Home On polynomial convergence to tangent cones for singular Kähler–Einstein metrics
Article Open Access

On polynomial convergence to tangent cones for singular Kähler–Einstein metrics

  • Junsheng Zhang EMAIL logo
Published/Copyright: July 11, 2025

Abstract

Let ( Z , p ) be a pointed Gromov–Hausdorff limit of non-collapsing Kähler–Einstein metrics with uniformly bounded Ricci curvature. We show that the singular Kähler–Einstein metric on 𝑍 is conical at 𝑝 if and only if C = W in Donaldson–Sun’s two-step degeneration theory, assuming curvature grows at most quadratically near 𝑝. Let ( X , p ) be a germ of an isolated log terminal algebraic singularity. Following Hein–Sun’s approach, we show that if C = W in the two-step stable degeneration of ( X , p ) and 𝒞 has a smooth link, then every singular Kähler–Einstein metric on 𝑋 with non-positive Ricci curvature and bounded potential is conical at 𝑝.

1 Introduction

In this paper, we study the geometry of singular Kähler–Einstein metrics. Such metrics arise from two main sources: Gromov–Hausdorff limits of smooth Kähler–Einstein metrics and pluripotential theory. Similar to other problems in geometric analysis, a crucial approach to understanding the structure of a singular Kähler–Einstein metric near its singularity involves examining its tangent cones. This process typically involves three steps:

  1. the existence of tangent cones;

  2. the uniqueness of tangent cones;

  3. relating the geometry of 𝑍 near 𝑝 with the geometry on the tangent cone at 𝑝.

In this paper, we make progress on step (3) under certain assumptions that guarantee the existence and uniqueness of tangent cones.

Firstly, we consider singular Kähler–Einstein metrics coming from the Gromov–Hausdorff limits. Let ( Z , p , d Z ) be a pointed Gromov–Hausdorff limit of complete non-collapsing Kähler–Einstein manifolds with uniformly bounded Ricci curvature. According to [9, 1, 10, 18], there is a decomposition Z = Z reg Z sing , where the regular part Z reg consists of points around which a neighborhood is a smooth Riemannian manifold. This Z reg is an open, connected, smooth manifold, and the restriction of d Z to Z reg is induced by a Kähler–Einstein metric ( g Z , J Z , ω Z ) , which is referred to as the singular Kähler–Einstein metric on 𝑍.

If one further assumes that the smooth Kähler–Einstein metrics are polarized, i.e., the Kähler forms are curvature forms of line bundles, and that the canonical bundle is homomorphically trivial for Ricci-flat metrics, then Donaldson and Sun [23, 24] showed that 𝑍 has the structure of a normal analytic space with log terminal singularities and the tangent cone 𝒞 at 𝑝 is unique. Furthermore, a two-step degeneration exists, proceeding from the local ring O Z , p through an intermediate K-semistable cone𝑊 to 𝒞. This is referred to as Donaldson–Sun’s two-step degeneration theory. It is an open problem whether Donaldson–Sun’s two-step degeneration theory holds without the polarization condition.

In this paper, we assume 𝑝 is an isolated singularity and that, near 𝑝, the curvature grows at most quadratically. That is, there exists a constant C > 0 such that the following holds near 𝑝:

(1.1) | Rm ( ) | C d Z ( p , ) 2 ;

however, the polarization condition is not required. The main result is as follows.

Theorem 1.1

Suppose that (1.1) holds. Then Donaldson–Sun’s two-step degeneration theory applies. Moreover, the singular Kähler–Einstein metric is conical at 𝑝 if and only if 𝒞 is isomorphic to 𝑊 as Fano cones in Donaldson–Sun’s two-step degeneration theory.

We refer to Section 2.3 for the definition of Fano cones and to Section 3 for further details on Donaldson–Sun’s two-step degeneration theory. We will simply use C = W to denote that 𝑊 is isomorphic to 𝒞 as Fano cones in the following. We remark that, assuming (1.1), Donaldson–Sun’s two-step degeneration theory in this context (without the polarized condition) can be directly achieved using results in [47] and following the original argument in [24]. The essence of the result lies in establishing the algebraic criterion for the polynomial convergence rate. We also remark that, as a consequence of the results in [16, 17], condition (1.1) is equivalent to stating that a tangent cone at 𝑝 has a smooth link, which is an a priori weaker condition.

In this paper, we use the terms “being conical” and “polynomially close to a Calabi–Yau cone” interchangeably, which means the following definition holds for some δ > 0 .

Definition 1.2

Let ( Z , J Z , p ) be a germ of normal isolated singularity. A Kähler metric ( g Z , J Z , ω Z ) is said to be conical of order δ > 0 at 𝑝 if there exist a Calabi–Yau cone with smooth link ( C , o , g C , ω C , J C ) and a diffeomorphism Φ : ( U o , o ) ( U p , p ) between neighborhoods U o and U p of 𝑜 and 𝑝 respectively such that, as r 0 , for all k 0 ,

(1.2) r k ( | g C k ( Φ g Z g C ) | g C + | g C k ( Φ J Z J C ) | + | g C k ( Φ ω Z ω C ) | g C ) = O ( r δ ) .

It is conjectured in [24] that both 𝑊 and 𝒞 depend only on the germ of the singularity ( Z , p ) , rather than the metric. For algebraic singularities, this conjecture has been confirmed and generalized by Li–Xu and Li–Wang–Xu; see [46, 45] and the references therein. By this result in algebraic geometry and considering cones over strictly K-semistable Fano manifolds, we obtain examples ( Z , p ) that are not conical.

Corollary 1.3

There are examples of ( Z , p ) , which are pointed Gromov–Hausdorff limits of non-collapsing Calabi–Yau metrics. While the tangent cone 𝒞 at 𝑝 has a smooth link and is unique, 𝑍 is not conical at 𝑝.

This highlights the sharpness of the logarithmic rate result proved by Colding–Minicozzi [17] within the context of Kähler geometry, as initially predicted by Hein–Sun [39]. We remark that, in [17], the logarithmic convergence rate is also established for tangent cones at infinity; however, it is proved in [53] that every complete Calabi–Yau metric with Euclidean volume growth and quadratic curvature decay is polynomially close to its tangent cone at infinity. This reveals distinctions between the two settings.

Then let us compare Theorem 1.1 with some results in the literature. In [39], assuming the germ ( Z , p ) is biholomorphic to ( C , o ) and 𝒞 has a smooth link, Hein–Sun proved that the ( Z , p ) is polynomially close to ( C , o ) , and indeed, in (1.2), Φ can be chosen to be a biholomorphism. In [15], assuming ( Z , p ) is a polarized limit and the germ ( Z , p ) is biholomorphic to ( C , o ) , Chiu–Székelyhidi proved a polynomial rate closeness on the level of Kähler potentials without assuming any smoothness on 𝒞. Although we assume 𝒞 has a smooth link, we are able to address the case where the germ ( Z , p ) is not biholomorphic to ( C , o ) and provide a characterization for the polynomial closeness. The direction from the polynomial closeness to C = W is established directly by constructing holomorphic functions using the Hörmander L 2 -method. The direction from C = W to polynomial closeness uses an approach similar to that in [53] and relies on the results in [39, 15]. See Section 5 for more details.

Then we move to singular Kähler–Einstein metrics coming from pluripotential theory. Singular Kähler–Einstein metrics with non-positive Ricci curvature on log terminal Kähler varieties have been constructed in [28]; see [5, 6, 43] and references therein for further studies. Understanding the metric behavior of the singular Kähler–Einstein metrics near the singular locus is a major open problem. The main progress in this direction is due to Hein–Sun [39]. Following the approach in [39], we extend their results to a broader class of singular Kähler–Einstein metrics.

Let ( X , p ) be a germ of an isolated log terminal algebraic singularity. Motivated by the results in [24], there is a well-developed theory for local stability of (Kawamata) log terminal singularities; see [44, 46, 45, 58] and the references therein. Specifically, this local stability theory establishes the existence of a K-semistable Fano cone 𝑊 and a K-polystable Fano cone 𝒞, canonically associated with the singularity. Then we have the following.

Theorem 1.4

Suppose 𝒞 is isomorphic to 𝑊 as Fano cones and 𝒞 has a smooth link; then every singular Kähler–Einstein metric with non-positive Ricci curvature and with bounded potential on the germ ( X , p ) is conical at 𝑝.

We remark that, as in Theorem 1.1, the isomorphism between 𝒞 and 𝑊 is also a necessary condition for the existence of a conical singular Kähler–Einstein metric. If the germ ( X , p ) itself is biholomorphic to a Calabi–Yau cone, then the assumption in the theorem is satisfied, and we can show that Φ can be chosen to be a biholomorphism. This makes this result a generalization of [39]. The proof follows the same idea as in [39], with some new key insights from [14, 45]. Chen–Wang’s compactness and regularity results in [14], generalizing Cheeger–Colding’s results, ensure that we can take limits for Kähler–Einstein metrics with mild singularities. Li–Wang–Xu’s result in [45] ensures that we can determine the tangent cone for the limiting metric from the underlying algebraic structure. See Section 7 for more details. Unfortunately, we are unable to prove similar results for singular Kähler–Einstein metrics with positive Ricci curvature. We will briefly discuss the related issues in Section 7.3.

This paper is organized as follows. In Section 2, we collect some definitions and preliminary results necessary for this paper. In Section 3, we give a review for Donaldson–Sun’s two-step degeneration theory. In Section 4, the direction from the polynomial convergence to C = W is proved. In Section 5, the other direction from C = W to polynomial convergence is established. In Section 6, we construct examples claimed in Corollary 1.3. In Section 7, we give the proof of Theorem 1.4. In the appendices, we give proofs of two results needed in this paper, which are also well-studied in the literature.

2 Preliminary

2.1 Weighted analysis near the vertex for a cone metric

Definition 2.1

( C , o , g C ) is said to be a (Riemannian) cone with a smooth link if C { o } = R > 0 × L for some compact manifold 𝐿 and g C = d r 2 + r 2 g L with 𝑟 the projection onto the first factor and g L is a Riemannian metric on 𝐿. The point 𝑜 is referred to as the vertex of the cone.

A cone 𝒞 admits a complete metric d C , which is the metric completion of the one induced by the Riemannian metric g C on C { o } . In the following, the symbol 𝑟 is used to represent the distance function to the vertex of the cone. For every s > 0 , let B s denote the ball centered at 𝑜 of radius 𝑠. For any k Z 0 , β R and α ( 0 , 1 ) , we define the Banach space C β k , α ( B s ) using the following weighted norms on a function 𝑓 defined on B s C :

(2.1) f C β k , α ( B s ) = j = 0 k sup B s | r β + j g C j f | + sup x B s ( r ( x ) β + k + α [ g C k f ] C 0 , α ( B s B ( x , 1 2 r ( x ) ) ) ) ,

where [ ] C 0 , α denotes the ordinary C 0 , α seminorm.

In the following, we will denote by Ψ ( ϵ ; ) any non-negative functions depending on 𝜖 and some additional parameters such that, when these parameters are fixed,

lim ϵ 0 Ψ ( ϵ ; ) = 0 .

Now fix s 0 ( 0 , 1 ) . Suppose there is another Riemannian metric 𝑔 on B s 0 such that, for all l 0 and s s 0 ,

(2.2) s l sup B s | g C l ( g g C ) | = Ψ ( s ; l ) .

Fix k 2 n + 1 and α ( 0 , 1 ) . Then (2.2) ensures that, for 𝑠 sufficiently small,

Δ g : C 2 + β k + 2 . α ( B s ) C β k , α ( B s )

is a bounded linear operator. Then the following result ensures that Δ g admits a right inverse, whose norm is uniformly bounded independently of 𝑠.

Let 0 = λ 0 < λ 1 λ 2 denote the eigenvalues of the Laplacian on the link 𝐿 of 𝒞 (listed with multiplicity), and

(2.3) μ i ± = m 2 2 ± ( m 2 ) 2 4 + λ i .

Let Γ = { μ i ± : i = 1 , } . Then we have

Proposition 2.2

For any given β ( 0 , 1 ) Γ , there exist s 1 s 0 and C > 0 such that, for all s s 1 , there is a bounded linear map T s : C β k , α ( B s ) C β + 2 k + 2 , α ( B s ) such that Δ g T s = Id and T s C .

Proof

This is a standard result, and for readers’ convenience, we give a sketch of the proof. One first shows that Δ g C admits right inverse with norm bounded independently of 𝑠. For any s ( 0 , 1 ] , there is a linear extension map E s : C β k , α ( B s ) C β k , α ( B 2 ) whose norm is bounded independently of 𝑠. The existence of E 1 follows from the local construction in [51], and then one can define E s for s ( 0 , 1 ) using scaling. Then, using separation of variables [39, Proposition 2.9], for f C β k , α ( B 2 ) , one can solve Δ g C u = f with 𝑢 depending linearly on 𝑓 and

u C 2 + β k , α ( B 1 ) C f C β k , α ( B 2 ) .

Then the existence of a uniformly bounded right inverse of Δ g C follows from this.

For a general Riemannian metric satisfying (2.2), one can write

Δ g = Δ g C + ( g g C ) g C 2 + g C g g C .

Then, acting on the function defined on B r , we have

Δ g Δ g C C sup s r ,  0 l k + 2 Ψ ( s ; l ) .

