\text {can}}$ on the base of such fiber spaces, satisfying

$$ \begin{align*}\text{Ric}(\omega_{\text{can}}) \ = \ \lambda \omega_{\text{can}} + \omega_{\text{WP}}.\end{align*} $$

Here, $\omega _{\text {WP}}$ is the Weil–Petersson metric measuring the variation in the complex structure of the smooth fibers of f.

To state the main results of the present manuscript, let $\tilde {f} : M^m \longrightarrow N^n$ be a Calabi–Yau fiber space, where we assume that $K_M$ is either holomorphically torsion or semi-ample. Denote by $\text {disc}(\tilde {f})$ the discriminant locus of $\tilde {f}$ , i.e., the set of point $p \in N$ over which the corresponding fiber $\tilde {f}^{-1}(p)$ is singular. We denote by $\mathcal {D}$ the divisorial component of $\text {disc}(\tilde {f})$ , and let $\pi : (Y, \mathcal {E}) \longrightarrow (N,\mathcal {D})$ be a log resolution such that $(Y, \mathcal {E})$ is log smooth. By making a base change along $\pi $ , we may construct a Calabi–Yau fiber space $f : X^m \longrightarrow Y^n$ such that Y is a smooth projective variety, and the discriminant locus of f is $\text {disc}(f) = \mathcal {E}$ ; in particular, $\text {disc}(f)$ has simple normal crossing support. Decompose $\mathcal {E}$ into irreducible components $\mathcal {E} = \bigcup _{i=1}^{\mu } \mathcal {E}_i$ , and let $s_i$ be local defining sections for $\mathcal {E}_i$ . Endow the associated line bundles $\mathcal {O}_Y(\mathcal {E}_i)$ with smooth Hermitian metrics $| \cdot |_{h_i}^2$ .

The main conjecture concerning the geometry of these twisted Kähler–Einstein metrics is the following [Reference Broder1, Reference Gross, Tosatti and Zhang12Reference Gross, Tosatti and Zhang14, Reference Li and Tosatti21, Reference Song, Tian and Zhang30, Reference Tosatti34Reference Tosatti37, Reference Tosatti and Zhang39, Reference Tosatti and Zhang41].

Conjecture 1.1

Let $(\mathcal {Z},d_{\mathcal {Z}})$ denote the metric completion of $(N \backslash \mathcal {D}, \text {dist}_{\omega _{\text {can}}})$ . Let

$$ \begin{align*}\Phi : (N \backslash \mathcal{D}, \text{dist}_{\omega_{\text{can}}}) \longrightarrow (\mathcal{Z},d_{\mathcal{Z}})\end{align*} $$

be the corresponding local isometric embedding with singular set ${S_{\mathcal {Z}} : = \mathcal {Z} - \Phi (N \backslash \mathcal {D})}$ . Then:

  1. (i) The real Hausdorff codimension of $S_{\mathcal {Z}}$ (inside $\mathcal {Z}$ ) is at least 2.

  2. (ii) $\left ( M, \text {dist}_{\omega _t} \right )$ converges to $(\mathcal {Z},d_{\mathcal {Z}})$ in the Gromov–Hausdorff topology.

  3. (iii) $\mathcal {Z}$ is homeomorphic to N.

In particular, (iii) implies that the (collapsed) Gromov–Hausdorff limit for these metrics carries an algebro-geometric structure.

Tosatti and Zhang [Reference Tosatti and Zhang41] showed that if the pullback of the twisted Kähler–Einstein metric to the birational model Y had (modulo logarithmic poles) at worst conical singularities, then parts (i) and (ii) of Conjecture 1.1 followed. Moreover, if $(N, \mathcal {D})$ is log smooth, then part (iii) also follows. The required estimate was relaxed by the author [Reference Broder1], showing the following.

Theorem 1.2

Suppose that, for any $\varepsilon>0$ , there is a constant $C=C(\varepsilon )>0$ and $d \in \mathbb {N}$ such that the twisted Kähler–Einstein metric affords the estimate

(1.1) $$ \begin{align} \pi^{\ast} \omega_{\text{can}} & \leq \frac{C}{| s_{\mathcal{F}} |^{2\varepsilon}} \left( 1 - \sum_{i=1}^{\mu} \log | s_i |_{h_i}^2 \right)^d \omega_{\text{cone}}, \end{align} $$

where $\omega _{\text {cone}}$ is a conical Kähler metric and $| s_{\mathcal {F}} |^2 : = \prod _j | s_{\mathcal {F}_j} |_{h_j}^2$ is a shorthand, and $\mathcal {F}_j$ denote the $\pi $ -exceptional divisors. Then statements (i) and (ii) of Conjecture 1.1 hold, and statement (iii) of Conjecture 1.1 holds if Y is smooth.

We refer to (1.1) as the conjectural partial second-order estimate. The partial second-order estimate with the divisorial pole large and not explicit was proved by Gross, Tosatti, and Zhang [Reference Gross, Tosatti and Zhang14]. There, the Aubin–Yau inequality is used, making use of the curvature lower bound on the reference conical metric obtained by Guenancia and Paun [Reference Guenancia and Paun16]. In general, the curvature of the reference conical metric is not bounded below, but is bounded above (see [Reference Jeffres, Mazzeo and Rubinstein19, Reference Lin and Shen22, Reference Rubinstein26]). Even in the case of a single smooth divisor, the computation in [Reference Jeffres, Mazzeo and Rubinstein19] shows that the holomorphic sectional curvature of the reference conical metric decays to $-\infty $ near the divisor.

The main theorem of the present manuscript is to prove the partial second-order estimate with an explicit expression for the divisorial pole.

Theorem 1.3

There are uniform constants $C, \delta _0, \Lambda , \varepsilon _1>0$ and $d \in \mathbb {N}$ such that, for a conical Kähler metric $\omega _{\text {cone}}$ , we have

(1.2) $$ \begin{align} \pi^{\ast} \omega_{\text{can}} & \leq \frac{C}{| s_{\mathcal{F}} |^{2\delta_0(\Lambda + \varepsilon_1)(n-1)}} \left( 1 - \sum_{i=1}^{\mu} \log | s_i |_{h_i} \right)^d \omega_{\text{cone}}. \end{align} $$

In contrast with the partial second-order estimate obtained in [Reference Gross, Tosatti and Zhang14], the divisorial pole is explicit. This will be made clear in the course of the proof (we avoid detailing them here since it becomes rather cumbersome to read). An important and interesting consequence of the explicit nature of the divisorial pole is the following corollary.

Corollary 1.4

Suppose that the holomorphic sectional curvature of the reference conical metric $\omega _{\text {cone}}$ is almost nonpositive in the sense that, for any $\varepsilon _0>0$ , we have $\text {HSC}(\omega _{\text {cone}}) \leq \varepsilon _0$ . Then there is some $\varepsilon>0$ such that

$$ \begin{align*}\pi^{\ast} \omega_{\text{can}} \ \leq \ \frac{C}{| s_{\mathcal{F}} |^{2\varepsilon}} \left( 1 - \sum_{i=1}^{\mu} \log | s_i |_{h_i}^2 \right)^d \omega_{\text{cone}}.\end{align*} $$

Remark 1.5

The proof of the partial second-order estimate makes use of a maximum principle argument applied to a function $\mathcal {Q}$ which we construct. As a consequence, it suffices to assume that the holomorphic sectional curvature is almost nonpositive at the point where $\mathcal {Q}$ achieves its maximum. In particular, we suspect that by modifying the function we consider here, the computation can be localized such that the maximum occurs sufficiently close to the divisor, where the holomorphic sectional curvature of $\omega _{\text {cone}}$ will be negative.

