Hostname: page-component-cb9f654ff-w5vf4 Total loading time: 0 Render date: 2025-08-15T13:30:25.580Z Has data issue: false hasContentIssue false

Kähler–Einstein metrics with positive curvature near an isolated log terminal singularity

Published online by Cambridge University Press:  13 August 2025

Vincent Guedj
Affiliation:
Institut de Mathématiques de Toulouse et Institut Universitaire de France, Université de Toulouse, 118 route de Narbonne, 31400 Toulouse, France vincent.guedj@math.univ-toulouse.fr
Antonio Trusiani
Affiliation:
Department of Mathematical Sciences, Chalmers University of Technology, Chalmers tvärgata 3, 41258 Göteborg, Sweden trusiani@chalmers.se
Rights & Permissions [Opens in a new window]

Abstract

We analyze the existence of Kähler–Einstein metrics of positive curvature in the neighborhood of a germ of a log terminal singularity (X,p). This boils down to solving a Dirichlet problem for certain complex Monge–Ampère equations. We establish a Moser–Trudinger inequality $(MT)_{\gamma}$ in subcritical regimes $\gamma<\gamma_{\rm crit}(X,p)$ and show the existence of smooth solutions in those cases. We show that the expected critical exponent $\tilde{\gamma}_{\rm crit}(X,p)=(({n+1})/{n}) \widehat{\mathrm{vol}}(X,p)^{1/n}$ can be expressed in terms of the normalized volume, an important algebraic invariant of the singularity.

Information

Type
Research Article
Creative Commons
Creative Common License - CCCreative Common License - BYCreative Common License - NC
This is an Open Access article, distributed under the terms of the Creative Commons Attribution-NonCommercial licence (https://creativecommons.org/licenses/by-nc/4.0/), which permits non-commercial re-use, distribution, and reproduction in any medium, provided the original article is properly cited. Written permission must be obtained prior to any commercial use.
Copyright
© The Author(s), 2025.

1. Introduction

Let (X,p) be a germ of an isolated singularity. We analyze the existence of local Kähler–Einstein metrics of positive curvature in a neighborhood of p. It follows from [Reference Berman, Boucksom, Eyssidieux, Guedj and ZeriahiBBEGZ19, Proposition 3.8] that the singularity has to be log terminal, a relatively mild type of singularity that plays a central role in birational geometry. We refer the reader to Definition 2.8 for a precise formulation and simply indicate here that a prototypical example is the vertex of the affine cone over a Fano manifold. Consider indeed

\[X=\{ z \in \mathbb C^{n+1},\ P(z)=0 \},\]

P a homogeneous polynomial of degree $d \in \mathbb N^*$ so that $H=\{ [z] \in \mathbb C\mathbb P^n,\ P(z)=0 \}$ is a smooth hypersurface of the complex projective space. Then (X,0) is log terminal if and only if H is Fano (which is equivalent here to $d <n+1$ ). Thus, log terminal singularities can be seen as a local analogue of Fano varieties.

Given a local embedding $(X,p)\hookrightarrow (\mathbb C^{N},0)$ , constructing such a local Kähler–Einstein metric boils down to solve a complex Monge–Ampère equation

\[(MA)_{\gamma,\phi,\Omega}\begin{cases}(dd^{c}\varphi)^{n}=\dfrac{e^{-\gamma\varphi}\,d\mu_{p}}{\int_{\Omega}e^{-\gamma \varphi} \,d\mu_{p}},\\\varphi_{\vert\partial\Omega}=\phi,\end{cases}\]

where $\Omega$ is a smooth neighborhood of p, $\phi$ is a smooth boundary data, $\mu_{p}$ is an adapted volume form (see Definition 2.9), and $\gamma>0$ is a parameter. We seek for a solution $\varphi \in {\mathcal C}^{\infty}(\Omega\setminus\{p\}) \cap {\mathcal{C}}^0(\overline{\Omega})$ which is strictly plurisubharmonic in $\Omega \setminus \{ p \}$ , so that $\omega_{\rm KE}:=dd^c \varphi$ is a Kähler form in $\Omega \setminus \{p\}$ satisfying the Einstein equation

\[\mathrm{Ric}(\omega_{\rm KE})=\gamma \omega_{\rm KE}.\]

An important motivation comes from the global study of positively curved Kähler–Einstein metrics $\omega_{\rm KE}$ on $\mathbb Q$ -Fano varieties. Such canonical singular metrics have been constructed in [Reference Berman, Boucksom, Eyssidieux, Guedj and ZeriahiBBEGZ19] and further studied in [Reference Berman, Boucksom and JonssonBBJ21, Reference Li, Tian and WangLTW21, Reference LiLi22], extending the resolution of the Yau–Tian–Donaldson conjecture [Reference Chen, Donaldson and SunCDS15] to this singular context. Despite recent important progress [Reference Hein and SunHS17, Reference DruelDru18, Reference Höring and PeternellHP19, Reference Bakker, Guenancia and LehnBGL22], the geometry of these singular metrics remains mysterious and one needs to better understand the asymptotic behavior of $\omega_{\rm KE}$ near the singularities.

We restrict the metric $\omega_{\rm KE}$ to a neighborhood of p and wish to analyze the behavior of its local potentials $\omega_{\rm KE}=dd^c \varphi_{\rm KE}$ near p. The latter solve a Monge–Ampère equation $(MA)_{\gamma,\phi,\Omega}$ , as can be seen by locally trivializing a representative of the first Chern class (after an appropriate rescaling). The boundary data are thus given by the solution $\varphi_{\rm KE}=\phi$ itself.

Studying the family of equations $(MA)_{\gamma,\phi,\Omega}$ we will give evidence that:

  • the possibility of solving $(MA)_{\gamma,\phi,\Omega}$ should be independent of $\Omega$ and $\phi$ ;

  • the largest exponent ${\gamma}_{\rm crit}(X,p)$ for which we can solve $(MA)_{\gamma,\phi,\Omega}$ should only depend on the algebraic nature of the log terminal singularity.

Following earlier works dealing with the case of compact Kähler varieties or the local smooth setting [Reference Berman, Boucksom, Guedj and ZeriahiBBGZ13, Reference Guedj, Kolev and YeganefarGKY13, Reference Berman and BerndtssonBB22, Reference Berman, Boucksom, Eyssidieux, Guedj and ZeriahiBBEGZ19], we develop a variational approach to solve these equations. A crucial role is played by

\[E_{\phi}(\varphi)= \frac{1}{n+1}\sum_{j=0}^{n}\int_{\Omega}(\varphi-\phi_{0})(dd^{c}\varphi)^{j}\wedge(dd^{c}\phi_{0})^{n-j},\]

the Monge–Ampère energy of $\varphi$ relative to a plurisubharmonic extension $\phi_0$ of $\phi$ . This energy is a primitive of the Monge–Ampère operator and a building block of the functional $F_{\gamma}$ whose Euler–Lagrange equation is $(MA)_{\gamma,\phi,\Omega}$ ,

\[\varphi \in {\mathcal T}_{\phi}(\Omega) \mapsto F_{\gamma}(\varphi)= E_{\phi}(\varphi)+\frac{1}{\gamma} \log \int_{\Omega} e^{-\gamma \varphi} \,d\mu_p \in \mathbb R.\]

Here ${\mathcal T}_{\phi}(\Omega)$ denotes the set of all plurisubharmonic functions $\varphi$ in $\Omega$ which are continuous on $\overline{\Omega}$ and such that $\varphi_{|\partial \Omega}=\phi$ .

In order to solve $(MA)_{\gamma,\phi,\Omega}$ one can try and extremize $F_{\gamma}$ by showing that it is a proper functional. Our first main result in this direction (Theorem 5.1) is the following Moser–Trudinger-type inequality.

Theorem A. For any $0<\gamma<(({n+1})/{n}) \alpha (X,\mu_p)$ , there exists $C_{\gamma}>0$ such that

\begin{equation}\tag{$MT_{\gamma}$}\bigg(\int_{\Omega} e^{-\gamma \varphi} \,d\mu_p \bigg)^{\!\!{1}/{\gamma}}\leq C_{\gamma} \exp (-E_{\phi}(\varphi)),\end{equation}

for all $\varphi \in {\mathcal T}_{\phi}(\Omega)$ .

The alpha invariant of the singularity (X,p) is defined by

\[\alpha (X,\mu_p):=\sup \bigg\{\alpha>0,\ \sup_{\varphi \in {\mathcal F}_1(\Omega)}\int_{\Omega} e^{-\alpha \varphi} \,d\mu_p <+\infty\bigg\},\]

where ${\mathcal F}_1(\Omega)$ denotes the set of plurisubharmonic functions $\varphi$ with $\phi$ -boundary values, whose Monge–Ampère mass is bounded by $\int_{\Omega} (dd^c \varphi)^n \leq 1$ .

When (X,p) is smooth, Theorem A has been obtained independently in [Reference Berman and BerndtssonBB22, Reference Guedj, Kolev and YeganefarGKY13] with $\alpha(X,\mu_p)=n$ (the normalizations and methods are quite different in these two works, but they eventually produce the same critical exponent).

We introduce

While Theorem A provides a lower bound for $\gamma_{\rm crit}(X,p)$ , we provide an upper bound in Theorem 4.5, which yields

\[\frac{n+1}{n} \alpha (X,\mu_p) \leq \gamma_{\rm crit}(X,p) \leq \frac{n+1}{n}\widehat{\rm vol}(X,p)^{1/n},\]

where $\widehat{\rm vol}(X,p)$ denotes the normalized volume of the singularity (X,p). This is an algebraic invariant of the singularity at p introduced by Chi Li in [Reference LiLi18], which has recently played a key role in the algebraic understanding of the moduli space of K-stable Fano varieties (see [Reference BlumBlu18, Reference LiuLiu18, Reference Li, Wang and XuLWX21, Reference Liu, Xu and ZhangLXZ22] and the references therein); we refer to Definition 2.15 for a precise definition.

When p is smooth then $\alpha(X,\mu_p)= \widehat{\rm vol}(X,p)^{1/n}=n$ by [Reference Åhag, Cegrell, Kołodziej, Pham and ZeriahiACKPZ09, Reference DemaillyDem09]. It is tempting to conjecture that the equality $\alpha(X,\mu_p)= \widehat{\rm vol}(X,p)^{1/n}$ always holds. We establish in § 5 the following partial bounds on $\alpha(X,\mu_p)$ .

Theorem B. The following inequalities hold:

\[\frac{n}{{\rm mult}(X,p)^{1-1/n}} \frac{{\rm lct}(X,p)}{1+{\rm lct}(X,p)}\leq \alpha(X,\mu_p) \leq \widehat{\rm vol}(X,p)^{1/n}.\]

Moreover, $\alpha(X,\mu_p)=\widehat{\rm vol}(X,p)^{1/n} $ if (X,p) is an admissible singularity.

Here ${\rm mult}(X,p)$ denotes the algebraic multiplicity of (X,p), while ${\rm lct}(X,p)$ is its log canonical threshold (see Definition 2.12). Having $\alpha(X,\mu_p)$ bounded from below is quite involved; we show that $\alpha(X,\mu_p)=\widehat{\rm vol}(X,p)^{1/n}$ when $n=2$ , but our lower-bound is not sharp when $n \geq 3$ unless (X,p) is an admissible singularity, a notion introduced in [Reference Li, Tian and WangLTW21]. The vertex of the affine cone over a smooth Fano manifold is an example of admissible singularity (see § 5).

Using analytic Green functions and Demailly’s comparison theorem, we provide in Propositions 5.6 and 5.8 evidence for the equality $\alpha(X,\mu_p) =\widehat{\rm vol}(X,p)^{1/n}$ . Appendix A uses an algebraic approach based on [Reference Boucksom, de Fernex and FavreBdFF12], to establish a stronger result than Proposition 5.8.

We note in Lemma 3.13 that if (MT γ ) holds, then $F_{\gamma}$ is coercive (a strong quantitative version of properness). When $\gamma <{\gamma}_{\rm crit}(X,p)$ , we then further show the existence of smooth solutions to $(MA)_{\gamma,\phi,\Omega}$ .

Theorem C. If $\gamma <\gamma_{\rm crit}(X,p)$ , then there exists a plurisubharmonic function $\varphi \in {\mathcal{C}}^{\infty}({\Omega} \setminus \{p\})$ which is continuous in $\overline{\Omega}$ with $\varphi_{|\partial \Omega} =\phi$ , and such that

\[(dd^{c}\varphi)^{n}=\frac{e^{-\gamma\varphi} \, d\mu_{p}}{\int_{\Omega}e^{-\gamma \varphi} \,d\mu_{p}}\quad\text{in}\ \Omega.\]

We expect the solution to be unique, at least when $\Omega$ is a generic Stein neighborhood of p. We refer the reader to [Reference Guedj, Kolev and YeganefarGKY13, Reference Berman and BerndtssonBB22] for partial results in this direction when p is a smooth point.

2. Preliminaries

2.1 Analysis on singular spaces

Let X be a reduced complex analytic space of pure dimension $n \ge 1$ . We let $X_{{reg}}$ denote the complex manifold of regular points of X and $ X_{\mathrm{sing}} := X \setminus X_{{reg}} $ be the set of singular points; this is an analytic subset of X of complex codimension ${\ge}1$ . We always assume in this article that:

  • $X_{\mathrm{sing}}=\{p\}$ consists of a single isolated point;

  • $X_{{reg}}$ is locally irreducible at p;

  • U is a fixed neighborhood of p and $j: U \hookrightarrow \mathbb C^N$ is a local embedding onto an analytic subset of $\mathbb C^N$ for some $N \ge 1$ .

As we are interested in the asymptotic behavior of Kähler–Einstein potentials near the singular point p, we shall identify X with j(U) in the following.

2.1.1 Plurisubharmonic functions

Using the local embedding j, it is possible to define the spaces of smooth forms on X as restriction of smooth forms of $\mathbb C^N$ . The notion of currents on X is defined by duality; the operators $\partial $ and $\bar{\partial}$ , d, $d^c$ and $dd^c$ are also well defined by duality (see [Reference DemaillyDem85] for more details).

Here $d=\partial+\overline{\partial}$ and $d^c=({1}/{4i\pi})(\partial-\overline{\partial})$ are real operators and $dd^c =({i}/{2\pi})\partial\overline{\partial}$ . With this normalization the function $z \in \mathbb C^n \mapsto \rho_{FS}(z)=\log[1+|z|^2] \in \mathbb R$ is smooth and plurisubharmonic (psh for short) in $\mathbb C^n$ , with

\[\int_{\mathbb C^n} (dd^c \rho_{FS})^n=1.\]

Definition 2.1. We say that a function $u : X \longrightarrow \mathbb R \cup \{-\infty\}$ is psh on X if it is the restriction of a psh function of $\mathbb C^N$ .

We let PSH(X) denote the set of all psh functions on X that are not identically $-\infty$ .

Recall that u is called weakly psh on X if it is locally bounded from above on X and its restriction to $X_{{reg}}$ is psh. One can extend it to X by $u^* (p) :=\limsup_{X_{{reg}} \ni y \to p}u (y)$ . Since X is locally irreducible, it follows from the work of Fornæss and Narasimhan [Reference Fornæss and NarasimhanFN80] that u is weakly psh if and only if $u^*$ is psh (see [Reference DemaillyDem85, Corollary 1.11]).

If $u \in PSH(X)$ , then u is upper semi-continuous on X and locally integrable with respect to the volume form

\[dV_X:=\omega_{\rm eucl}^n \wedge [X].\]

Here [X] denotes the current of integration along X and $\omega_{\rm eucl}:=\sum_{j=1}^N i \, dz_j\wedge d\overline{z_j}$ is the euclidean Kähler form. In particular, $dd^c u$ is a well-defined current of bidegree (1,1) which is positive.

2.1.2 Pseudoconvex domains and boundary data

Following [Reference Fornæss and NarasimhanFN80] we say that X is Stein if it admits a ${\mathcal C}^2$ -smooth strongly psh exhaustion.

Definition 2.2. A domain $\Omega \Subset X$ is strongly pseudoconvex if it admits a negative ${\mathcal{C}}^2$ -smooth strongly psh exhaustion, i.e. a function $\rho$ strongly psh in a neighborhood $\Omega'$ of $\overline{\Omega}$ such that $\Omega := \{ x \in \Omega'; \rho (x)< 0\}$ , $d\rho \neq 0$ on $\partial \Omega$ , and for any $c < 0$ ,

\[\Omega_c := \{x \in \Omega'; \rho (x) < c\} \Subset \Omega\]

is relatively compact.

We are interested in solving a Dirichlet problem for some complex Monge–Ampère equations in a bounded strongly pseudoconvex domain $\Omega=\{\rho<0\}$ , with given boundary data $\phi \in{\mathcal C}^{\infty}(\partial \Omega)$ .

Definition 2.3. Given $\phi \in {\mathcal C}^{\infty}(\partial \Omega)$ , we fix $\phi_0$ a psh function in $\Omega$ which is ${\mathcal C}^{\infty}$ -smooth near $\overline{\Omega}$ and such that $\phi_0{|\partial \Omega}=\phi$ .

Such an extension can be obtained as follows: we pick $\tilde{\phi}$ an arbitrary ${\mathcal{C}}^2$ -smooth extension to $\overline{\Omega}$ , and then consider $\phi_0:=\tilde{\phi}+A \rho$ , for A so large that $\phi_0$ is ${\mathcal C}^2$ -smooth and psh in $\overline{\Omega}$ . All quantities introduced in the remainder of the paper are essentially independent of the particular choice of the extension.

2.1.3 Monge–Ampère operators

The complex Monge–Ampère operator $(dd^c \cdot)^n$ acts on a smooth psh functions $\varphi$ . When $X=\mathbb C^n$ , it boils down to

\[(dd^c \varphi)^n=c_n \det \bigg(\frac{\partial^2 \varphi}{\partial z_j \overline{\partial} z_k}\bigg) \omega_{\rm eucl}^n,\]

where $c_n>0$ is a normalizing constant.

2.1.4 Bounded functions

Following [Reference Bedford and TaylorBT82] this operator can be extended to the class $PSH(X) \cap L^{\infty}_{\rm loc}$ by using approximation by smooth psh functions: given $\varphi \in PSH(X) \cap L^{\infty}_{\rm loc}$ , there exists a unique positive Radon measure $\mu_{\varphi}$ on X such that for any sequence $(\varphi_j)$ of smooth psh functions decreasing to $\varphi$ , one has

\[\mu_{\varphi}=\lim (dd^c \varphi_j)^n,\]

where the limit holds in the weak sense. One then sets $(dd^c \varphi)^n:=\mu_{\varphi}$ .

Definition 2.4. We set

\[\mathcal{T}^{\infty}_{\phi}(\Omega):=\{\varphi\in SPSH(\Omega)\cap {\mathcal C}^{\infty}(\overline{\Omega}): \varphi_{|\partial \Omega}=\phi\},\]

where $SPSH(\Omega)$ is the set of strictly psh functions, and

\[\mathcal{T}_{\phi}(\Omega):=\bigg\{\varphi\in PSH(\Omega)\cap {\mathcal C}^0(\overline{\Omega}):\varphi_{|\partial \Omega}=\phi,\ \int_{\Omega}(dd^{c}\varphi)^{n}<+\infty\bigg\},\]

This latter class has been introduced by Cegrell in [Reference CegrellCeg98]; it can be used as a psh version of test functions (in the sense of distributions), as well as a building block for finite-energy classes of mildly unbounded functions.

Lemma 2.5. Any $\varphi \in \mathcal{T}_{\phi}(\Omega)$ is a quasi-decreasing limit of functions in $\mathcal{T}^{\infty}_{\phi}(\Omega)$ .

Proof. Fix a local embedding $X \hookrightarrow \mathbb C^N$ . A function $\varphi \in \mathcal{T}_{\phi}(\Omega)$ is the restriction of an ambient continuous psh function $\psi$ . We use standard convolution in $\mathbb C^N$ to find a sequence of smooth strictly psh functions $\psi_j$ decreasing to $\psi$ . Consider $\varphi_j={\psi_j}_{|X}-\varepsilon_j$ , where $0 <\varepsilon_j$ goes to zero so that $\varphi_j<\phi_0$ near $\partial \Omega$ (the functions ${\psi_j}_{|X}$ uniformly converge to $\varphi$ by continuity). Set $\tilde{\varphi_j}:=\tilde{\max}(\varphi_j,A_j \rho+\phi_0)$ , where $\tilde{\max}$ is a regularized maximum, then $\varphi_j \in \mathcal{T}^{\infty}_{\phi}(\Omega)$ converges to $\varphi$ as $A_j \rightarrow +\infty$ .

2.1.5 Mildly unbounded functions

The complex Monge–Ampère operator can be defined for mildly unbounded psh functions. We refer the reader to [Reference CegrellCeg04, Reference BłockiBlo06] for the case of smooth domains in $\mathbb C^n$ ; their analysis easily extends to our context.

Definition 2.6. We let $\mathcal{F}(\Omega)$ denote the set of all functions $\varphi\in PSH(\Omega)$ which are decreasing limit of a sequence of functions $\varphi_{j}\in \mathcal{T}_{\phi}(\Omega)$ such that

\[\sup_{j}\int_{\Omega}(dd^{c}\varphi_{j})^{n}<+\infty.\]

The operator $(dd^c \cdot )^n$ is well defined on $ \mathcal{F}(\Omega)$ , continuous along monotonic sequences, and yields Radon measures $(dd^c \varphi)^n$ which have finite mass in $\Omega$ . We endow $ \mathcal{F}(\Omega)$ with the $L^1$ -topology. Let us stress that the operator $\varphi \mapsto(dd^c \varphi)^n$ is not continuous for the $L^1$ -topology, but the class $ \mathcal{F}(\Omega)$ enjoys the following useful compactness property.