The right-hand side is small when r 1 . Then the conclusion follows from standard functional analysis. ∎

The following result ensures that we can get a pointwise bound of a function from the integral bound and the proof follows from standard elliptic regularity and a rescaling argument.

Proposition 2.3

Let 𝑔 be a Riemannian metric on B s 0 satisfying (2.2). Suppose 𝑢 is a function defined on B s 0 and, for some τ , τ R and for all l 0 ,

| g l ( Δ g u ) | g C l r τ l 2 and B s 0 u 2 r 2 τ 2 n d Vol g 1 .

Then, for γ = min ( τ , τ ) and any k > 0 , there exist constants A k > 0 such that

| g k u | g A k r γ k on B s 0 .

2.2 Harmonic functions on Calabi–Yau cones

A cone 𝒞 has a natural scaling vector field r r . A tensor 𝑇 on 𝒞 is called 𝜇-homogeneous for some μ R if L r r T = μ T . We remark that this condition implies that | j T | r μ + p q j with respect to the cone metric for all j N 0 if 𝑇 is 𝑝-fold covariant and 𝑞-fold contravariant. In general, if a tensor 𝑇 satisfies | T | r λ as r 0 , then we will say 𝑇 has growth rate of 𝜆. A Kähler cone 𝒞 is a cone with a parallel complex structure J C . For a Kähler cone 𝒞, the Reeb vector field ξ C = J C ( r r ) is a holomorphic Killing vector field.

Definition 2.4

A Calabi–Yau cone with smooth link is a tuple ( C , o , g C , J C , ω C ) such that

  • ( C , o , g C ) is a Riemannian cone,

  • ( g C , J C , ω C ) defines a Kähler Ricci-flat metric on C { o } .

The following lemma, which is proved in [39, Theorem 2.14], building on the early work [11], plays an important role in this paper.

Theorem 2.5

Theorem 2.5 ([39])

Let 𝒞 be a Calabi–Yau cone with a smooth link.

  1. If 𝑢 is a real-valued 𝜇-homogeneous harmonic function with μ > 0 , then μ 1 . If 1 < μ < 2 , then 𝑢 is pluriharmonic. If μ = 2 , then u = u 1 + u 2 , where u 1 and u 2 are 2-homogeneous, u 1 is pluriharmonic and u 2 is 𝜉-invariant.

  2. The space of all holomorphic vector fields that commute with r r can be written as p J C p , where 𝔭 is spanned by r r and by the gradient fields of the 𝜉-invariant 2-homogeneous harmonic functions. All elements of J C p are Killing fields.

2.3 Polarized affine cones and Fano cones

In this section, we fix some basic definitions following [19, 46]. Let 𝑋 be a normal affine variety with coordinate ring R ( X ) , and let 𝜉 be a non-zero (real) holomorphic vector field on 𝑋 which generates a compact torus of automorphisms 𝕋 on 𝑋 with a unique a fixed point. Then we have a weight decomposition

R ( X ) = χ Γ R χ ( X ) .

Here Γ Lie ( T ) is the weight lattice of the 𝕋-action, that is, R χ ( X ) { 0 } and f R χ ( X ) if and only if L V η f = 1 χ , η f for every η Lie ( T ) . Here we use V η to denote the holomorphic vector field induced by 𝜂 and , to denote the natural pairing between Lie ( T ) and its dual.

Definition 2.6

A polarized affine cone is a pair ( X , ξ ) such that χ , ξ > 0 for all non-zero χ Γ . Such a 𝜉 is called a Reeb field or polarization.

Definition 2.7

A polarized affine cone ( X , ξ ) is said to be a Fano cone if it has log terminal singularities.

Remark 2.8

We use the terminology polarized affine cone instead of polarized affine variety as in [19] to emphasize the grading structure on the coordinate ring R ( X ) . It is shown in [19] that if 𝑋 has only an isolated singularity, then a Fano cone is equivalent to a polarized affine cone ( X , ξ ) with a trivializing section Ω of m K X for some m > 0 such that

L ξ Ω = 1 λ Ω

for some λ > 0 .

A Calabi–Yau cone with smooth link naturally admits a Fano cone structure, as we will briefly review here. For a detailed proof, please refer to [57, 24]. The Reeb vector field ξ = J C ( r r ) is a holomorphic Killing vector field on 𝒞, which generates a compact torus 𝕋-action on 𝒞. We have the following:

  1. Polynomial growth holomorphic functions form a finitely generated ring R ( C ) which defines a normal affine variety structure on 𝒞. Moreover, there is a positive grading induced by the action of 𝕋, on the coordinate ring

    R ( C ) = μ S R μ ( C ) ,

    where S = R 0 { α , ξ : α Γ } and Γ Lie ( T ) is the weight lattice of the 𝕋-action, with 𝒮 a discrete set, which is called the holomorphic spectrum of 𝒞. In the decomposition, f R μ ( C ) if and only if 𝑓 is 𝜇-homogeneous, i.e., L ξ f = 1 μ f . For f R ( C ) , its degree deg ( f ) is defined to be the smallest number 𝑑 such that f μ d R μ ( C ) and it is direct to show that this is the same as the growth order of 𝑓 at infinity.

  2. Since 𝒞 has a smooth link, there exists a 𝕋-invariant, nowhere vanishing holomorphic section Ω of m K C for some positive integer 𝑚 such that

    (2.4) ω C n = ( c n Ω Ω ̄ ) 1 m .

    For this, note that the metric on the link 𝐿 of 𝒞 has positive Ricci curvature and hence C { o } has finite fundamental group and therefore the flat bundle m K C admits a parallel section for some m 1 as the line bundle K C admits a flat connection induced by the Calabi–Yau metric. Then we can adjust the constant to guarantee (2.4) holds. Since ω C n has finite volume in a neighborhood of 𝑜, (2.4) implies that 𝒞 has log terminal singularity at 𝑜 (see [28, Lemma 6.4]).

Remark 2.9

Let ( Z , p , d Z ) be a pointed Gromov–Hausdorff limit of a sequence of complete non-collapsing Kähler metrics with Ricci curvature uniformly bounded below, and let ( C , o ) be a tangent cone of 𝑍 at 𝑝. In this general setting (without assuming the polarized condition and only assuming Ricci curvature lower bound), it is proved in [48] that 𝒞, which typically has non-isolated singularities, always admits a polarized affine cone structure.

2.4 K-stability for Fano cones

General definitions and systematic studies for (log) Fano cone singularities are available, as discussed in works such as [19, 42, 46, 45, 58] and references therein. To streamline our discussion and avoid excessive terminology, we adhere to the following definitions, which coincide with those established in the existing literature.

Definition 2.10

A Fano cone ( C , ξ ) is said to be K-polystable if it admits a Calabi–Yau cone structure.

Here, in general, a Calabi–Yau cone structure means that there exists a smooth radius function 𝑟, i.e., L J ξ r = r over C reg that is locally bounded on 𝒞, and on C reg , the Kähler form ω = 1 ̄ r 2 satisfies Ric ( ω ) = 0 and L J ξ ω = 2 ω . When 𝒞 has only an isolated singularity, this coincides with the one given in Definition 2.4.

Definition 2.11

A Fano cone ( X , ξ X ) is said to be K-semistable if it admits an equivariant degeneration to a K-polystable Fano cone ( C , ξ ) .

Let 𝐷 be a Fano manifold and consider the affine variety

K D × : = Spec ( m 0 H 0 ( D , m K D ) ) ,

which is obtained by blowing down the zero section of K D geometrically. Then K D × admits a standard Reeb vector field 𝜉 induced from the fiberwise rotation on K D . It is direct to show that ( K D × , ξ ) is K-polystable if and only if 𝐷 admits a Kähler–Einstein metric.

2.5 Hörmander L 2 -method

The following version of the Hörmander L 2 -estimate is used and we refer to [22, Chapter VIII] for a proof.

Theorem 2.12

Let 𝑀 be an 𝑛-dimensional complex manifold, which admits a complete Kähler metric. Let ω M be any Kähler metric on 𝑀 and let φ M be a smooth function with 1 ̄ φ M + Ric ( ω M ) Υ ω M for a continuous non-negative function Υ. Let 𝑞 be a positive integer. Then, for any ( 0 , q ) -form 𝜂 on 𝑀 with ̄ η = 0 and M Υ 1 | η | ω M 2 e φ M ω M n < , there exists a ( 0 , q 1 ) -form 𝜁 on 𝑀 satisfying ̄ ζ = η and with estimate

M | ζ | ω M 2 e φ M ω M n M q 1 Υ 1 | η | ω M 2 e φ M ω M n .

3 Donaldson–Sun two-step degeneration theory

Let ( Z , p , d Z ) be a pointed Gromov–Hausdorff limit of complete non-collapsing Kähler–Einstein manifolds with uniformly bounded Ricci curvature. This means that there is a sequence of complete Kähler manifolds ( X i , h i ) and points q i X i satisfying Ric ( h i ) = λ i h i with | λ i | 1 and Vol ( B ( q i , 1 ) ) κ for some constant 𝜅 independent of 𝑖 such that ( X i , q i , h i ) converges to ( Z , p , d Z ) in the pointed Gromov–Hausdorff topology. Indeed, since we are interested in the local geometry of 𝑍 near 𝑝, completeness of ( X i , h i ) is not important as long as B ( q i , 1 ) is compactly contained in X i .

Note that we do not assume the Kähler–Einstein metrics are polarized; therefore, Donaldson–Sun’s theory cannot be directly applied here. We assume the following curvature condition holds near 𝑝:

(3.1) | Rm ( ) | C d Z ( p , ) 2 .

Liu’s result [47, Proposition 2.5] is the crucial step in establishing Donaldson–Sun’s two-step degeneration theory in our context. The curvature condition (3.1) is needed there to obtain pointwise bound from integral bound for functions. It is worth noting that, although Liu’s work primarily concerns tangent cones at infinity, the proof can be directly applied here without much modification.

Let ( Z j , p j , J j , g j , ω j , d j ) denote the rescaling of ( Z , p , J Z , g Z , ω Z , d Z ) by a factor 2 j , j Z 1 . This means that we take J j = J , p j = p , g j = 2 2 j g Z , ω j = 2 2 j ω Z and d j = 2 j d Z . Let r j be the distance to p j with respect to d j . Recall that we denote by Ψ ( ϵ ; ) any non-negative functions depending on 𝜖 and some additional parameters such that, when these parameters are fixed, lim ϵ 0 Ψ ( ϵ ; ) = 0 .

Proposition 3.1

Proposition 3.1 ([47])

Given any r 0 > 0 , for sufficiently large 𝑗, we can find a plurisubharmonic function 𝑢 on B ( p j , r 0 ) with

| u r j 2 | Ψ ( j 1 ) r 0 2 , | u | 2 4 r j 2 Ψ ( j 1 ) r 0 2 , 1 ̄ u ( 1 Ψ ( j 1 ) ) ω j on B ( p j , r 0 ) B ( p j , Ψ ( j 1 ) r 0 ) .

Proposition 3.1 implies that there are (arbitrary small) neighborhoods of 𝑝 in 𝑍, which have strictly pseudoconvex boundary. In the regular part of 𝑍, we have smooth convergence of Kähler–Einstein metrics ( X i , q i , h i ) ; therefore, there are neighborhoods of q i with strictly pseudoconvex boundary, converging in the Gromov–Hausdorff sense. Then we can use the Hörmander L 2 -estimate to construct holomorphic functions in a neighborhood of q i ; therefore, repeating the argument in [23, 24], one can show that 𝑍 has a normal complex analytic space structure.

Let ( C , o , J C , g C , ω C , d C ) denote a tangent cone of ( Z , p , g Z , J Z , ω Z , d Z ) at 𝑝. As a consequence of (3.1), 𝒞 is a Calabi–Yau cone with a smooth link. Since 𝑍 admits a normal analytic space structure with an isolated singularity 𝑝, we can also do the Hörmander L 2 -method in a neighborhood of p Z . In particular, we have the following [23].

Proposition 3.2

Let λ > 0 , and 𝑓 be a holomorphic function defined on B ( o , λ ) C . Then, for 𝑗 sufficiently large, there exist holomorphic functions f j defined on B ( p j , λ / 2 ) such that f j converges uniformly to 𝑓.

Donaldson–Sun’s two-step degeneration theory can be established following their original argument [24]. In particular, the tangent cone 𝒞 is unique. This theory describes the Gromov–Hausdorff convergence of Z j to 𝒞 using complex analytic data and we summarize their results as follows.

Let B j denote the unit ball in Z j centered at p j , and 𝐵 the unit ball in 𝒞 centered at 𝑜. Fix a distance d j on B j B that realizes the Gromov–Hausdorff convergence of B j to 𝐵. More precisely, this means that the Hausdorff distance between B j and 𝐵 under d j is Ψ ( j 1 ) and d j ( p j , p ) = Ψ ( j 1 ) . Moreover, for any compact set 𝐾 contained in the regular part of 𝐵, we can find, for large enough 𝑗, open embeddings χ j of an open neighborhood of 𝐾 into B j such that d j ( x , χ j ( x ) ) = Ψ ( j 1 ) for all x K and ( χ j g j , χ j J j ) converges smoothly over 𝐾 to ( g C , J C ) .