Combining these results, we have the following corollary:

Corollary 1.6

Suppose that the holomorphic sectional curvature of the reference conical metric is almost nonpositive. Then:

  1. (i) The real Hausdorff codimension of $S_{\mathcal {Z}}$ (inside $\mathcal {Z}$ ) is at least 2.

  2. (ii) $\left ( M, \text {dist}_{\omega _t} \right )$ converges to $(\mathcal {Z},d_{\mathcal {Z}})$ in the Gromov–Hausdorff topology.

Moreover, if the base of the Calabi–Yau fiber space is smooth and the divisorial component of the discriminant locus has simple normal crossings, then $\mathcal {Z}$ is homeomorphic to N.

We note that in [Reference Gross, Tosatti and Zhang14], Gross, Tosatti and Zhang obtained the partial second-order estimate with the divisorial pole being a large constant $A>0$ , which is, unfortunately, not explicit. The present work owes substantially to [Reference Gross, Tosatti and Zhang14], from which we build here. We also note that there has been some recent progress on understanding the Gromov–Hausdorff limit by Li and Tosatti in [Reference Li and Tosatti21].

2 Previously known results

2.1 Kähler–Ricci flow

Hamilton’s Ricci flow has proved itself to be a natural candidate for a geometric flow which deforms a fixed Kähler metric to a canonical metric. Indeed, the Ricci flow is known to preserve the Kähler property of the metric; the Ricci flow starting from a Kähler metric $\omega _0$ is therefore referred to as the (normalized) Kähler–Ricci flow:

(2.1) $$ \begin{align} \frac{\partial \omega_t}{\partial t} &= - \text{Ric}(\omega_t) - \omega_t, \hspace{1cm} \omega_t \vert_{t=0} = \omega_0. \end{align} $$

Cao [Reference Cao5] showed that starting with any initial reference Kähler metric $\omega _0$ on a Calabi–Yau manifold, the Kähler–Ricci flow converges smoothly to the Ricci-flat Kähler metric in the polarization $[\omega _0]$ . Similar results hold for the Kähler–Ricci flow on canonically polarized manifolds, i.e., compact Kähler manifolds with ample canonical bundle. For Fano manifolds which admit a Kähler–Einstein metric, or more generally, admit a Kähler–Ricci soliton, the Kähler–Ricci flow converges exponentially fast to the Kähler–Einstein metric or Kähler–Ricci soliton. From the work of Tian and Zhang [Reference Tian and Zhang32], it is known that the maximal existence time T for the Kähler–Ricci flow on a compact Kähler manifold X is determined by cohomological data:

(2.2) $$ \begin{align} T \ = \ \sup \{ t \in \mathbb{R} : [\omega_0] + t [K_X ]> 0 \}. \end{align} $$

In particular, this gives a sharp local existence theorem for the Kähler–Ricci flow, and the Kähler–Ricci flow encounters finite-time singularities if and only if the flow intersects the boundary of the Kähler cone in finite time. If the canonical bundle is nef, then it follows from (2.2) that the Kähler–Ricci flow exists for all time. The resulting solutions, in this case, are referred to as longtime solutions of the Kähler–Ricci flow.

In [Reference Song and Tian28, Reference Song and Tian29], Song and Tian outlined a program for the study of the longtime solutions of the Kähler–Ricci flow on compact Kähler manifolds with semi-ample canonical bundle. The additional assumption of $K_X$ being semi-ample is fruitful since it endows X with the structure of the total space of a Calabi–Yau fiber space. As before, a fiber space is understood to mean a surjective holomorphic map $f : X \longrightarrow Y$ with connected fibers from a compact Kähler manifold X onto a normal irreducible, reduced, projective variety Y. Such a map is said to be a Calabi–Yau fiber space if the smooth fibers $X_y:= f^{-1}(y)$ are Calabi–Yau manifolds in the sense that $K_{X_y}$ is holomorphically torsion.

2.2 Sequences of Ricci-flat Kähler metrics

In [Reference Tosatti35], Tosatti laid the foundational framework, building on [Reference Song and Tian28], for the study of noncollapsed and collapsed limits of Ricci-flat Kähler metrics. Here, let X be a Calabi–Yau manifold and denote by $\mathcal {K}$ the Kähler cone of X. The Kähler cone is an open salient convex cone in the finite-dimensional vector space $H^{1,1}(X, \mathbb {R})$ . With respect to the metric topology induced by any norm on $\mathcal {K}$ , we let $\overline {\mathcal {K}}$ denote the closure of the Kähler cone in $H^{1,1}(X, \mathbb {R})$ and denote by $\partial \mathcal {K}$ its boundary. Fix a nonzero class $\alpha _0$ on the boundary of the Kähler cone,Footnote 1 i.e., a nef class, which we may assume to have the form $\alpha _0 = f^{\ast }[\omega _Y]$ for some Kähler metric $\omega _Y$ on Y. Given a polarization $[\omega _0]$ on M, we may consider the path

(2.3) $$ \begin{align} \alpha_t = \alpha_0 + t[\omega_0], \end{align} $$

for $0 < t \leq 1$ . Yau’s solution of the Calabi conjecture furnishes a bijection between the Ricci-flat Kähler metrics and the points of $\mathcal {K}$ . Hence, given that for each $t> 0$ , the class $\alpha _t$ is Kähler, Yau’s theorem endows $\alpha _t$ with a unique Ricci-flat Kähler representative $\omega _t$ . To understand the behavior of these metrics, therefore, we can vary the complex structure, keeping the cohomological (or symplectic) data fixed, or keep the complex structure fixed and vary the cohomological data; of course, one can also vary both pieces of data, but we will not discuss that here. The former leads to the study of large complex structure limits, which is important in mirror symmetry, but here we will treat only the latter, given its intimate link with the Kähler–Ricci flow and canonical metrics. Therefore, we focus on the problem of understanding the behavior of the metrics $\omega _t$ as the cohomology class $\alpha _t$ degenerates to the boundary of the Kähler cone.

The metrics $\omega _t$ are given by solutions to the Monge–Ampère equation

(2.4) $$ \begin{align} \omega_t^m &= (f^{\ast} \omega_Y + t \omega_X + \sqrt{-1} \partial \overline{\partial} \varphi_t)^m \ = \ c_t t^{m-n} \omega_X^m, \hspace{1cm} \sup_X \varphi_t =0, \end{align} $$

where the constants $c_t$ are bounded away from $0$ and $+\infty $ , and converge as $t\longrightarrow 0$ .

Some remarks concerning the choice of path (2.3) are in order: the reason for considering such a path was initially provided by the fact that this was the path chosen by Gross and Wilson [Reference Gross and Wilson15] in their study of elliptically fibered K3 surfaces with $\text {I}_1$ -singular fibers. However, the motivation for sticking with such a path, however, is the formal analog with the Kähler–Ricci flow (cf. (2.2)). Much of the behavior of the Kähler–Ricci flow can be understood from the study of these sequences of Ricci-flat Kähler metrics (and vice versa). Hence, it has become standard practice to treat both contexts simultaneously, and this practice will be maintained here.