Proposition 2.7. The set $\mathcal{F}_1(\Omega)=\{\varphi\in\mathcal{F}(\Omega);\int_{\Omega}(dd^{c}\varphi)^{n}\leq 1\}$ is compact.

This is shown in [Reference ZeriahiZer09, Observation A.3] for smooth domains, and the same proof applies in our mildly singular context. Let us stress that the Monge–Ampère operator cannot be defined for all psh functions: there is, for example, no reasonable way to make sense of $(dd^c \log |z_1|)^n$ . A consequence of Proposition 2.7 is that one cannot approximate such a function by a decreasing sequence of psh functions with prescribed boundary values and uniformly bounded Monge–Ampère masses.

2.2 Adapted volume form

2.2.1 Log terminal singularities

Let Y be a connected normal complex variety such that $K_Y$ is $\mathbb Q$ -Cartier near $p \in Y$ . One can consider the $dd^c$ -cohomology class of $-K_Y$ , denoted by $c_1(Y)$ .

Given a log-resolution $\pi: \tilde{Y} \to Y$ of (Y,p), there exists a unique $\mathbb Q$ -divisor $\sum_i a_i E_i$ whose push-forward to Y is 0 and with

\[K_{\tilde{Y}}=\pi^*(K_Y)+\sum_i a_i E_i.\]

Definition 2.8. The coefficient $a_i\in\mathbb Q$ is the discrepancy of Y along $E_j$ . One says that p is a log terminal singularity if $a_j>-1$ for all j.

It is classical that this condition is independent of the choice of resolution. In the remainder of this article we assume that:

  • the singularity (X,0) is log terminal;

  • $Y=\Omega$ is a strongly pseudoconvex neighborhood of $0=p \in X$ ;

  • the canonical bundle $K_{\Omega}$ is $\mathbb Q$ -Cartier and $rK_{\Omega}=0$ for some $r\in\mathbb N$ .

Definition 2.9 [Reference Eyssidieux, Guedj and ZeriahiEGZ09, Definition 6.5]. Fix $\sigma$ a nowhere-vanishing holomorphic section of $rK_{\Omega}$ , and h a smooth hermitian metric of $ K_{\Omega}$ , then

\[\mu_{p}=\lambda \frac{(c_{n}\sigma\wedge \bar{\sigma})^{1/r}}{|\sigma|^{2/r}_{h^r}}\]

is an adapted measure, where $\lambda >0$ is a positive normalizing constant.

Observe that $\mu_p$ is independent of the choice of $\sigma$ , and

\[dd^c \log \mu_{p}=-\Theta_h(K_{\Omega})\]

is the curvature of h, as follows from the Poincaré–Lelong formula.

The measure $\mu_p$ has finite mass by [Reference Eyssidieux, Guedj and ZeriahiEGZ09, Lemma 6.4]: let $\pi: \tilde{\Omega}\to \Omega$ be a resolution of $(\Omega,0)$ , then

\[\pi^{*}\mu_{p}=\lambda \prod_{j=1}^{M}\lvert s_{E_{j}} \rvert^{2a_{j}} \, dV_{\tilde{\Omega}},\]

where $dV_{\tilde{\Omega}}$ is a smooth volume form on $\tilde{\Omega}$ , $E_{1},\ldots,E_{M}$ are exceptional divisors, $s_{E_j}$ are holomorphic sections such that $E_j=(s_{E_j}=0)$ , and

\[rK_{\tilde{\Omega}}=\pi^{*}(rK_{\Omega})+r\sum_{j=1}^{M}a_{j}E_{j}=r\sum_{j=1}^{M}a_{j}E_{j}.\]

Thus $\tilde{f}=\prod_{j=1}^{M}\lvert s_{E_{j}} \rvert^{2a_{j}}$ belongs to $L^s(dV_{\tilde{\Omega}})$ for some $s>1$ , as p is log terminal.

Definition 2.10. We choose $\lambda=\lambda_{\Omega}$ so that $\mu_p$ is a probability measure in $\Omega$ .

The results to follow are independent of this (convenient) normalization.

2.2.2 Ricci curvature

Let $\omega$ be a positive closed current of bidegree (1,1) in $\Omega$ with bounded local potentials. Its top power $\omega^n$ is well defined as explained in § 2.1.3. If $\omega^n$ is absolutely continuous with respect to $dV_X$ , then we set

\[{\rm Ric}(\omega):=-dd^c \log \omega^n.\]

Definition 2.11. We say that $\omega$ is a Kähler–Einstein metric if it satisfies

\[{\rm Ric}(\omega)=\gamma \omega\]

for some $\gamma \in \mathbb R$ .

In this article, we are mainly interested in the case when $\gamma >0$ . We choose the hermitian metric $h \equiv 1$ for $K_{\Omega}$ , so that $\Theta_h=0$ . Since

\[{\rm Ric}(\omega)={\rm Ric}(\mu_p)-dd^c \log (\omega^n/\mu_p),\]

the above Kähler–Einstein equation is equivalent, writing $\omega=dd^c \varphi$ , to

\[(dd^c \varphi)^n=e^{-\gamma \varphi} e^{w} \mu_p,\]

where w is a pluriharmonic function in $\Omega$ . Changing $\varphi$ in $\varphi-w/\gamma$ and then $\varphi$ in $t \varphi$ (observe that ${\rm Ric}(t \omega)={\rm Ric}(\omega)$ for any $t>0$ ), we can normalize $\omega$ by $\int_{\Omega} \omega^n=1$ and reduce to

\[(dd^c \varphi)^n= \frac{e^{-\gamma \varphi} \mu_p}{\int_{\Omega} e^{-\gamma \varphi} \mu_p}.\]

Seeking for a Kähler–Einstein metric thus leads one to solve $(MA)_{\gamma,\phi,\Omega}$ .

Conversely solving $(MA)_{\gamma,\phi,\Omega}$ will produce a Kähler–Einstein metric $\omega=dd^c\varphi$ , if we can establish enough regularity of the solution $\varphi$ .

2.2.3 Log canonical threshold

We consider the density $f=\mu_p/dV_X$ . It is related to the density $\tilde{f}$ in a resolution by

\[\pi^* \mu_p=f \circ \pi \cdot \pi^* \, dV_X=\tilde{f} \, dV_{\tilde{\Omega}}.\]

An analytic expression for f is obtained as follows. Recall that $dV_X=\omega_{\rm eucl}^n \wedge[X]$ , where $\omega_{\rm eucl}$ denotes the euclidean Kähler form on $\mathbb C^{N}$ . Set $dz_{I}=dz_{i_{1}}\wedge \cdots \wedge dz_{i_{n}}, $ where $1 \leq i_1< \cdots <i_n \leq N$ . There exists germs of holomorphic functions $f_{I}\in \mathcal{O}_{\Omega,0}$ such that $(dz_{I})^{r}=f_{I}\sigma $ since $\sigma$ is a local generator of $rK_{X}$ . In particular, the volume form $dV_{X}:=\omega_{\rm eucl}^{n}\wedge [\Omega]$ is comparable to $(\sum_{I}\lvert f_{I}\rvert^{{2}/{r}})\mu_{p},$ i.e.

\[\mu_{p}=f \, dV_{X},\quad \text{with}\ f\sim\bigg(\sum_{I}\lvert f_{I}\rvert^{{2}/{r}}\bigg)^{-1}.\]

The germs of holomorphic functions $f_{I}$ generate an ideal $\mathcal{I}_{p}^{r}$ , where $\mathcal{I}^{r}$ is an ideal sheaf associated to the singularities of (X,p). In particular,

\[\pi^{-1}\mathcal{I}^{r}\cdot \mathcal{O}_{\tilde{\Omega}}=\mathcal{O}_{\tilde{\Omega}}\bigg(-r\sum_{j=1}^{M}b_{j}E_{j}\bigg)\]

for coefficients $b_{j}\in\mathbb N$ such that $ f\circ \pi\sim \prod_{j=1}^{M}\lvert s_{E_{j}}\rvert^{-2b_{j}}$ .

Definition 2.12. The log canonical threshold of (X,p) is given by

\[\mathrm{lct}(X,p):=\inf_{j\in{1,\ldots,M}} \frac{a_{j}+1}{b_{j}}.\]

We let the reader check that the definition is independent of the choice of resolution, and that $\mathrm{lct}(X,\mathcal{I})\in (0,n]$ . One can equivalently use the following point of view: if $\mathcal{I}$ is a general ideal sheaf,

(2.1) \begin{equation}\mathrm{lct}(X,\mathcal{I}):=\inf_{E/X}\frac{A_{X}(E)}{\mathrm{ord}_{E}(\mathcal{I})}\end{equation}

where $A_{X}(E):=1+\mathrm{ord}_{E}(K_{Y/X}) $ is the log-discrepancy of E, and the infimum is over all prime divisors E on resolutions Y of X. When $\mathcal{I}$ is supported at p we can restrict in (2.1) to consider prime divisors centered at p.

Example 2.13. The ordinary double point (ODP) $X=\{z \in \mathbb C^{n+1},\ \sum_{j=0}^n z_j^2=0\} $ is the simplest isolated log terminal singularity which is not a quotient singularity when $n \geq 3$ (when $n=2$ , log terminal singularities are precisely the singularities of the form $X=\mathbb C^2/G$ , $G \subset GL(2,\mathbb C)$ a finite subgroup).

In this case $\mathcal{I}^2=(z_1^2,\ldots,z_n^2)$ . Indeed the n-forms

\[\sigma_j:=\frac{(dz_0\wedge \cdots\wedge \widehat{dz_j}\wedge \cdots \wedge dz_n)^2}{z_j^2}=-\frac{(dz_0\wedge \cdots \wedge \widehat{dz_j}\wedge \cdots \wedge dz_n)^2}{\sum_{k\neq j} z_k^2},\]

defined on $U_j:=\{z_j\neq 0\}$ , glue together to give a local generator $\sigma$ of $2 K_X$ (note that $\sum_{j=0}^n z_j \, dz_j=0$ ). In particular, $|f_I|^{2/r}=|z_j|^2$ where $j=[0,n] \setminus I$ , $r=2$ and

\[\mu_{p} \sim \frac{1}{\sum_{j=0}^n \lvert z_j \rvert^2} \, dV_X.\]

If $\pi: \rm Bl_0 \mathbb C^{n+1}\to \mathbb C^{n+1}$ denotes the blow-up at 0, E the exceptional divisor, and F the restriction of E to Y, the strict transform of X, we obtain

\[\pi^{-1}\mathcal{I}^2\cdot \mathcal{O}_Y=\mathcal{O}_Y(-2F)\quad \text{and}\quad\pi^* \mu_p= \lvert s_F \rvert^{2(n-2)} \, dV_Y\]

for a smooth volume form $dV_Y$ . Thus, ${\rm lct}(X,p)={\rm lct}(X,\mathcal{I})=n-1$ .

We will need the following result which connects ${\rm lct}(X,p)$ and the integrability properties of the density $f=\mu_p/dV_X$ .

Lemma 2.14. The density $f=\mu_p/dV_X$ belongs to $L^r(dV_X)$ for $r<1+{\rm lct}(X,p)$ .

Proof. Let $\pi:\tilde{\Omega} \rightarrow \Omega$ be a resolution of the singularity. Recall that

\[f \circ \pi \sim \prod_{j=1}^{M}\lvert s_{E_{j}} \rvert^{-2b_{j}}\quad \text{and}\quad\tilde{f}=\prod_{j=1}^{M}\lvert s_{E_{j}} \rvert^{2a_{j}},\quad \text{hence}\quad\pi^* dV_X \sim \prod_{j=1}^{M}\lvert s_{E_{j}} \rvert^{2(a_{j}+b_j)} \, dV_{\tilde{\Omega}}.\]

It follows that $\int_{{\Omega}} f^r \,dV_X \sim \int_{\tilde{\Omega}} \prod_{j=1}^{M}\lvert s_{E_{j}} \rvert^{2(a_{j}+b_j)-2rb_j} \,dV_{\tilde{\Omega}}<+\infty $ if and only if $r<({1+a_j+b_j})/{b_j}$ for all j, which yields the statement since ${\rm lct}(X,p)=\inf_j(({1+a_j})/{b_j})$ .

2.3 Normalized volume

The (Hilbert–Samuel) multiplicity of an ideal $\mathcal{I}$ supported at p is defined as

\[\mathrm{e}(X,\mathcal{I}):=\lim_{m\to +\infty}\frac{l(\mathcal{O}_{X,p}/\mathcal{I}^{m})}{m^{n}/n!}\]

where l denotes the length of an Artinian module.

Given a divisor E over X centered at p, the volume of E over $p\in X$ is

\[\mathrm{vol}_{X,p}(E):=\lim_{m\to +\infty}\frac{l(\mathcal{O}_{X,p}/\mathfrak{a}_{m}(E))}{m^{n}/n!}\]

where $\mathfrak{a}_{m}(E):=\{f\in \mathcal{O}_{X,p}: \mathrm{ord}_{E}(f)\geq m\}$ (see [Reference Ein, Lazarsfeld and SmithELS03]).

Definition 2.15 [Reference LiLi18]. The normalized volume of $p\in X$ is

\[\widehat{\mathrm{vol}}(X,p):=\inf_{E/X}\widehat{\mathrm{vol}}_{X,p}(E),\]

where the infimum runs over all prime divisors E over X centered at p, and

\[\widehat{\mathrm{vol}}_{X,p}(E):=A_{X}(E)^{n}\cdot\mathrm{vol}_{X,p}(E)\]

is the normalized volume of E over $(x\in X)$ .

We shall need the following important result.

Theorem 2.16 [Reference LiuLiu18, Theorem 27]. Let (X,p) be a log terminal singularity of complex dimension $\dim_{\mathbb C} X=n$ . Then

\[\widehat{\mathrm{vol}}(X,p)=\inf_{\mathcal{I}\, \mathrm{supported}\,\mathrm{at}\, p} \mathrm{lct}(X,\mathcal{I})^{n}\cdot \mathrm{e}(X,\mathcal{I}).\]

Observe that the quantity $\mathrm{lct}(X,\mathcal{I})^{n}\cdot \mathrm{e}(X,\mathcal{I})$ is invariant under rescaling $\mathcal{I}\to \mathcal{I}^{r}$ , $r\in\mathbb{N}$ . One can actually only consider coherent ideal sheaves supported at p. Indeed any ideal $\mathcal{I}$ supported at p is associated to a closed subscheme Z such that $\mathrm{Supp}\,Z=\{p\}$ (see [Reference HartshorneHar77, Corollary II.5.10]), while any ideal associated to a closed subscheme is coherent [Reference HartshorneHar77, Proposition II.5.9].

Example 2.17. Consider again $X=\{z \in \mathbb C^{n+1},\ \sum_{j=0}^n z_j^2=0\}$ . Recall that $\mathcal{I}^2=(z_1^2,\ldots,z_n^2)$ is the ideal sheaf associated to the adapted measure, and that the ideal $\mathcal{I}^2$ corresponds to 2F where F is the exceptional divisor in the blow-up at p. In particular, $A_X(F)=n-1$ .

We observe here that $e(X,\mathcal{I}^2)=2^{n+1}$ and $\widehat{{\rm vol}}_{X,p}(F)=2(n-1)^n$ since

\[l(\mathcal{O}_{X,p}/\mathcal{I}^{2m})=l(\mathcal{O}_{X,p}/\mathfrak{a}_{2m}(F))=2^{n+1}\frac{m^n}{n!}+O(m^{n-1}).\]

In [Reference LiLi18, Example 5.3] it is further shown that F is a minimizer for the normalized volume of $p\in X$ , i.e. that $\widehat{{\rm vol}}(X,p)=2(n-1)^n$ .

3. A variational approach

A variational approach for solving degenerate complex Monge–Ampère equations has been developed in [Reference Berman, Boucksom, Guedj and ZeriahiBBGZ13] in the context of compact Kähler manifolds. It notably applies to the construction of singular Kähler–Einstein metrics of non-positive curvature. This has been partially adapted to smooth pseudoconvex domains of $\mathbb C^n$ in [Reference Åhag, Cegrell and CzyzACC12].

The case of positive curvature is notoriously more difficult, as illustrated by the resolution of the Yau–Tian–Donaldson conjecture by Chen, Donaldson and Sun [Reference Chen, Donaldson and SunCDS15]. It has been treated extensively in [Reference Berman, Boucksom, Eyssidieux, Guedj and ZeriahiBBEGZ19], and eventually lead to an alternative solution of the Yau–Tian–Donaldson conjecture for Fano varieties [Reference Berman, Boucksom and JonssonBBJ21, Reference Li, Tian and WangLTW21, Reference LiLi22]. Adapting [Reference Berman, Boucksom, Eyssidieux, Guedj and ZeriahiBBEGZ19] to our local singular context, we develop in this section a variational approach for solving the equation

(3.1) \begin{equation}(MA)_{\gamma,\phi,\Omega}\begin{cases}(dd^{c}\varphi)^{n}=\dfrac{e^{-\gamma\varphi} \, d\mu_{p}}{\int_{\Omega}e^{-\gamma \varphi} \,d\mu_{p}},\\\varphi_{\vert\partial\Omega}=\phi.\end{cases}\end{equation}

3.1 Monge–Ampère energy

3.1.1 Smooth tests

Fix $\Omega=\{\rho<0\}$ and $\phi$ as described previously, and

\[\mathcal{T}_{\phi}^{\infty}(\Omega)=\{\varphi\in SPSH(\Omega)\cap\mathcal{C}^{\infty}(\overline{\Omega}): \varphi_{\vert \partial\Omega}\equiv \phi\}.\]

Recall that $\phi_{0}\in \mathcal{C}^{\infty}(\bar{\Omega})\cap PSH(\Omega)$ denotes a smooth psh extension of $\phi$ to $\bar{\Omega}$ . We set $\omega:=dd^{c}\phi_{0}$ . This is a semi-positive form, which can be assumed to be Kähler. However, if $\phi \equiv 0$ , we can equally well take $\phi_0 \equiv 0$ and get $\omega \equiv 0$ .

Definition 3.1. We call $E_{\phi}(\varphi):=({1}/({n+1}))\sum_{j=0}^{n}\int_{\Omega}(\varphi-\phi_{0})(dd^{c}\varphi)^{j}\wedge(dd^{c}\phi_{0})^{n-j}$ the $\phi$ -relative Monge–Ampère energy of $\varphi \in \mathcal{T}_{\phi}^{\infty}(\Omega)$ .

While the formula depends on the choice of $\phi_0$ , it follows from Lemma 3.2 that the difference of two such relative energies is constant:

\[E_{\phi_1}(\varphi)-E_{\phi_0}(\varphi)=E_{{\phi_1}}(\phi_0).\]

For $\phi_{0}=0$ , the formula reduces to $E(\varphi):=E_{0}(\varphi)=({1}/({n+1}))\int_{\Omega}\varphi(dd^{c}\varphi)^{n}$ .

This definition is motivated by the fact the $E_{\phi}$ is a primitive of the Monge–Ampère operator for smooth psh functions with $\phi$ -boundary values.

Lemma 3.2. Fix $\varphi \in \mathcal{T}_{\phi}^{\infty}(\Omega),\ v \in {\mathcal D}({\Omega})$ . Then $\varphi+tv \in\mathcal{T}_{\phi}^{\infty}(\Omega)$ for t small, and

\[{\frac{d}{dt} }_{|t=0}E_{\phi}(\varphi+tv)=\int_{\Omega} v (dd^c \varphi)^n.\]

In particular, $\varphi \mapsto E_{\phi}(\varphi)$ is increasing.

Here ${\mathcal D}({\Omega})$ denotes the space of smooth functions with compact support in $\Omega$ .

Proof. Fix $\varphi \in \mathcal{T}_{\phi}^{\infty}(\Omega)$ and $v \in {\mathcal D}({\Omega})$ . Since v is smooth with compact support, the function $\pm v+C\rho$ is psh for $C>0$ large enough, while $\varphi-\varepsilon\rho$ is psh for $\varepsilon>0$ small enough. It follows that $\varphi+tv$ is psh for t small enough.

Set $\omega=dd^c \phi_0$ . The function $\psi_t=\varphi-\phi_0+tv$ has zero boundary values, and

\[E_{\phi}(\varphi+tv)=\frac{1}{n+1}\sum_{j=0}^{n}\int_{\Omega}\psi_t (\omega+dd^{c}\psi_t)^{j}\wedge \omega^{n-j}.\]

It follows from Stokes theorem, as all functions involved in the integration by parts are identically zero on $\partial \Omega$ , that

\begin{align*}& (n+1) {\frac{d}{dt} }E_{\phi}(\varphi+tv) \\&\quad =\sum_{j=0}^{n}\int_{\Omega} \dot{\psi_t} (\omega+dd^{c}\psi_t)^{j}\wedge \omega^{n-j}+\sum_{j=1}^{n}\int_{\Omega} j {\psi_t} \,dd^c \dot{\psi_t} \wedge (\omega+dd^{c}\psi_t)^{j-1}\wedge \omega^{n-j} \\&\quad =\sum_{j=0}^{n}\int_{\Omega} \dot{\psi_t} (\omega+dd^{c}\psi_t)^{j}\wedge \omega^{n-j}+\sum_{j=1}^{n}\int_{\Omega} j \dot{\psi_t} \,dd^c {\psi_t} \wedge (\omega+dd^{c}\psi_t)^{j-1}\wedge \omega^{n-j} \\&\quad =\sum_{j=0}^{n}\int_{\Omega} (j+1) \dot{\psi_t} (\omega+dd^{c}\psi_t)^{j}\wedge \omega^{n-j}-\sum_{j=1}^{n}\int_{\Omega} j \dot{\psi_t} (\omega+dd^{c}\psi_t)^{j-1}\wedge \omega^{n-j+1} \\&\quad = (n+1) \int_{\Omega} \dot{\psi_t} (\omega+dd^{c}\psi_t)^n,\end{align*}

writing $dd^c \psi_t=(\omega+dd^c \psi_t)-\omega$ in the third line, and then distributing and relabelling so as to obtain a telescopic series. The formula follows for $t=0$ .