It is proved in [24] that there exist holomorphic embeddings

F : ( C , o ) ( C N , 0 ) and F j : ( B j , p j ) ( C N , 0 )

such that

  1. F = ( h 1 , , h N ) , where each h i is a d i -homogeneous holomorphic function with d i > 0 . Under the embedding F , ξ = J C ( r r ) extends to a linear vector field on C N of the form Re ( 1 i d i z i z i ) , which we also denote by 𝜉. In particular, the dilation action Λ : r 2 r on 𝒞 extends to the diagonal linear transformation of C N given by

    Λ ( z 1 , , z N ) = ( 2 d 1 z 1 , , 2 d N z N ) .

  2. F j = Λ j F j 1 | B j = Λ j Λ 2 F 1 | B j , where Λ j G ξ . Here G ξ denotes the subgroup of GL ( N ; C ) consisting of elements that commute with the actions generated by 𝜉. Moreover, Λ j Λ as j .

  3. d j ( x j , x ) 0 if and only if F j ( x j ) F ( x ) C N .

Moreover, there is an intrinsic intermediate K-semistable cone𝑊 associated to ( Z , p ) that can be characterized as follows. Let O p be the ring of germs of holomorphic functions on 𝑍 at 𝑝. For any non-zero function f O p , one defines its order of vanishing

d KE ( f ) : = lim r 0 log sup B ( p , r ) | f | log r .

Such a limit exists and is in 𝒮, the holomorphic spectrum of 𝒞. We list them in order as 0 = d 0 < d 1 < . Then, for any k 0 , one can define an ideal

I k = { f O x : d KE ( f ) d k } .

We obtain a filtration O p = I 0 I 1 and an associated graded ring k 0 I k / I k + 1 . Then the intermediate K-semistable cone is defined to be

W : = Spec ( k 0 I k / I k + 1 ) ,

which is a normal affine variety and also admits a natural 𝕋-action, which defines a polarized affine cone structure on 𝑊.

The relation between 𝑊 and 𝒞 can be described as follows. Let W j denote the weighted tangent cone of F j ( B j ) with respect to the weight ( d 1 , , d N ) . It is proved in [24] that each W j is isomorphic to 𝑊 as polarized affine cones. As Λ j G ξ , commuting with 𝜉, we have W j = Λ j ( W j 1 ) . Moreover, W j converge to F ( C ) in a certain multi-graded Hilbert scheme Hilb (see [24]). It is shown in [24] that transverse automorphisms of the cone 𝒞, i.e., automorphisms of 𝒞 that preserve the Reeb vector field ξ = J C ( r r ) , form a reductive complex Lie group. Then a variant of Luna’s slice theorem, proved by Donaldson [27], implies the existence of a one-parameter subgroup λ ( t ) of G ξ such that C = lim t 0 λ ( t ) . W . That is, 𝑊 admits an equivariant degeneration to 𝒞. Since 𝒞 has only an isolated singularity and 𝑊 is a cone, it follows that 𝑊 also has only an isolated singularity. Since 𝒞 has an isolated log terminal singularity and has a quotient singularity in dimension 2, we know that 𝑊 also has log terminal singularity by [40, Proposition 9.1.9 and Theorem 9.1.19]. Therefore, 𝑊 admits a Fano cone structure, justifying the terminology intermediate K-semistable cone.

4 Polynomial convergence implies C = W

Suppose ( g Z , J Z , ω Z ) is conical of order δ > 0 ; we show in this section that 𝒞 is isomorphic to 𝑊 as Fano cones. This is achieved using the Hörmander L 2 -method to construct J Z -holomorphic functions, whose rescalings converge directly to holomorphic functions on 𝒞.

Let Φ denote the diffeomorphism given in Definition 1.2. Using the diffeomorphism Φ, we will identify the neighborhood U p of 𝑝 with U o , which, without loss of generality, we may assume to be 𝐵, the unit ball in 𝒞 centered at the vertex. We recall that d j denotes the rescaling of d Z by a factor 2 j , and B j denotes the unit ball in Z j , and Λ denotes the dilation r 2 r on 𝒞. Then, using (1.2), one can easily show the following result.

Lemma 4.1

There exists a constant C > 0 such that, for all x , y B and j N ,

| d j ( Λ j . x , Λ j . y ) d C ( x , y ) | C 2 j δ .

As mentioned in [24], the notion of convergence of holomorphic functions depends on the choice of the metric on B j B , which realizes the pointed Gromov–Hausdorff convergence. Moreover, the argument in [24] works for any such choice. The above lemma implies that

Λ j : ( B , d C ) ( B j , d j )

realizes the pointed Gromov–Hausdorff convergence as j . Therefore, in the following, when we talk about the convergence of holomorphic functions, we are considering convergence under this Gromov–Hausdorff approximation.

Let 𝜓 denote the plurisubharmonic function log r 2 ( log r 2 ) 1 2 on 𝒞. Note that

1 J C ̄ J C ψ = ( 1 + 1 2 ( log r 2 ) 1 2 ) 1 ̄ r 2 4 1 r ̄ r r 2 + 1 r ̄ r r 2 ( log r 2 ) 3 2 .

Then it follows that

(4.1) 1 J C ̄ J C ψ 1 ( log r 2 ) 3 2 r 2 ω C on { r 2 < e 1 } .

Since ( g Z , J Z , ω Z ) is Kähler–Einstein and conical of order δ > 0 , as a consequence of (4.1), we know that, in a neighborhood of 𝑜, for any ϵ > 0 , there exists a constant c ϵ > 0 such that

(4.2) 1 J Z ̄ J Z ψ + Ric ( ω Z ) c ϵ r 2 ϵ ω Z .

There are homogeneous holomorphic functions ( h 1 , , h N ) on 𝒞, generating R ( C ) and giving an embedding F = ( h 1 , , h N ) : C C N such that the Reeb vector field 𝜉 extends to a linear diagonal vector field Re ( 1 i d i z i z i ) and the dilation Λ on 𝒞 extends to a diagonal dilation on C N , Λ ( z 1 , , z N ) = ( 2 d 1 z 1 , , 2 d N z N ) , where d i = deg ( h i ) > 0 .

Then, as ( g Z , J Z , ω Z ) is conical of order δ > 0 , we obtain that, for all l 0 ,

(4.3) | ω Z l ( ̄ J Z h i ) | ω Z = O ( r d i 1 l + δ ) .

We choose a neighborhood 𝑈 of 𝑜 with smooth strictly pseudoconvex boundary with respect to the complex structure J Z such that (4.2) holds on 𝑈, and choose the weight

ψ i = ( d i + δ 2 + n ) ψ

and the background Kähler metric ω Z . Using Theorem 2.12, we can solve the ̄ J Z -equation

(4.4) ̄ J Z u i = ̄ J Z h i ,

and get a solution u i of (4.4) with the estimate

(4.5) U | u i | 2 e ( d i + δ 2 + n ) ψ ω Z n < .

Let f i = h i u i . Then we have the following.

Lemma 4.2

For any i { 1 , , N } , 2 j d i f i ( Λ j . x ) h i ( x ) uniformly as j .

Proof

As h i is homogeneous of degree d i , we have 2 j d i h i ( Λ j . x ) = h i ( x ) . To prove the lemma, it suffices to show that 2 j d i u i ( Λ j . x ) 0 . This follows from (4.3), (4.5) and Proposition 2.3. ∎

This lemma says that, with respect to the Gromov–Hausdorff approximation Λ j , the holomorphic function 2 j d i f i | B j converges to h i | B on 𝒞 as j . Therefore, it implies that C = W in Donaldson–Sun’s two-step degeneration theory.

5 C = W implies polynomial convergence

The overall idea of the proof is similar to [53]. Firstly, using C = W , we get an almost Kähler–Einstein metric ( ω ̃ , J C , g ̃ ) in a neighborhood of o C , which is (by construction) polynomially close to the singular Kähler–Einstein metric ( ω Z , J Z , g Z ) . Then, solving a complex Monge–Ampère equation, we get a Calabi–Yau metric ( ω ̄ , J C , g ̄ ) in a neighborhood of o C , which is again polynomially close to the singular Kähler–Einstein metric ( ω Z , J Z , g Z ) . Finally, we can apply Chiu–Székelyhidi’s result [15] to get that ( ω ̄ , J C , g ̄ ) is polynomially close to the Calabi–Yau cone metric ( ω C , J C , g C ) .

We follow the notation in Section 3. Moreover, in the following, we identify 𝒞 with its image F ( C ) C N , and as we only care about the local geometry near 𝑝, we also identify 𝑍 with its image F 1 ( B 1 ) C N .

5.1 Construction of the diffeomorphism

The following lemma is proved in [39, Lemma 3.6]. We include its proof for readers’ convenience. Recall that G ξ denotes the subgroup of GL ( N ; C ) consisting of elements that commute with the actions generated by 𝜉. Let 𝐺 denote the subgroup of G ξ which leaves 𝒞 invariant.

Lemma 5.1

Lemma 5.1 ([39])

If 𝑊 is isomorphic to 𝒞 as polarized affine cones, then we can make W j equal to 𝒞 for all 𝑗 in Donaldson–Sun’s two-step degeneration theory.

Proof

Recall that we identify 𝒞 with F ( C ) C N and both W j and F ( C ) are in a Hilbert scheme Hilb . Since 𝑊 is isomorphic to 𝒞 and W j is isomorphic to 𝑊 as polarized affine cones, there exists g j G ξ such that W j = g j ( C ) . Moreover, W j converges to 𝒞 in Hilb ; then a variant of Luna’s slice theorem, proved by Donaldson [27], implies that there exist h j G ξ with h j Id and h j ( W j ) = h j ( g j ( C ) ) = C . Replacing F j with h j F j , we can therefore guarantee that W j = C . ∎

The following lemma can be proved in a similar way to [53, Proposition 3.10]. It plays an essential role for constructing Kähler metrics by pulling back their potentials. We fix a Kähler cone metric ω ξ on C N with the Reeb vector field given by 𝜉. Such a metric always exists; see [38, Lemma 2.2]. Although, when restricted to 𝒞, ω ξ is different from ω C , they share the same Reeb vector field. This implies that ω C and ω ξ are uniformly comparable and they define the same Banach space using (2.1) with comparable norms on it. Denote by r ξ = d ω ξ ( 0 , ) the radial function on C N defined by ω ξ .

Lemma 5.2

There exists a diffeomorphism Φ from a neighborhood of o C to a neighborhood of p Z such that, for some δ 0 > 0 and all k 0 , as r ξ 0 ,

| ω ξ k ( Φ J Z J C ) | ω ξ = O ( r ξ δ 0 k ) .

Proof

Recall that we identify a neighborhood of p Z with a neighborhood of

p F 1 ( B 1 ) C N .

We define Φ 1 to be the normal projection map to 𝒞, i.e., for 𝑥 in a neighborhood of 𝑝, we let Φ 1 ( x ) be the unique point in 𝒞 that is closest to 𝑥 with respect to the metric ω ξ . We need to show that this is well-defined in a neighborhood of p Z and satisfies the desired properties.

As shown in Section 3, 𝑍 is a normal complex analytic space with an isolated singularity at 𝑝. Then we know that there exist holomorphic functions f 1 , , f m defined in a neighborhood of 0 C N and a neighborhood 𝑉 of 𝑝 in 𝑍 such that 𝑉 is the common zero of f i and rank ( d f 1 , , d f m ) = N n on V p . Let g i be the initial term of f i with respect to the weight induced by 𝜉, i.e., the homogeneous term of smallest weight in the Taylor expansion of f i . We may assume f i generate the ideal of 𝒞, which is equal to the weighted tangent cone of ( F 1 ( B 1 ) , p ) with respect to 𝜉. Let w i denote the weight of f i .

For l 1 , we denote by A l the annulus in C N defined by 2 l r ξ 2 l + 1 . By the conical nature of ω ξ , it suffices to consider the normal projection from Λ l ( F 1 ( B 1 ) A l + 1 ) A 1 to 𝒞 for l 1 . Then the defining functions of Y l : = Λ l ( F 1 ( B 1 ) A l + 1 ) A 1 can be chosen to be f i , l ( z ) = 2 w i l f i ( Λ l . z ) . As the weights of 𝜉 form a discrete set and g i is the initial term of f i with respect to this weight, we know that there exist some holomorphic functions e i l ( z ) with sup | z | 1 | e i , l ( z ) | uniformly bounded independently of 𝑖 and 𝑙 and constants δ i > 0 such that f i , l ( z ) = g i ( z ) + 2 l δ i e i , l ( z ) . Let δ 0 = min i δ i .

Claim

For 𝑙 sufficiently large, there are finitely many open sets U γ C N such that C U γ is a cover of 𝒞, and for each 𝛾, Y l U γ is a graph of holomorphic functions over C U γ , that is, there are holomorphic functions h γ , l : U γ W C N with

| ω ξ k ( h γ , l id ) | ω ξ C k 2 l δ 0

for all k 0 and Y l γ Im ( h γ , l ) .