2.3 Twisted Kähler–Einstein metrics

The first systematic approach to the study of these collapsed limits of Ricci-flat Kähler metrics (and the Kähler–Ricci flow in this setting) was given by Song and Tian [Reference Song and Tian28]. They showed that the metrics $\omega _t$ converge to the pullback of a metric $\omega _{\text {can}}$ on the base of the Calabi–Yau fiber space, which satisfies an elliptic equation away from the discriminant locus:

(2.5) $$ \begin{align} \text{Ric}(\omega_{\text{can}}) &= \lambda \omega_{\text{can}} + \omega_{\text{WP}}, \end{align} $$

where $\omega _{\text {WP}}$ is the Weil–Petersson metric measuring the variation in the complex structure of the smooth fibers. A precise description of $\omega _{\text {WP}}$ was given by Tian [Reference Tian31], where he showed that $\omega _{\text {WP}}$ is the curvature form of the Hodge metric on $f_{\ast } \Omega _{M/N}$ . Let us note that in (2.5), $\lambda =0$ for sequences of Ricci-flat Kähler metrics, and $\lambda =-1$ for the longtime solution of the Kähler–Ricci flow (see [Reference Song and Tian28, Reference Song and Tian29, Reference Tosatti35]). In [Reference Tosatti35], it was shown that $\omega _t \to f^{\ast } \omega $ in the $\mathcal {C}_{\text {loc}}^{1,\gamma }(M^{\circ })$ topology of Kähler potentials for all $0 < \gamma < 1$ . In [Reference Tosatti, Weinkove and Yang38], this convergence was improved to local uniform convergence and to $\mathcal {C}_{\text {loc}}^{\alpha }(M^{\circ })$ in [Reference Hein and Tosatti18]. If the generic fiber is a (finite quotient) of a torus, or if the family is isotrivial in the sense that all smooth fibers are biholomorphic, the convergence can be improved to the $\mathcal {C}_{\text {loc}}^{\infty }(M^{\circ })$ topology (see [Reference Gross, Tosatti and Zhang12, Reference Hein and Tosatti17, Reference Tosatti and Zhang40] and [Reference Hein and Tosatti18], respectively).

If $\kappa = n$ , then $K_M$ is nef and big. In this case, it is known [Reference Tian and Zhang32, Reference Tsuji42] that the Kähler–Ricci flow converges weakly to the canonical metric on N, which is smooth on the regular part of Y. Furthermore, in [Reference Song27], it is shown that the metric completion of the Kähler–Einstein metric on $N^{\circ }$ is homeomorphic to N. If $0 < \kappa < n$ , then $K_M$ is no longer big, and much less is known; it is this case that we consider here. It is expected that the Ricci curvature of the metrics along the flow remains uniformly bounded on compact subsets of $M^{\circ }$ (see, e.g., [Reference Tian and Zhang33]). This is known to be the case when the generic fiber is a (finite quotient of a) torus. Recently, in [Reference Fong and Lee10], the higher-order estimates in [Reference Hein and Tosatti18] were used to obtain the uniform bound on $\text {Ric}(\omega _t)$ when the smooth fibers are biholomorphic. Assuming this uniform bound on the Ricci curvature of the metrics along the flow, one can formulate Conjecture 1.1 for the Kähler–Ricci flow.

In [Reference Tosatti and Zhang41], Tosatti and Zhang initiated a program to attack Conjecture 1.1 by understanding the nature of the singularities of the twisted Kähler–Einstein metric $\omega _{\text {can}}$ near the discriminant locus. They showed that if the canonical metric afforded the following $\mathcal {C}^2$ -estimate:

(2.6) $$ \begin{align} \pi^{\ast} \omega_{\text{can}} & \leq C \left( 1 - \sum_{i=1}^{\mu} \log | s_i |_{h_i}^2 \right)^d \omega_{\text{cone}}, \end{align} $$

then parts (i) and (ii) of Conjecture 1.1 are true. That is, to prove the conjecture, it suffices to show that (modulo some logarithmic poles) the canonical metric is at worst conical near the discriminant locus. For part (iii), Tosatti and Zhang required the additional assumption that the base of the Calabi–Yau fiber space is smooth and the divisorial component of the discriminant locus has simple normal crossings (in particular, they require the resolution $\pi $ to be the identity).

3 Proof of the main theorem

Let us recall the setup from [Reference Gross, Tosatti and Zhang14]: let $\tilde {f} : M^m \longrightarrow N^n$ be a Calabi–Yau fiber space with discriminant locus $\text {disc}(\tilde {f})$ . Let $\mathcal {D}$ denote the divisorial component of $\text {disc}(\tilde {f})$ , and let $\pi : (Y,\mathcal {E}) \longrightarrow (N, \mathcal {D})$ be log resolution such that Y is smooth, $\mathcal {E} = \pi ^{-1}(\text {disc}(\tilde {f}))$ has simple normal crossings, and $\pi : Y^{\circ } = Y \backslash \mathcal {E} \longrightarrow N \backslash \mathcal {D}$ is an isomorphism. Let $\tau : X \to M \times _N Y$ be a birational morphism which induces the Calabi–Yau fiber space $f : X \longrightarrow Y$ with X smooth, $f^{-1}(\mathcal {E})$ a divisor with simple normal crossings, and $\text {disc}(f) = \mathcal {E}$ . The twisted Kähler–Einstein metric $\omega _{\text {can}}$ (2.5) is given by

$$ \begin{align*}\omega_{\text{can}} = \omega_N + \sqrt{-1} \partial \overline{\partial} \varphi\end{align*} $$

for some Kähler metric $\omega _N$ (understood in the sense of Kähler spaces if N is not smooth; see, e.g., [Reference Moishezon24]) and a continuous $\omega _N$ -plurisubharmonic function $\varphi $ . Let ${\pi : (Y, \mathcal {E}) \longrightarrow (N,\mathcal {D})}$ be a log resolution as before. The pulled-back metric $\pi ^{\ast } \omega _{\text {can}}$ then satisfies

$$ \begin{align*} (\pi^{\ast} \omega_{\text{can}})^n &= (\pi^{\ast} \omega_N + \sqrt{-1} \partial \overline{\partial} (\pi^{\ast} \varphi))^n. \end{align*} $$

Recall that $\mathcal {E} = \bigcup _{i=1}^{\mu } \mathcal {E}_i$ is a decomposition of $\mathcal {E}$ into irreducible components. For $\alpha _i \in (0,1] \cap \mathbb {Q}$ , we may associate a conical metric $\omega _{\text {cone}}$ with cone angle $2\pi \alpha _i$ along $\mathcal {E}_i$ . In more detail, $\omega _{\text {cone}}$ is smooth on $Y \backslash \mathcal {E}$ , and in any adapted coordinate system, $\omega _{\text {cone}}$ is uniformly equivalent to the model metric

$$ \begin{align*}\sqrt{-1} \sum_{i=1}^k \frac{dz_i \wedge d\overline{z}_i}{| z_i |^{2(1-\alpha_{j_i})}} + \sqrt{-1} \sum_{i=k+1}^n dz_i \wedge d\overline{z}_i.\end{align*} $$

Proposition 3.1

[Reference Gross, Tosatti and Zhang14, Theorem 2.3]. There is a constant $C>0$ , positive integers $d \in \mathbb {N}$ , $0 \leq p \leq \mu $ , and rational numbers $\beta _i> 0$ ( $1 \leq i \leq p$ ), $0 < \alpha _i \leq 1$ ( $p+1 \leq i \leq \mu $ ), such that on $Y^{\circ }$ ,

$$ \begin{align*} C^{-1} \prod_{j=1}^p | s_j |_{h_j}^{2\beta_j} \omega_{\text{cone}}^n \ \leq \ \pi^{\ast} \omega_{\text{can}}^n \ \leq \ C \prod_{j=1}^p | s_j |_{h_j}^{2\beta_j} \left( 1 - \sum_{i=1}^{\mu} \log | s_i |_{h_i}^2 \right)^d \omega_{\text{cone}}^n, \end{align*} $$

where $\omega _{\text {cone}}$ is a conical Kähler metric with cone angle $2\pi \alpha _i$ along components $\mathcal {E}_i$ with $p+1 \leq i \leq \mu $ .