In short, the derivative of $E_{\phi}$ is the complex Monge–Ampère operator $(dd^c \varphi)^n$ which is a positive measure. It follows that $\varphi \mapsto E_{\phi}(\varphi)$ is increasing.

3.1.2 Continuous setting

The previous result extends to the case of continuous psh functions that are not necessarily strictly psh. Recall that

\[\mathcal{T}_{\phi}(\Omega):=\bigg\{\varphi\in PSH(\Omega)\cap C^{0}(\bar{\Omega}),\ \varphi_{|\partial \Omega}=\phi\text{ and } \int_{\Omega} (dd^c \varphi)^n <+\infty\bigg\}.\]

We would like to extend Lemma 3.2 to this less-regular setting. As $\varphi+tv$ is not necessarily psh, we need to project it onto the cone of all psh functions. The following result will thus be useful.

Lemma 3.3. Fix $\varphi\in \mathcal{T}_{\phi}(\Omega)$ and $f \in {\mathcal D}({\Omega})$ . Then $P(\varphi+f)\in\mathcal{T}_{\phi}(\Omega)$ where

\[P(\varphi+f):=\sup\{\psi \in PSH(\Omega),\ \psi \leq \varphi+f\}.\]

Moreover, $(dd^{c}P(\varphi+f))^{n}$ is supported on the contact set $\{P(\varphi+f)=\varphi+f\}$ .

Proof. Since $\varphi+f$ is bounded and continuous, it is classical to check that the envelope $P(\varphi+f)$ is a well-defined psh function. As f has compact support, one moreover checks that $P(\varphi+f)$ is continuous on $\partial\Omega$ with $P(\varphi+f)_{|\partial\Omega}=\varphi_{|\partial\Omega}=\phi$ .

Solving Dirichlet problems in small ‘balls’ not containing the singular point, it follows from a balayage argument that the Monge–Ampère measure of the envelope $(dd^{c}P(\varphi+f))^{n}$ is supported on the contact set $\{P(\varphi+f)=\varphi+f\}$ .

We extend $E_{\phi}(\cdot)$ to $\mathcal{T}_{\phi}(\Omega)$ by monotonicity, setting

\[E_{\phi}(\varphi):=\inf\{E_{\phi}(\psi),\ \psi \in\mathcal{T}_{\phi}^{\infty}(\Omega)\text{ and } \varphi \leq \psi\}.\]

It has been observed by Berman and Boucksom (in the setting of compact Kähler manifolds [Reference Berman and BoucksomBB10]) that $E_{\phi} \circ P$ is still differentiable, with $(E_{\phi} \circ P)'=E_{\phi}'\circ P$ . This result extends to our local singular setting.

Proposition 3.4. Fix $\varphi\in \mathcal{T}_{\phi}(\Omega)$ and $f\in\mathcal{D}(\Omega)$ . Then $t\to E_{\phi}(P(\varphi+tf))$ is differentiable and

\[{\frac{d}{dt}}_{|t=0}E_{\phi}(P(\varphi+tf)) =\int_{\Omega}f(dd^{c}\varphi)^{n}.\]

Proof. The proof is very similar to that in the compact case, we provide it as a courtesy to the reader. Set $\varphi_{t}:=P(\varphi+tf)$ . By Lemma 3.5 we have

(3.2) \begin{equation}\int_{\Omega}(\varphi_{t}-\varphi)(dd^{c}\varphi_{t})^{n}\leq E_{\phi}(\varphi_{t})-E_{\phi}(\varphi)\leq \int_{\Omega}(\varphi_{t}-\varphi)(dd^{c}\varphi)^{n}.\end{equation}

Since $\varphi_t-\varphi \leq tf$ , the second inequality yields

\[\limsup_{t\to 0^{+}}\frac{E_{\phi}(\varphi_{t})-E_{\phi}(\varphi)}{t}\leq \int_{X}f(dd^{c}\varphi)^{n},\]

and $\liminf_{t\to 0^{-}}(({E_{\phi}(\varphi_{t})-E_{\phi}(\varphi)})/{t})\geq\int_{X}f(dd^{c}\varphi)^{n}$ .

It follows from Lemma 3.3 that $(dd^{c}\varphi_{t})^{n}$ is supported on $\{\varphi_{t}=\varphi+tf\}$ , hence the first inequality in (3.2) yields

\[\int_{\Omega}\frac{\varphi_{t}-\varphi}{t}(dd^{c}\varphi_{t})^{n}=\int_{\Omega}f(dd^{c}\varphi_{t})^{n}.\]

Now $(dd^{c}\varphi_{t})^{n}\to (dd^{c}\varphi)^{n}$ weakly since $\varphi_{t}\to \varphi$ uniformly, therefore

\[\liminf_{t\to 0^{+}}\frac{E_{\phi}(\varphi_{t})-E_{\phi}(\varphi)}{t}\geq\liminf_{t\to 0^{+}}\int_{\Omega}f(dd^{c}\varphi_{t})^{n}=\int_{\Omega}f(dd^{c}\varphi)^{n},\]

and

\[\limsup_{t\to0^{-}}\frac{E_{\phi}(\varphi_{t})-E_{\phi}(\varphi)}{t}\leq\limsup_{t\to 0^{-}}\int_{\Omega}f(dd^{c}\varphi_{t})^{n}=\int_{\Omega}f(dd^{c}\varphi)^{n}.\]

Lemma 3.5. For any $\varphi_{1},\varphi_{2}\in \mathcal{T}_{\phi}(\Omega)$ ,

(3.3) \begin{equation}\int_{\Omega}(\varphi_{1}-\varphi_{2})(dd^{c}\varphi_{1})^{n}\leq E_{\phi}(\varphi_{1})-E_{\phi}(\varphi_{2})\leq\int_{\Omega}(\varphi_{1}-\varphi_{2})(dd^{c}\varphi_{2})^{n},\end{equation}

while if $\varphi_{1}\leq \varphi_{2}$ , then

(3.4) \begin{equation}E_{\phi}(\varphi_{1})-E_{\phi}(\varphi_{2})\leq \frac{1}{n+1}\int_{X}(\varphi_{1}-\varphi_{2})(dd^{c}\varphi_{1})^{n}.\end{equation}

The energy is continuous along decreasing sequence in $\mathcal{T}_{\phi}(\Omega)$ .

Proof. It follows from Stokes theorem that

\[E_{\phi}(\varphi_{1})-E_{\phi}(\varphi_{2})=\frac{1}{n+1}\sum_{j=0}^{n}\int_{\Omega}(\varphi_{1}-\varphi_{2})(dd^{c}\varphi_{1})^{j}\wedge (dd^{c}\varphi_{2})^{n-j}\]

and

\[\int_{\Omega}(\varphi_{1}-\varphi_{2})(dd^{c}\varphi_{1})^{j+1}\wedge(dd^{c}\varphi_{2})^{n-j-1}\leq \int_{\Omega}(\varphi_{1}-\varphi_{2})(dd^{c}\varphi_{1})^{j}\wedge (dd^{c}\varphi_{2})^{n-j},\]

for any $j=0,\ldots,n-1$ . The desired inequalities follow.

Let $\varphi_{j}\in \mathcal{T}_{\phi}(\Omega)$ be a decreasing sequence converging to $\varphi \in\mathcal{T}_{\phi}(\Omega)$ . We obtain

\[0\leq E_{\phi}(\varphi_{j})-E_{\phi}(\varphi)\leq\int_{\Omega}(\varphi_{j}-\varphi)(dd^{c}\varphi)^{n}\to 0\]

as $j\to +\infty$ by the monotone convergence theorem.

3.1.3 Finite energy class

Let $PSH_{\phi}(\Omega)$ denote the set of decreasing limits of functions in $\mathcal{T}_{\phi}(\Omega)$ . We extend $E_{\phi}$ to $PSH_{\phi}(\Omega)$ by monotonicity, setting

\[E_{\phi}(\varphi):=\inf\{E_{\phi}(\psi),\ \psi \in \mathcal{T}_{\phi}(\Omega)\text{ and } \varphi \leq \psi\}.\]

Definition 3.6. We set $\mathcal{E}^1(\Omega):=\left\{ \varphi \in PSH_{\phi}(\Omega); \; E_{\phi}(\varphi)>-\infty \right\}$ .

This ‘finite energy class’ has been introduced and studied intensively by Cegrell for smooth domains of $\mathbb C^n$ . His analysis extends to our mildly singular context. We summarize here the key facts that we shall need.

Theorem 3.7 (Cegrell). The complex Monge–Ampère operator $(dd^c \cdot)^n$ and the energy $E_{\phi}$ are well defined on the class $\mathcal{E}^1(\Omega)$ . Moreover:

  • functions in $\mathcal{E}^1(\Omega)$ have zero Lelong numbers;

  • the sets $\mathcal{G}_b(\Omega)=\{ \varphi \in \mathcal{E}^1(\Omega),\ -b \leq E_{\phi}(\varphi) \}$ are compact for all $b \in \mathbb R$ ;

  • Lemma 3.5 holds if $\varphi_1,\varphi_2 \in \mathcal{E}^1(\Omega)$ ;

  • if $\mu$ is a non-pluripolar probability measure such that $\mathcal{E}^1(\Omega)\subset L^1(\mu)$ , then there exists a unique function $v \in \mathcal{E}^1(\Omega) \cap\mathcal{F}_1(\Omega)$ such that $\mu=(dd^c v)^n$ .

We refer the reader to [Reference CegrellCeg98, Theorems 3.8, 7.2 and 8.2] for the proof of these results when $\Omega$ is smooth.

3.2 Ding functional

3.2.1 Euler–Lagrange equation

The Ding functional is

\[F_{\gamma}(\varphi):=E_{\phi}(\varphi)+\frac{1}{\gamma}\log \int_{\Omega}e^{-\gamma \varphi}\,d\mu_{p}.\]

Proposition 3.8. If $\varphi$ maximizes $F_{\gamma}$ over $\mathcal{T}_{\phi}(\Omega)$ , then $\varphi$ solves the complex Monge–Ampère equation (3.1).

Proof. Assume that $\varphi$ maximizes $F_{\gamma}$ over $\mathcal{T}_{\phi}(\Omega)$ , fix $f \in\mathcal{D}(\Omega)$ , and set $\varphi_{t}:=P(\varphi+tf)$ . Then

\[E_{\phi}(\varphi_{t})+\frac{1}{\gamma}\log \int_{\Omega}e^{-\gamma(\varphi+tf)}\,d\mu_{p}\leq F_{\gamma}(\varphi_{t})\leq F_{\gamma}(\varphi),\]

i.e. the function $t\to E_{\phi}(\varphi_{t})+({1}/{\gamma})\log\int_{\Omega}e^{-\gamma(\varphi+tf)}\,d\mu_{p}$ reaches its maximum at $t=0$ . Combining Proposition 3.4 and Lemma 3.9, we obtain

\[0=\frac{d}{dt}\bigg(E_{\phi}(\varphi_{t})+\frac{1}{\gamma}\log\int_{\Omega}e^{-\gamma(\varphi+tf)}\,d\mu_{p}\bigg)=\int_{\Omega}f\bigg((dd^{c}\varphi)^{n}-\frac{e^{-\gamma\varphi}\,d\mu_{p}}{\int_{\Omega}e^{-\gamma\varphi}\,d\mu_{p}}\bigg),\]

i.e. $\varphi$ solves (3.1).

Lemma 3.9. Fix $\varphi \in \mathcal{T}_{\phi}(\Omega)$ , $f \in {\mathcal D}(\Omega)$ , and set $\psi_t:=\varphi+tf$ . Then

\[\frac{d}{dt}\bigg(\log \int_{\Omega}e^{-\gamma\psi_t}\,d\mu_{p}\bigg)_{\!\!|t=0}=-\gamma\frac{\int_{\Omega}fe^{-\gamma\varphi}\,d\mu_{p}}{\int_{\Omega}e^{-\gamma\varphi}\,d\mu_{p}}.\]

Proof. By the chain rule, it is enough to observe that

\[\frac{\int_{\Omega}e^{-\gamma \psi_t}\,d\mu_{p}-\int_{\Omega}e^{-\gamma \varphi}\,d\mu_{p}}{t}=-\int_{\Omega}e^{-\gamma\varphi}\bigg(\frac{1-e^{-t\gamma f}}{t} \bigg)\,d\mu_{p}\]

and to apply the Lebesgue dominated convergence theorem to conclude.

3.2.2 Coercivity

In order to solve (3.1), one is lead to try and maximize $F_{\gamma}$ . We show in § 6 that when $F_{\gamma}$ is coercive, the complex Monge–Ampère equation (3.1) admits a solution $\varphi\in \mathcal{T}_{\phi}(\Omega)$ which is smooth away from p.

Definition 3.10. The functional $F_{\gamma}$ is coercive if there exists $A,B>0$ such that

\[F_{\gamma}(\varphi)\leq A E_{\phi}(\varphi)+B\]

for all $\varphi\in \mathcal{T}_{\phi}(\Omega)$ .

We observe in Lemma 3.13 that $E_{\phi}(\varphi) \leq C(\phi_0)$ is bounded from above, uniformly in $\varphi \in \mathcal{T}_{\phi}(\Omega)$ . In particular, if $F_{\gamma}$ is coercive with slope $A>0$ , then it is coercive for any $A'\in (0,A]$ . We can thus assume, without loss of generality, that $A\in(0,1)$ . The coercivity property is then equivalent to

\[\frac{1}{\gamma}\log \int_{\Omega}e^{-\gamma\varphi}\,d\mu_{p}\leq (1-A)(-E_{\phi}(\varphi))+B,\]

or, equivalently, to the following Moser–Trudinger inequality

\[\bigg(\int_{X}e^{-\gamma \varphi}\,d\mu_{p}\bigg)^{{1}/{\gamma}} \leq C e^{(1-A)(-E_{\phi}(\varphi))}.\]

We summarize these observations in the following.

Proposition 3.11. Fix $\gamma>0$ . The following properties are equivalent:

  1. (i) $F_{\gamma}$ is coercive;

  2. (ii) there exists $C_{\gamma}>0$ and $a\in (0,1)$ such that for all $\varphi \in\mathcal{T}_{\phi}(\Omega)$ ,

    \[\bigg(\int_{\Omega}e^{-\gamma\varphi}\,d\mu_{p}\bigg)^{\!\!1/\gamma}\leq C_{\gamma} e^{-a E_{\phi}(\varphi)}.\]

It follows from Hölder inequality (and the normalization $\mu_p(\Omega)=1$ ) that if

(3.5) \begin{equation}\bigg(\int_{\Omega}e^{-\gamma\varphi}\,d\mu_{p}\bigg)^{\!\!1/\gamma}\leq Ce^{-E_{\phi}(\varphi)}\end{equation}

holds for $\gamma>0$ , then it also holds for $\gamma'<\gamma$ . We thus introduce the following critical exponent.

Definition 3.12. We set

Lemma 3.13. The functional $E_{\phi}$ is bounded from above on $\mathcal{T}_{\phi}(\Omega)$ . Moreover:

  • if $F_{\gamma}$ is coercive, then $\gamma \leq \gamma_{\rm crit}(X,p)$ ;

  • conversely, if $\gamma < \gamma_{\rm crit}(X,p)$ , then $F_{\gamma}$ is coercive.

Proof. Consider $\tilde{\phi_0}=P(\phi):=\sup \{ \psi, \psi \in \mathcal{T}_{\phi}(\Omega)\}$ . This is the largest psh function in $\Omega$ such that $\tilde{\phi_0} =\phi$ on $\partial \Omega$ . The reader can check that it is continuous on ${\overline \Omega}$ and satisfies $(dd^c \tilde{\phi_0})^n=0$ in $\Omega$ . If $\varphi \in \mathcal{T}_{\phi}(\Omega)$ , then $\varphi \leq \tilde{\phi_0}$ , hence $E_{\tilde{\phi_0}}(\varphi) \leq 0$ . Thus,

\[E_{{\phi_0}}(\varphi) = E_{\tilde{\phi_0}}(\varphi)+E_{\phi_0}(\tilde{\phi_0})\leq E_{\phi_0}(\tilde{\phi_0}),\]

hence $E_{{\phi_0}}(\varphi)$ is uniformly bounded from above independently of the choice of $\phi_0$ .

Similarly, the coercivity of $F_{\gamma}$ or the inequality (3.5) do not depend on the choice of $\phi_0$ . In the remainder of this proof we thus assume that $\phi_0=P(\phi)$ . Since $E_{\phi}(\varphi) \leq 0$ in this case, it follows from Proposition 3.11 that if $F_{\gamma}$ is coercive, then (3.5) holds, hence $\gamma \leq \gamma_{\rm crit}(X,p)$ .

Conversely, assume $\gamma < \gamma_{\rm crit}(X,p)$ . Fix $\gamma<\gamma'<\gamma_{\rm crit}(X,p)$ and $\lambda=\gamma/\gamma'<1$ . We can assume that $\lambda$ is close to 1. We assume first that $\phi \equiv 0$ . For $\varphi \in \mathcal{T}_{0}(\Omega)$ we observe that $\lambda \varphi \in\mathcal{T}_{0}(\Omega)$ , with $E_0(\lambda \varphi)=\lambda^{n+1} E_0(\varphi)$ . The Moser–Trudinger (3.5) applied to $(\gamma',\lambda \varphi)$ thus yields

\[\bigg(\int_{\Omega}e^{-\gamma\varphi}\,d\mu_{p}\bigg)^{\!\!1/\gamma}\leq C_{\gamma'} e^{-\lambda^n E_0(\varphi)},\]

so that $F_{\gamma}$ is coercive.

We now treat the general case, replacing the condition $\phi \equiv 0$ by $(dd^c \phi_0)^n \equiv0$ . For $\varphi \in \mathcal{T}_{\phi}(\Omega)$ we observe that $\varphi_{\lambda}=\lambda \varphi+(1-\lambda)\phi_0 \in \mathcal{T}_{\phi}(\Omega)$ , with $\varphi_{\lambda}-\phi_0=\lambda(\varphi-\phi_0) \leq 0$ and

\begin{align*}(n+1) E_{\phi_0}(\varphi_{\lambda}) &=\lambda \sum_{j=0}^{n-1} \int_{\Omega} (\varphi-\phi_0) (dd^c \varphi_{\lambda})^j \wedge (dd^c \phi_0)^{n-j} \\&= \lambda \sum_{k=1}^n \sum_{j=k}^{n}\bigg(\begin{array}{c}j \\k \end{array}\bigg)\lambda^k (1-\lambda)^{j-k}\int_{\Omega} (\varphi-\phi_0) (dd^c \varphi)^k \wedge (dd^c \phi_0)^{n-k}.\end{align*}

Now $\sum_{j=k}^{n} \big(\begin{smallmatrix} j \\ k \end{smallmatrix}\big) \lambda^k(1-\lambda)^{j-k} \leq a <1$ for all $1 \leq k \leq n$ , since $\lambda<1$ can be chosen arbitrarily close to 1. Thus, $E_{\phi_0}(\varphi_{\lambda}) \geq a \lambda E_{\phi_0}(\varphi)$ and the result follows as previously by applying the Moser–Trudinger inequality (3.5) to $\varphi_{\lambda}$ .

3.3 Invariance

We give here some evidence that the critical exponent $\gamma_{\rm crit}(X,p)$ should be independent of the domain $\Omega$ and the boundary values $\phi$ .

3.3.1 Enlarging the domain

We first reduce to the case of zero boundary values.

Proposition 3.14. Let $\Omega_2$ be a smooth strongly pseudoconvex domain containing $\overline{\Omega}$ . If the Moser–Trudinger inequality holds for $(\gamma,\Omega_2,0)$ , then it holds for $(\gamma,\Omega,\phi)$ .

Proof. Consider indeed $\varphi \in {\mathcal T}_{\phi}(\Omega)$ and set

\[\varphi_2:=\sup \{u \in {\mathcal T}_0(\Omega_2),\ \text{such that}\ u \leq \varphi \text{ in } \Omega\}.\]

The family ${\mathcal F}$ of such functions is non-empty, as it contains $A \rho_2$ for some large $A>1$ , where $\rho_2$ is a psh defining function for $\Omega_2$ . Moreover, ${\mathcal F}$ is uniformly bounded from above by 0, so the upper envelope $\varphi_2$ is well defined and psh, as ${\mathcal F}$ is compact. Finally, $\varphi_2 \geq A \rho_2$ , hence $\varphi_2$ has zero boundary values, and $\varphi_2$ is lower semi-continuous, as an envelope of continuous functions, thus $\varphi_2 \in {\mathcal T}_0(\Omega_2)$ .

Since $\varphi_2 \leq \varphi$ in $\Omega$ , we observe that

\[\int_{\Omega} e^{-\gamma \varphi} \,d\mu_p \leq \int_{\Omega_2} e^{-\gamma \varphi_2} \,d\mu_p.\]

Our claim will follow if we can show that, on the other hand, $E_0(\varphi_2) \geq E_{\phi}(\varphi)$ .