To prove the claim, fix a point z C A 1 . Then we can find a neighborhood U C N such that C U is given by the zero set of N n terms g α and rank ( d f 1 , , d f m ) = N n on 𝑈. For simplicity of notation, we may assume these are g 1 , , g N n . Shrinking 𝑈 if necessary and using implicit function theory, we can find local holomorphic coordinates { ζ 1 , , ζ N } such that ζ α = g α for α = 1 , , N n . Now we have f α , l = g α ( z ) + 2 l δ α e α l ( z ) . It follows that, for 𝑙 large, the common zero set of { f α , l } α = 1 N n is a smooth complex submanifold in 𝑈, so in particular, it agrees with Y l U , by shrinking 𝑈 if necessary. Using the local coordinates { ζ 1 , , ζ N } and implicit function theorem, it is easy to see that Y l U is contained in the image of a holomorphic function h l : U C N such that | ω ξ k ( h l id ) | ω ξ C k 2 l δ 0 for all k 0 . Since A 1 is compact, the claim follows.

Since the normal injectivity radius of C A 1 is uniformly bounded, there is a tubular neighborhood 𝒩 of C A 1 such that the normal projection map Π : N C is smooth. It follows from the claim that, for 𝑙 large, Y l N . So Φ is well-defined and smooth in a neighborhood of 𝑝. It is straightforward to check that it satisfies the desired derivative bounds. The estimates on the complex structures follow from the fact that h γ , l is holomorphic and ω ξ is a Kähler metric. ∎

5.2 Construction of the almost Kähler–Einstein metrics

Using the uniqueness of the tangent cone at p Z , one can show that, see for example [53, Lemma 3.1], a diffeomorphism Φ 0 : U o U p such that, for all k 0 ,

(5.1) lim s 0 sup B s s k ( | g C k ( Φ 0 g Z g C ) | + | g C k ( Φ 0 J Z J C ) | + | g C k ( Φ 0 ω Z ω C ) | ) = 0 .

In particular, we know that, on a neighborhood of 𝑝, there exists a smooth function r Z which is comparable to the distance function d Z ( p , ) . Recall that we use 𝑟 to denote the radial function on 𝒞 defined by the Calabi–Yau cone metric and r ξ to denote the radial function on C N defined by ω ξ , and we know that 𝑟 and r ξ are comparable on 𝒞.

Repeating the proof of [53, Proposition 3.5], we get the following rough estimates. For simplicity of notation, we omit the pullback notation and the restriction notation, and naturally view ω Z and ω ξ as Kähler forms on F 1 ( B 1 ) in the following.

Lemma 5.3

In a neighborhood of p F 1 ( B 1 ) , for all δ > 0 and k 1 , we have

C δ 1 r δ ω Z ω ξ C δ r δ ω Z , | ω Z k ω ξ | ω Z C δ , k r δ k , C δ 1 r Z 1 + δ r ξ C δ r Z 1 δ .

Note that we do not claim the uniform equivalence between ω Z and ω ξ . Nonetheless, we know the error is smaller than any polynomial order and this suffices for our applications. Then we show that we can find a Kähler potential for ω Z with an almost quadratic order estimate.

Lemma 5.4

For any ϵ > 0 , there exists a function ψ ϵ in a neighborhood of 𝑝 such that ω Z = 1 ̄ φ ϵ and | ω Z k φ ϵ | = O ( r Z 2 ϵ k ) for all k 0 .

Proof

Firstly, we show that ω Z is a 𝑑-exact 2-form in a neighborhood of 𝑝. As a consequence of (5.1), we know that a neighborhood of 𝑝 in F 1 ( B 1 ) is diffeomorphic to ( 0 , 1 ) × Y , some smooth compact manifold 𝐿, and ω Z is a closed 2-form with norm bounded with respect to the cone metric g C = d r 2 + r 2 g L . Note that ω Z defines a cohomology class 𝛼 in H 2 ( L , R ) . The fact that the norm ω Z is bounded with respect to g C implies that, in the class 𝛼, we can find smooth representatives which are O ( r 2 ) with respect to the fixed Riemannian metric g L . Therefore, this cohomology class 𝛼 has to be 0. Therefore, we know that ω Z is a 𝑑-exact 2-form in a neighborhood of 𝑝.

Secondly, we show that ω Z = d η for some real 1-form 𝜂 with

| ω Z k η | = O ( r Z 1 k ) for all k 0 .

This is a standard integration argument. Write ω Z = d r α 1 + α 2 with r α 1 = r α 2 = 0 . Then we define η = 0 r α 1 d r . Using the fact that ω Z is 𝑑-exact, one can directly check that ω Z = d η . Then the estimate follows from (5.1).

Let 𝜓 denote the plurisubharmonic function log r ξ 2 ( log r ξ 2 ) 1 2 on C N . Then, by Lemma 5.3, we know that, in a neighborhood of 𝑜, for any ϵ > 0 , there exists a constant c ϵ > 0 such that

1 J Z ̄ J Z ψ c ϵ r 2 ϵ ω Z .

By Theorem 2.12, choosing the weight as ( n ϵ + 2 ) ψ and using that ω Z is Kähler–Einstein, we can solve the ̄ equation with integral estimate for φ ϵ , ̄ J Z φ ϵ = η 0 , 1 . As in Section 4, using Proposition 2.3, we can get a pointwise estimate for φ ϵ . ∎

Let Φ and δ 0 be the diffeomorphism and the constant obtained in Lemma 5.2. Choosing ϵ = δ 0 2 in Lemma 5.4, we obtain a Kähler potential φ = φ ϵ . Then we can define a Kähler metric ( g ̃ , ω ̃ , J C ) in a neighborhood of 𝑜 by

ω ̃ = 1 J C ̄ J C Φ ( φ ) , g ̃ ( , ) = ω ̃ ( , J C ) .

As a consequence of Lemma 5.2, Lemma 5.3 and Lemma 5.4, we obtain that, in a neighborhood of 𝑜, there exists δ 1 > 0 such that, for all k 0 , we have

(5.2) | g ̃ k ( Φ ω Z ω ̃ ) | = O ( r δ 1 k ) and | g ̃ k ( Φ g Z g ̃ ) | = O ( r δ 1 k ) .

5.3 Solving complex Monge–Ampère equations locally

Next we are going to solve some complex Monge–Ampère equations via Banach fixed point theorem to construct Calabi–Yau metrics in a neighborhood of o C . For this, we need to specify the Banach spaces that we are working on. Combining (5.1), Lemma 5.2 and the construction of ( g ̃ , ω ̃ ) , we know that there exists a self-diffeomorphism Φ 1 of U o such that

(5.3) lim s 0 sup B s s k ( | g ̃ k ( g ̃ Φ 1 g C ) | + | g ̃ k ( J C Φ 1 J C ) | + | g ̃ k ( ω ̃ Φ 1 ω C ) | ) = 0 ,

and for all δ > 0 , there exists constant C δ such that

(5.4) C δ 1 r 1 + δ Φ 1 r C δ r 1 δ .

In the following, Banach spaces are defined with respect to the cone metric Φ 1 g C .

Proposition 5.5

There exists a δ 2 > 0 such that, for sufficiently small 𝑠, there exists φ C 2 + δ 2 k , α ( B s ) such that

ω ̄ = ω ̃ + 1 J C ̄ J C φ

is a Calabi–Yau metric.

Proof

Since ω Z is Kähler–Einstein, i.e., Ric ( ω Z ) = λ ω Z for some λ R , by (5.2), we know that

| g ̃ k Ric ω ̃ | = O ( r δ 1 2 k ) .

Then, by a similar argument to Lemma 5.4, we obtain that there exists a smooth function f ̃ such that

Ric ( ω ̃ ) = 1 ̄ f ̃ and | g ̃ k f ̃ | = O ( r δ 1 2 k ) .

Then we want to solve

( ω ̃ + 1 ̄ φ ) n = e f ̃ ω ̃ n .

Fix δ 2 ( 0 , 1 ) Γ and δ 2 δ 1 4 . Here Γ is the set in Proposition 2.2. Fix k 2 n + 2 and α ( 0 , 1 ) . Let us define

B = { u C 2 + δ 2 k + 2 , α ( B r 0 ) : u C 2 + δ 2 k + 2 , α ( B r 0 ) ϵ 0 } ,

where ϵ 0 is chosen to be sufficiently small such that ω ̃ + 1 ̄ u is uniformly equivalent to ω ̃ 0 and r 0 is to determined later. Let us consider the operator

F : B C δ 2 k , α ( B r 0 ) , u log ( ω ̃ + 1 ̄ u ) n e f ̃ ω ̃ n .

Then we have

F ( 0 ) C δ 2 k , α ( B r 0 ) r 0 δ 1 / 4 f ̃ C δ 1 / 2 k , α ( B r 0 ) C 0 r 0 δ 1 / 4 ,
Q ( u ) Q ( v ) C δ 2 k , α ( B r 0 ) C ( u C 2 k + 2 , α ( B r 0 ) + v C 2 k + 2 , α ( B r 0 ) ) u v C 2 + δ 2 k + 2 , α ( B r 0 ) C r 0 δ 2 ( u C 2 + δ 2 k + 2 , α ( B r 0 ) + v C 2 + δ 2 k + 2 , α ( B r 0 ) ) u v C 2 + δ 2 k + 2 , α ( B r 0 ) ,
where Q ( u ) = F ( u ) F ( 0 ) d 0 F ( u ) .

By estimate (5.3), we can apply Proposition 2.2. Let 𝒯 denote the right inverse of Δ g ̃ . Choosing r 0 small, we can apply the Banach fixed point theorem to the operator

N ( u ) : = T ( F ( 0 ) Q ( u ) ) .

Therefore, we obtain a solution φ C 2 + δ 2 k + 2 , α ( B r 0 ) . ∎

Let ω ̄ : = ω ̃ + 1 J C ̄ J C φ and g ̄ : = ω ̄ ( , J C ) . Combining Proposition 5.5 and estimates (5.3) and (5.4), we know that there exists δ 3 > 0 such that, for all k 0 ,

(5.5) | g ̃ k ( g ̃ g ̄ ) | + | g ̃ k ( ω ̃ ω ̄ ) | = O ( r δ 3 k ) .

Although, in our setting, ( ω ̄ , J C , g ̄ ) may not be the pointed Gromov–Hausdorff limit of smooth polarized Kähler–Einstein metrics, we still have the following:

  • the regular part is geodesically convex since the space has only an isolated singularity and the metric is asymptotic to a cone with smooth link near this singularity;

  • Donaldson–Sun’s two-step degeneration theory still applies. Indeed, by estimate (5.2) and (5.5), we obtain that the tangent cone of g ̄ at 𝑜 still exists and is equal to 𝒞 and

    | Rm g ̄ | = O ( r 2 ) .

    Therefore, we can apply the discussion given in Section 3.

With these two properties, the argument in [15] still works for the Calabi–Yau metric ( ω ̄ , J C , g ̄ ) and therefore we get the following polynomial closeness result.

Theorem 5.6

Theorem 5.6 ([15])

There exist a constant α > 0 and a biholomorphism

Ψ : ( U ̃ o , J C ) ( U o , J C )

such that, for all k 0 , as r 0 , we have | g C k ( Ψ g ̄ g C ) | = O ( r α k ) .

Then the polynomial closeness between ( ω Z , J Z , g Z ) and ( ω C , J C , g C ) follows from (5.2), (5.5) and Theorem 5.6.

6 Examples

We construct examples of singular Kähler–Einstein metrics, which have a unique tangent cone with smooth link, yet are not polynomially close to their tangent cones. In [12], a Riemannian metric with G 2 holonomy, in particular Ricci-flat, and exhibiting only a logarithmic rate of convergence to the cone is constructed. A key feature of our examples is that they can also be realized as Gromov–Hausdorff limits of non-collapsing Calabi–Yau metrics, and therefore in particular showing that the result of Colding–Minicozzi [17] is sharp in a certain sense. This was first pointed out in [39].

Let 𝐷 be an ( n 1 ) -dimensional Fano manifold and let Ω be a nowhere vanishing holomorphic volume form Ω on K D , the canonical bundle of 𝐷. Let h D denote a Hermitian metric with negative curvature form, ξ K D ; then ψ = ξ h D 2 is a plurisubharmonic function on K D and it is strictly plurisubharmonic on K D D . Then, for any δ > 0 , { ψ < δ } is a domain containing 𝐷, with a strictly pseudoconvex boundary. Therefore, there is a holomorphic map π : ( K D , D ) ( K D × , o ) which blows down the zero section. This map 𝜋 can be viewed as a crepant resolution of the isolated singularity. We may assume K D × is holomorphically embedded in C N . Let ω 0 denote the restriction of the Euclidean metric to K D × . Using 𝜋, we identify it with an semipositive form on K D .

Fix a small δ > 0 such that U = { ψ < δ } has smooth strictly pseudoconvex boundary. Fix a Kähler form 𝜂 on K D . Then, for 0 < ϵ < 1 , we consider the following Monge–Ampère equations with Dirichlet boundary condition:

(6.1) { ( ω 0 + ϵ η + 1 ̄ φ ϵ ) n = 1 n 2 Ω Ω ̄ , φ ϵ | U = 0 .

The following result is proved in [7, 29, 32], following the fundamental work [8, 41]. For readers’ convenience, we give a sketch of the proof and refer to [7, 29, 32] for details. Let ω ϵ denote ω 0 + ϵ η + 1 ̄ φ ϵ when a solution of (6.1) exists.