Consider the smooth nonnegative function

$$ \begin{align*}H \ : = \ \prod_{j=1}^p | s_j |_{h_j}^{2\beta_j},\end{align*} $$

where the product is over those $\pi $ -exceptional components (which we denote by $\mathcal {F}$ ) on which the volume form $\pi ^{\ast }\omega _{\text {can}}^{n}$ degenerates. Away from the exceptional divisor $\mathcal {F}$ ,

$$ \begin{align*} \sqrt{-1} \partial \overline{\partial} \log(H) &= -\sum_{j=1}^p \beta_j \Theta^{(\mathcal{F}_j,h_j)}, \end{align*} $$

where $\Theta ^{(\mathcal {F}_j,h_j)}$ is the curvature form of the Hermitian metric $h_j$ on $\mathcal {F}_j$ . Pulling back the Monge–Ampère equation to Y via $\pi $ , we get the degenerate, singular Monge–Ampère equation:

(3.1) $$ \begin{align} \pi^{\ast} \omega_{\text{can}}^n &= (\pi^{\ast} \omega_N + \sqrt{-1} \partial \overline{\partial}(\pi^{\ast}\varphi))^{n} \ \ = \ \ H \psi \frac{\omega_Y^n}{\prod_{i=p+1}^{\mu} | s_i |_{h_i}^{2(1-\alpha_i)}}, \end{align} $$

where $\omega _N$ is the reference form on N and $\omega _Y$ is a reference form on Y. From Proposition 3.1, the function

$$ \begin{align*}\psi \ : = \ \frac{\pi^{\ast} \omega_{\text{can}}^{n} \prod_i | s_i |_{h_i}^{2(1-\alpha_i)}}{H \omega_{Y}^{n}},\end{align*} $$

which is smooth on $Y^{\circ } \ = \ Y \backslash E$ , has the following growth control:

(3.2) $$ \begin{align} C^{-1} \ \leq \ \psi \ \leq \ C \left( 1 - \sum_{i=1}^{\mu} \log | s_i |_{h_i}^2 \right)^d. \end{align} $$

Away from the divisor $\mathcal {E}$ , compute

(3.3) $$ \begin{align} \sqrt{-1} \partial \overline{\partial} \log(1/\psi) &= \text{Ric}(\pi^{\ast} \omega_{\text{can}}) - \text{Ric}(\omega_{Y}) + \sum_i (1-\alpha_i) \Theta^{(\mathcal{E}_i,h_i)} + \sqrt{-1} \partial \overline{\partial} \log(H) \nonumber \\ & \geq - \text{Ric}(\omega_{Y}) + \sum_i (1-\alpha_i) \Theta^{(\mathcal{E}_i,h_i)} - \sum_j \beta_j \Theta^{(\mathcal{F}_j,h_j)}, \end{align} $$

where the right-hand side of (3.3) is smooth on all of Y. In light of (3.2), an old result of Grauert and Remmert [Reference Grauert and Remmert11] tells us that $\log (1/\psi )$ extends to a quasi-plurisubharmonic function with vanishing Lelong numbers on all of Y, satisfying (3.3) everywhere in the sense of currents. Apply Demailly’s regularization technique [Reference Demailly6], such that we can approach $\log (1/\psi )$ from above by a decreasing sequence of smooth functions $u_j$ which afford the lower bound

$$ \begin{align*} \sqrt{-1} \partial \overline{\partial} u_j & \geq - \text{Ric}(\omega_{Y}) + \sum_i (1-\alpha_i) \Theta^{(\mathcal{E}_i,h_i)} - \sum_j \beta_j \Theta^{(\mathcal{F}_j,h_j)}- \frac{1}{j} \omega_{Y}, \end{align*} $$

on all of Y. Similarly, since $\sqrt {-1} \partial \overline {\partial } \log (H) \geq - C \omega _{Y}$ (in the sense of currents) on Y, we may regularize $\log (H)$ by smooth functions $v_k$ with $\sqrt {-1} \partial \overline {\partial } v_k \geq -C \omega _{Y}$ and $v_k \leq C$ on Y. Out of this, we obtain a regularization of (3.1):

$$ \begin{align*}\omega_{j,k}^{n} \ = \ c_{j,k} e^{v_k -u_j} \frac{\omega_{Y}^{n}}{\prod_i | s_i |_{h_i}^{2(1-\alpha_i)}},\end{align*} $$

where

$$ \begin{align*}\omega_{j,k} \ = \ \pi^{\ast} \omega_N + \frac{1}{j} \omega_{Y} + \sqrt{-1} \partial \overline{\partial} \varphi_{j,k}\end{align*} $$

is a cone metric (which exists by the main theorems of [Reference Guenancia and Paun16, Reference Jeffres, Mazzeo and Rubinstein19]) with the same cone angles as $\omega _{\text {cone}}$ . In particular, the metrics $\omega _{j,k}$ are smooth on $Y^{\circ }$ and there is a uniform (independent of j) bound on the $L^p$ -norm of $\omega _{j,k}^{n}/\omega _{Y}^{n}$ , for some $p>1$ . By [Reference Eyssidieux, Guedj and Zeriahi9], it follows that

$$ \begin{align*}\sup_Y | \varphi_{j,k} | \ \leq \ C,\end{align*} $$

and since $e^{v_k -u_j} \longrightarrow H \psi $ , we conclude from the stability theorem in [Reference Eyssidieux, Guedj and Zeriahi9] that as $j,k \longrightarrow \infty $ , we have $c_{j,k} \longrightarrow 1$ , $\varphi _{j,k} \longrightarrow \pi ^{\ast } \varphi $ in $\mathcal {C}^0(Y)$ , where $\pi ^{\ast } \varphi $ solves (3.1). Fix a reference metric $\omega _Y$ on Y and consider partial regularizations

(3.4) $$ \begin{align} \omega_j = \pi^{\ast} \omega_N + \frac{1}{j} \omega_Y + \sqrt{-1} \partial \overline{\partial} \varphi_j \end{align} $$

given by solutions of the Monge–Ampère equation

(3.5) $$ \begin{align} \omega_j^n &= c_j H e^{-u_j} \frac{\omega_Y^n}{\prod_{i=1}^{\mu} | s_i |_{h_i}^{2(1-\alpha_i)}}. \end{align} $$

Proof of Theorem 1.3

For positive constants $\delta , \delta _0, \Lambda _{\delta , j}, \varepsilon , \varepsilon _0, \varepsilon _1, b, C_1>0$ , we will apply the maximum principle to the quantity

$$ \begin{align*} \mathcal{Q} & : = \log \text{tr}_{\omega_j}(\omega_{\text{cone}}) + \delta \log(H) + ( (\Lambda_{\delta,j} + \varepsilon_0) b - C_1 \delta) \eta - (\Lambda_{\delta,j} + \varepsilon_1) \varphi_j \\ & \quad - \delta_0(\Lambda_{\delta,j} + \varepsilon_0) \log | s_{\mathcal{F}} |^2 + \varepsilon \log | s_{\mathcal{E}} |^2. \end{align*} $$