If $\varphi$ is smooth one can show, by adapting standard techniques, that:

  • $\varphi_2$ is ${\mathcal C}^{1,\overline{1}}$ -smooth in $\Omega_2 \setminus \{p\}$ ;

  • $(dd^c \varphi_2)^n=0$ in $\Omega_2 \setminus \Omega$ and $(dd^c \varphi_2)^n={{\bf 1}}_{\{ \varphi_2=\varphi\}} (dd^c \varphi)^n$ in $\overline{\Omega}$ .

Assuming $\phi \geq 0$ and $\phi_0=\sup \{\psi,\ \psi \in {\mathcal T}_{\phi}(\Omega) \}$ , we infer

\begin{align*}E_0(\varphi_2) &= \frac{1}{n+1}\int_{\Omega} \varphi_2 (dd^c \varphi_2)^n= \frac{1}{n+1}\int_{\Omega} {{\bf 1}}_{\{ \varphi_2 =\varphi\}} \varphi (dd^c \varphi)^n \\&\geq \frac{1}{n+1}\int_{\Omega} \varphi (dd^c \varphi)^n\geq \frac{1}{n+1}\int_{\Omega} (\varphi-\phi) (dd^c \varphi)^n \geq E_{\phi}(\varphi).\end{align*}

To get rid of the assumption $\phi \geq 0$ , we observe that the Moser–Trudinger inequality holds for given boundary data $\phi$ if and only if does so for $\phi+c$ , for any $c \in \mathbb R$ (by changing $\varphi$ in $\varphi+c$ ).

Using Lemma 2.5, one can uniformly approximate $\varphi$ by a sequence of smooth $\varphi_{j}\in {\mathcal T}_{\phi}(\Omega)$ . The corresponding sequence $\varphi_{2,j}$ uniformly converges to $\varphi_2$ , and we obtain the desired inequality by passing to the limit in $E_0(\varphi_{2,j}) \geq E_{\phi}(\varphi_{j})$ .

3.3.2 Rescaling

We now assume that $\phi=0$ and reformulate the coercivity property after an appropriate rescaling. Observe that for any $\lambda>0$ , the map

\[\varphi\in\mathcal{T}_{0}(\Omega) \mapsto \lambda \varphi \in \mathcal{T}_{0}(\Omega)\]

is a homeomorphism. This allows us to reformulate the Moser–Trudinger inequality.

Proposition 3.15 The following statements are equivalent:

  1. (a) $F_{\gamma}$ is coercive;

  2. (b) there exists $C>0,\ B\in(0,1)$ such that for all $\varphi\in\mathcal{T}_{0}(\Omega)$ , $\int_{\Omega}e^{-\varphi}\,d\mu_{p}\leq Ce^{-({B}/{\gamma^{n}})E_{0}(\varphi)}$ .

In particular, we can define the following critical exponent.

Definition 3.16. We set

\[\beta_{\rm crit}:=\inf \bigg\{\beta>0;\ \sup_{\varphi\in\mathcal{T}_{0}(\Omega)}\bigg(\int_{\Omega}e^{-\varphi}\,d\mu_{p}/e^{-\beta E_{0}(\varphi)}\bigg)<+\infty\bigg\}.\]

Note that $\gamma_{\rm crit}^{n}=1/\beta_{\rm crit}$ , hence it follows from the previous analysis that $F_{\gamma}$ is coercive if and only if $\gamma< \beta_{\rm crit}^{-1/n}$ . When $p\in X$ is smooth, it has been shown in [Reference Guedj, Kolev and YeganefarGKY13, Theorem 9] and independently [Reference Berman and BerndtssonBB22, Theorem 1.5] that

\[\beta_{\rm crit}(\Omega)=\frac{1}{(n+1)^{n}},\]

or, equivalently, that $\gamma_{\rm crit}(\Omega)=n+1$ . In particular, it does not depend on $\Omega$ .

We extend this independence to the case when p is the vertex of a cone over a Fano manifold.

Proposition 3.17. Assume that (X,p) is the affine cone over a Fano manifold Z embedded in a projective space by the linear system $|-rK_Z|$ for $r\in \mathbb N$ such that $L=rK_Z^*$ is very ample, and fix $\lambda\in\mathbb C^*$ . The Moser–Trudinger inequality holds for $(\gamma,\Omega,0)$ if and only if it does so for $(\gamma,\lambda \Omega, 0)$ .

Proof. Let $L=rK_Z^*$ , let $D_{\lambda}$ denote the dilatation $z \mapsto \lambda z$ and set $\Omega_{\lambda}=D_{\lambda}( \Omega)$ . We blow up p to obtain a resolution $f:Y \rightarrow X$ , where Y is the total space of $L^*$ and the exceptional divisor E is the zero section of $L^*$ .

Recall that $K_Y=f^*K_X+aE$ , where a is the discrepancy of Y along E. The adjunction formula yields $(K_Y+E)_{|E}=K_E$ , hence $K_Z^*=(a+1)L$ . In particular, $a=-1+1/r$ and (X,p) is log terminal. The fibration $\pi:Y=L^* \rightarrow Z$ yields $K_Y=\pi^*(K_Z+L)$ , hence $f^*K_X=\pi^*(K_Z+L)-aE$ .

Since $\pi^*(K_Z+L)$ is $\mathbb C^*$ -invariant, we can cook up an adapted volume form $\mu_p=\mu_1 \cdot\mu_E$ with $D_{\lambda}^* \mu_1=\mu_1$ while $D_{\lambda}^* \mu_E=|\lambda|^{2a} \mu_E$ . For $\varphi\in {\mathcal T}_0(\Omega_{\lambda})$ we set $\varphi_{\lambda}=\varphi \circ D_{\lambda} \in {\mathcal T}_0(\Omega)$ and observe that

\[|\lambda |^{2a}\int_{\Omega} e^{-\gamma \varphi_{\lambda}} \,d\mu_p=\int_{\Omega_{\lambda}} e^{-\gamma \varphi} \,d\mu_p,\]

while $E_{\Omega,0}(\varphi_{\lambda})=E_{\Omega_{\lambda},0}(\varphi)$ . The conclusion follows.

We conjecture in § 5 that $ \gamma_{\rm crit}(X,p) \stackrel{?}{=}(({n+1})/{n}) \widehat{\mathrm{vol}}(X,p)^{1/n} $ and give partial results towards this equality, which again suggest that $\gamma_{\rm crit}(X,p) $ should be independent of $(\Omega,\phi)$ . In the whole article, we therefore use the notation $\gamma_{\rm crit}(X,p) $ instead of the more precise, and heavy, $\gamma_{\rm crit}(X,p,\Omega,\phi)$ .

4. Upper bound for the coercivity

The purpose of this section is to establish the following upper bound:

\[\gamma_{\rm crit}(X,p) \leq \frac{n+1}{n}\widehat{\mathrm{vol}}(X,p)^{1/n}.\]

Adapting the proof of [Reference Berman and BerndtssonBB17, Theorem 1.6], we will construct approximate Green’s functions to test the thresholds in the Moser–Trudinger inequality.

4.1 Functions with algebraic singularities

Let $\mathcal{I}$ be a coherent ideal sheaf, and $f_{1},\ldots,f_{N}\in \mathcal{O}_{X,p}$ be local generators of $\mathcal{I}_{p}$ . The psh function

\[\varphi_{\mathcal{I}}:=\log\bigg(\sum_{i=1}^{N}\lvert f_{i}\rvert^{2}\bigg)\]

is well defined near p, with algebraic singularities encoded in $\mathcal{I}$ .

Proposition 4.1. Let $\mathcal{I}$ be a coherent ideal sheaf supported at p. Then

\[e(X,\mathcal{I})=\int_{\{p\}}(dd^c\varphi_\mathcal{I})^n\]

and

where $\Omega$ is any (small) neighborhood of $p\in X$ .

These algebraic quantities are thus independent of the choice of generators.

Proof. The equality $e(X,\mathcal{I})=\int_{\{p\}}(dd^c\varphi_{\mathcal{I}})^n$ is classical when $\Omega$ is smooth (see, e.g., [Reference DemaillyDem12, Lemma 2.1]), and the proof can be adapted to the singular context (see [Reference DemaillyDem85, Chapter 4]).

Let $\pi:\tilde{\Omega}\to \Omega$ be a local log resolution of the ideal $(X,\mathcal{I})$ , i.e. a composition of blow-ups such that $ \pi^{*}\mu_{p}=\prod_{j=1}^{N}\lvert s_{E_{j}}\rvert^{2a_{j}}dV_{\tilde{\Omega}} $ and

\[\pi^{-1}\mathcal{I}\cdot \mathcal{O}_{\tilde{\Omega}}=\mathcal{O}_{\tilde{\Omega}}\bigg(-\sum_{j=1}^{M}b_{j}E_{j}\bigg),\]

where $b_{j}\in\mathbb N$ , $a_{j}\in\mathbb Q_{>-1}$ , and $E_{1},\ldots,E_{M}$ have simple normal crossings. Observe that

\[\int_{\Omega}e^{-\alpha \varphi_{\mathcal{I}}}\,d\mu_{p}\sim\int_{\tilde{\Omega}}\frac{\prod_{j=1}^{M}\lvert s_{E_{j}}\rvert^{2a_{j}}}{\prod_{j=1}^{M}\lvert s_{E_{j}}\rvert^{2b_{j}\alpha}}\,dV_{\tilde{\Omega}}=\int_{\tilde{\Omega}}\prod_{j=1}^{M}\lvert s_{E_{j}} \rvert^{2(a_{j}-\alpha b_{j})}\,dV_{\tilde{\Omega}},\]

is finite if and only if $a_{j}-\alpha b_{j}>-1$ for any $j=1,\ldots,M$ , i.e. if and only if

\[\alpha< \inf_{j=1,\ldots,M}\frac{a_{j}+1}{b_{j}}=\mathrm{lct}(X,\mathcal{I}),\]

as recalled in Definition 2.12.

4.2 Approximate Green functions

The functions $\lambda \varphi_{\mathcal{I}}$ play the role of Green functions adapted to the singularity (X,p). We show here how to approximate them from above by smooth functions with prescribed boundary values.

Lemma 4.2. Let $\mathcal{I}$ be a coherent ideal sheaf supported at p, and let $f_{1},\ldots,f_{m}$ denote local generators of $\mathcal{I}$ . Fix an open set $\Omega'\Subset \Omega$ . There exists a family $\{\varphi_{\mathcal{I},\lambda,\epsilon}\}_{\lambda>0, \epsilon> 0}\in PSH(\Omega)\cap\mathcal{C}^{\infty}(\bar{\Omega})$ such that:

  1. (i) $\varphi_{\mathcal{I},\lambda,\epsilon\,|\,\partial\Omega}=\phi$ for any $\lambda>0,\epsilon\in [0,1]$ ;

  2. (ii) $\varphi_{\mathcal{I},\lambda,\epsilon}=\lambda\log\big(\sum_{j=1}^{m}\lvert f_{j}\rvert^{2}+\epsilon^{2}\big)+\phi_{0}$ in $\Omega'$ ;

  3. (iii) $\varphi_{\mathcal{I},\lambda,\epsilon}\searrow\varphi_{\mathcal{I},\lambda,0}=:\varphi_{\mathcal{I},\lambda}$ as $\epsilon\searrow 0$ for any $\lambda>0$ fixed.

Proof. Without loss of generality we can assume that $\sum_{j=1}^{m}\lvert f_{j}\rvert^{2}\leq 1/e-1$ in $\Omega$ . Let $\rho$ be a smooth psh exhaustion for $\Omega$ and fix $0< r\ll 1$ , $0<\delta\ll1$ small enough. There exists $A>0$ big enough and relatively compact open sets $B_{r}(0)\Subset\Omega'\Subset \tilde{\Omega}\Subset \Omega$ such that

while

. We infer that

\[u_{\mathcal{I},\epsilon}:=\begin{cases}A\rho & \mathrm{on}\ \Omega\setminus \tilde{\Omega},\\\displaystyle\max_{\delta}\bigg(\log\bigg(\sum_{j=1}^{m}\lvert f_{j}\rvert^{2}+\epsilon^{2}\bigg), A\rho\bigg) &\mathrm{on}\ \tilde{\Omega}\setminus \Omega',\\\displaystyle\log\bigg(\sum_{j=1}^{m}\lvert f_{j}\rvert^{2}+\epsilon^{2}\bigg) & \mathrm{on}\ \Omega',\end{cases}\]

is a decreasing family (in $\epsilon\in[0,1]$ ) of psh functions which are smooth in $\bar{\Omega}\setminus\{p\}$ (smooth in $\bar{\Omega}$ for $\epsilon>0$ ) and which are identically 0 on $\partial \Omega$ . Here $\max_{\delta}(\cdot,\cdot)$ denotes the regularized maximum. The lemma follows by setting $ \varphi_{\mathcal{I},\lambda,\epsilon}:=\lambda u_{\mathcal{I},\epsilon}+\phi_{0}$ .

We now compute the asymptotic behavior, as $\varepsilon$ decreases to 0, of the quantities involved in the expected Moser–Trudinger inequality.

Lemma 4.3. Let $\mathcal{I}$ and $\{\varphi_{\mathcal{I},\lambda,\epsilon}\}_{\epsilon\in(0,1]}\subset\mathcal{T}_{\phi}(\Omega)$ be as in Lemma 4.2. Then, for any $\gamma>0, \lambda>0$ fixed there exists a constant $C_{\lambda,\gamma}\in\mathbb R$ (independent of $\epsilon$ ) such that

(4.1) \begin{equation}C_{\lambda,\gamma}+(\gamma\lambda-\mathrm{lct}(X,\mathcal{I}))\log\epsilon^{-2}\leq \log\int_{\Omega}e^{-\gamma \varphi_{\mathcal{I},\lambda,\epsilon}}\,d\mu_{p}\end{equation}

for all $0<\epsilon<\epsilon_0$ .

Proof. Taking a log resolution $\pi: Y\to X$ we obtain

\begin{align*}\int_{\Omega}e^{-\gamma \varphi_{\mathcal{I,\lambda,\epsilon}}}\,d\mu_{p}&\geq \int_{\Omega'}\frac{1}{(\sum_{j=1}^{m}\lvert f_{j}\rvert^{2}+\epsilon^{2})^{\gamma\lambda}}\,d\mu_{p} \\&\geq C_{1}\int_{\pi^{-1}(\Omega')}\frac{\prod_{j=1}^{M}\lvert s_{E_{j}}\rvert^{2a_{j}}}{(\prod_{j=1}^{M}\lvert s_{E_{j}} \rvert^{2b_{j}}+\epsilon^{2})^{\gamma\lambda}}\,dV_{\pi^{-1}(\Omega')},\end{align*}

where $C_{1}$ is a uniform constant (independent on $\epsilon$ ). We set

\[f:=\frac{\prod_{j=1}^{M}\lvert s_{E_{j}}\rvert^{2a_{j}}}{(\prod_{j=1}^{M}\lvert s_{E_{j}}\rvert^{2b_{j}}+\epsilon^{2})^{\gamma\lambda}}.\]

We can assume without loss of generality that $\mathrm{lct}(X,\mathcal{I})=({a_{1}+1})/{b_{1}}$ . Pick $x\in E_{1}, x\notin E_{j},j=2,\ldots,M$ . We can find $0<r\ll 1$ so small that $B_{r}(x)\cap E_{j}=\emptyset$ for any $j=2,\ldots,M$ . We choose holomorphic coordinates $(z_{1},\ldots,z_{n})$ centered at x such that $E_{1}=\{z_{1}=0\}$ . Thus, setting $a:=a_{1},b:=b_{1}$ and $c:=\gamma\lambda$ we get

\[\int_{\pi^{-1}(\Omega')}f\,dV_{\pi^{-1}(\Omega')}\geq C_{2}\int_{B_{r}(0)}\frac{\lvert z_{1} \rvert^{2a}}{(\lvert z_{1}\rvert^{2b}+\epsilon^{2})^{c}}\,d\lambda(z)=C_{3}\int_{0}^{r}\frac{u^{2a+1}}{(u^{2b}+\epsilon^{2})^{c}}\,du,\]

where $C_{2}, C_{3}$ are uniform constants. If $c\leq ({a+1})/{b}$ (i.e. $\gamma\lambda\leq\mathrm{lct}(X,\mathcal{I})$ ), then

\[\int_{0}^{r}\frac{u^{2a+1}}{(u^{2b}+\epsilon^{2})^{c}}\,du\geq\int_{0}^{1}\frac{u^{2a+1}}{(u^{2b}+1)^{c}}=:C_{4}\]

and (4.1) trivially follows. If $c>({a+1})/{b}$ , then the substitution $v:=u/\epsilon^{1/b}$ yields

\begin{align*}\int_{0}^{r}\frac{u^{2a+1}}{(u^{2b}+\epsilon^{2})^{c}}\,du&= \epsilon^{-2(c-(({a+1})/{b}))}\int_{0}^{r/\epsilon^{1/b}}\frac{v^{2a+1}}{(v^{2b}+1)^{c}} \\&\geq \epsilon^{-2(\gamma\lambda-\mathrm{lct}(X,\mathcal{I}))}\int_{0}^{r}\frac{v^{2a+1}}{(v^{2b}+1)^{c}}\,dv.\end{align*}

The lemma follows.

Lemma 4.4. Let $\mathcal{I}$ and $\varphi_{\mathcal{I},\lambda,\epsilon}\in\mathcal{T}_{\phi}(\Omega)$ be as in Lemma 4.2. There exist positive constants $\{C_{\ell,\lambda}\}_{\ell\in\mathbb N,\lambda>0}$ and a family of functions $F_{\ell}:(0,1]\to \mathbb R_{>0}$ such that

\[-E_{\phi}(\varphi_{\mathcal{I},\lambda,\epsilon})\leq C_{\ell,\lambda}+\frac{\lambda^{n+1}}{n+1}F_{\ell}(\epsilon)\log\epsilon^{-2},\]

for any $\epsilon \in (0,1]$ , where:

  • $\{C_{\ell,\lambda}\}_{\ell\in\mathbb N,\lambda>0}$ is independent of $\epsilon\in(0,1]$ ;

  • $F_{\ell}(\epsilon)\to F_{\ell}(0)=:e_{\ell}>0$ as $\epsilon\searrow 0$ ;

  • $e_{\ell}\searrow e(X,\mathcal{I})$ as $\ell\to +\infty$ .

Proof. We take a sequence $\{\Omega_{\ell}\}_{\ell\in\mathbb N}$ of open sets such that $\Omega_{\ell+1}\Subset\Omega_{\ell}$ for any $\ell\in\mathbb N$ and such that $\bigcap_{\ell\in\mathbb N}\Omega_{\ell}=\{p\}$ . Since $\Omega_{\ell}\subset \Omega'$ (same notation as Lemma 4.2) for $\ell\in\mathbb N$ big enough, we obtain

(4.2) \begin{align}-E_{\phi}(\varphi_{\mathcal{I},\lambda,\epsilon}) &=\frac{1}{n+1}\sum_{j=0}^{n}\int_{\Omega}(\phi_{0}-\varphi_{\mathcal{I},\lambda,\epsilon})(dd^{c}\varphi_{\mathcal{I},\lambda,\epsilon})^{j}\wedge (dd^{c}\phi_{0})^{n-j} \nonumber\\[8pt]&= \frac{1}{n+1}\sum_{j=0}^{n}\int_{\Omega\setminus\Omega_{\ell}}(\phi_{0}-\varphi_{\mathcal{I},\lambda,\epsilon})(dd^{c}\varphi_{\mathcal{I},\lambda,\epsilon})^{j}\wedge (dd^{c}\phi_{0})^{n-j} \nonumber\\[8pt]&\quad -\frac{1}{n+1}\sum_{j=0}^{n}\!\int_{\Omega_{\ell}} \!\lambda^{j+1} \log\!\bigg(\sum_{k=1}^{m}\lvert f_{k}\rvert^{2}+\epsilon^{2}\!\bigg)\!\bigg(\!dd^{c}\log\bigg(\sum_{k=1}^{m}\lvert f_{k}\rvert^{2}+\epsilon^{2}\bigg)\!\bigg)^{\!\!j} \wedge (dd^{c}\phi_{0})^{n-j}. \end{align}

The first term on the right-hand side of (4.2) is uniformly bounded in $\epsilon\in[0,1]$ , for $\lambda>0,\ell\in\mathbb N$ fixed, since $\{\varphi_{\mathcal{I},\lambda,\epsilon}\}_{\epsilon\in [0,1]}$ is a continuous family of smooth functions on $\Omega\setminus \Omega_{\ell}$ . We let $C_{\ell,\lambda}$ denote a uniform upper bound for this quantity.