Proposition 6.1

Proposition 6.1 ([7, 29, 32])

For any ϵ ( 0 , 1 ) , equation (6.1) admits a unique smooth solution φ ϵ . There exist constants 𝐶 independent of 𝜖 such that

φ ϵ L C and C 1 ω 0 ω ϵ C 1 n 2 Ω Ω ̄ ω 0 n ω 0 .

Moreover, for any k 0 and compact subset K U ̄ D , there exist constants C K , k such that

(6.2) ω 0 k φ ϵ L ( K ) C K , k .

Proof

The uniqueness of the solution follows from the maximal principle for complex Monge–Ampère equations. Note that A ( ψ δ ) is subsolution of (6.1) for 𝐴 sufficiently large, depending on 𝜖. Therefore, by [7, Theorem A], we know that (6.1) admits a smooth solution for any ϵ ( 0 , 1 ) .

The uniform L bound for φ ϵ is proved in [29, 32]. The comparison between ω 0 and ω ϵ can be derived from the Chern–Lu’s inequality and the uniform L bound for φ ϵ , since ω 0 is the restriction of the Euclidean metric to 𝑈 and hence has a uniform upper bound on the bisectional curvature. See [21, 29] and references therein for the uniform higher-order estimate (6.2). ∎

As we are going to take a pointed Gromov–Hausdorff limit of ω ϵ , we fix a point p D and establish the following uniform non-collapsing result.

Lemma 6.2

There exists a 𝐶 independent of 𝜖 such that

sup x U d ω ϵ ( p , x ) C .

Therefore, there exists a κ > 0 such that, for every r < min { 1 , dist ( x , U ) } , we have

Vol ( B ( x , r ) , ω ϵ ) κ r 2 n .

Proof

Suppose, on the contrary, that

sup x U d ω ϵ ( p , x ) as ϵ 0 .

As we have a uniform estimate for ω ϵ away from 𝐷, we know that d ω ϵ ( p , U ) as ϵ 0 . Note that we have Ric ( ω ϵ ) = 0 and

U ω ϵ n = U 1 n 2 Ω Ω ̄

has a uniform upper bound. This contradicts a theorem of Yau [50, Theorem I.4.1]. The second statement follows from the Bishop–Gromov volume comparison. ∎

Passing to a subsequence, we may assume ω ϵ converges in C loc ( U ̄ D ) to a Kähler form ω . Moreover, we may assume ( U ̄ , p , ω ϵ ) converges in the pointed Gromov–Hausdorff sense to ( Z , p , d Z ) . Then an argument similar to [53, Section 5] implies the following. See also [52].

Proposition 6.3

𝑍 coincides with the metric completion of ( U ̄ D , ω ) and is naturally homeomorphic to π ( U ̄ ) .

Therefore, a neighborhood of 𝑝 in 𝑍 is Stein and hence admits a complete Kähler metric. Moreover, by the estimate in Proposition 6.1, we know that the Kähler form ω Z on the Gromov–Hausdorff limit 𝑍 admits a bounded potential φ Z , and moreover, we have the gradient estimate for φ Z ; see [53, Lemma 5.2]. Therefore, we can view the space as being polarized by the trivial line bundle equipped with the Hermitian metric e φ Z . This allows one to extend Donaldson–Sun’s two-step degeneration theory to the pointed space ( Z , p , d Z ) ; see [30] for a detailed argument and see also [53, Section 5] for a related discussion. Then we get a 𝐾-semistable cone 𝑊 and its unique tangent cone 𝒞 at 𝑝. Both 𝑊 and 𝒞 depend only on the local algebraic structure of ( K D × , o ) by [46, 45], in particular, if we choose 𝐷 to be a smooth Fano manifold which does not admit a Kähler–Einstein metric, but admit a degeneration to a smooth Fano manifold D that admits a Kähler–Einstein metric. Such examples exist; in particular, some deformations of a Mukai–Umemura 3-fold are among them. See [25], [55, Section 7] and [26, Section 5]. Then we know that W = K D × K D × = C . More precisely, as 𝐷 degenerates to D , K D × admits an equivariant degeneration to K D × , which admits a Calabi–Yau cone metric. Therefore, K D × with the standard Reeb vector field is K-semistable. Moreover, as 𝐷 does not admit Kähler–Einstein metrics, K D (with the standard Reeb vector field) is not K-polystable. Therefore, the result in Section 4 in this paper implies that the singular Calabi–Yau metric near 𝑝 is not polynomially close to its tangent cone 𝒞.

7 Local non-positive Kähler–Einstein metrics

Let ( X , p ) be a germ of an isolated log terminal algebraic singularity. There is a well-developed theory for local stability of (Kawamata) log terminal singularities; see [44, 46, 45, 58] and the references therein. Roughly speaking, this local stability theory says that every (Kawamata) log terminal singularity has a two-step degeneration to a K-polystable Fano cone. The first degeneration is canonical and is induced by the unique (up to scaling) valuation that minimizes the normalized volume. Through this degeneration, we get a K-semistable Fano cone ( W , ξ W ) . Secondly, ( W , ξ W ) admits an equivariant degeneration to a K-polystable Fano cone ( C , ξ ) . In the following, we will call this process the two-step stable degeneration of a log terminal (algebraic) singularity.

In this paper, we consider the special case that 𝑊 is isomorphic to 𝒞 as Fano cones. For simplicity, we will denote this by C = W . We further assume that 𝒞 has only an isolated singularity.

We have the following setup, similar to that in Section 3. The Fano cone ( C , ξ ) admits a Calabi–Yau cone structure, that is, there exists a Kähler Ricci-flat cone metric ω C on 𝒞 with ξ = J C ( r r ) . There exist embeddings ( C , o ) ( C N , 0 ) and ( X , p ) ( C N , 0 ) such that, under the embedding, ξ = J C ( r r ) extends to a linear vector field on C N of the form

Re ( 1 i d i z i z i ) ,

which we also denote by 𝜉, where d i R > 0 for all 𝑖. Moreover, the weighted tangent cone of ( X , p ) with respect to 𝜉 coincides with 𝒞.

Assume we have a singular non-positive Kähler–Einstein metric with bounded potential on the germ ( X , p ) . This means that there exist a neighborhood 𝑈 of p X and a function

φ KE C ( U { p } ) L ( U )

such that the Kähler form ω KE = 1 ̄ φ KE defined on U { p } satisfies

(7.1) Ric ( ω KE ) = λ ω KE

for some constant λ 0 . By rescaling, we may assume λ { 1 , 0 } . Since we only care about the behavior of ω KE near 𝑝, in the following, we will shrink 𝑈 if necessary without explicitly mentioning this. Let us repeat the statement of Theorem 1.4 here.

Theorem 7.1

Suppose C = W and it has only an isolated singularity; then every singular non-positive Kähler–Einstein metric with bounded potential on the germ ( X , p ) is conical at 𝑝.

The same proof as in Section 4 implies that the isomorphism between 𝒞 and 𝑊 as Fano cones is also a necessary condition for the existence of a conical singular Kähler–Einstein metric. To show its sufficiency, we will closely follow Hein–Sun’s continuity argument. For the openness, we need to deal with the issue that ( X , p ) is not biholomorphic to 𝒞. For the closeness, we need Theorem 1.1 in this paper and the results [14, 46, 45], which allow us to remove the smoothability assumption in Hein–Sun’s work.

We start with the following result, which gives the initial metric we start with.

Lemma 7.2

On U { p } , there exists a Kähler form ω 0 = 1 ̄ φ 0 , which is Ricci-flat near 𝑝 and polynomially close to the Calabi–Yau cone 𝒞 and φ 0 | U = φ KE | U .

Proof

Using the diffeomorphism Φ obtained using Lemma 5.2 to pull back Kähler potentials and solving Monge–Ampère equations as we did in Section 5, we can prove that, in a neighborhood of 𝑝, there exists a conical Calabi–Yau metric ω ̃ 0 = 1 ̄ φ ̃ 0 . As before, we view 𝑟 as a smooth function in a neighborhood of 𝑝 via the diffeomorphism. Fix a small positive constant δ 1 such that both ω KE and ω ̃ 0 are well-defined in a neighborhood of { r δ 1 } and r 2 is strictly plurisubharmonic on it. Then we are going to use a scaling and cut-off argument similar to that in [2] to obtain the desired Kähler form ω 0 . Without loss of generality, we may assume min { r = δ 1 } φ KE = 10 . As φ KE is bounded, then we can choose 𝐴 sufficiently large such that φ 1 : = φ KE + A ( r 2 δ 1 2 ) 0 on the region { r δ 1 2 } . Then choose ϵ 1 δ 1 such that φ 1 2 on { δ 1 ϵ 1 r δ 1 } . Then fix two cut-off functions χ 1 and χ 2 which are smooth non-negative functions on ℝ such that

  • χ 1 ( t ) equals a constant for t 1 2 and χ 1 ( t ) t for t 3 2 and χ 1 , χ 1 ′′ 0 ,

  • χ 2 ( t ) 1 for t δ 1 ϵ 1 and χ 2 ( t ) 0 for t δ 1 ϵ 1 2 .

Then one can directly show that, for ϵ 2 sufficiently small,

ω 0 : = 1 ̄ ( ϵ 2 χ 2 ( r ) φ ̃ 0 + χ 1 ( φ 1 ) )

is a Kähler form and satisfies the required properties when we choose U = { r δ 1 } . ∎

Since ω KE = 1 ̄ φ KE is Kähler–Einstein (7.1) and ω 0 is Ricci-flat near 𝑝, there exists a pluriharmonic function f 1 in a neighborhood of 𝑝 (see [31]) such that

( 1 ̄ φ KE ) n = e λ φ KE + f 1 ω 0 n ,

where λ { 1 , 0 } . Then consider the following continuous path: we are looking for functions φ t C ( U p ) L ( U ) satisfying

(7.2) { ( 1 ̄ φ t ) n = exp ( λ t ( φ t + f 1 ) ) ω 0 n , φ t | U = φ KE .

Let ω t = 1 ̄ φ t . Let 𝐼 be the subset of [ 0 , 1 ] such that ω t is polynomially close to 𝒞 near 𝑝. Then, by Lemma 7.2, we have 0 I . We are going to show that 𝐼 is both open and closed, and therefore I = [ 0 , 1 ] . Assuming this, then by the uniqueness result [33, Theorem 1.4], we obtain φ 1 = φ KE and therefore ω KE is polynomially close to 𝒞 near 𝑝.

7.1 Openness

We first show that the set 𝐼 is open. This is analogous to [39, Theorem 2.19].

Theorem 7.3

Let ω T = 1 ̄ φ T be a solution to (7.2) for some T [ 0 , 1 ] . Suppose ω T is polynomially close to 𝒞; then there exists τ > 0 such that, for all t ( T τ , T + τ ) , (7.2) admits a solution ω t = 1 ̄ φ t , which is also polynomially close to 𝒞.

Proof

By assumption, there exist a δ > 0 and a diffeomorphism Φ from a neighborhood of 𝑝 in 𝑈 to a neighborhood of o C such that, for all k 0 ,

(7.3) | g C k ( Φ J X J C ) | + | g C k ( Φ ω T ω C ) | + | g C k ( Φ g T g C ) | = O ( r δ k ) .

Fix α ( 0 , 1 ) and a positive number β < min { δ , μ 1 + , 1 } such that ( 2 , 2 + β ) { μ i + } = , where μ i + are given in (2.3). Let C 2 + β 2 , α denote the weighted Banach space defined using the cone metric ( Φ 1 ) g C . More precisely, this means the following. We fix a neighborhood V U of 𝑝, where Φ is defined, and fix a cut-off function 𝜒 on 𝑈 which equals 1 in a neighborhood of 𝑝 and has support in 𝑉. Then a function f C 2 + β k + 2 , α means Φ ( χ f ) C 2 + β k + 2 , α , which is defined by (2.1) using g C and ( 1 χ ) f C 2 , α with respect to a fixed smooth metric on 𝑈. Let C 2 + β , 0 2 , α denote the Banach subspace of C 2 + β k + 2 , α , consisting of functions which have zero boundary value.

Let P H denote the vector space of homogeneous J C -pluriharmonic real-valued functions with growth rate in ( 0 , 2 ] . It is direct to show that these J C -pluriharmonic functions are the real part of J C -holomorphic functions [39, Lemma 2.13]. As J X is polynomially close to J C , using the Hörmander L 2 -method as we did in Section 4, we can construct J X -pluriharmonic functions with growth rate in ( 0 , 2 ] that converge to homogeneous J C -pluriharmonic functions. Let 𝒫 denote a vector space spanned by dim P H -dimensional J X -pluriharmonic functions with growth rate in ( 0 , 2 ] . The space 𝒫 is not unique and any choice of such a space works for the following argument. Let ℋ denote the space of all 𝜉-invariant 2-homogeneous Δ g C -harmonic real-valued functions on 𝒞. Transverse automorphisms of the cone 𝒞, i.e., automorphisms of 𝒞 that preserve the Reeb vector field ξ = J C ( r r ) , form a complex Lie group. Let 𝐺 denote the connected component of this Lie group. Consider spaces of functions

X = C 2 + β , 0 2 , α χ P χ 1 2 ( r 2 ϕ r 2 ) and Y = C β 0 , α ,

where ϕ G , and using the diffeomorphism Φ, we view χ 1 2 ( r 2 ϕ r 2 ) as functions in a neighborhood of 𝑝. It is shown in the proof of [39, Theorem] that 𝒳 admits a C 1 Banach manifold structure, and moreover, we have

T 0 X = C 2 + β , 0 2 , α χ P χ 1 2 r 2 χ H .