Here, we fix a conical Kähler metric $\omega _{\text {cone}}$ with cone angle $2\pi \alpha _i$ along $\mathcal {E}_i$ (an irreducible component of $\mathcal {E})$ . For a reference Kähler metric $\omega _Y$ , we set $\omega _{\text {cone}} = \omega _Y + \sqrt {-1} \partial \bar {\partial } \eta $ , where $\eta : = C^{-1} \sum _i \left ( | s_i |_{h_i}^2 \right )^{\alpha _i}$ for some sufficiently large $C>0$ . The constants will be described throughout the proof. We start with the following observation: from [Reference Gross, Tosatti and Zhang14, Proposition 5.1], there is a j-dependent constant $C_j$ such that $\text {tr}_{\omega _{\text {cone}}}(\omega _j) \leq C_j$ . From the complex Monge–Ampère equation, this implies that for any $\delta \in (0,1)$ ,

$$ \begin{align*} \text{tr}_{\omega_j}(\omega_{\text{cone}}) & \leq \frac{C_j^{n-1}}{c_j H e^{-u_j}} \ \leq \ \frac{C_j^{n-1}}{c_j H^{\delta} e^{-u_j}}. \end{align*} $$

In particular, there is a j-dependent constant (abusively denoted by the same symbol) $C_j$ such that

$$ \begin{align*}H^{\delta} e^{-u_j} \text{tr}_{\omega_j}(\omega_{\text{cone}}) \ \leq \ C_j.\end{align*} $$

This will allow us to apply the maximum principle to terms involving $\text {tr}_{\omega _j}(\omega _{\text {cone}})$ so long as we scale $\text {tr}_{\omega _j}(\omega _{\text {cone}})$ by $H^{\delta } e^{-u_j}$ and add a term which decays to $-\infty $ along the divisor $\mathcal {E}$ (at an arbitrary rate). This latter term is given by $\varepsilon \log | s_{\mathcal {E}} |^2$ , where ${0 < \varepsilon \leq \frac {1}{j}}$ . To compute $\Delta _{\omega _j} \log \text {tr}_{\omega _j}(\omega _{\text {cone}})$ , we observe that the complex Monge–Ampère equation gives the following control of the Ricci curvature of $\omega _j$ :

$$ \begin{align*} \text{Ric}(\omega_j) & = - \sqrt{-1} \partial \bar{\partial} \log(H)+ \sqrt{-1} \partial \bar{\partial}u_j + \text{Ric}(\omega_Y) + \sqrt{-1} \partial \bar{\partial} \log \prod |s_i |_{h_i}^{2(1-\alpha_i)} \\ & \geq - \frac{1}{j} \omega_Y \ \geq \ - \frac{C}{j} \omega_{\text{cone}}. \end{align*} $$

Let $g,h$ be metrics underlying $\omega _j$ and $\omega _{\text {cone}}$ , respectively. Then, for any holomorphic map $f : (Y^{\circ }, \omega _j) \to (Y^{\circ }, \omega _{\text {cone}})$ , with $(f_i^{\alpha })$ denoting the local expression for $\partial f$ , we have (see, e.g., [Reference Broder2, Reference Broder3, Reference Yang and Zheng44])

$$ \begin{align*} \Delta_{\omega_j} | \partial f |^2 & = | \nabla \partial f |^2 + \text{Ric}_{k \overline{\ell}}^g g^{k \overline{q}} g^{p \overline{\ell}} h_{\alpha \overline{\beta}} f_p^{\alpha} \overline{f_q^{\beta}} - R^h_{\alpha \overline{\beta} \gamma \overline{\delta}} g^{i \overline{j}} f_i^{\alpha} \overline{f_j^{\beta}} g^{p \overline{q}} f_p^{\gamma} \overline{f_q^{\delta}}. \end{align*} $$

The lower bound on the Ricci curvature of g implies that

$$ \begin{align*} \text{Ric}_{k \overline{\ell}}^g g^{k \overline{q}} g^{p \overline{\ell}} h_{\alpha \overline{\beta}} f_p^{\alpha} \overline{f_q^{\beta}} & \geq - \frac{C}{j} h_{\gamma \overline{\delta}} g^{k \overline{q}} g^{p \overline{\ell}} h_{\alpha \overline{\beta}} f_p^{\alpha} \overline{f_q^{\beta}}. \end{align*} $$

Choose the frame such that at the point where we are computing, we have $f_i^{\alpha } = \lambda _i \delta _i^{\alpha }$ , $g_{i\overline {j}} = \delta _{ij}$ and $h_{\alpha \overline {\beta }} = \delta _{\alpha \overline {\beta }}$ . Then

$$ \begin{align*} - \frac{C}{j} h_{\gamma \overline{\delta}} f_k^{\gamma} \overline{f_{\ell}^{\delta}} g^{k \overline{q}} g^{p \overline{\ell}} h_{\alpha \overline{\beta}} f_p^{\alpha} \overline{f_q^{\beta}} &= - \frac{C}{j} \delta_{\gamma \delta} \lambda_k \delta_k^{\gamma} \lambda_{\ell} \delta_{\ell}^{\delta} \delta^{kq} \delta^{p\ell} \delta_{\alpha \beta} \lambda_p \delta_p^{\alpha} \lambda_q \delta_q^{\beta} \ = \ - \frac{C}{j} \lambda_{\alpha}^4. \end{align*} $$

Since $\sum _{\alpha } \lambda _{\alpha }^4 \leq \left ( \sum _{\alpha } \lambda _{\alpha }^2 \right )^2 = | \partial f |^4$ , we deduce that

$$ \begin{align*} \text{Ric}_{k \overline{\ell}}^g g^{k \overline{q}} g^{p \overline{\ell}} h_{\alpha \overline{\beta}} f_p^{\alpha} \overline{f_q^{\beta}} & \geq - \frac{C}{j} \text{tr}_{\omega_j}(\omega_{\text{cone}})^2. \end{align*} $$

From [Reference Jeffres, Mazzeo and Rubinstein19, Reference Lin and Shen22], we know that there is a uniform constant $K>0$ such that $\text {HBC}(\omega _{\text {cone}}) \leq K$ . This, in particular, implies that the holomorphic sectional curvature of the cone metric $\omega _{\text {cone}}$ affords the (positive) upper bound ${\text {HSC}(\omega _{\text {cone}}) \leq K}$ for $K>0$ . Royden’s polarization argument [Reference Broder3, Reference Broder4, Reference Lu23, Reference Royden25] implies that

$$ \begin{align*} R^h_{\alpha \overline{\beta} \gamma \overline{\delta}} g^{i \overline{j}} f_i^{\alpha} \overline{f_j^{\beta}} g^{p \overline{q}} f_p^{\gamma} \overline{f_q^{\delta}} & \leq K \text{tr}_{\omega_j}(\omega_{\text{cone}})^2. \end{align*} $$

Hence, by Royden’s Schwarz lemma [Reference Broder4, Reference Lu23, Reference Royden25], we see that

$$ \begin{align*} \Delta_{\omega_j} \log \text{tr}_{\omega_j}(\omega_{\text{cone}}) &\geq - \left( \frac{C}{j} + K \right) \text{tr}_{\omega_j}(\omega_{\text{cone}}), \end{align*} $$

at any point away from the divisor $\mathcal {E}$ . Therefore,

$$ \begin{align*}& \Delta_{\omega_j} (\log \text{tr}_{\omega_j}(\omega_{\text{cone}}) + \delta \log(H) - u_j ) \geq - \left( \frac{C}{j} + K \right) \text{tr}_{\omega_j}(\omega_{\text{cone}}) \\ & \quad+ \delta \text{tr}_{\omega_j}(\sqrt{-1} \partial \bar{\partial} \log(H)) - \Delta_{\omega_j} u_j, \end{align*} $$

at any point away from the divisor $\mathcal {E}$ . Since $\sqrt {-1} \partial \bar {\partial } \log (H) \geq - C_1 \omega _Y$ , and adding $u_j$ to the quantity being differentiated (which will not affect the maximum principle), we see that