The second term on the right-hand side of (4.2) is bounded from above by

\begin{align*}& -\frac{1}{n+1}\sum_{j=0}^{n}\int_{\Omega_{\ell}} \lambda^{j+1} \log\bigg(\sum_{k=1}^{m}\lvert f_{k}\rvert^{2}+\epsilon^{2}\bigg) \bigg(dd^{c}\log\bigg(\sum_{k=1}^{m}\lvert f_{k}\rvert^{2}+\epsilon^{2}\bigg)\bigg)^{\!\!j}\wedge (dd^{c}\phi_{0})^{n-j} \\[8pt]&\quad \leq \frac{\lambda^{n+1}}{n+1}\log\epsilon^{-2}\sum_{j=0}^{n}\int_{\Omega_{\ell}}\bigg(dd^{c}\log\bigg(\sum_{k=1}^{m}\lvert f_{k}\rvert^{2}+\epsilon^{2}\bigg)\bigg)^{\!\!j}\wedge (dd^{c}\phi_{0})^{n-j}.\end{align*}

We set

\[F_{\ell}(\epsilon):=\sum_{j=0}^{n}\int_{\Omega_{\ell}}\bigg(dd^{c}\log\bigg(\sum_{k=1}^{m}\lvert f_{k}\rvert^{2}+\epsilon^{2}\bigg)\bigg)^{\!\!j}\wedge(dd^{c}\phi_{0})^{n-j}.\]

Observe that, for $j=0,\ldots,n-1$ ,

\[\lim_{\ell\nearrow +\infty}\lim_{\epsilon \searrow 0}\int_{\Omega_{\ell}}\bigg(dd^{c}\log\bigg(\sum_{k=1}^{m}\lvert f_{k}\rvert^{2}+\epsilon^{2}\bigg)\bigg)^{\!\!j}\wedge(dd^{c}\phi_{0})^{n-j}=0,\]

since the ideal sheaf $\mathcal{I}$ generated by $f_{1},\ldots,f_{m}$ is supported at one point, while

\[\int_{\Omega_{\ell}}\bigg(dd^{c}\log\bigg(\sum_{k=1}^{m}\lvert f_{k}\rvert^{2}+\epsilon^{2}\bigg)\bigg)^{\!\!n}\to e_{\ell}\]

as $\epsilon\searrow 0$ , where $e_{\ell}\geq e(X,\mathcal{I})$ and $e_{\ell}\searrow\int_{\{p\}}(dd^c\varphi_{\mathcal{I},1})^n$ as $\ell \nearrow +\infty$ .

Proposition 4.1 yields $\int_{\{p\}}(dd^c\varphi_{\mathcal{I},1})^n=e(X,\mathcal{I})$ , ending the proof.

4.3 The upper bound

We are now ready for the proof of the following result.

Theorem 4.5. Let (X,p) be a an isolated log terminal singularity. Then

\[\gamma_{\rm crit}\leq \frac{n+1}{n}\widehat{\mathrm{vol}}(X,p)^{1/n}.\]

Proof. Fix $\gamma<\gamma_{\rm crit}$ and $C_{1}>0$ such that

(4.3) \begin{equation}\frac{1}{\gamma}\log\int_{\Omega}e^{-\gamma \varphi}\,d\mu_{p}\leq C_{1}-E_{\phi}(\varphi)\end{equation}

for any $\varphi\in\mathcal{T}_{\phi}(\Omega)$ .

Fix $\mathcal{I}$ coherent ideal sheaf supported at p, and let $\{\varphi_{\mathcal{I},\lambda,\epsilon}\}_{\lambda>0,\epsilon\in(0,1]}\in\mathcal{T}_{\phi}(\Omega)$ as defined in Lemma 4.2. Evaluating (4.3) at $\{\varphi_{\mathcal{I},\lambda,\epsilon}\}_{\epsilon\in(0,1]}$ yields

\begin{align*}C_{\gamma,\lambda}+\bigg(\lambda-\frac{\mathrm{lct}(X,\mathcal{I})}{\gamma}\bigg)\log \epsilon^{-2}&\leq \frac{1}{\gamma}\log\int_{\Omega}e^{-\gamma\varphi_{\mathcal{I},\lambda,\epsilon}}\,d\mu_{p} \\&\leq C_{1}-E_{\phi}(\varphi_{\mathcal{I},\lambda,\epsilon}) \\&\leq C_{1}+ C_{N,\lambda}+\frac{\lambda^{n+1}}{n+1}F_{N}(\epsilon)\log \epsilon^{-2}\end{align*}

for any $N\in\mathbb N,\epsilon\in (0,1]$ thanks to Lemmas 4.3 and 4.4. We infer

\[\bigg(\lambda-\frac{\mathrm{lct}(X,\mathcal{I})}{\gamma}-\frac{\lambda^{n+1}}{n+1}F_{N}(\epsilon)\bigg)\log\epsilon^{-2}\leq C_{1}+C_{N,\lambda}-C_{\gamma,\lambda},\]

hence

(4.4) \begin{equation}\lambda-\frac{\lambda^{n+1}}{n+1}e_{N}\leq \frac{\mathrm{lct}(X,\mathcal{I})}{\gamma}\end{equation}

for any $N\in\mathbb N, \lambda>0$ since $F_{N}(\epsilon)\to e_{N}$ as $\epsilon\searrow 0$ (Lemma 4.4).

The function $g_{N}: \lambda \in (0,+\infty)\mapsto \lambda-({\lambda^{n+1}}/({n+1}))e_{N} \in\mathbb R$ reaches its maximum at $\lambda_{N,M}:=1/e_{N}^{1/n}$ . It follows therefore from (4.4) that

\[\gamma\leq \frac{\mathrm{lct}(X,\mathcal{I})}{g_{N}(\lambda_{N,M})}=\frac{n+1}{n}\mathrm{lct}(X,\mathcal{I})e_{N}^{1/n}.\]

Now $e_{N}\searrow e(X,\mathcal{I})$ as $N\to +\infty$ by Lemma 4.4, hence

\[\gamma\leq \frac{n+1}{n}\mathrm{lct}(X,\mathcal{I})e(X,\mathcal{I})^{1/n}.\]

Since this holds for any coherent ideal sheaf $\mathcal{I}$ supported at p, we obtain

\[\gamma\leq \frac{n+1}{n}\inf_{\mathcal{I}}\mathrm{lct}(X,\mathcal{I})e(X,\mathcal{I})^{1/n}=\frac{n+1}{n}\widehat{\mathrm{vol}}(X,p)^{1/n},\]

where the equality follows from Theorem 2.16.

5. Moser–Trudinger inequality

5.1 Uniform integrability versus Moser–Trudinger inequality

Recall that

\[\alpha(X,\mu_p):=\sup\bigg\{\alpha>0,\\sup_{\varphi\in\mathcal{F}_1(\Omega)}\int_{\Omega}e^{-\alpha \varphi}\,d\mu_{p}<+\infty\bigg\}.\]

This uniform integrability index is a local counterpart to Tian’s celebrated $\alpha$ -invariant, introduced in [Reference TianTia87] in the quest for Kähler–Einstein metrics on Fano manifolds. We refer to [Reference Demailly and KollàrDK01, Reference DemaillyDem09, Reference ZeriahiZer09, Reference Åhag, Cegrell, Kołodziej, Pham and ZeriahiACKPZ09, Reference Demailly and PhamDP14, Reference Guan and ZhouGZ15, Reference PhamPha18] for some contributions to the local study of analogous invariants.

In this section we prove Theorem A, which can be seen as a local analogue of [Reference Berman, Boucksom, Eyssidieux, Guedj and ZeriahiBBEGZ19, Proposition 4.13].

Theorem 5.1. One has $\gamma_{\rm crit}(X,p)\geq (({n+1})/{n})\alpha(X,\mu_p)$ .

When (X,p) is smooth then $\alpha(X,\mu_p)=n$ and this statement is equivalent (after an appropriate rescaling) to [Reference Berman and BerndtssonBB22, Theorem 1.5] and [Reference Guedj, Kolev and YeganefarGKY13, Theorem 9].

Together with Theorem 4.5, we would obtain the precise value

\[\gamma_{\rm crit}(X,p) \stackrel{?}{=} \frac{n+1}{n} \widehat{\mathrm{vol}}(X,p)^{1/n}\]

if we knew that $\alpha(X,\mu_p)=\widehat{\mathrm{vol}}(X,p)^{1/n}$ . We establish in § 5.2 the bound $\alpha(X,\mu_p) \leq \widehat{\mathrm{vol}}(X,p)^{1/n}$ and analyze the reverse inequality in § 5.3.

5.1.1 Entropy

We let $\mathcal{P}(\Omega)$ denote the set of probability measures on $\Omega$ . Given two measures $\mu,\nu\in\mathcal{P}(\Omega)$ , the relative entropy of $\nu$ with respect to $\mu$ is

\[H_{\mu}(\nu):=\int_{\Omega}\frac{d\nu}{d\mu}\log \frac{d\nu}{d\mu} \,d\mu=\int_{\Omega}\log\frac{d\nu}{d\mu} \,d\nu\]

if $\nu$ is absolutely continuous with respect to $\mu$ , and as $H_{\mu}(\nu):=+\infty$ otherwise.

Given $\mu\in\mathcal{P}(X)$ , the relative entropy $H_{\mu}(\cdot)$ is the Legendre transform of the convex functional $g \in \mathcal{C}^{0}(\Omega)\cap L^{\infty}(\Omega) \mapsto \log\int_{\Omega}e^{g}\,d\mu \in \mathbb R$ , i.e.

\[H_{\mu}(\nu)=\sup_{g\in\mathcal{C}^{0}(\Omega)\cap L^{\infty}(\Omega)}\bigg(\int_{\Omega}g\,d\nu-\log\int_{\Omega}e^{g}\,d\mu\bigg).\]

We shall need the following duality result.

Lemma 5.2 [Reference Berman, Boucksom, Eyssidieux, Guedj and ZeriahiBBEGZ19, Lemma 2.11]. Fix $\mu\in\mathcal{P}(\Omega)$ . Then

\[\log\int_{\Omega}e^{g}\,d\mu=\sup_{\nu\in\mathcal{P}(\Omega)}\bigg(\int_{\Omega}g \,d\nu-H_{\mu}(\nu)\bigg)\]

for each lower semi-continuous function $g:\Omega\to \mathbb R\cup \{+\infty\}$ .

Recall that we have normalized the adapted volume form so that $\mu_{p}\in\mathcal{P}(\Omega)$ .

Corollary 5.3. Fix $0<\alpha<\alpha(X,\mu_p)$ . Then there exists $C_{\alpha}>0$ such that

\[H_{\mu_{p}}(\nu)\geq -\alpha\int_{\Omega}\varphi \,d\nu -C_{\alpha}\]

for all $\varphi\in\mathcal{F}_1(\Omega)$ and for all $\nu\in\mathcal{P}(\Omega)$ such that $H_{\mu_{p}}(\nu)<+\infty$ .

Proof. This follows from Lemma 5.2 applied to $g=-\alpha \varphi$ and $\mu=\mu_p$ . By the definition of $\alpha(X,\mu_p)$ , we obtain $-\log\int_{\Omega}e^{-\alpha \varphi}\,d\mu_p \geq -C_{\alpha}$ .

This corollary shows, in particular, that $\mathcal{F}_{1}(\Omega)\subset L^{1}(\nu)$ for any probability measure $\nu\in\mathcal{P}(X)$ with finite $\mu_p$ -entropy. Since the measure $\nu$ is moreover non-pluripolar, the following result is a consequence of Theorem 3.7.

Proposition 5.4. Fix $\nu\in\mathcal{P}(\Omega)$ such that $H_{\mu_{p}}(\nu)<+\infty$ . Then there exists a unique $v\in \mathcal{F}_{1}(\Omega) \cap \mathcal{E}^1(\Omega)$ such that

\[\nu=(dd^{c}v)^{n}.\]

5.1.2 Proof of Theorem 5.1

The proof is similar to the derivation of the Moser–Trudinger inequality from Brezis–Merle inequality by Berman and Berndtsson, see [Reference Berman and BerndtssonBB22, § 4.2]. Fix $\varphi\in\mathcal{T}_{\phi}(\Omega)$ and $0<\alpha<\alpha(X,\mu_p)$ . By Lemma 5.2 for any $\epsilon>0$ there exists $\nu_{\epsilon}\in\mathcal{P}(\Omega)$ such that $H_{\mu_{p}}(\nu_{\epsilon})<+\infty$ and

(5.1) \begin{equation}\log\int_{\Omega}e^{-(({n+1})/{n})\alpha \varphi}\,d\mu_{p}\leq \epsilon-\frac{n+1}{n}\alpha\int_{\Omega}\varphi \,d\nu_{\epsilon}-H_{\mu_{p}}(\nu_{\epsilon}).\end{equation}

Proposition 5.4 ensures the existence of $v_{\epsilon}\in\mathcal{F}_{1}(\Omega)\cap \mathcal{E}^1(\Omega)$ such that $\nu_{\epsilon}=(dd^{c}v_{\epsilon})^{n}$ . It follows, moreover, from Corollary 5.3 thatd

(5.2) \begin{equation}H_{\mu_{p}}(\nu_{\epsilon})\geq -\alpha\int_{\Omega}v_{\epsilon} \,d\nu_{\epsilon}-C_{\alpha}.\end{equation}

Combining (5.1) and (5.2) we obtain

(5.3) \begin{equation}\log\int_{\Omega}e^{-(({n+1})/{n})\alpha \varphi}\,d\mu_{p}\leq \epsilon+C_{\alpha}-\frac{n+1}{n}\alpha\int_{\Omega}\varphi \,d\nu_{\epsilon}+\alpha\int_{\Omega}v_{\epsilon} \,d\nu_{\epsilon}.\end{equation}

We observe that

\begin{align*}& -\frac{n+1}{n}\alpha\int_{\Omega}\varphi \,d\nu_{\epsilon}+ \alpha\int_{\Omega}v_{\epsilon}\,d\nu_{\epsilon}= \frac{n+1}{n}\alpha \int_{\Omega}(v_{\epsilon}-\varphi) (dd^c v_{\epsilon})^n-\frac{\alpha}{n} \int_{\Omega} v_{\epsilon} (dd^c v_{\epsilon})^n \\&\quad \leq -\frac{n+1}{n}\alpha E_{\phi}(\varphi)+\frac{\alpha}{n} \bigg\{(n+1) E_{\phi}(v_{\varepsilon})- \int_{\Omega} v_{\epsilon} (dd^c v_{\epsilon})^n\bigg\}\end{align*}

by using Lemma 3.5 (the latter has been stated for functions in ${\mathcal T}_{\phi}(\Omega)$ , it easily extends to the class ${\mathcal F}_{1}(\Omega)\cap\mathcal{E}^1(\Omega)$ by approximation). Since $v_{\varepsilon} \leq \phi_0$ and $E_{\phi}(\phi_0)=0$ , the same lemma ensures

\[(n+1) E_{\phi}(v_{\varepsilon})- \int_{\Omega} v_{\epsilon} (dd^c v_{\epsilon})^n\leq -\int_{\Omega} \phi_0 (dd^c v_{\varepsilon})^n \leq -\inf_{\Omega} \phi_0,\]

using that $\nu_{\epsilon}=(dd^c v_{\epsilon})^n$ is a probability measure. Altogether this yields

\[\log \int_{\Omega}e^{-(({n+1})/{n})\alpha\varphi}\,d\mu_{p} \leq \epsilon+C_{\alpha} -\frac{\alpha}{n} \inf_{\Omega} \phi_0-\frac{n+1}{n}\alpha E_{\phi}(\varphi).\]

Letting $\epsilon\searrow 0$ we conclude that

\[\bigg(\int_{\Omega}e^{-(({n+1})/{n})\alpha\varphi}\,d\mu_{p}\bigg)^{\!\!{n}/({(n+1)\alpha}) }\leq C_{\alpha}' e^{-E_{\phi}(\varphi)}\]

for any function $\varphi\in\mathcal{T}_{\phi}(\Omega)$ . Thus, $\gamma_{\rm crit}(X,p) \geq(({n+1})/{n})\alpha(X,\mu_p)$ .

5.2 Upper bound on the $\alpha$ -invariant

Definition 5.5. We set

\[\tilde{\alpha}(X,\mu_p):=\inf\{c(\varphi),\ \varphi \in \mathcal{F}_1(\Omega)\},\]

where $c(\varphi):=\sup\{c>0;\ \int_{\Omega}e^{-c\varphi}\,d\mu_{p}<+\infty\}$ .

5.2.1 Bounding the $\alpha$ -invariant by the normalized volume

Proposition 5.6. One has $\alpha(X,\mu_p) \leq \tilde{\alpha}(X,\mu_p) \leq \widehat{\mathrm{vol}}(X,p)^{1/n}$ .

Proof. It follows from the definition that $\alpha(X,\mu_p) \leq \tilde{\alpha}(X,\mu_p)$ .

For any $\epsilon>0$ and $\mathcal{I}$ coherent ideal sheaf supported at 0, the function

\[\psi_{\mathcal{I},\epsilon}:=\psi_{\mathcal{I},\lambda,\epsilon},\quad \text{with}\ \lambda=\bigg(\frac{1-\epsilon}{e(X,\mathcal{I})}\bigg)^{\!\!1/n}\]

given by Lemma 5.7, belongs to $\mathcal{F}_{1}(\Omega)$ and yields

\[\tilde{\alpha}(X,\mu_p)\leq c(\psi_{\mathcal{I},\epsilon})=\frac{1}{(1-\epsilon)^{1/n}}\mathrm{lct}(X,\mathcal{I})e(X,\mathcal{I})^{1/n}.\]

The latter equality is a consequence of Proposition 4.1. We conclude the proof by taking the infimum over all ${\mathcal I}$ and letting $\epsilon\searrow 0$ .

Lemma 5.7. Let $\mathcal{I}$ be a coherent ideal sheaf supported at p. Then, for any $\lambda, \epsilon>0$ there exists a function $\psi_{\mathcal{I},\lambda,\epsilon}\in \mathcal{F}(\Omega)$ such that:

  1. (i) $\psi_{\mathcal{I},\lambda,\epsilon}=\lambda\log(\sum_{j=1}^{m}\lvert f_{j}\rvert^{2})$ near 0 for local generators $f_{1},\ldots,f_{m}$ of $\mathcal{I}$ ;

  2. (ii) $\lambda^{n}e(X,\mathcal{I}) \leq\int_{\Omega}(dd^{c}\psi_{\mathcal{I},\lambda,\epsilon})^{n} \leq\lambda^{n}e(X,\mathcal{I})+\epsilon$ .

Proof. Assume that $\phi_0$ is the maximal psh extension of $\phi$ to $\Omega$ , i.e. the largest psh function in $\Omega$ which lies below $\phi$ on $\partial \Omega$ . It satisfies $(dd^c \phi_0)^n=0$ in $\Omega$ .

Fix $f_{1},\ldots,f_{m}$ local generators of the ideal $\mathcal{I}$ and set $\psi_{\lambda}:=\lambda\log (\sum_{j=1}^{m}\lvert f_{j} \rvert^{2})$ . We can assume without loss of generality that the $f_j$ are well defined in $\Omega$ and normalized so that $\psi_{\lambda}\leq \phi_0-1$ in $\Omega$ . For $r>0$ , we consider

\[\varphi_{r}:=\sup \{u \in PSH(\Omega),\, u \leq \psi_{\lambda} \text{ in } B(r)\ \text{and}\, u \leq \phi_0 \text{ in } \Omega\}.\]

The corresponding family of psh functions is non-empty as it contains $\psi_{\lambda}$ . For $A>1$ large enough, the function

\[w_r=\left\{\begin{array}{ll}\psi_{\lambda} & \text{ in } B(r) ,\\\max(\psi_{\lambda}, A \rho +\phi_0) & \text{ in } \Omega \setminus B(r),\end{array}\right.\]

is psh and coincides with $A \rho+\phi_0$ near $\partial \Omega$ . It follows that:

  • $\varphi_r \in PSH(\Omega)$ with $\varphi_r =\phi$ on $\partial \Omega$ ;

  • $\varphi_r \equiv \psi_{\lambda}$ in B(r), hence $\lambda^{n}e(X,\mathcal{I}) \leq\int_{\Omega} (dd^{c} \varphi_r)^n$ ;

  • $(dd^{c} \varphi_r)^n=0$ in $\Omega \setminus \overline{B}(r)$ (balayage argument).

The family $r \mapsto \varphi_r$ increases, as $r>0$ decreases to 0, to some psh limit $\varphi$ whose Monge–Ampère measure $(dd^c \varphi)^n$ is concentrated at the origin. It follows from Bedford–Taylor continuity theorem that $(dd^c \varphi)^n $ is the weak limit of $(dd^c \varphi_r)^n \geq\lambda^{n}e(X,\mathcal{I}) \delta_0$ , hence $(dd^c \varphi)^n \geq \lambda^{n}e(X,\mathcal{I})\delta_0$ . Conversely, $\psi_{\lambda} \leq \varphi$ near 0, hence Demailly’s comparison theorem ensures that

\[(dd^c \varphi)^n(0) \leq (dd^c \psi_{\lambda})^n(0) \leq \lambda^{n}e(X,\mathcal{I}),\]

whence equality. Thus, $\phi_{\mathcal{I},\lambda,\epsilon}:=\varphi_{r_{\varepsilon}}$ satisfies the required properties.

5.2.2 Normalized volume versus uniform integrability

Proposition 5.8. One has $\tilde{\alpha}(X,\mu_p)= \widehat{\mathrm{vol}}(X,p)^{1/n}$ .

We refer the reader to Appendix A for a more algebraic approach based on [Reference Boucksom, de Fernex and FavreBdFF12], which moreover provides a slightly stronger result.

When (X,p) is smooth, it follows from [Reference Demailly and KollàrDK01] that $\tilde{\alpha}(X,\mu_p)={\alpha}(X,\mu_p)$ . The situation is, however, more subtle in the singular context (see § 5.3.2).