Then there exist a neighborhood 𝒰 of ( 0 , T ) in X × R and a neighborhood 𝒱 of 0 in 𝒴 such that the following map M : U V is well-defined and is C 1 :

M ( φ ) : = log ( ω T + 1 ̄ φ ) ω 0 n + λ t ( φ T + φ + f 1 ) .

Indeed, we just need to show f 1 C β 0 , α . Since f 1 is J X -pluriharmonic, in particular, it is ω T -harmonic. Then, by the gradient estimate, we know | ω T f 1 | C and hence f 1 C β 0 , α by the polynomial closeness (7.3). The linearization of ℳ at ( 0 , T ) along the first component is given by

Δ ω T + λ T Id : C 2 + β , 0 2 , α χ P χ 1 2 r 2 χ H C β 0 , α .

This is invertible as shown in Appendix A. The implicit function theorem implies that (7.2) has a solution ω t = ω T + 1 ̄ ψ t with ψ t U for some τ > 0 and all t ( T τ , T + τ ) .

It remains to prove that ω t is polynomially close to 𝒞 near 𝑝. By construction, ψ t has three terms, that is, ψ t = u 1 + χ u 2 + 1 2 ( r 2 ϕ r 2 ) , where ϕ G is a transverse automorphisms of 𝒞. Then we are going to show that Φ ϕ 1 gives the desired diffeomorphism. Note that we have the following:

(7.4) g C k ( ( Φ ϕ 1 ) J X J C ) = ( ϕ 1 ) ( ϕ g C k ( Φ J X J C ) ) ,
g C k ( ( Φ ϕ 1 ) ω t ω C ) = ( ϕ 1 ) ( ϕ g C k ( Φ ω t ϕ ω C ) ) .
As g C and ϕ g C are Riemannian cone metrics with the same scaling vector field, the weighted analysis with respect to the metric g C is comparable to that with respect to ϕ g C . This can be seen for example by writing ϕ g = g ( θ , ) . Then we have L r r θ = 0 , and as a consequence, we have | g C j θ | r j . Then (7.4) follows from (7.3) and a direct computation. Since u 1 C 2 + β , 0 2 , α and u 2 is J X -plurisubharmonic, and due to (7.3), we know that

Φ ω t ϕ ω C = ( Φ ω T ω C ) + d η

for some 1-form 𝜂 satisfying

(7.5) | g C j η | C ( | g C j ( Φ ( J X d u 1 ) ) | + | g C j ( Φ J X J C ) ( d ( r 2 ϕ r 2 ) ) | ) .

Then, using scaled Schauder estimates, we can get higher-order estimates for u 1 and hence for Φ ω t ϕ ω C using (7.3) and (7.5). ∎

7.2 Closeness

In this section, we are going to show the set 𝐼, which consists of 𝑡 such that ω t is polynomially close to 𝒞 near 𝑝, is closed, i.e., if t i I and t i T , then T I . Firstly, we recall the following uniform estimate proved in [21, 29, 32].

Proposition 7.4

There exists a constant C 0 independent of 𝑡 such that φ t L C 0 . Moreover, for any k 0 and compact subset K U ̄ { p } , there exist constants C K , k such that

(7.6) ω 0 k φ t L ( K ) C K , k .

Next we want to obtain a uniform diameter upper bound for ω t . As we have the uniform estimate (7.6) away from 𝑝, it is sufficient to show that diam ( V , ω t ) has a uniform upper bound, where 𝑉 is a small neighborhood of 𝑝. Notice that we can choose 𝑉 small such that ω t is Kähler–Einstein on 𝑉. In the Calabi–Yau setting, i.e., λ = 0 , this diameter upper bound can be proved in the same way as Lemma 6.2. For negative Kähler–Einstein metrics, the uniform diameter bound can be proved following [36]. For readers’ convenience, we give some details in Appendix B.

Proposition 7.5

There exists a constant independent of 𝑡 such that

diam ( U , ω t ) C .

Since the metric spaces ( U , ω t ) , t I , have only an isolated singularity and are asymptotic to a cone with a smooth link near the singularity, we know that their regular part is geodesically convex. As illustrated in [14, Proposition 2.3], the Bishop–Gromov volume comparison still holds for the metrics defined by ω t ; therefore, this uniform diameter upper bound implies volume non-collapsing, that is, there exists a κ > 0 such that, for every r < min { 1 , dist ( x , U ) } , we have

Vol ( B ( x , r ) , ω t ) κ r 2 n .

We can then apply the compactness theorem of Chen–Wang [14], which generalizes the results of Cheeger–Colding [9, 10] to metric spaces with mild singularities. Note that we have uniform estimates for ω t away from the point 𝑝 (see Proposition 7.4), so the boundary of 𝑈 does not introduce additional complications in the discussion of the convergence of ω t . To be more precise, the results in [14] are stated only for complete Calabi–Yau metrics; however, it is straightforward to verify that the same conclusions also hold for the local Kähler–Einstein metrics considered here, provided there are uniform estimates away from the singularities.

As part of their main result [14, Theorem 1.1], we obtain the following. We refer to [14] for the precise definition of the convergence used below. We simply remark here that the C ̂ -convergence implies pointed Gromov–Hausdorff convergence.

Theorem 7.6

Theorem 7.6 ([14])

For each sequence ω t i , by taking subsequence if necessary, we have

( U , p , ω t i ) C ̂ ( U , p , d ) .

Moreover, we have the following.

  1. ( U , d ) is a length space and has a regular-singular decomposition U = R S such that there exists a Kähler structure ( g , J ) on ℛ and 𝒮 is closed subset and has Hausdorff codimension at least 4.

  2. The distance structure induced by g is the same as the restriction of d and ℛ is geodesically convex.

  3. Every tangent cone of x S is a metric cone and satisfies the same properties (1) and (2) described above.

Since we have uniform estimates of ω t away from 𝑝, passing to a subsequence, we also know that

ω t i = 1 ̄ φ t i 1 ̄ φ T in C loc ( U p ) .

Let g T denote the Riemannian metric corresponding to the Kähler form 1 ̄ φ T . Then, using the argument in [24], [53, Section 5], we obtain the following.

Proposition 7.7

The limit ( U , g , J ) coincides with the metric completion of ( U p , g T , J X ) near 𝑝, which is naturally homeomorphic to 𝑈.

Then Donaldson–Sun’s two-step degeneration theory still applies in this setting. Indeed, since the Kähler form admits a bounded potential φ T , and [14, Proposition 2.24] provides a gradient estimate for φ T , and a neighborhood of 𝑝 is Stein with all tangent cones being metric cones with codimension 4 singularities, we can regard the space as being polarized by the trivial line bundle equipped with the Hermitian metric e φ T and apply Hörmander’s L 2 -estimate as in [24]. See [30] for a detailed argument and see also [53, Section 5] for a related discussion. Let C denote the tangent cone of ( g , J ) at 𝑝. Then, by [45], this C has to be isomorphic to 𝒞, which has an isolated singularity. Then the argument in Section 5 can be applied to ( g , J ) = ( ω T , J X ) to get polynomial closeness and therefore T I .

7.3 Positive singular Kähler–Einstein metrics

The continuity method does not work for singular Kähler–Einstein metrics with positive Ricci curvature, due to the lack of the openness and uniqueness; see [34, Section 5] for a related discussion. See [35] for using a variational method to study the existence of singular Kähler–Einstein metrics in a neighborhood of log terminal singularities.

It is very likely that, in the global setting, we can avoid such an issue. Let ω KE denote a singular Kähler–Einstein metric on a ℚ-Fano variety 𝑋, which has only isolated singularities and C = W for every point in X sing . To show that ω KE is conical, one may need two steps.

The first step is to construct a Kähler metric on 𝑋 with non-negative Ricci curvature and polynomially close to Calabi–Yau cones near singular points of 𝑋. Let θ 0 c 1 ( K X 1 ) be a smooth Kähler form. Solving a complex Monge–Ampère locally as we did in Section 5 and using the cut-off and gluing trick in [2], we obtain a Kähler form α 0 = θ 0 + 1 ̄ ψ 0 , which is Ricci-flat and polynomially close to Calabi–Yau cones near singularities of 𝑋, and a non-negative (1,1)-form η c 1 ( K X 1 ) , which is zero in a neighborhood of X sing and coincides with θ 0 outside a neighborhood of X sing . Then we know that there exists a function 𝑓, which is smooth on X reg and pluriharmonic in a neighborhood of X sing and satisfies

Ric ( α 0 ) η = 1 ̄ f .

We consider the first continuous path

(7.7) ( θ 0 + 1 ̄ ψ t ) n = c t e t f α 0 n ,

where c t is a normalization constant. Note that, along the path (7.7), the Kähler metrics are Ricci-flat in a neighborhood of X sing and one can adapt the argument in Section 7 to show that every metric in this continuous path is polynomially close to Calabi–Yau cones near X sing . As an output, we get a Kähler metric

α : = θ 0 + 1 ̄ ψ 1 ,

with Ric ( α ) = η and polynomially close to Calabi–Yau cones near X sing .

The second continuity path is the classical one of Aubin,

(7.8) ( α + 1 ̄ φ t ) n = e h t φ t α n .

Here ℎ is a smooth function on X reg , and in a neighborhood of X sing , it can be decomposed as h 0 + h 1 , where h 0 is pluriharmonic and h 1 satisfies | α k h 1 | = O ( r 2 k ) , and satisfies

Ric ( α ) α = 1 ̄ h and X ( e h 1 ) α n = 0 .

We want to show that, along this continuous path, every metric is conical. Let ω t be a solution of (7.8); then Ric ( ω t ) = t ω t + ( 1 t ) α . Openness for t ( 0 , 1 ) can be proved in a similar vein to Aubin’s work [3], and openness at t = 0 can be proved using a trick of Székelyhidi [54, Section 3.5]. Darvas [20] proves that the existence of a singular Kähler–Einstein metric implies that the 𝐾-energy is proper in a certain sense. Given that the functional J χ is convex along smooth geodesics when 𝜒 is a semipositive form proved by Chen [13], and holomorphic vector fields generate smooth geodesics in the space of Kähler potentials, one may adapt the argument in [56, Section 2] to obtain the uniform C 0 bound for φ t and hence the higher-order estimate away from X sing . To show that ω 0 is conical, we need to use another continuous path

(7.9) ( α + 1 ̄ u t ) n = c t e t h α n .

Then one may generalize the compactness and regularity results in [9, 10, 14] to non-collapsing Kähler metrics with isolated singularities and Ricci curvature lower bound, and then establish the closeness of being conical in the continuous paths (7.8) and (7.9).

Award Identifier / Grant number: DMS-2304692

Funding statement: The author is partially supported by NSF Grant DMS-2304692.

A Laplacian for conical metrics

The result in this appendix is standard and we just aim to give a detailed and elementary proof of the result we used in Section 7.1. First we recall some notation. Let ( g C , J C ) denote a Calabi–Yau cone metric and let ( g , J ) denote a conical metric of order 𝛿 in the sense of Definition 1.2 defined on B 2 , the ball of radius 2 centered at the vertex of the cone. The weighted Hölder spaces are defined by (2.1) and C 2 + β , 0 k + 2 , α is the Banach subspace of C 2 + β k + 2 , α , consisting of functions which have zero boundary value.

Let P H denote the vector space of homogeneous J C -pluriharmonic real-valued functions with growth rate in ( 0 , 2 ] . It is direct to show that these J C -pluriharmonic functions are the real part of J C -holomorphic functions [39, Lemma 2.13]. As 𝐽 is polynomially close to J C , using the Hörmander L 2 -method as we did in Section 4, we can construct 𝐽-pluriharmonic functions with growth rate in ( 0 , 2 ] that converge to homogeneous J C -pluriharmonic functions. Let 𝒫 denote a vector space spanned by dim P H -dimensional 𝐽-pluriharmonic functions with growth rate in ( 0 , 2 ] . The space 𝒫 is not unique and any choice of such a space works for the following argument. Let ℋ denote the space of all 𝜉-invariant 2-homogeneous Δ g C -harmonic real-valued functions on 𝒞.

Let 0 = λ 0 < λ 1 λ 2 denote the eigenvalues of the Laplacian on the link 𝐿 of 𝒞 (listed with multiplicity), and let ϕ 0 , ϕ 1 , ϕ 2 , denote an associated orthonormal basis of eigenfunctions. Then r μ i ± ϕ i are homogeneous harmonic functions on 𝒞, where

μ i ± = m 2 2 ± ( m 2 ) 2 4 + λ i .

Fix α ( 0 , 1 ) and a positive number β < min { δ , μ 1 + , 1 } such that ( 2 , 2 + β ) { μ i + } = . Fix a cut-off function 𝜒 such that χ 1 near the vertex and Supp ( χ ) B 1 . The main result of this appendix is the following. Note that, due to the choice of the parameters and ( g , J ) being conical, the operator is well-defined.