$$ \begin{align*} \Delta_{\omega_j} (\log \text{tr}_{\omega_j}(\omega_{\text{cone}}) + \delta \log(H)) & \geq - \left( \frac{C}{j} + K \right) \text{tr}_{\omega_j}(\omega_{\text{cone}}) - \delta C_1 \text{tr}_{\omega_j}(\omega_Y), \end{align*} $$

at any point away from $\mathcal {E}$ . We know that the conical metric is given by $\omega _{\text {cone}} = \omega _Y + \sqrt {-1} \partial \bar {\partial } \eta $ . Hence,

$$ \begin{align*} \Delta_{\omega_j}(\log \text{tr}_{\omega_j}(\omega_{\text{cone}}) +\delta \log(H) - \delta C_1 \eta + \varepsilon \log | s_{\mathcal{E}} |^2) & \geq -\kern-1pt \left( \frac{C}{j} + K + C_1 \delta \right) \kern-1pt\text{tr}_{\omega_j}(\omega_{\text{cone}}), \end{align*} $$

and we adopt the notation $\Lambda _{\delta ,j} : = \frac {C}{j} + K + C_1 \delta $ . To ensure the positivity of the coefficient of $\text {tr}_{\omega _j}(\omega _{\text {cone}})$ , let $\varepsilon _1>0$ be a positive constant to be determined later. Then

$$ \begin{align*} &\Delta_{\omega_j}( \log \text{tr}_{\omega_j}(\omega_{\text{cone}}) + \delta \log(H) - C_1 \delta \eta + \varepsilon \log | s_{\mathcal{E}} |^2 - (\Lambda_{\delta,j} + \varepsilon_1) \varphi_j) \\ &\quad\geq - \Lambda_{\delta,j} \text{tr}_{\omega_j}(\omega_{\text{cone}}) - (\Lambda_{\delta,j} + \varepsilon_1)n + (\Lambda_{\delta,j} +\varepsilon_1) \text{tr}_{\omega_j}(\pi^{\ast} \omega). \end{align*} $$

The pullback $\pi ^{\ast } \omega $ is degenerate in the directions tangent to the divisor. Let $\delta _0>0$ be the suitably small constant such that $\pi ^{\ast } \omega - \delta _0 \Theta ^{(\mathcal {F},h)}$ is positive-definite. Furthermore, we let $b>0$ be the suitably small constant such that $\pi ^{\ast } \omega - \delta _0 \Theta ^{(\mathcal {F},h)} + b \sqrt {-1} \partial \bar {\partial } \eta $ has the same cone angles as $\omega _{\text {cone}}$ . Hence, we can find a constant $c_0>0$ such that

$$ \begin{align*}\pi^{\ast} \omega - \delta_0 \Theta^{(\mathcal{F},h)} + b \sqrt{-1} \partial \bar{\partial} \eta \ \geq \ c_0 \omega_{\text{cone}}.\end{align*} $$

Hence, with $\mathcal {Q}$ defined above, we see that

$$ \begin{align*} \Delta_{\omega_j} \mathcal{Q} & \geq - \Lambda_{\delta, j} \text{tr}_{\omega_j}(\omega_{\text{cone}}) \kern1.3pt{-}\kern1.3pt (\Lambda_{\delta,j} + \varepsilon_1) n + (\Lambda_{\delta,j} + \varepsilon_1) \text{tr}_{\omega_j}(\pi^{\ast} \omega - \delta_0 \Theta^{(\mathcal{F},h)} + b \sqrt{-1} \partial \bar{\partial} \eta) \\ & \geq (c_0(\Lambda_{\delta,j} + \varepsilon_1) - \Lambda_{\delta,j}) \text{tr}_{\omega_j}(\omega_{\text{cone}}) - (\Lambda_{\delta,j} + \varepsilon_1) n. \end{align*} $$

If $c_0 \geq 1$ , then $c_0 (\Lambda _{\delta ,j} + \varepsilon _1) - \Lambda _{\delta ,j} \geq c_0 \varepsilon _0$ , and we take $\varepsilon _1>0$ arbitrary. For instance, if we take $\varepsilon _1 = c_0^{-1}$ , then

$$ \begin{align*} \Delta_{\omega_j} \mathcal{Q} & \geq \text{tr}_{\omega_j}(\omega_{\text{cone}}) - (\Lambda_{\delta,j} + c_0^{-1}) n. \end{align*} $$

By the maximum principle, at the point where $\mathcal {Q}$ achieves its maximum, we have

$$ \begin{align*}\text{tr}_{\omega_j}(\omega_{\text{cone}}) \ \leq \ n(\Lambda_{\delta,j} + c_0^{-1}) n \ \leq \ C,\end{align*} $$

for some uniform constant $C>0$ . On the other hand, if $0 < c_0 < 1$ , then we choose $\varepsilon _1 = c_0^{-1}(1 + \Lambda _{\delta ,j}(1-c_0))$ . In this case, we see that

$$ \begin{align*} \Delta_{\omega_j}(\mathcal{Q}) & \geq \text{tr}_{\omega_j}(\omega_{\text{cone}}) - c_0^{-1}n(1+\Lambda_{\delta,j}), \end{align*} $$

and by the maximum principle, at the point where $\mathcal {Q}$ achieves its maximum,

$$ \begin{align*}\text{tr}_{\omega_j}(\omega_{\text{cone}}) \ \leq \ c_0^{-1} n(1+\Lambda_{\delta,j}).\end{align*} $$

Both estimates imply that $\mathcal {Q} \leq C$ for some uniform constant $C>0$ . Hence, we have the estimate

$$ \begin{align*} \text{tr}_{\omega_j}(\omega_{\text{cone}}) & \leq \frac{C}{H^{\delta} | s_{\mathcal{F}} |^{2\delta_0(\Lambda_{\delta,j}+\varepsilon_1)}} \end{align*} $$

everywhere on $Y^{\circ }$ . From the complex Monge–Ampère equation, we see that

$$ \begin{align*} \text{tr}_{\omega_{\text{cone}}}(\omega_j) & \leq \text{tr}_{\omega_j}(\omega_{\text{cone}})^{n-1} \frac{\omega_j^n}{\omega_{\text{cone}}^n} \ \leq \ \frac{CH e^{u_j}}{| s_{\mathcal{F}} |^{2\delta_0(\Lambda_{\delta,j} + \varepsilon_1)(n-1)} H^{\delta(n-1)}}. \end{align*} $$

Choosing $\delta < \frac {1}{n-1}$ , we see that

$$ \begin{align*} \text{tr}_{\omega_{\text{cone}}}(\omega_j) & \leq \frac{Ce^{-u_j}}{| s_{\mathcal{F}} |^{2\delta_0(\Lambda_{\delta,j} + \varepsilon_1)(n-1)}}, \end{align*} $$

and letting $j \to \infty $ , we obtain the partial second-order estimate

$$ \begin{align*} \pi^{\ast} \omega_{\text{can}} & \leq \frac{C}{| s_{\mathcal{F}} |^{2\delta_0(\Lambda_{\delta,j} + \varepsilon_1)(n-1)}} \left( 1 - \sum_{i=1}^{\mu} \log | s_i |_{h_i}^2 \right)^d \omega_{\text{cone}}. \end{align*} $$

Proof of Corollary 1.4

Suppose the holomorphic sectional curvature of $\omega _{\text {cone}}$ is almost nonpositive, i.e., for any $\varepsilon _0>0$ , we have $\text {HSC}(\omega _{\text {cone}}) \leq \varepsilon _0$ , then

$$ \begin{align*} 2\delta_0(\Lambda_{\delta,j} + \varepsilon_1)(n-1) & \leq 2\delta_0 \left( \frac{C}{j} + \varepsilon_0 + C_1 \delta \right)(n-1). \end{align*} $$

The constant $\delta \in (0, \frac {1}{n-1})$ can be made arbitrarily small, and therefore, as $j \to \infty $ , we have

$$ \begin{align*} 2\delta_0 \left( \frac{C}{j} + \varepsilon_0 + C_1 \delta \right) (n-1) \to 2 \delta_0(n-1)(\varepsilon_0 + C_1 \delta). \end{align*} $$

In particular, if the holomorphic sectional curvature of the conical Kähler metric is bounded above by an arbitrarily small positive constant, then the multiplicity of the divisorial pole can be made arbitrarily small.