Proof. By Proposition 5.6 it suffices to show that $\tilde{\alpha}(X,\mu_p)\geq\widehat{{\rm vol}}(X,p)^{1/n}$ , i.e. $\int_\Omega e^{-\alpha \varphi}\,d\mu_p<+\infty$ for all $\varphi\in \mathcal{F}_1(\Omega)$ and $\alpha<\widehat{\mathrm{vol}}(X,p)^{1/n}=\inf_{\mathcal{I}} {\rm lct}(X,\mathcal{I})^n e_p({\mathcal I})$ .

In a log resolution $\pi:\tilde{\Omega}\to \Omega$ , this boils down to $\int_{\tilde{\Omega}}e^{-\alpha \varphi\circ \pi}\prod_{i=1}^M \lvert s_i\rvert_{h_i}^{2a_i}\,dV<+\infty$ , where $s_i$ are holomorphic sections defining simple normal crossing exceptional divisors $E_1,\ldots,E_M$ , $K_{\tilde{\Omega}/\Omega}=\sum_{j=1}^M a_i E_i$ and where dV is a smooth volume form. The log terminal condition ensures that $a_i>-1$ for all $i=1,\ldots,M$ .

As $\alpha<\widehat{\mathrm{vol}}(X,p)^{1/n}\leq n$ , the integrability condition is equivalent to show that for any point $x\in \bigcup_{i=1}^M E_i$ there exists a small ball B(x,r) such that

(5.4) \begin{equation}\int_{B(x,r)}e^{-\alpha \varphi\circ \pi}\prod_{i=1}^M \lvert s_i\rvert_{h_i}^{2a_i} \,dV<+\infty.\end{equation}

Set $U:=\sum_{i: a_i\geq 0} a_i\log \lvert s_i \rvert_{h_i}^2$ , $V:=\alpha \varphi\circ \pi$ and $W:=-\sum_{i: a_i<0}a_i\log \lvert s_i\rvert_{h_i}^2$ . By [Reference Berman, Boucksom and JonssonBBJ21, Theorem B.5] the condition (5.4) holds if and only if there exists $\epsilon>0$ such that

(5.5) \begin{equation}\nu(U\circ g, F)+A_{\tilde{\Omega}}(F)\geq (1+\epsilon) \nu(V\circ g, F)+(1+\epsilon)\nu(W\circ g, F)\end{equation}

for any F prime divisor over $\tilde{\Omega}$ with center in a small ball $\overline{B(x,r')}\subset B(x,r)$ , i.e. $F\subset \Omega'$ for $g:\Omega'\to\tilde{\Omega}$ modification. Observe that

\begin{align*}\nu(U\circ g,F)+A_{\tilde{\Omega}}(F)-\nu(W\circ g, F)&= \mathrm{ord}_F( g^*K_{\tilde{\Omega}/\Omega})+1+\mathrm{ord}_F( K_{\Omega'/\tilde{\Omega}}) \\&= 1+\mathrm{ord}_F (K_{\Omega'/\Omega})=A_{\Omega}(F).\end{align*}

Thus, (5.5) becomes

(5.6) \begin{equation}\alpha(1+\epsilon)\leq \frac{A_{\Omega}(F)-\epsilon\nu(W\circ g,F)}{\nu(\varphi\circ \pi\circ g,F)}.\end{equation}

As $a_i>-1$ for all i, [Reference Berman, Boucksom and JonssonBBJ21, Theorem B.5] ensures the existence of $a>0$ such that $A_{\tilde{\Omega}}(F)\geq (1+a)\nu(W\circ g,F) $ for any prime divisor F over $\tilde{\Omega}$ as above. Thus,

\[A_{\tilde{\Omega}}(F)\leq A_\Omega(F)+\nu(W\circ g, F)\leq \frac{1}{1+a}A_{\tilde{\Omega}}(F)+A_\Omega(F),\]

and $\nu(W\circ g,F)\leq ({1}/({1+a}))A_{\tilde{\Omega}}(F)\leq ({1}/{a})A_\Omega(F)$ . Therefore, (5.6) holds if

(5.7) \begin{equation}\alpha(1+\epsilon)\leq \frac{a-\epsilon}{a}\frac{A_\Omega(F)}{\nu(\varphi\circ \pi\circ g,F)}.\end{equation}

Since $\varphi \in {\mathcal F}_1(\Omega)$ , it follows from the comparison theorem of Demailly [Reference DemaillyDem85, Theorem 4.2] that for a coherent ideal sheaf $\mathcal{I}$ supported at $p\in \Omega$ ,

(5.8) \begin{equation}1\geq \int_\Omega (dd^c \varphi)^n\geq \nu_\mathcal{I}(\varphi,p)^n\int_{\mathbb C^N}(dd^cf_\mathcal{I})^n\wedge [X]=\nu_\mathcal{I}(\varphi,p)^n e_p(\mathcal{I}),\end{equation}

where $f_\mathcal{I}= \log (\sum_i\lvert f_i\rvert^2)$ for generators $\{f_i\}_i$ of $\mathcal{I}$ and

\[\nu_\mathcal{I}(\varphi,p):=\sup\{s>0: \varphi\leq s f_\mathcal{I}+ O(1)\}=\min_{G}\frac{\nu(\varphi\circ\pi\circ g,G)}{\mathrm{ord}_G \mathcal{I}},\]

where $\pi\circ g: \Omega'\to \Omega$ is a log resolution for $\mathcal{I}$ and the minimum is over all exceptional divisors of $\Omega'\to \Omega$ . Lemma 5.9 ensures that for any prime divisor F and $\delta>0$ there exists an ideal $\mathcal{I}$ such that

\begin{align*}\frac{A_\Omega(F)}{\nu(\varphi\circ \pi\circ p,F)}&\geq (1-\delta)\frac{A_\Omega(F)}{\mathrm{ord}_F\mathcal{I}}(\nu_\mathcal{I}(\varphi,p))^{-1}\geq(1-\delta) \frac{A_\Omega(F)}{\mathrm{ord}_F\mathcal{I}} e_p(\mathcal{I})^{1/n} \\&\geq (1-\delta){\rm lct}(X,\mathcal{I})e_p(\mathcal{I})^{1/n}\geq (1-\delta)\widehat{\mathrm{vol}}(X,p)^{1/n}.\end{align*}

Thus, (5.7) holds if $\alpha(1+\epsilon)\leq(({(a-\epsilon)(1-\delta)})/{a})\widehat{\mathrm{vol}}(X,p)^{1/n}$ , concluding the proof.

Lemma 5.9. Fix $\varphi\in \mathcal{F}_1(\Omega)$ and $F\subset \Omega'$ prime divisor such that $\pi\circ g(F)=p$ . For any $\epsilon>0$ , there exists a coherent ideal sheaf $\mathcal{I}$ supported at p such that

\[\nu_\mathcal{I}(\varphi,p)\geq(1-\epsilon)\frac{\nu(\varphi\circ \pi\circ g,F)}{\mathrm{ord}_F (\mathcal{I})}.\]

Proof. Let $c:=\nu(\varphi\circ \pi\circ g,F)$ and for $c^{\prime}\in \mathbb Q, c^{\prime}\leq c$ , set

\[\mathcal{A}_{mc^{\prime}}(F):=\{f\in \mathcal{O}_{X,p}: \mathrm{ord}_F (f\circ \pi\circ g)\geq mc^{\prime}\}\]

for $m\in \mathbb N$ divisible enough. Then $\mathcal{A}_{mc^{\prime}}(F)$ is an ideal sheaf and

(5.9) \begin{equation}\limsup_{m\to +\infty} \frac{\mathrm{ord}_F (\mathcal{A}_{mc^{\prime}}(F))}{m}=c^{\prime}.\end{equation}

In particular, if $\varphi_{mc^{\prime}}\in \mathrm{PSH}(B(p,r))$ has algebraic singularities along $\mathcal{A}_{mc^{\prime}}(F)$ , then for any $\epsilon>0$ , $\varphi_{mc^{\prime}}$ is less singular than $({mc^{\prime}}/({c-\epsilon}))\varphi$ around p if $m\geq m_1(\epsilon)\gg 1$ . For any G exceptional divisor on $\Omega'$ and $m\geq m_1(\epsilon)$ we infer

(5.10) \begin{equation}\frac{\nu(\varphi\circ \pi\circ g,G)}{\mathrm{ord}_{G}(\mathcal{A}_{mc^{\prime}}(F))/m}=\frac{\nu(\varphi\circ \pi\circ g,G)}{\nu(\varphi_{mc^{\prime}}\circ \pi\circ g,G)/m}\geq \frac{c-\epsilon}{c^{\prime}}.\end{equation}

On the other hand, (5.9) implies that there exists $m_0(\epsilon)\geq m_1(\epsilon)\gg 1$ with

(5.11) \begin{equation}\frac{\nu(\varphi\circ \pi\circ g,F)}{\mathrm{ord}_F(\mathcal{A}_{m_0c^{\prime}}(F))/m_0}\leq \frac{c}{c^{\prime}-\epsilon}.\end{equation}

Combining (5.10) and (5.11) we obtain

\begin{align*}\min_{G}\frac{\nu(\varphi\circ \pi\circ g, G)}{\mathrm{ord}_G(\mathcal{A}_{m_0c^{\prime}}(F))/m_0}\geq \frac{c-\epsilon}{c^{\prime}} &\geq \bigg(1-\epsilon\frac{c+c^{\prime}}{cc^{\prime}}\bigg)\frac{c}{c^{\prime}-\epsilon} \\&\geq \bigg(1-\epsilon\frac{c+c^{\prime}}{cc^{\prime}}\bigg)\frac{\nu(\varphi\circ \pi\circ g,F)}{\mathrm{ord}_F(\mathcal{A}_{m_0c^{\prime}}(F))/m_0}.\end{align*}

Since $c^{\prime}$ and $\epsilon$ are arbitrary, and $x \mapsto f(x)=({c+x})/{cx}$ is decreasing, we deduce that for any $\epsilon>0$ there exists $c^{\prime}\in\mathbb Q$ and $m_0=m_0(c,c^{\prime},\epsilon)$ such that

\[\nu_{\mathcal{A}_{m_0c^{\prime}}(F)}(\varphi,p)\geq (1-\epsilon)\frac{\nu(\varphi\circ \pi\circ g,F)}{\mathrm{ord}_F(\mathcal{A}_{m_0c^{\prime}}(F))}.\]

Setting $\mathcal{A}:=\mathcal{A}_{m_0c^{\prime}}(F)$ concludes the proof.

5.3 Lower bounds on the $\alpha$ -invariant

We provide in this section two effective (but not sharp) lower bounds on $\alpha(X,\mu_p)$ .

5.3.1 Using projections on n-planes

A result of Skoda ensures that $e^{- \varphi}$ is integrable if the Lelong numbers of $ \varphi$ are small enough (see [Reference Guedj and ZeriahiGZ17, Theorem 2.50]). This has been largely extended by Demailly and Zeriahi who provided uniform integrability results for functions $\varphi \in {\mathcal F}_1(\Omega)$ (see [Reference DemaillyDem09, Reference Åhag, Cegrell, Kołodziej, Pham and ZeriahiACKPZ09]). In this section, we extend these results to our singular setting.

Theorem 5.10. One has

\[\alpha(X,\mu_p) \geq \frac{n}{{\rm mult}(X,p)^{1-1/n}} \frac{{\rm lct}(X,p)}{1+{\rm lct}(X,p)}.\]

Proof. Recall that $\mu_p=fdV_X$ with $f \in L^r(dV_X)$ . The exponent $r>1$ has been estimated in Lemma 2.14. Using Hölder inequality, we thus obtain

\[\alpha(X,\mu_p) \geq \frac{{\rm lct}(X,p)}{1+{\rm lct}(X,p)} \alpha(\Omega,dV_X).\]

The remainder of the proof consists of establishing the lower bound

\[\alpha(\Omega,dV_X) \geq\frac{n}{{\rm mult}(X,p)^{1-1/n}}.\]

Recall that $dV_X=\omega_{\rm eucl}^n \wedge [X]$ , where $\omega_{\rm eucl}$ denotes the euclidean Kähler form. Thus, $dV_X=\sum_I (\pi_I)^*(dV_I)$ , where $I=(i_1,\ldots,i_n)$ is a n-tuple, $\pi_I:\mathbb C^N\rightarrow \mathbb C_I^n$ denotes the linear projection on $\mathbb C_I^n$ , and $dV_I$ is the euclidean volume form on $\mathbb C_I^n$ . We choose coordinates in $\mathbb C^N$ so that each projection map $\pi_I: \Omega\rightarrow \Omega_I \subset \mathbb C^n$ is proper. For $\varphi \in {\mathcal F}_1(\Omega)$ , we obtain

\[\int_{\Omega} e^{-\alpha \varphi} \,dV_X = \sum_I \int_{\Omega_I} (\pi_I)_* (e^{-\alpha \varphi}) \,dV_I\leq {\rm mult}(X,p) \sum_I \int_{\Omega_I} e^{-\alpha (\pi_I)_* \varphi} \,dV_I.\]

We assume here, without loss of generality, that $\varphi \leq 0$ , and use the (sub-optimal) inequality $(\pi_I)_* (e^{-\alpha \varphi}) \leq {\rm mult}(X,p) e^{-\alpha (\pi_I)_* \varphi}$ . The function $\varphi_I:=(\pi_I)_* \varphi$ is psh in $\Omega_I=\pi_I(\Omega)$ , with boundary values $(\pi_I)_*(\phi)$ . We claim that

(5.12) \begin{equation}\int_{\Omega_I} (dd^c \varphi_I)^n \leq {\rm mult}(X,p)^{n-1}.\end{equation}

Once this is established, it follows from the main result of [Reference Åhag, Cegrell, Kołodziej, Pham and ZeriahiACKPZ09] that for all $0<\varepsilon$ small enough, there exists $C_{\varepsilon}>0$ independent of $\varphi$ such that

\[\int_{\Omega_I} e^{- (({n-\varepsilon})/{{\rm mult}(X,p)^{1-1/n}}) \varphi_I} \,dV_I \leq C_{\varepsilon},\]

which yields the desired lower bound $\alpha(\Omega,dV_X) \geq ({n}/({{\rm mult}(X,p)^{1-1/n}}))$ .

It remains to check (5.12). We decompose $\varphi_I(z)=\sum_{i=1}^m \varphi(x_i)$ , where $m={\rm mult}(X,p)$ and $x_1,\ldots,x_m$ denote the preimages of z counted with multiplicities. The assumption on the Monge–Ampère mass of $\varphi$ reads

\[\sum_{i=1}^m \int (dd^c \varphi)^n(x_i) \leq 1.\]

We set $a_i^n:=\int (dd^c \varphi)^n(x_i)$ and use [Reference CegrellCeg04, Corollary 5.6] to estimate

\[\int (dd^c \varphi_I)^n = \sum_{i_1,\ldots,i_n=1}^m \int dd^c \varphi(x_{i_1}) \wedge\cdots \wedge dd^c \varphi (x_{i_n}) \leq \sum_{i_1,\ldots,i_n=1}^m a_{i_1}\cdots a_{i_n}=\bigg(\sum_{i}^m a_i\bigg)^{\!\!n}.\]

The latter sum is maximized when $a_1=\cdots=a_m=m^{-1/n}$ , yielding (5.12).

Example 5.11. Let $X=\{z \in \mathbb C^{n+1},\ F(z)=0\}$ be the $A_k$ -singularity, where $F(z)=z_0^{k+1}+z_1^2+\cdots z_n^2$ . Arguing as we have done for the ODP $(k=1)$ , one can check that $\mu_p \sim{dV_X}/{\|F'\|^2}$ so that ${\rm mult}(X,p)=2$ and ${\rm lct}(X,p)=n-2+{2}/({k+1})$ . Now

\[\widehat{\mathrm{vol}}(A_k,p)^{1/n}=\begin{cases}2^{1/n}\bigg(\dfrac{n-2}{n-1}\bigg)^{\!\!1-1/n}n & \mathrm{if}\ \dfrac{k+1}{2}\geq \dfrac{n-1}{n-2},\\[16pt](k+1)^{1/n}\bigg(\dfrac{(n-2)(k+1)+2}{k+1}\bigg) & \mathrm{if}\ \dfrac{k+1}{2}<\dfrac{n-1}{n-2},\end{cases}\]

as computed by Li in [Reference LiLi18, Example 5.3]. For $n \gg 1$ , the lower bound provided by Theorem 5.10 is thus short of a factor $2={\rm mult}(X,p)$ by comparison with the expected lower bound $\widehat{\mathrm{vol}}(A_k,p)^{1/n}$ .

5.3.2 Using resolutions

Proposition 5.12. Let $\pi:\tilde{\Omega}\to \Omega$ be a resolution of singularities with simple normal crossing, and let $\{a_i\}_{i=1,\ldots,M}$ be the discrepancies. Then

\[\alpha(X,\mu_p)\geq \frac{\widehat{\mathrm{vol}}(X,p)^{1/n}}{1+(\max_i a_i)_+}.\]

In particular, if the singularity is ‘admissible’, then $\alpha(X,\mu_p)=\widehat{\mathrm{vol}}(X,p)^{1/n}$ .

Following [Reference Li, Tian and WangLTW21, Definition 1.1] we say here that (X,p) is an admissible singularity if there exists a resolution $\pi:\tilde{X} \rightarrow X$ (with snc exceptional divisor $E=\sum_jE_j$ and $\pi$ -ample divisor $-\sum b_j E_j$ , $b_j \in \mathbb Q^+$ ) such that the discrepancies $a_i \in(-1,0]$ are all non-positive. Recall that:

  • any two-dimensional log terminal singularity is admissible;

  • the vertex of the affine cone over a Fano manifold embedded in a projective space by the linear system associated to a multiple of the anticanonical bundle is admissible (cf. the proof of Proposition 3.17);

  • (X,p) is admissible if it is $\mathbb Q$ -factorial and admits a crepant resolution.

Theorem B from the introduction follows from the combination of Proposition 5.6, Theorem 5.10 and Proposition 5.12.

Proof. We seek $\alpha>0$ such that

(5.13) \begin{equation}\sup_{\varphi\in \mathcal{F}_1(\Omega)}\int_{\tilde{\Omega}} e^{-\alpha \varphi\circ \pi}\prod_{i=1}^M \lvert s_i\rvert_{h_i}^{2a_i} \,dV<+\infty.\end{equation}

If all the $a_i$ are non-positive we can use [Reference Demailly and KollàrDK01, Main Theorem] to show that $\alpha(X,\mu_p)=\tilde{\alpha}(X,\mu_p)$ , hence $\alpha(X,\mu_p)=\widehat{\mathrm{vol}}(X,p)^{1/n}$ by Proposition 5.8. Indeed, assume that there exists $\gamma>0$ such that $\alpha(X,\mu_p)<\gamma<\tilde{\alpha}(X,\mu_p)$ . By definition, we can find $\psi_j \in \mathcal{F}_1(\Omega)$ such that $\int_{\Omega} e^{-\gamma \psi_j}\,d\mu_p \rightarrow +\infty$ . Extracting and relabelling, we can assume that $\psi_j \rightarrow \psi$ in $L^1$ with $c(\psi)>\gamma$ . The psh functions $\varphi_j=\psi_j+\gamma^{-1}\sum_{i=1}^M(-a_i) \log\lvert s_i\rvert_{h_i}^2$ converge to $\varphi=\psi+\gamma^{-1}\sum_{i=1}^M (-a_i) \log \lvert s_i\rvert_{h_i}^2$ in $L^1(\tilde{\Omega})$ and $c(\varphi)>\gamma$ . It follows therefore from [Reference Demailly and KollàrDK01, Theorem 0.2.2] that

\[\int_{\Omega} e^{-\gamma \psi_j} \,d\mu_p=\int_{\tilde{\Omega}} e^{-\gamma \varphi_j} \,dV \longrightarrow\int_{\tilde{\Omega}} e^{-\gamma \varphi} \,dV <+\infty,\]

contradicting the assumption $\int_{\Omega} e^{-\gamma \psi_j} \,d\mu_p \rightarrow +\infty$ .

In general, we set $U:=\sum_{i:a_i>0}a_i\log \lvert s_i\rvert^2_{h_i}$ and $W:=-\sum_{i:a_i\leq 0}a_i \log \lvert s_i \rvert^2_{h_i}$ . Using [Reference Demailly and KollàrDK01, Main Theorem], we obtain

(5.14) \begin{equation}\alpha(X,\mu_p)\geq \inf_{\varphi\in \mathcal{F}_1(\Omega)}c_W(\varphi\circ \pi),\end{equation}

where

\[c_W(\varphi\circ\pi):=\sup\bigg\{\alpha>0: \int_{\tilde{\Omega}}e^{-\alpha\varphi\circ \pi- W}\,dV<+\infty\bigg\}\]

is the twisted complex singularity exponent. It then remains to estimate $c_W(\varphi\circ\pi)$ for a fixed $\varphi\in \mathcal{F}_1(\Omega)$ . As $\pi^*d\mu_p=e^{U-W}dV$ , Hölder inequality yields

(5.15) \begin{equation}\int_{\tilde{\Omega}}e^{-\alpha \varphi\circ \pi-W}\,dV\leq \bigg(\int_{\tilde{\Omega}}e^{-p'\alpha\varphi\circ \pi}\pi^{*}\,d\mu_p\bigg)^{\!\!1/p'}\bigg(\int_{\tilde{\Omega}}e^{(1-q')U-W}\,dV\bigg)^{\!\!1/q'}.\end{equation}

Set $A:=(\max_i a_i)_+>0$ . The second factor on the right-hand side of (5.15) is finite for any $q'<(({A+1})/{A})$ , while the first factor on the right-hand side gives the condition $p'\alpha< \tilde{\alpha}=\widehat{\mathrm{vol}}(X,p)^{1/n}$ . We infer $ c_W(\varphi\circ \pi)\geq({\widehat{\mathrm{vol}}(X,p)^{1/n}})/({1+A})$ , which concludes the proof.