Proposition A.1

For any t 0 , the following operator is invertible:

Δ g t Id : C 2 + β , 0 k + 2 , α ( B 1 ) χ P χ 1 2 r 2 χ H C β k , α ( B 1 ) .

Proof

By the maximal principle, this operator is injective; therefore, it is sufficient to show the surjectivity. It is enough to show that Δ g : C 2 + β , 0 k + 2 , α C β k , α is a Fredholm operator with index bounded below by dim ( P ) dim ( H ) 1 .

Note that, by the Schauder estimate, the operator

Δ g C : C 2 + β , 0 k + 2 , α C β k , α

has closed image. Let 𝑉 denote the vector space spanned by r μ i + ϕ i , μ i + ( 0 , 2 ] . Then we are going to show that

Δ g C : C 2 + β , 0 k + 2 , α χ V C β k , α

is an isomorphism. It suffices to show the surjectivity. By Proposition 2.2, for any f C β k , α , there exists u ̄ C 2 + β k + 2 , α such that

Δ g C u ̄ = f and u ̄ C 2 + β k + 2 , α ( B 1 ) C f C β k , α ( B 1 ) .

Expand u ̄ into Fourier series

u ̄ = μ i + > 0 u ̄ i ( r ) ϕ i ( y ) .

Consider the following function 𝑢:

μ i + > 2 ( u ̄ i ( r ) u ̄ i ( 1 ) r μ i + ) ϕ i ( y ) + μ i + ( 0 , 2 ] ( u ̄ i ( r ) ( 1 χ ) u ̄ ( 1 ) r μ i + ) ϕ i ( y ) + χ μ i + ( 0 , 2 ] u ̄ i ( 1 ) r μ i + ϕ i .

Then we know u C 2 + β , 0 k + 2 , α and Δ g C u = Δ g C u ̄ = f . By Theorem 2.5, we have

dim V = dim ( P ) + dim ( H ) + 1 ;

therefore, Δ g C has index dim ( P ) dim ( H ) 1 .

It is proved in [49, Theorem 6.1] that Δ g is Fredholm and has the same index as Δ g C . We give a sketch of the proof here. As 𝑔 is conical, by writing g ( , ) = g C ( θ , ) and considering g s : = g C ( θ s , ) for s [ 0 , 1 ] , we have a smooth family of conical metrics connecting g C and 𝑔. As the index of Fredholm operators is a continuous function. it is enough to show that, for every conical metric 𝑔, Δ g : C 2 + β , 0 k + 2 , α ( B 1 ) C β k , α ( B 1 ) is Fredholm.

Solving the Dirichlet problem on B 1 B ϵ , taking limits and using a Schauder estimate, we know that there exists a bounded operator P 1 : C β k , α C 0 , 0 k + 2 , α such that Δ g P 1 = Id . For r ( 0 , 1 ) , let ρ r denote the cut-off function which equals 1 on B 1 B r and has support on B 1 B r 2 . Then consider the operator

Δ : = Δ g C + ( 1 ρ r ) ( Δ g Δ g C ) .

As 𝑔 is conical, Δ Δ g C has sufficiently small operator norm if 𝑟 is sufficiently small. Fix an 𝑟 sufficiently small such that Δ is Fredholm. Let P 2 denote a Fredholm inverse of Δ , that is, both Id Δ P 2 and Id P 2 Δ are compact operators. We are going to use P 1 and P 2 to construct a Fredholm inverse of Δ g , thereby proving it is Fredholm.

Let χ 1 be a cut-off function, which has support on B B r 4 and equals 1 on supp ( ρ r 2 ) . Let χ 2 be a cut-off function, which has support on B r 2 and equals 1 on supp ( 1 ρ r 2 ) . Then one can check the following operator:

T ( f ) : = χ 1 P 1 ( ρ r 2 f ) + χ 2 P 2 ( ( 1 ρ r 2 ) f ) ,

making Id Δ g T and Id T Δ g compact operators. Let us check this property for the former and the latter is similar. First notice that the inclusion operator C β l , α C β k , α is compact for l > k . Then

f Δ g T ( f ) = f χ 1 ρ r 2 f + χ 2 Δ g P 2 ( ( 1 ρ r 2 ) f ) + K ( f )

for some compact operator 𝐾. Notice that, on supp ( χ 2 ) , Δ = Δ g , since Id Δ P 2 is a compact operator, we obtain that

f Δ g T ( f ) = f χ 1 ρ r 2 f + χ 2 ( 1 ρ r 2 ) f + K ( f )

for some compact operator K . Note that χ 1 1 on supp ( ρ r 2 ) and χ 2 equals 1 on supp ( 1 ρ r 2 ) ; therefore, we get that Id Δ g T = K is a compact operator. ∎

B Diameter bounds for Kähler manifolds with boundary

The results in this section follow essentially from [37, 36]. For readers’ convenience, we give a sketch of the proof and mention some modifications needed for manifolds with boundary. Results here are not optimal, but are sufficient for our purpose.

We recall the setting considered in Section 7. That is, we have a normal Stein space ( U , p ) with an isolated log terminal singularity, embedded into ( C N , 0 ) with U being strictly pseudoconvex. We have a smooth function f 1 which is pluriharmonic in a neighborhood of 𝑝. For some λ { 1 , 0 } , we have a family of Kähler metrics with uniformly bounded potential ω t = ̄ φ t , t [ 0 , 1 ] , satisfying

{ ( 1 ̄ φ t ) n = exp ( λ t ( φ t + f 1 ) ) ω 0 n , φ t | U = φ KE .

Moreover, we know that each ω t is conical at 𝑝. The main result in this appendix is the following uniform diameter estimate.

Proposition B.1

There exists a constant 𝐶 independent of 𝑡 such that diam ( U , ω t ) C .

In the following, we will omit the index 𝑡 and just write ω t as 𝜔. It is clear from the statement and proof that the estimates in the following results are uniform for 𝑡 and only depend on the estimates in (B.1). Let ω E = 1 ̄ φ E be the restriction of the Euclidean metric on C N and let e F = ω 0 n / ω E n . By our assumption, we know that there exist a constant 𝐴 and p > 1 such that

(B.1) φ t L + f 1 L A , U e p F ω E n A .

Green functions and Sobolev inequalities

Firstly, we recall some basic properties for Green functions on manifolds with boundary [4, Chapter 4]. Let G ω ( x , y ) denote the Green function for the Laplacian Δ ω on 𝑋 which is a manifold with boundary. It satisfies the following.

  1. G ω is defined and smooth on X ̄ × X ̄ { ( x , x ) : x X } and G ω ( x , y ) = G ω ( y , x ) .

  2. G ω ( x , y ) 0 and it equals zero if and only if x X or y X .

  3. For every u C 2 ( X ̄ ) , we have

    (B.2) u ( x ) = X G ω ( x , y ) Δ ω u ( y ) d V X n G ω ( x , y ) u d S = X G ω , u d V X n G ω u d S ,

    where n is the unit out-normal vector of the boundary and d S is the volume element induced by the Riemannian metric on the boundary.

Let U δ denote an exhaustion of U { p } with smooth boundary. For example, we can choose

U δ = U { | z | > δ } .

Let G ω , δ ( x , y ) denote the Green function for the Laplacian Δ ω on the region U δ .

Lemma B.2

There exist constants ϵ 0 = ϵ 0 ( n , p ) and C 1 = C 1 ( n , p , A , ϵ 0 ) such that, for all δ > 0 , we have

(B.3) sup x U δ U δ G ω , δ ( x , ) 1 + ϵ 0 ω n C 1 .

Proof

Step 1.  Let 𝑣 solve the Laplacian equation Δ ω v = 1 in U δ and v = 0 on U δ . By the maximal principle, we have v L 2 n φ t L A . Applying the Green formula (B.2) to 𝑣, we obtain

(B.4) sup x U δ U δ G ω , δ ( x , ) ω n A .

Step 2.  Fix a large number 𝑘 and consider a smooth positive function H k which is a smoothing of min { G ω , δ ( x , ) , k } . Then, by (B.4) and (B.1), and Hölder inequality, we obtain that there exist constants ϵ 0 , A 0 that depend only on 𝑝, 𝑛 and 𝐴 such that

(B.5) U ( H k n ϵ 0 e F ) p ω E n A 0 .

Let ψ k be the solution of the following complex Monge–Ampère equation:

{ ( 1 ̄ ψ k ) n = H k n ϵ 0 e F ω E n in U , ψ k | U = 0 .

Then, by [32] and (B.5), we have a uniform L bound on ψ k . Let v k solve the equation

Δ ω v k = H k ϵ 0 in U δ and v = 0 on U δ .

By the maximal principle, v k L C ψ k L C . Applying the Green formula to v k and letting k , we obtain (B.3). ∎

Proposition B.3

There exist constants q = q ( n , p ) and C 2 = C 2 ( ϵ 0 , n , p , C 1 , q ) such that, for u W 0 1 , 2 ( U δ ) , we have

(B.6) ( U δ | u | 2 q ω n ) 1 q C U δ | u | ω 2 ω n .

Proof

It is sufficient to prove this for u C c ( U δ ) . Applying (B.2) to a constant function and then applying the Green formula (B.2) to ( G ω , δ ( x , ) + 1 ) β , for any β > 0 , we get

(B.7) sup x U δ U δ | y G ω , δ ( x , ) | ω 2 ( G ω , δ ( x , ) + 1 ) 1 + β ω n 1 β .

Applying (B.2) to 𝑢 and using Hölder inequality and (B.7), we obtain that, for any β , q > 0 , we have

| u ( x ) | 2 q 1 β q ( U δ ( G ω , δ ( x , y ) + 1 ) 1 + β | u ( y ) | 2 ω n ( y ) ) q .

Integrating over U δ and using Minkowski inequality, one obtains that

(B.8) ( U δ | u | 2 q ω n ) 1 q 1 β U δ ( U δ ( G ω , δ ( x , y ) + 1 ) ( 1 + β ) q ω n ( x ) ) 1 q | u | 2 ω n ( y ) .

Let ϵ 0 be the constant obtained in Lemma B.2 and choose β = ϵ 0 2 and q = 1 + ϵ 0 1 + ϵ 0 / 2 > 1 . Then (B.6) follows from (B.8) and Lemma (B.2). ∎

Diameter bound

Then we are ready to prove Proposition B.1 following [37, 36].

Proof of Proposition B.1

Recall that we have uniform high regularity estimate for ω t away from 𝑝 and each ω t is conical at 𝑝. Fix a small neighborhood 𝑉 of 𝑝. Without loss of generality, we may assume dist ( x , U ) 1 for any x V .

Step 1.  Let ρ : R [ 0 , ) be a cut-off function with ρ = 1 on ( , 1 / 2 ) and ρ = 0 on [ 1 , ) . For any x V { p } and r ( 0 , 1 ) , let u ( y ) = ρ ( d ω ( x , y ) / r ) . We are going to show

(B.9) ( B ω ( x , r ) | u | 2 q ω n ) 1 q C 2 B ω ( x , r ) | u | ω 2 ω n ,

where 𝑞 and C 2 are the constants in Proposition B.3. If p B ̄ ω ( x , r ) , then this follows from Proposition B.3 directly as B ω ( x , r ) U δ for some 𝛿 sufficiently small. In general, we can use a cut-off argument as follows. Since ω t is conical at 𝑝, for every ϵ > 0 sufficiently small and compact set F U { p } , there exists a cut-off function η ϵ : U [ 0 , 1 ] such that η ϵ 1 on 𝐹, η ϵ = 0 in a neighborhood of 𝑝 and moreover

(B.10) U | η ϵ | ω 2 ω n < ϵ .

Since η ϵ u has support on U { p } , we obtain

( B ω ( x , r ) | η ϵ u | 2 q ω n ) 1 q C 2 B ω ( x , r ) | ( η ϵ u ) | ω 2 ω n .

Since both 𝑢 and u are bounded, letting ϵ 0 and using (B.10), we get (B.9).

Step 2.  By (B.9), we get, for x V { p } and r ( 0 , 1 ) ,

( Vol ω ( B ω ( x , r / 2 ) ) ( r / 2 ) 2 q q 1 ) 1 q C Vol ω ( B ω ( x , r ) ) r 2 q q 1 .

Applying this to the radii r m = 2 m r and iterating the estimates, one obtains that there exists a constant κ > 0 such that

(B.11) Vol ω ( B ω ( x , r ) ) κ r 2 q q 1 .

Approximating 𝑝 by a sequence x i V { p } , one can show the same estimate holds for balls centered at 𝑝.

Step 3.  We can argue by contradiction to show that diam ( U , ω t ) has a uniform upper bound. Suppose not. Since 𝜔 has uniform estimate on U V , we can only find a geodesic contained in 𝑉 whose length can be arbitrary large. Then this would contradict (B.11) as we have uniform volume upper bound for ω t . ∎

Acknowledgements

The author expresses his gratitude to his advisor Song Sun for the constant support and valuable suggestions. He thanks Xin Fu for discussing the results in [29]. He also thanks the anonymous referee for careful reading of the paper and helpful suggestions. Part of this paper was completed during his visit to IASM at Zhejiang University, which he would like to thank for the hospitality.