Remark 3.2

The assumption that the holomorphic sectional curvature of the conical metric is arbitrarily small is not as restrictive as one may expect. We suspect that by appropriately localizing the function $\mathcal {Q}$ used in the maximum principle, one should be able to ensure that the holomorphic sectional curvature of $\omega _{\text {cone}}$ is almost nonpositive at the point where $\mathcal {Q}$ achieves its maximum. To see this, consider the case where the divisor $\mathcal {E}$ consists of a single component, locally described by $\{ z_1 =0 \}$ . Then the computation in [Reference Jeffres, Mazzeo and Rubinstein19, Appendix] shows that

$$ \begin{align*}R_{1 \overline{1} 1 \overline{1}} \sim -\alpha^2(\alpha-1)^2 | z |^{-2(2-\alpha)}.\end{align*} $$

In other words, if the function $\mathcal {Q}$ can be tweaked such that the maximum occurs sufficiently close to $\mathcal {E}$ , but not on $\mathcal {E}$ , then the holomorphic sectional curvature of $\omega _{\text {cone}}$ should be negative at this point, and the argument goes through.

3.1 The holomorphic sectional curvature

The main theorem, although primarily concerned with Conjecture 1.1, fits into a more general program the author has initiated. There are the three primary aspects of the holomorphic sectional curvature:

  1. (i) Symmetries—The holomorphic sectional curvature of a Kähler metric (or more generally, Kähler-like metric) determines the curvature tensor entirely.

  2. (ii) Value distribution of curves—A compact Kähler manifold with negative holomorphic sectional curvature is Kobayashi hyperbolic (every entire curve $\mathbb {C} \to X$ is constant). On the other hand, a compact Kähler manifold with positive holomorphic sectional curvature is rationally connected (any two points lie in the image of a rational curve $\mathbb {P}^1 \to X$ ) [Reference Yang43].

  3. (iii) Singularities—The holomorphic sectional curvature influences the singularities of the geometry.

The first facet of the holomorphic sectional curvature is well known to all complex geometers. The assertions in statement (ii) are also well known, but it is worth emphasizing a particular subtlety concerning (ii): let $(X, \omega )$ be a Hermitian manifold with a Hermitian metric of negative (Chern) holomorphic sectional curvature ${{}^c \text {HSC}(\omega ) \leq - \kappa <0}$ . Then X is Brody hyperbolic in the sense that every entire curve $\mathbb {C} \to X$ is constant. The proof is an elementary application of the Schwarz lemma. Suppose that there is a nonconstant map $f : \mathbb {C} \to X$ . Then $\Delta _{\omega _{\mathbb {C}}} | \partial f |^2 \geq - {}^c R(\partial f, \bar {\partial f}, \partial f, \bar {\partial f})$ , and by the maximum principle, f is constant. The key point here, however, is that for holomorphic curves, there is only one direction to consider. For holomorphic maps of higher rank (e.g., $f : \mathbb {C}^2 \to X$ ), it is not clear whether the holomorphic sectional curvature of an arbitrary Hermitian metric gives any control. This is the reason for the introduction of the real bisectional curvature [Reference Lee and Streets20, Reference Yang and Zheng44] and the Schwarz bisectional curvatures [Reference Broder2, Reference Broder3].

At this point, the reader may be perplexed, since we used the Schwarz lemma in the proof of the main theorem and this involves a map of rank $>1$ . But here, we note that Royden’s polarization argument [Reference Broder3, Reference Broder4, Reference Royden25] is used. In particular, for holomorphic maps into Kähler manifolds, the holomorphic sectional curvature gives suitable control. In other words, we see the Schwarz lemma as an incarnation of aspects (i) and (ii).

The third aspect is far less understood (emphasized by the vague nature of the statement in (iii)). The main theorem of the present manuscript provides evidence for the relationship, but it is certainly not the only evidence. Recall that Demailly [Reference Demailly7] constructed an example of a projective Kobayashi hyperbolic surface that does not admit a Hermitian metric of negative Chern holomorphic sectional curvature. The construction relies on the following algebraic hyperbolicity criterion [Reference Demailly7].

Theorem 3.3

(Demailly). Let $(X, \omega )$ be a compact Hermitian manifold with ${}^c \text {HSC}_{\omega } \leq \kappa _0$ for some $\kappa _0 \in \mathbb {R}$ . If $f : \mathcal {C} \to X$ is a nonconstant holomorphic map from a compact Riemann surface $\mathcal {C}$ of genus g, then

$$ \begin{align*} 2g-2 & \geq -\frac{\kappa_0}{2\pi} \deg_{\omega}(\mathcal{C}) - \sum_{p \in \mathcal{C}} (m_p-1). \end{align*} $$

We note that, in [Reference Demailly7] (see also [Reference Diverio8]), the constant is assumed to be nonpositive, but the formula holds more generally (see, e.g., [Reference Broder4]). The example is constructed as a fibration (with hyperbolic base and fiber) with a fiber sufficiently singular to violate the above algebraic hyperbolicity criterion.

Acknowledgment

I am grateful to my Ph.D. advisors Ben Andrews and Gang Tian for their support and encouragement. I am tremendously grateful to Tian for posing this problem to me, and to Valentino Tosatti for his support, interest, and encouragement.

Footnotes

The author was partially supported by an Australian Government Research Training Program (RTP) Scholarship and funding from the Australian Government through the Australian Research Council’s Discovery Projects funding scheme (project DP220102530).

1 This, of course, implies that the dimension of $H^{1,1}(X, \mathbb {R})$ is at least 2. Otherwise, the only class on the boundary is the zero class, in which case everything is trivial.