As the proof shows, the main obstruction to proving the equality $\alpha(X,\mu_p)=\tilde{\alpha}(X,\mu_p)=\widehat{\mathrm{vol}}(X,p)^{1/n}$ is the lack of a Demailly–Kollàr result on complex spaces. Resolving the singularities, one ends up with a twisted version of Demailly and Kollàr’s problem on a smooth manifold. It is known that the general form of such a problem has a negative answer [Reference PhamPha14, Remark 1.3].

6. Ricci inverse iteration

In this section, we prove Theorem C from the introduction. The strategy is similar to that of [Reference Guedj, Kolev and YeganefarGKY13, Theorem 1], with a singular twist.

We fix $\gamma < \gamma_{\rm crit}(X,p)$ and consider, for $j \in \mathbb N$ , the sequence of functions $\varphi_j\in PSH(\Omega)$ defined by induction as follows: pick $\varphi_0 \in {\mathcal T}_{\phi}^{\infty}(\Omega)$ a smooth initial data, and let $\varphi_{j+1} \in PSH(\Omega) \cap {\mathcal{C}}^0(\overline{\Omega}) \cap {\mathcal C}^{\infty}(\overline{\Omega} \setminus \{p\}) $ be the unique solution to

\[(dd^c \varphi_{j+1})^n=\frac{e^{-\gamma \varphi_j} \mu_p}{\int_{\Omega} e^{-\gamma \varphi_j} \mu_p}\]

with boundary values ${\varphi_{j+1}}_{|\partial \Omega}=\phi$ . The existence and regularity of $\varphi_j$ off the singular locus follows from [Reference FuFu23, Theorem 1.4], while the continuity of $\varphi_j$ near p is a consequence of [Reference Guedj, Guenancia and ZeriahiGGZ23, Theorem A].

We are going to establish uniform a priori estimates on arbitrary derivatives of the $\varphi_j$ in $\overline{\Omega} \setminus \{p\}$ , thus $(\varphi_j)$ admits ‘smooth’ cluster values. We show that the functional $F_{\gamma}$ is constant on the set ${\mathcal K}$ of these cluster points, so that any such $\psi$ is a solution of the Monge–Ampère equation

\[(dd^c \psi)^n=\frac{e^{-\gamma \psi} \mu_p}{\int_{\Omega} e^{-\gamma \psi} \mu_p}\]

with boundary values ${\psi}_{|\partial \Omega}=\phi$ .

6.1 Uniform estimates

Proposition 6.1. There exists $C_0>0$ such that $\|\varphi_j\|_{L^{\infty}(\Omega)} \leq C_0$ for all $j \in \mathbb N$ .

This uniform estimate relies crucially on a technique introduced by Kolodziej in [Reference KołodziejKol98], which has been extended to this singular setting in [Reference Guedj, Guenancia and ZeriahiGGZ23].

Proof. We assume without loss of generality that $\phi_0$ is the maximal psh extension of $\phi$ in $\Omega$ . In particular, $\varphi_j \leq \phi_0$ for all $j \in \mathbb N$ , and $E_{\phi}(\varphi_j) \leq E_{\phi}(\phi_0)=0$ . Our task is to establish a uniform lower bound $\varphi_j \geq -C_0$ .

The assumption $\gamma < \gamma_{\rm crit}(X,p)$ ensures, by Lemma 3.13, that the functional $F_{\gamma}$ is coercive, and in particular there exist $0 <a<1$ and $0<b$ such that

\[F_{\gamma}(\varphi_j) \leq a E_{\phi}(\varphi_j)+b\]

for all $j \in \mathbb N$ . It follows from [Reference Guedj, Kolev and YeganefarGKY13, Proposition 12] (exactly the same proof applies here) that $j \mapsto F_{\gamma}(\varphi_j)$ is increasing, hence

\[F_{\gamma}(\varphi_0) \leq F_{\gamma}(\varphi_j) \leq a E_{\phi}(\varphi_j) +b \leq b,\]

showing that the energies $(E_{\phi}(\varphi_j))$ are uniformly bounded, $-b' \leq E_{\phi}(\varphi_j) \leq0$ .

The corresponding family ${\mathcal G}_{b'}$ of psh functions with $\phi$ -boundary values and energy bounded by b’ is compact, and all its members have zero Lelong number at all points in $\Omega$ (see Theorem 3.7). Passing through a resolution, one can thus invoke Skoda’s uniform integrability theorem [Reference Guedj and ZeriahiGZ17, Theorem 2.50] to conclude that the densities $e^{-\gamma \varphi_j}$ are uniformly in $L^r(dV_X)$ for any $r>1$ .

Now $\mu_p=fdV_X$ with $f \in L^{1+\varepsilon}$ for some $\varepsilon>0$ since (X,p) is log terminal. Hölder inequality thus ensures that the densities $g_j:={e^{-\gamma \varphi_j} f}/{\int_{\Omega}e^{-\gamma \varphi_j} \,d\mu_p}$ are uniformly in $L^{1+\varepsilon'}(dV_X)$ for some $0<\varepsilon'<\varepsilon$ .

It therefore follows from [Reference Guedj, Guenancia and ZeriahiGGZ23, Proposition 1.8] (an extension of the main result of [Reference KołodziejKol98] to the setting of pseudoconvex subsets of a singular complex space) that the $\varphi_j$ are uniformly bounded.

6.2 ${\mathcal C}^2$ -estimates

In this section we establish the following a priori estimates.

Proposition 6.2. For any compact subset K of $\overline{\Omega} \setminus \{p\}$ , there exists a constant $C_2(K)>0$ such that for all $j \in \mathbb N$ ,

\[0 \leq \sup_K \Delta_{\omega_X} \varphi_j \leq C_2(K).\]

Here $\Delta_{\omega_X}h:=n (({dd^c h \wedge \omega_X^{n-1}})/{\omega_X^n})$ denotes the Laplace operator with respect to the Kähler form $\omega_X$ . Such an estimate goes back to the regularity theory developed in [Reference Caffarelli, Kohn, Nirenberg and SpruckCKNS85]. The strategy of the proof is similar to that of [Reference Guedj, Kolev and YeganefarGKY13, Theorem 15], with a twist due to the presence of the singular point p.

Proof. To obtain these estimates, one considers a resolution of the singularity $\pi:\tilde{\Omega}\rightarrow \Omega$ . We let $E=\bigcup_{\ell=1}^m E_{\ell}$ denote the exceptional divisor and let:

  • $s_{\ell}$ denote a holomorphic section of ${\mathcal O}(E_{\ell})$ such that $E_{\ell}=(s_{\ell}=0)$ ;

  • $b_{\ell}$ be positive rational numbers such that $-\sum_{\ell} b_{\ell}E_{\ell}$ is $\pi$ -ample;

  • $h_{\ell}$ denote a smooth hermitian metric of ${\mathcal O}(E_{\ell})$ and $K \gg 1$ such that

Observe that the function $\rho':=K \rho \circ \pi+\sum_{\ell=1}^m b_{\ell}\log|s_{\ell}|_{h_{\ell}}^2$ is strictly psh in $\tilde{\Omega}$ , with $dd^c \rho' \geq \beta$ and $\rho'(z) \rightarrow -\infty$ as $z \rightarrow E$ .

Recall that $\pi^* \mu_p=\Pi_{\ell=1}^m |s_{\ell}|^{2a_{\ell}} dV_{\tilde{\Omega}}$ with $a_{\ell}>-1$ , and set $|s|^2=\Pi_{\ell=1}^m |s_{\ell}|^{2b_{\ell}}$ . We are going to show that there exist uniform constants $C_2>0, m \in \mathbb N$ such that

(6.1) \begin{equation}0 \leq |s|^{2m} |\Delta_{\beta} \varphi_{j}| (z) \leq C_2\end{equation}

for all $j \in \mathbb N,\ z \in \overline{\Omega}$ , from which Proposition 6.2 follows. Slightly abusing notation, we still denote here by $\varphi_j$ the function $\varphi_j \circ \pi$ .

We approximate $\varphi_j$ by the smooth solutions $\varphi_{j,\varepsilon}$ of the Dirichlet problem

(6.2) \begin{equation}\begin{cases}\displaystyle(\varepsilon \beta+ dd^c\varphi_{j+1,\varepsilon})^n=\frac{e^{-\gamma \varphi_{j,\varepsilon}}\prod_{l=1}^m (\lvert s_l \rvert_{h_l}^2+\varepsilon^2)^{a_l}}{c_j} \, dV_{\tilde{\Omega}},\\\varphi_{j+1,\varepsilon|\partial \tilde{\Omega}}=\phi,\end{cases}\end{equation}

with $\varphi_{0,\varepsilon}=\varphi_0$ and $c_j=\int_{\Omega} e^{-\gamma \varphi_j} \,d\mu_p$ . We are going to establish a priori estimates on these smooth approximants, whose existence is guaranteed by [Reference Guan and LiGL10, Theorem 1.1]. We then show that $\varphi_{j,\varepsilon}$ converges to $\varphi_j$ as $\varepsilon$ decreases to zero.

Step 1. We first claim that for all $j,\varepsilon$ ,

(6.3) \begin{equation} \sup_{\partial \tilde{\Omega}} |\nabla \varphi_{j+1,\varepsilon}| \leq A_{1,j,\varepsilon},\end{equation}

where $A_{1,j,\varepsilon}>0$ only depends on an upper-bound on $\|\varphi_{j,\varepsilon}\|_{L^{\infty}(\tilde{\Omega})}$ .

Let $\Phi^-$ be a smooth psh extension of $-\phi$ to a neighborhood of $\overline{\Omega}$ . Observe that $\varphi_{j+1,\varepsilon}+\Phi^-\circ \pi$ is $\beta$ -psh in $\tilde{\Omega}$ , with zero boundary values. Thus, $\varphi_{j+1,\varepsilon}+\Phi^-\circ \pi \leq u$ , where u is the smooth solution in $\tilde{\Omega}$ to the Laplace equation $\Delta_{\beta} u=-n$ with zero boundary values. We infer $\varphi_{j+1,\varepsilon} \leq\psi_1:=u-\Phi^-\circ \pi$ in $\tilde{\Omega}$ .

We now construct a psh function $\psi_2 \leq \varphi_{j+1,\varepsilon}$ with $\phi$ -boundary values and such that $\sup_{\partial \tilde{\Omega}} |\psi_2|$ is controlled from above by $\|\varphi_{j,\varepsilon}\|_{L^{\infty}(\tilde{\Omega})}$ . The upper bound on $\sup_{\partial \tilde{\Omega}}|\nabla \varphi_{j+1,\varepsilon}|$ thus follows from the inequalities $\psi_2 \leq \varphi_{j,\varepsilon} \leq \psi_1$ .

Recall that $\pi^* \mu_p =\Pi_{\ell=1}^m |s_{\ell}|^{2a_{\ell}} dV_{\tilde{\Omega}}$ . We let $P\subset [1,m]$ denote the subset of indices such that $-1<a_{\ell} <0$ . For $\delta>0$ small enough, we observe that $v:=\rho'+\delta \sum_{\ell \in P} |s_{\ell}|^{2\delta}$ is strictly psh in $\tilde{\Omega}$ and satisfies, in $\tilde{\Omega} \setminus E$ ,

\[dd^c v \geq c\bigg\{\beta+\sum_{\ell \in P}\frac{i \,ds_{\ell} \wedge d\overline{s_{\ell}}}{|s_{\ell}|^{2(1-\delta})}\bigg\}\]

for some $c>0$ , hence $(dd^c v)^n \geq c^{\prime} \pi^* \mu_p$ . Replacing v by $\lambda_{j,\varepsilon} v$ , we obtain

\[(\varepsilon \beta+dd^c \lambda_{j,\varepsilon} v)^n \geq \lambda_{j,\varepsilon}^n (dd^c v)^n \geq\frac{e^{-\gamma \varphi_{j,\varepsilon}} \prod_{l=1}^m (\lvert s_l\rvert_{h_l}^2+\varepsilon^2)^{a_l}}{c_j} \, dV_{\tilde{\Omega}},\]

for some $\lambda_{j,\varepsilon}>0$ which only depends on an upper bound on $\|\varphi_{j,\varepsilon}\|_{L^{\infty}(\tilde{\Omega})}$ . In other words, $\lambda_{j,\varepsilon} v$ is a subsolution to the Monge–Ampère equation in $\tilde{\Omega} \setminus E$ .

We modify $\lambda_{j,\varepsilon} v$ near $\partial \tilde{\Omega}$ to produce a subsolution with the right boundary values. Let $\chi$ be a cut-off function which is 1 near E and has compact support in $\tilde{\Omega}$ . The function $\psi_2=\chi \lambda_{j,\varepsilon} v+(1-\chi) \phi_0+A \rho \circ \pi$ satisfies all our requirements for $A>0$ large enough. Note, however, that it is only locally bounded in $\tilde{\Omega} \setminus E$ .

Finally, consider $\max(\psi_2,\varphi_{j,\varepsilon})$ . This is a subsolution of the Dirichlet problem which is globally bounded in $\tilde{\Omega}$ . It follows from the maximum principle that $\max(\psi_2,\varphi_{j,\varepsilon}) \leq \varphi_{j,\varepsilon}$ , hence $\psi_2 \leq \max(\psi_2,\varphi_{j,\varepsilon}) \leq \varphi_{j,\varepsilon}$ .

Step 2. We next claim that there exist constants $A_2, A_{3,j+1,\varepsilon}>0$ such that

(6.4) \begin{equation}\sup_{\tilde{\Omega}}[\lvert s\rvert^{2A_2}_h\lvert\nabla \varphi_{j+1,\varepsilon}\rvert^2_{\beta}] \leq A_{3,j+1,\varepsilon},\end{equation}

where $ A_{3,j+1,\varepsilon}$ only depends on an upper bound on $\lVert \varphi_{k,\varepsilon}\rVert_{L^\infty(\tilde{\Omega})}$ for $k\leq j+1$ .

Proof. The proof is a variant of [Reference Datar, Fu and SongDFS23, Proposition 2.2], which itself relies on previous estimates due to Blocki and Phong-Sturm.

As we work in $\tilde{\Omega}\setminus\mathrm{Supp}(E)$ , we identify $\beta$ with $dd^c (K\rho\circ\pi +\log \lvert s\rvert^2_h)$ . Replacing $\varphi_{j+1,\epsilon}$ by $\tilde{\varphi}_{j+1,\epsilon}:=\varphi_{j+1,\epsilon}-(K\rho \circ \pi+\log\lvert s \rvert^2_h)$ , (6.2) becomes

(6.5) \begin{equation}\begin{cases}\displaystyle ((1+\epsilon)\beta+ dd^c \tilde{\varphi}_{j+1,\epsilon})^n=c_j^{-1}e^{-\gamma \varphi_{j,\epsilon}}\prod_{l=1}^m(\lvert s_l\rvert^2_{h_l}+\epsilon^2)^{a_l} \, dV_{\tilde{\Omega}},\\\tilde{\varphi}_{j+1,\epsilon|\partial \tilde{\Omega}}=\phi-\log\lvert s \rvert^2_h.\end{cases}\end{equation}

As

\[\lvert\lvert\nabla\varphi_{j,\epsilon}\rvert_\beta -\lvert\nabla\tilde{\varphi}_{j,\epsilon}\rvert_\beta\rvert\leq\frac{\lvert \nabla \lvert s \rvert^2_h\rvert_\beta}{\lvert s\rvert_h} +C,\]

to get the estimate (6.4) for $\tilde{\varphi}_{j,\epsilon}$ it is enough to prove by induction that there exists positive constants $B_2, B_{3,j+1,\varepsilon}$ such that

(6.6) \begin{equation}\sup_{\tilde{\Omega}} [\lvert s\rvert_h^{2B_2}\lvert \nabla \tilde{\varphi}_{j+1,\varepsilon}\rvert^2_\beta]\leq\max\bigg\{\sup_{\tilde{\Omega}}[\lvert s \rvert_h^{2B_2}\lvert \nabla\tilde{\varphi}_{j,\varepsilon}\rvert_\beta^2], B_{3,j+1,\varepsilon}\bigg\},\end{equation}

where $B_2$ is uniform in $j,\varepsilon$ while $ B_{3,j+1,\varepsilon}$ only depends on upper bounds on $\lVert\varphi_{j+1,\varepsilon}\rVert_{L^\infty(\tilde{\Omega})}, \lVert \varphi_{j,\varepsilon} \rVert_{L^\infty(\tilde{\Omega})}$ , and where $\tilde{\varphi}_{0,\varepsilon}:=-(K\rho\circ \pi + \log \lvert s \rvert^2_h)$ . To lighten notation we rewrite the equation

(6.7) \begin{equation}\begin{cases}\displaystyle(\beta_\epsilon+ dd^c u)^n=e^{-v}\prod_{l=1}^m(\lvert s_l\rvert^2_{h_l}+\epsilon^2)^{a_l}\beta_\epsilon^n,\\ u_{|\partial \tilde{\Omega}}=\tilde{\phi},\end{cases}\end{equation}

where $\beta_\epsilon:=(1+\epsilon)\beta$ is a non-degenerate smooth family of Kähler forms. Note that (6.6) becomes

(6.8) \begin{equation}\sup_{\tilde{\Omega}} [\lvert s\rvert_h^{2B_2}\lvert \nabla u\rvert^2_\beta]\leq\max\bigg\{\sup_{\tilde{\Omega}}[\lvert s \rvert_h^{2B_2}\lvert\nabla(v/\gamma-\log\lvert s \rvert_h^2-f_\epsilon)\rvert_\beta^2], B_{3,j+1,\varepsilon}\bigg\},\end{equation}

where $\{f_\epsilon\}_{\epsilon>0}$ is a non-degenerate smooth family. In the estimates that follow we indicate with $C_i$ all the constants under control, i.e. that depend on a upper bound on $\lVert \varphi_{j+1,\epsilon}\rVert_{L^\infty(\tilde{\Omega})}, \lVert\varphi_{j,\epsilon} \rVert_{L^\infty(\Omega)}$ . Observe that $\lVert u+\log\lvert s\rvert^2_h\rVert_{L^\infty(\tilde{\Omega})}, \lVert v\rVert_{L^\infty(\tilde{\Omega})}$ and $\sup_{\partial\tilde{\Omega}}\lvert \nabla u \rvert$ are under control. The constant $B_{3,j+1,\varepsilon}$ in (6.8) will clearly depend on the $C_i$ . We indicate with $D_i$ all the constants uniform in $j,\varepsilon$ , which will be used to determine the uniform constant $B_2$ in (6.8).

We denote by $\Delta_\epsilon$ and $\Delta'_\epsilon$ the Laplacian operators with respect to $\beta_\epsilon$ and to $\eta_{\epsilon}:=\beta_\epsilon + dd^c u$ , respectively. Consider

\[H:=\log \lvert \nabla u\rvert_{\beta_\epsilon}^2+\log \lvert s \rvert_h^{2k}-G(u),\]

where $G(x)=Ax-{B}/({x+C+1})$ for C chosen so that $u\geq -C$ , while $A>0, B>0$ to be determined later. The constants A,k are chosen to be uniform in $j,\varepsilon$ while B is under control. If H reaches its maximum at $x_M$ , then

(6.9) \begin{equation}\lvert \nabla u \rvert_{\beta_\epsilon}^2\lvert s\rvert_h^{2(k+A)}\leq C_1(\lvert \nabla u \rvert_{\beta_\epsilon}^2\lvert s\rvert_h^{2(k+A)})(x_M)\end{equation}

for a constant $C_1$ under control.