References

[1] M. T. Anderson, Convergence and rigidity of manifolds under Ricci curvature bounds, Invent. Math. 102 (1990), no. 2, 429–445. 10.1007/BF01233434Search in Google Scholar

[2] C. Arezzo and C. Spotti, On cscK resolutions of conically singular cscK varieties, J. Funct. Anal. 271 (2016), no. 2, 474–494. 10.1016/j.jfa.2016.04.025Search in Google Scholar

[3] T. Aubin, Réduction du cas positif de l’équation de Monge–Ampère sur les variétés kählériennes compactes à la démonstration d’une inégalité, J. Funct. Anal. 57 (1984), no. 2, 143–153. 10.1016/0022-1236(84)90093-4Search in Google Scholar

[4] T. Aubin, Some nonlinear problems in Riemannian geometry, Springer Monogr. Math., Springer, Berlin 1998. 10.1007/978-3-662-13006-3Search in Google Scholar

[5] R. J. Berman, S. Boucksom, P. Eyssidieux, V. Guedj and A. Zeriahi, Kähler-Einstein metrics and the Kähler–Ricci flow on log Fano varieties, J. reine angew. Math. 751 (2019), 27–89. 10.1515/crelle-2016-0033Search in Google Scholar

[6] R. J. Berman and H. Guenancia, Kähler–Einstein metrics on stable varieties and log canonical pairs, Geom. Funct. Anal. 24 (2014), no. 6, 1683–1730. 10.1007/s00039-014-0301-8Search in Google Scholar

[7] S. Boucksom, Monge–Ampère equations on complex manifolds with boundary, Complex Monge–Ampère equations and geodesics in the space of Kähler metrics, Lecture Notes in Math. 2038, Springer, Heidelberg (2012), 257–282. 10.1007/978-3-642-23669-3_7Search in Google Scholar

[8] L. Caffarelli, J. J. Kohn, L. Nirenberg and J. Spruck, The Dirichlet problem for nonlinear second-order elliptic equations. II. Complex Monge–Ampère, and uniformly elliptic, equations, Comm. Pure Appl. Math. 38 (1985), no. 2, 209–252. 10.1002/cpa.3160380206Search in Google Scholar

[9] J. Cheeger and T. H. Colding, On the structure of spaces with Ricci curvature bounded below. I, J. Differential Geom. 46 (1997), no. 3, 406–480. 10.4310/jdg/1214459974Search in Google Scholar

[10] J. Cheeger, T. H. Colding and G. Tian, On the singularities of spaces with bounded Ricci curvature, Geom. Funct. Anal. 12 (2002), 873–914. 10.1007/PL00012649Search in Google Scholar

[11] J. Cheeger and G. Tian, On the cone structure at infinity of Ricci flat manifolds with Euclidean volume growth and quadratic curvature decay, Invent. Math. 118 (1994), no. 3, 493–571. 10.1007/BF01231543Search in Google Scholar

[12] G. Chen, Rate of asymptotic convergence near an isolated singularity of G 2 manifold, J. Geom. Anal. 28 (2018), no. 4, 3139–3170. 10.1007/s12220-017-9952-7Search in Google Scholar

[13] X. Chen, The space of Kähler metrics, J. Differential Geom. 56 (2000), no. 2, 189–234. 10.4310/jdg/1090347643Search in Google Scholar

[14] X. Chen and B. Wang, Space of Ricci flows (II)—Part A: Moduli of singular Calabi–Yau spaces, Forum Math. Sigma 5 (2017), Paper No. e32. 10.1017/fms.2017.28Search in Google Scholar

[15] S.-K. Chiu and G. Székelyhidi, Higher regularity for singular Kähler–Einstein metrics, Duke Math. J. 172 (2023), no. 18, 3521–3558. 10.1215/00127094-2022-0107Search in Google Scholar

[16] T. H. Colding, Ricci curvature and volume convergence, Ann. of Math. (2) 145 (1997), no. 3, 477–501. 10.2307/2951841Search in Google Scholar

[17] T. H. Colding and W. P. Minicozzi, II, On uniqueness of tangent cones for Einstein manifolds, Invent. Math. 196 (2014), no. 3, 515–588. 10.1007/s00222-013-0474-zSearch in Google Scholar

[18] T. H. Colding and A. Naber, Sharp Hölder continuity of tangent cones for spaces with a lower Ricci curvature bound and applications, Ann. of Math. (2) 176 (2012), no. 2, 1173–1229. 10.4007/annals.2012.176.2.10Search in Google Scholar

[19] T. Collins and G. Székelyhidi, Sasaki–Einstein metrics and K–stability, Geom. Topol. 23 (2019), no. 3, 1339–1413. 10.2140/gt.2019.23.1339Search in Google Scholar

[20] T. Darvas, Metric geometry of normal Kähler spaces, energy properness, and existence of canonical metrics, Int. Math. Res. Not. IMRN 2017 (2017), no. 22, 6752–6777. 10.1093/imrn/rnw203Search in Google Scholar

[21] V. Datar, X. Fu and J. Song, Kähler–Einstein metrics near an isolated log-canonical singularity, J. reine angew. Math. 797 (2023), 79–116. 10.1515/crelle-2022-0095Search in Google Scholar

[22] J.-P. Demailly, Complex analytic and differential geometry, Université de Grenoble, Grenoble 1997. Search in Google Scholar

[23] S. Donaldson and S. Sun, Gromov–Hausdorff limits of Kähler manifolds and algebraic geometry, Acta Math. 213 (2014), no. 1, 63–106. 10.1007/s11511-014-0116-3Search in Google Scholar

[24] S. Donaldson and S. Sun, Gromov–Hausdorff limits of Kähler manifolds and algebraic geometry, II, J. Differential Geom. 107 (2017), no. 2, 327–371. 10.4310/jdg/1506650422Search in Google Scholar

[25] S. K. Donaldson, A note on the 𝛼-invariant of the Mukai–Umemura 3-fold, preprint (2007), https://arxiv.org/abs/0711.4357. Search in Google Scholar

[26] S. K. Donaldson, Kähler geometry on toric manifolds, and some other manifolds with large symmetry, Handbook of geometric analysis. No. 1, Adv. Lect. Math. (ALM) 7, International Press, Somerville (2008), 29–75. Search in Google Scholar

[27] S. K. Donaldson, Stability, birational transformations and the Kahler–Einstein problem, Surveys in differential geometry. Vol. XVII, Surv. Differ. Geom. 17, International Press, Boston (2012), 203–228. 10.4310/SDG.2012.v17.n1.a5Search in Google Scholar

[28] P. Eyssidieux, V. Guedj and A. Zeriahi, Singular Kähler–Einstein metrics, J. Amer. Math. Soc. 22 (2009), no. 3, 607–639. 10.1090/S0894-0347-09-00629-8Search in Google Scholar

[29] X. Fu, Dirichlet problem for complex Monge–Ampère equation near an isolated KLT singularity, Commun. Contemp. Math. 25 (2023), no. 5, Article ID 2250019. 10.1142/S0219199722500195Search in Google Scholar

[30] X. Fu, Uniqueness of tangent cone of Kähler–Einstein metrics on singular varieties with crepant singularities, Math. Ann. 388 (2024), no. 3, 3229–3258. 10.1007/s00208-023-02602-0Search in Google Scholar

[31] H. Grauert and R. Remmert, Plurisubharmonische Funktionen in komplexen Räumen, Math. Z. 65 (1956), 175–194. 10.1007/BF01473877Search in Google Scholar

[32] V. Guedj, H. Guenancia and A. Zeriahi, Continuity of singular Kähler–Einstein potentials, Int. Math. Res. Not. IMRN 2023 (2023), no. 2, 1355–1377. 10.1093/imrn/rnab294Search in Google Scholar

[33] V. Guedj, H. Guenancia and A. Zeriahi, Strict positivity of Kähler–Einstein currents, Forum Math. Sigma 12 (2024), Paper No. e68. 10.1017/fms.2024.54Search in Google Scholar

[34] V. Guedj, B. Kolev and N. Yeganefar, Kähler–Einstein fillings, J. Lond. Math. Soc. (2) 88 (2013), no. 3, 737–760. 10.1112/jlms/jdt031Search in Google Scholar

[35] V. Guedj, A. Trusiani and S. Boucksom, Kähler–Einstein metrics with positive curvature near an isolated log terminal singularity, preprint (2023), https://arxiv.org/abs/2306.07900. Search in Google Scholar

[36] B. Guo, D. H. Phong, J. Song and J. Sturm, Sobolev inequalities on Kähler spaces, preprint (2023), https://arxiv.org/abs/2311.00221. Search in Google Scholar

[37] B. Guo, D. H. Phong, J. Song and J. Sturm, Diameter estimates in Kähler geometry, Comm. Pure Appl. Math. 77 (2024), no. 8, 3520–3556. 10.1002/cpa.22196Search in Google Scholar

[38] W. He and S. Sun, Frankel conjecture and Sasaki geometry, Adv. Math. 291 (2016), 912–960. 10.1016/j.aim.2015.11.053Search in Google Scholar

[39] H.-J. Hein and S. Sun, Calabi–Yau manifolds with isolated conical singularities, Publ. Math. Inst. Hautes Études Sci. 126 (2017), 73–130. 10.1007/s10240-017-0092-1Search in Google Scholar

[40] S. Ishii, Introduction to singularities, Springer, Tokyo 2018. 10.1007/978-4-431-56837-7Search in Google Scholar

[41] S. Kołodziej, The complex Monge–Ampère equation, Acta Math. 180 (1998), no. 1, 69–117. 10.1007/BF02392879Search in Google Scholar

[42] C. Li, Notes on weighted Kähler–Ricci solitons and application to Ricci-flat Kähler cone metrics, preprint (2021), https://arxiv.org/abs/2107.02088. Search in Google Scholar

[43] C. Li, 𝐺-uniform stability and Kähler–Einstein metrics on Fano varieties, Invent. Math. 227 (2022), no. 2, 661–744. 10.1007/s00222-021-01075-9Search in Google Scholar

[44] C. Li and Y. Liu, Kähler–Einstein metrics and volume minimization, Adv. Math. 341 (2019), 440–492. 10.1016/j.aim.2018.10.038Search in Google Scholar

[45] C. Li, X. Wang and C. Xu, Algebraicity of the metric tangent cones and equivariant K-stability, J. Amer. Math. Soc. 34 (2021), no. 4, 1175–1214. 10.1090/jams/974Search in Google Scholar

[46] C. Li and C. Xu, Stability of valuations: Higher rational rank, Peking Math. J. 1 (2018), no. 1, 1–79. 10.1007/s42543-018-0001-7Search in Google Scholar

[47] G. Liu, Compactification of certain Kähler manifolds with nonnegative Ricci curvature, Adv. Math. 382 (2021), Article ID 107652. 10.1016/j.aim.2021.107652Search in Google Scholar

[48] G. Liu and G. Székelyhidi, Gromov–Hausdorff limits of Kähler manifolds with Ricci curvature bounded below II, Comm. Pure Appl. Math. 74 (2021), no. 5, 909–931. 10.1002/cpa.21900Search in Google Scholar

[49] R. B. Lockhart and R. C. McOwen, Elliptic differential operators on noncompact manifolds, Ann. Sc. Norm. Super. Pisa Cl. Sci. (4) 12 (1985), no. 3, 409–447. Search in Google Scholar

[50] R. Schoen and S.-T. Yau, Lectures on differential geometry. Vol. 2, International Press, Cambridge 1994. Search in Google Scholar

[51] R. T. Seeley, Extension of C functions defined in a half space, Proc. Amer. Math. Soc. 15 (1964), 625–626. 10.1090/S0002-9939-1964-0165392-8Search in Google Scholar

[52] J. Song, Riemannian geometry of Kähler–Einstein currents, preprint (2014), https://arxiv.org/abs/1404.0445. Search in Google Scholar

[53] S. Sun and J. Zhang, No semistability at infinity for Calabi–Yau metrics asymptotic to cones, Invent. Math. 233 (2023), no. 1, 461–494. 10.1007/s00222-023-01187-4Search in Google Scholar

[54] G. Székelyhidi, An introduction to extremal Kähler metrics, Grad. Stud. Math. 152, American Mathematical Society, Providence 2014. Search in Google Scholar

[55] G. Tian, Kähler–Einstein metrics with positive scalar curvature, Invent. Math. 130 (1997), no. 1, 1–37. 10.1007/s002220050176Search in Google Scholar

[56] G. Tian, Existence of Einstein metrics on Fano manifolds, Metric and differential geometry, Progr. Math. 297, Birkhäuser/Springer, Basel (2012), 119–159. 10.1007/978-3-0348-0257-4_5Search in Google Scholar

[57] C. van Coevering, Examples of asymptotically conical Ricci-flat Kähler manifolds, Math. Z. 267 (2011), no. 1–2, 465–496. 10.1007/s00209-009-0631-7Search in Google Scholar

[58] Z. Zhuang, Stability of klt singularities, Handbook of geometry and topology of singularities VII, Springer, Cham (2025), 501–551. 10.1007/978-3-031-68711-2_10Search in Google Scholar

Received: 2024-09-03
Revised: 2025-06-03
Published Online: 2025-07-11
Published in Print: 2025-09-01

© 2025 the author(s), published by De Gruyter

This work is licensed under the Creative Commons Attribution 4.0 International License.

Downloaded on 21.9.2025 from https://www.degruyterbrill.com/document/doi/10.1515/crelle-2025-0044/html
Scroll to top button