References

Broder, K., Twisted Kähler–Einstein metrics and collapsing. Preprint, 2021. arXiv:2003.14009 Google Scholar
Broder, K., The Schwarz lemma in Kähler and non-Kähler geometry. Preprint, 2022. arXiv:2109.06331, to appear in Asian J. Math.CrossRefGoogle Scholar
Broder, K., The Schwarz lemma: an Odyssey. Rocky Mountain J. Math. 52(2022), no. 4, 11411155.Google Scholar
Broder, K., Complex manifolds of hyperbolic and non-hyperbolic-type. Ph.D. thesis, Australian National University, Beijing International Center for Mathematics Research, 2022.Google Scholar
Cao, H. D., Deformations of Kähler metrics to Kähler–Einstein metrics on compact Kähler manifolds . Invent. Math. 81(1985), no. 2, 359372.CrossRefGoogle Scholar
Demailly, J.-P., Regularization of closed positive currents and intersection theory . J. Algebraic Geom. 1(1992), no. 3, 361409.Google Scholar
Demailly, J.-P., Algebraic criteria for Kobayashi hyperbolic projective varieties and jet differentials . In: Algebraic geometry—Santa Cruz 1995, Proceedings of Symposia in Pure Mathematics, 62, American Mathematical Society, Providence, RI, 1997, pp. 285360.CrossRefGoogle Scholar
Diverio, S., Kobayashi hyperbolicity, negativity of the curvature and positivity of the canonical bundle. Preprint, 2022. arXiv:2011.11379, to appear in a forthcoming monograph of the SMF in the series “Panoramas et Synthèses.”Google Scholar
Eyssidieux, P., Guedj, V., and Zeriahi, A., Singular Kähler–Einstein metrics . J. Amer. Math. Soc. 22(2009), 607639.CrossRefGoogle Scholar
Fong, F. T.-Z. and Lee, M.-C., Higher-order estimates of long-time solutions to the Kähler–Ricci flow. Preprint, 2020. arXiv:2001.11555 CrossRefGoogle Scholar
Grauert, H. and Remmert, R., Plurisubharmonische Funktionen in komplexen Räumen . Math. Z. 65(1956), 175194.CrossRefGoogle Scholar
Gross, M., Tosatti, V., and Zhang, Y., Collapsing of abelian fibered Calabi–Yau manifolds . Duke Math. J. 162(2013), no. 3, 517551.CrossRefGoogle Scholar
Gross, M., Tosatti, V., and Zhang, Y., Gromov–Hausdorff collapsing of Calabi–Yau manifolds . Comm. Anal. Geom. 24(2016), no. 1, 93113.CrossRefGoogle Scholar
Gross, M., Tosatti, V., and Zhang, Y., Geometry of twisted Kähler–Einstein metrics and collapsing . Comm. Anal. Geom. 380(2020), 14011438.Google Scholar
Gross, M. and Wilson, P. M. H., Large complex structure limits of K3 surfaces . J. Differential Geom. 55(2000), no. 3, 475546.CrossRefGoogle Scholar
Guenancia, H. and Paun, M., Conic singularities metrics with prescribed Ricci curvature: general cone angles along simple normal crossing divisors . J. Differential Geom. 103(2016), no. 1, 1557.CrossRefGoogle Scholar
Hein, H. J. and Tosatti, V., Remarks on the collapsing of torus fibered Calabi–Yau manifolds . Bull. Lond. Math. Soc. 47(2015), no. 6, 10211027.Google Scholar
Hein, H. J. and Tosatti, V., Higher-order estimates for collapsing Calabi–Yau metrics. Preprint, 2020. arXiv:1803.06697 CrossRefGoogle Scholar
Jeffres, T., Mazzeo, R., and Rubinstein, Y. A., Kähler–Einstein metrics with edge singularities . Ann. of Math. (2) 183(2016), no. 1, 95176.CrossRefGoogle Scholar
Lee, M. C. and Streets, J., Complex manifolds with negative curvature operator . Int. Math. Res. Not. IMRN 2021(2021), 1852018528.CrossRefGoogle Scholar
Li, Y. and Tosatti, V., On the collapsing of Calabi–Yau manifolds and Kähler–Ricci flows. Preprint, 2021. arXiv:2107.00836 Google Scholar
Lin, A. and Shen, L., Conic Kähler–Einstein metrics along normal crossing divisors on Fano manifolds . J. Funct. Anal. 275(2018), no. 2, pp. 300328.CrossRefGoogle Scholar
Lu, Y.-C., Holomorphic mappings of complex manifolds . J. Differential Geom. 2(1968), 299312.CrossRefGoogle Scholar
Moishezon, B. G., Singular Kählerian spaces . In: Manifolds—Tokyo 1973 (proceedings of the international conference on manifolds and related topics in topology, Tokyo, 1973), University of Tokyo Press, Tokyo, 1975, pp. 343351.Google Scholar
Royden, H. L., Ahlfors–Schwarz lemma in several complex variables . Comment. Math. Helv. 55(1980), no. 4, 547558.CrossRefGoogle Scholar
Rubinstein, Y., Smooth and singular Kähler–Einstein metrics . In: Geometric and spectral analysis, Contemporary Mathematics: Centre de Recherches Mathématiques Proceedings, 630, American Mathematical Society, Providence, RI, 2014, pp. 45138.Google Scholar
Song, J., Riemannian geometry of Kähler–Einstein currents. Preprint, 2014. arXiv:1404.0445 Google Scholar
Song, J. and Tian, G., The Kähler–Ricci flow on surfaces of positive Kodaira dimension . Invent. Math. 170 (2007), no. 3, 609653.CrossRefGoogle Scholar
Song, J. and Tian, G., Canonical measures and Kähler–Ricci flow . J. Amer. Math. Soc. 25 (2012), no. 2, 303353.CrossRefGoogle Scholar
Song, J., Tian, G., and Zhang, Z., Collapsing behavior of Ricci-flat Kähler metrics and long time solutions of the Kähler–Ricci flow. Preprint, 2019. arXiv:1904.08345 Google Scholar
Tian, G., Smoothness of the universal deformation space of compact Calabi–Yau manifolds and its Petersson–Weil metric . In: Mathematical aspects of string theory (San Diego, CA, 1986), Advanced Series in Mathematical Physics, vol. 1, World Science Publishing, Singapore, 1987, pp. 629646.CrossRefGoogle Scholar
Tian, G. and Zhang, Z., On the Kähler–Ricci flow on projective manifolds of general type . Chinese Ann. Math. Ser. B 27(2006), no. 2, 179192.CrossRefGoogle Scholar
Tian, G. and Zhang, Z., Relative volume comparison of Ricci flow and its applications. Preprint, 2018. arXiv:1802.09506 Google Scholar
Tosatti, V., Limits of Calabi–Yau metrics when the Kähler class degenerates . J. Eur. Math. Soc. (JEMS) 11(2009), no. 4, 755776.CrossRefGoogle Scholar
Tosatti, V., Adiabatic limits of Ricci-flat Kähler metrics . J. Differential Geom. 84(2010), no. 2, 427453.CrossRefGoogle Scholar
Tosatti, V., Degenerations of Calabi–Yau metrics . Acta Phys. Pol. B Proc. Suppl. 4(2011), 495.CrossRefGoogle Scholar
Tosatti, V., Calabi–Yau manifolds and their degenerations . Ann. N. Y. Acad. Sci. 1260(2012), 813.CrossRefGoogle Scholar
Tosatti, V., Weinkove, B., and Yang, X., The Kähler–Ricci flow, Ricci-flat metrics and collapsing limits . Amer. J. Math. 140 (2018), no. 3, 653698.CrossRefGoogle Scholar
Tosatti, V. and Zhang, Y., Triviality of fibered Calabi–Yau manifolds without singular fibers . Math. Res. Lett. 21(2014), no. 4, 905918.CrossRefGoogle Scholar
Tosatti, V. and Zhang, Y., Infinite time singularities of the Kähler–Ricci flow . Geom. Topol. 19(2015), no. 5, 29252948.CrossRefGoogle Scholar
Tosatti, V. and Zhang, Y., Collapsing hyperkähler manifolds . Ann. Sci. Éc. Norm. Supér. 53(2020), no. 3, 751786.CrossRefGoogle Scholar
Tsuji, H., Existence and degeneration of Kähler–Einstein metrics on minimal algebraic varieties of general type . Math. Ann. 281(1988), no. 1, 123133.CrossRefGoogle Scholar
Yang, X., RC-positivity, rational connectedness and Yau’s conjecture . Camb. J. Math. 6(2018), no. 2, 183212.CrossRefGoogle Scholar
Yang, X. and Zheng, F., On real bisectional curvature for Hermitian manifolds . Trans. Amer. Math. Soc. 371(2019), no. 4, 27032718.CrossRefGoogle Scholar