As $u+\log\lvert s\rvert^2_h$ is smooth on $\tilde{\Omega}$ , we ensure that $H(x)\simeq(k+A-1)\log\lvert s\rvert_h^2\to -\infty$ as $x\to \mathrm{Supp}(E_j)$ by imposing $k\geq 1$ . If H reaches its maximum on $\partial \tilde{\Omega}$ , then we are done since $\sup_{\partial\tilde{\Omega}}\lvert\nabla u \rvert$ is under control. From now on we thus suppose that H reaches its maximum in $\tilde{\Omega}\setminus\{s=0\}$ . A direct computation [Reference Phong, Song and SturmPSS12, (5.11) and (5.20)] yields

(6.10) \begin{align}\Delta'_\epsilon \log \lvert \nabla u\rvert^2_{\beta_\epsilon} &\geq\frac{2\mathrm{Re}\langle \nabla v+\sum_{l=1}^m a_l\nabla \log(\lvert s_l \rvert^2_{h_l}+\epsilon^2),\nabla u\rangle_{\beta_{\epsilon}}}{\lvert \nabla u\rvert_{\beta_\epsilon}^2}-\Lambda \mathrm{tr}_{\eta_\epsilon}\beta_\epsilon \nonumber\\&\quad +2\mathrm{Re}\bigg\langle \frac{\nabla \lvert \nabla u\rvert_{\beta_\epsilon}^2}{\lvert\nabla u\rvert_{\beta_\epsilon}^2},\frac{\nabla u}{\lvert \nabla u\rvert_{\beta_\epsilon}^2}\bigg\rangle_{\!\!\eta_\epsilon}-2\mathrm{Re}\bigg\langle\frac{\nabla \lvert \nabla u\rvert_{\beta_\epsilon}^2}{\lvert\nabla u\rvert_{\beta_\epsilon}^2},\frac{\nabla u}{\lvert \nabla u\rvert_{\beta_\epsilon}^2}\bigg\rangle_{\!\!\beta_\epsilon} ,\end{align}

where $\Lambda$ denotes a (uniform in $\varepsilon$ ) lower bound on the holomorphic bisectional curvature of $\beta_\epsilon$ . At the point where H reaches its maximum we obtain

\[\frac{\nabla\lvert\nabla u\rvert_{\beta_\epsilon}^2}{\lvert\nabla u\rvert^2_{\beta_\epsilon}}=\nabla \log \lvert\nabla u\rvert^2_{\beta_\epsilon}=-\nabla (\log \lvert s \rvert_h^{2k}-G(u))=-\frac{k\nabla\lvert s \rvert^2_h }{\lvert s\rvert^2_h}+G'(u)\nabla u,\]

hence

\begin{align*}& 2\mathrm{Re}\bigg\langle \frac{\nabla \lvert \nabla u\rvert_{\beta_\epsilon}^2}{\lvert\nabla u\rvert_{\beta_\epsilon}^2},\frac{\nabla u}{\lvert\nabla u\rvert_{\beta_\epsilon}^2} \bigg\rangle_{\!\!\eta_\epsilon}-2\mathrm{Re}\bigg\langle\frac{\nabla \lvert \nabla u\rvert_{\beta_\epsilon}^2}{\lvert\nabla u\rvert_{\beta_\epsilon}^2},\frac{\nabla u}{\lvert\nabla u\rvert_{\beta_\epsilon}^2} \bigg\rangle_{\!\!\beta_\epsilon} \\&\quad = 2k\mathrm{Re}\bigg\langle \frac{\nabla \lvert s \rvert^2_h}{\lvert s \rvert^2_h},\frac{\nabla u}{\lvert \nabla u \rvert^2_{\beta_\epsilon}} \bigg\rangle_{\!\!\beta_\epsilon} -2k\mathrm{Re}\bigg\langle \frac{\nabla \lvert s \rvert^2_h}{\lvert s \rvert^2_h},\frac{\nabla u}{\lvert \nabla u \rvert^2_{\beta_\epsilon}} \bigg\rangle_{\!\!\eta_\epsilon}+2G'(u)\frac{\lvert \nabla u\rvert^2_{\eta_\epsilon}}{\lvert \nabla u\rvert^2_{\beta_\epsilon}}-2G'(u)\\&\quad \geq 2k\mathrm{Re}\bigg\langle \frac{\nabla \lvert s \rvert^2_h}{\lvert s \rvert^2_h},\frac{\nabla u}{\lvert \nabla u \rvert^2_{\beta_\epsilon}} \bigg\rangle_{\!\!\beta_\epsilon} -2k\mathrm{Re}\bigg\langle \frac{\nabla \lvert s \rvert^2_h}{\lvert s \rvert^2_h},\frac{\nabla u}{\lvert \nabla u \rvert^2_{\beta_\epsilon}} \bigg\rangle_{\!\!\eta_\epsilon}-2G'(u),\end{align*}

using the monotonicity of G(x) in the last inequality. By (6.9) and asking $k\geq2$ , we can assume that $\lvert s \rvert_h^2 \lvert \nabla u\rvert_{\beta_\epsilon}\geq 1$ at $x_M$ . Thus,

\[\bigg\lvert 2\mathrm{Re}\bigg\langle \frac{\nabla \lvert s \rvert^2_h}{\lvert s \rvert^2_h},\frac{\nabla u}{\lvert \nabla u \rvert^2_{\beta_\epsilon}} \bigg\rangle_{\!\!\beta_\epsilon}\!\bigg\rvert\leq 2\bigg\lvert \mathrm{Re}\bigg\langle \nabla \lvert s \rvert_h^2,\frac{\nabla u}{\lvert \nabla u\rvert_{\beta_\epsilon}} \bigg\rangle_{\!\!\beta_\epsilon}\! \bigg\rvert\leq D_1\]

and

\begin{align*}\bigg\lvert 2\mathrm{Re}\bigg\langle \frac{\nabla \lvert s \rvert^2_h}{\lvert s \rvert^2_h},\frac{\nabla u}{\lvert \nabla u \rvert^2_{\beta_\epsilon}} \bigg\rangle_{\!\!\eta_\epsilon}\!\bigg\rvert&\leq 2 \bigg\lvert \mathrm{Re}\bigg\langle \nabla \lvert s \rvert_h^2,\frac{\nabla u}{\lvert \nabla u\rvert_{\beta_\epsilon}} \bigg\rangle_{\!\!\eta_\epsilon}\! \bigg\rvert\leq \lvert \nabla \lvert s\rvert^2_h \rvert_{\eta_\epsilon}^2+\frac{\lvert s\rvert_h^4\lvert\nabla u\rvert_{\eta_\epsilon}^2}{\lvert s\rvert_h^4\lvert \nabla u\rvert_{\beta_\epsilon}^2} \\&\leq \lvert \nabla \lvert s\rvert^2_h \rvert_{\beta_\epsilon}^2\mathrm{tr}_{\eta_\epsilon}{\beta_\epsilon}+\lvert s\rvert_h^4\lvert \nabla u\rvert^2_{\eta_\epsilon}.\end{align*}

We infer that at $x=x_M$ ,

\begin{align*}& 2\mathrm{Re}\bigg\langle \frac{\nabla \lvert \nabla u\rvert_{\beta_\epsilon}^2}{\lvert\nabla u\rvert_{\beta_\epsilon}^2},\frac{\nabla u}{\lvert\nabla u\rvert_{\beta_\epsilon}^2} \bigg\rangle_{\!\!\eta_\epsilon}-2\mathrm{Re}\bigg\langle \frac{\nabla \lvert \nabla u\rvert_{\beta_\epsilon}^2}{\lvert\nabla u\rvert_{\beta_\epsilon}^2},\frac{\nabla u}{\lvert\nabla u\rvert_{\beta_\epsilon}^2} \bigg\rangle_{\!\!\beta_\epsilon} \\&\quad \geq -kD_1-k\lvert \nabla \lvert s\rvert^2_h \rvert_{\beta_\epsilon}^2\mathrm{tr}_{\eta_\epsilon}\beta_\epsilon-k\lvert s\rvert_h^4\lvert \nabla u\rvert^2_{\eta_\epsilon}-2G'(u) \\&\quad \geq -kD_1-kD_2\mathrm{tr}_{\eta_\epsilon}\beta_\epsilon-k\lvert s\rvert_h^4 \lvert \nabla u \rvert^2_{\eta_\epsilon}-2G'(u),\end{align*}

which is the first estimate of the right-hand side in (6.10).

Next, as we want to prove (6.8), as a consequence of (6.9) and of $\lvert \nabla u\rvert^2_\beta\leq D_3 \lvert \nabla u \rvert^2_{\beta_\varepsilon}$ in the estimate that follows we can assume that

\[D_3C_1\lvert \nabla u\rvert^2_{\beta_\varepsilon} \geq \max\{\lvert\nabla (v/\gamma- \log\lvert s \rvert^2_h-f_\varepsilon) \rvert^2_\beta, 1\}\]

at the point $x_M$ . We deduce

\begin{align*}& \bigg\lvert \frac{2\mathrm{Re}\langle \nabla v+\sum_{l=1}^m \nabla a_l\log(\lvert s_l \rvert^2_{h_l}+\epsilon^2), \nabla u\rangle_{\beta_\epsilon}}{\lvert\nabla u \rvert_{\beta_\epsilon}^2} \bigg\rvert\\&\quad \leq D_4+\frac{\lvert \nabla (v-\gamma\log\lvert s\rvert_h^2-\gamma f_\varepsilon)\rvert_{\beta_\epsilon}^2}{\lvert \nabla u \rvert_{\beta_\epsilon}^2}+\frac{D_5}{\lvert \nabla u \rvert_{\beta_\varepsilon}^2}\lvert s\rvert_h^{-2}+\frac{D_6}{\lvert \nabla u\rvert^2_{\beta_\epsilon}}\sum_{l=1}^m \lvert s_l \rvert_{h_l}^{-2}\leq C_2 + C_3\lvert s\rvert_h^{-2M}\end{align*}

for $M:={1}/{\min_lb_l}$ so that $Mb_l\geq 1$ for any l. The previous inequalities yield

(6.11) \begin{equation}\Delta'_\epsilon\log \lvert \nabla u \rvert^2_{\beta_\epsilon}\geq -kD_1-C_2-C_3 \lvert s\rvert_h^{-2M}-(kD_2+\Lambda)\mathrm{tr}_{\eta_\epsilon}\beta_\epsilon-k\lvert s\rvert_h^4\lvert \nabla u\rvert^2_{\eta_\epsilon}-2G'(u).\end{equation}

Moreover,

\[-\Delta'_\epsilon G(u)=-G'(u)\Delta'_\epsilon u-G''(u)\lvert\nabla u \rvert^2_{\eta_\epsilon}=G'(u) \mathrm{tr}_{\eta_\epsilon}\beta_\epsilon-nG'(u)-G''(u)\lvert \nabla u\rvert_{\eta_\epsilon}^2\]

and $\Delta'_\epsilon\log \lvert s \rvert^{2k}_h\geq -kD_7\mathrm{tr}_{\eta_\epsilon}\beta_\epsilon$ . Together with (6.11) we obtain

\[\Delta'_\epsilon H\geq (G'-kD_2-\Lambda-kD_7)\mathrm{tr}_{\eta_\epsilon}\beta_\epsilon-(n+2)G'-(G''+k\lvert s\rvert_h^4)\lvert\nabla u\rvert^2_{\eta_\epsilon}-kD_1-C_2-C_3\lvert s \rvert^{-2M}_h.\]

Taking $k=M(n+1)+1$ , this can be rewritten

(6.12) \begin{equation}\Delta'_\epsilon H\geq (G'-D_8)\mathrm{tr}_{\eta_\epsilon}\beta_\epsilon-(n+2)G'-(G''+D_9\lvert s\rvert^4_h)\lvert\nabla u\rvert^2_{\eta_\epsilon}-C_4\lvert s \rvert_h^{-2M}.\end{equation}

We now define $G(x):=(D_8+1)x-{B}/({x+C+1})$ , where $B>0$ is so large that

\[\frac{2B}{(u+C+1)^3}-D_9\lvert s\rvert^4_h\geq \lvert s\rvert^2_h\]

at $x_M$ . Note that B can be chosen such that it only depends on $C, D_9$ and on $\lVert u+\log\lvert s\rvert^2_h\rVert_{L^{\infty}(\tilde{\Omega})}$ , i.e. it is under control. From (6.12) we deduce at $x_M$

\[0\geq \Delta'_\epsilon H\geq \mathrm{tr}_{\eta_\epsilon}\beta_\epsilon+\lvert s \rvert^2_h\lvert \nabla u \rvert^2_{\eta_\epsilon}-C_5\lvert s \rvert^{-2M}_h.\]

This yields $\mathrm{tr}_{\eta_\epsilon}\beta_\epsilon\leq C_5\lvert s \rvert_h^{-2M} \text{ and }\lvert\nabla u \rvert_{\eta_\epsilon}^2\leq C_5\lvert s \rvert_h^{-2M-2}, $ hence

\[\lvert\nabla u \rvert^2_{\beta_\epsilon}\leq \lvert\nabla u \rvert^2_{\eta_\epsilon}\mathrm{tr}_{\beta_\epsilon}\eta_\epsilon\leq\lvert \nabla u \rvert^2_{\eta_\varepsilon}(\mathrm{tr}_{\eta_\epsilon}\beta_\epsilon)^{n-1} \bigg(\frac{\eta_\varepsilon^n}{\beta_\varepsilon^n}\bigg)\leq C_6 \lvert s\rvert^{-2M}_h\lvert \nabla u \rvert^2_{\eta_\varepsilon} (\mathrm{tr}_{\eta_\epsilon}\beta_\epsilon)^{n-1}\leq C_7 \lvert s \rvert_h^{-2k},\]

where we also used [Reference Guedj and ZeriahiGZ17, Lemma 14.4], the Monge–Ampère equation (6.7) and the fact that $\prod_{l=1}^m(\lvert s_l\rvert^2_{h_l}+\varepsilon^2)^{a_l}\leq D_{10} \lvert s\rvert_h^{-2M}$ as $a_l>-1$ . From (6.9) we deduce $ \lvert\nabla u\rvert^2_{\beta_\epsilon}\lvert s\rvert^{2(k+D_8+1)}_h\leq C_8$ . As $\{\beta_\epsilon\}_{\epsilon>0}$ is a non-degenerate continuous family of Kähler forms converging to $\beta$ as $\epsilon\to 0$ , we get

\[\lvert s\rvert^{2(k+D_8+1)}_h\lvert\nabla u \rvert^2_\beta\leq\max\bigg\{\sup_{\tilde{\Omega}}[\lvert s \rvert_h^{2(k+D_8+1)} \lvert\nabla(v/\gamma-\log\lvert s\rvert^2-f_\varepsilon) \rvert_\beta^2], C_9\bigg\},\]

i.e. (6.8), which concludes the proof by setting $B_2:=k+D_8+1$ , $B_{3,j+1,\varepsilon}:=C_9$ .

Step 3. Fix V a small neighborhood of $\partial \tilde{\Omega}$ (intersected with $\overline{\tilde{\Omega}}$ ). We claim that

(6.13) \begin{equation}\sup_{\partial \tilde{\Omega}} |\Delta_{\beta} \varphi_{j,\varepsilon}| \leq C_V[1+\sup_V |\nabla \varphi_{j,\varepsilon}|^2],\end{equation}

for some uniform constant $C_V$ independent of $j,\varepsilon$ . This follows from a long series of estimates established in [Reference Guedj, Kolev and YeganefarGKY13, Lemma 18] (which itself was adapting the technique developed by [Reference Caffarelli, Kohn, Nirenberg and SpruckCKNS85]) when $\mu_p$ and $\Omega$ are smooth. The statement of [Reference Guedj, Kolev and YeganefarGKY13, Lemma 18] mentions $\sup_{\tilde{\Omega}} |\nabla \varphi_j|^2$ ; however, the arguments only involve:

  • local reasonings in a small fixed neighborhood of the boundary;

  • –smoothness of $\mu_p$ in this neighborhood and pseudoconvexity of $\partial \tilde{\Omega}$ .

Step 4. We now show that there exist constants $m,B_{3,j,\varepsilon}>0$ such that

(6.14) \begin{equation}\sup_{\tilde{\Omega}} |s|^{2m} |\Delta_{\beta} \varphi_{j,\varepsilon}| \leq B_{3,j,\varepsilon} \bigg[1+\sup_{\partial \tilde{\Omega}} |\Delta_{\beta} \varphi_{j,\varepsilon}|\bigg],\end{equation}

where $B_{3,j,\varepsilon}$ only depends on an upper bound on $\|\varphi_{k,\varepsilon}\|_{L^{\infty}(\tilde{\Omega})}$ , for $k \leq j$ . This is a variant of [Reference Guedj, Kolev and YeganefarGKY13, Lemma 17], for which we provide a detailed proof.

We set $\omega_j:=\varepsilon\beta+dd^c \varphi_{j,\varepsilon} $ and observe that

\[\omega_j^n=e^{\psi_{\varepsilon}-\varphi_{j-1,\varepsilon}-c^{\prime}_{j-1}} \beta^n,\]

where $\psi_{\varepsilon}$ is a difference of quasi-psh functions in $\tilde{\Omega}$ such that $e^{\psi_{\varepsilon}}\leq c_1 |s|^{-2a}$ and $dd^c \psi_{\varepsilon} \geq -c_1 |s|^{-2} \beta$ in $\tilde{\Omega}$ , for some uniform constants $a,c_1>0$ . We consider

\[H_j:=\log {\rm Tr}_{\beta}(\omega_j)+\varphi_{j-1,\varepsilon} -A \varphi_{j,\varepsilon} +A \rho',\]

where $A>0$ is chosen below. We use here the classical notation

\[{\rm Tr}_{\eta}(\omega):=n \frac{\omega \wedge \eta^{n-1}}{\eta^n}\quad \text{and}\quad\Delta_{\eta}(h):=n \frac{dd^c h \wedge \eta^{n-1}}{\eta^n}.\]

Either $H_j$ reaches its maximum on $\partial \tilde{\Omega}$ and we are done, or it reaches its maximum at some point $x_j \in \tilde{\Omega} \setminus E$ since $\rho \rightarrow -\infty$ along E. We are going to estimate $\Delta_{\omega_j} H_j$ from below and use the fact that $0 \geq\Delta_{\omega_j} H_j(x_j)$ .

It follows from [Reference SiuSiu87] that

\[\Delta_{\omega_j} \log {\rm Tr}_{\beta}(\omega_j) \geq-\frac{{\rm Tr}_{\beta}( {\rm Ric}(\omega_j))}{{\rm Tr}_{\beta}(\omega_j)}-B {\rm Tr}_{\omega_j}(\beta),\]

where $-B$ is a lower bound on the holomorphic bisectional curvature of $\beta$ . Now

\[-{\rm Ric}(\omega_j)=-{\rm Ric}(\beta)+dd^c (\psi_{\varepsilon}-\varphi_{j-1,\varepsilon})\geq -\omega_{j-1}-\frac{A_1}{|s|^2} \beta\]

in $\tilde{\Omega} \setminus E$ . Moreover, ${\rm Tr}_{\beta}(\omega_{j-1}) \leq {\rm Tr}_{\beta}(\omega_j) {\rm Tr}_{\omega_j}(\omega_{j-1})$ , hence

\[\Delta_{\omega_j} \log {\rm Tr}_{\beta}(\omega_j) \geq -{\rm Tr}_{\omega_j}(\omega_{j-1})-\frac{nA_1}{|s|^2{\rm Tr}_{\beta}(\omega_j)} -B {\rm Tr}_{\omega_j}(\beta).\]

Using that $dd^c \rho' \geq \beta$ , we obtain

\[\Delta_{\omega_j} H_j \geq -An+(A-B) {\rm Tr}_{\omega_j}(\beta)-\frac{n A_1}{|s|^2{\rm Tr}_{\beta}(\omega_j)}.\]

Using the classical inequality $n [{\rm Tr}_{\omega_j}(\beta)]^{n-1} \geq (\beta^n/\omega_j^n) {\rm Tr}_{\beta}(\omega_j)$ , we infer

(6.15) \begin{equation}\Delta_{\omega_j} H_j \geq -An+ c(A-B) e^{{-\psi_{\varepsilon}}/({n-1})}[{\rm Tr}_{\beta}(\omega_j)]^{{1}/({n-1})}-\frac{n A_1}{|s|^2{\rm Tr}_{\beta}(\omega_j)}.\end{equation}

Let us stress that the constant c depends here on an upper bound on $\|\varphi_{j-1,\varepsilon}\|_{L^{\infty}(\tilde{\Omega})}$ .

We fix A so large that $A>B$ and $\psi_{\varepsilon}+A\rho' \leq c_1'-a \log|s|^2+A \rho'$ is bounded from above. At the point $x_j$ we obtain $0 \geq \Delta_{\omega_j} H_j$ , therefore:

  • either $|s|^2 {\rm Tr}_{\beta}(\omega_j) \leq 1$ , hence $H_j(x_j) \leq (\varphi_{j-1,\varepsilon} -A\varphi_{j,\varepsilon} +A \rho')(x_j) \leq C$ ;

  • or $|s|^2{\rm Tr}_{\beta}(\omega_j) \geq 1$ and (6.15) yields ${\rm Tr}_{\beta}(\omega_j) \leq C' e^{\psi_{\varepsilon}(x_j)}$ , hence

    \[H_j(x_j) \leq \psi_{\varepsilon}(x_j)+A\rho'(x_j)+C'' \leq C'''.\]

Thus, $H_j$ is uniformly bounded from above in both cases, and (6.14) follows (we use here an upper bound on $\|\varphi_{j-1,\varepsilon}\|_{L^{\infty}(\tilde{\Omega})}$ and $\|\varphi_{j,\varepsilon}\|_{L^{\infty}(\tilde{\Omega})}$ ).

Step 5. We finally show by induction on j that $\varphi_{j,\varepsilon}$ uniformly converges towards $\varphi_j$ as $\varepsilon$ decreases to 0. There is nothing to prove for $j=0$ since $\varphi_{0,\varepsilon}=\varphi_0$ .

For $j=1$ , it follows from (a slight generalization of) [Reference Guedj, Guenancia and ZeriahiGGZ23, Proposition 1.8] that $\|\varphi_{1,\varepsilon}\|_{L^{\infty}(\tilde{\Omega})} \leq C_1$ is bounded uniformly in $\varepsilon>0$ . Proceeding by induction, we similarly obtain that for all $j \in \mathbb N$ ,

\[\|\varphi_{j,\varepsilon}\|_{L^{\infty}(\tilde{\Omega})} \leq C_j\]

is bounded uniformly in $\varepsilon>0$ . By previous steps, the family $(\varphi_{j,\varepsilon})_{\varepsilon}$ is relatively compact in ${\mathcal C}^{1,\alpha}$ for all $0<\alpha < 1$ . Any cluster point $\psi_j$ , as $\varepsilon\rightarrow 0$ , is a solution of

\[(dd^c \psi_{j+1})^n=\frac{e^{-\gamma \psi_j} \mu_p}{c_j}\]

with boundary values ${\psi_{j+1}}_{|\partial \Omega}=\phi$ , hence $\psi_j=\varphi_j$ by uniqueness. Thus, $\varphi_{j,\varepsilon}$ converges to $\varphi_j$ as $\varepsilon$ decreases to zero, and the convergence is moreover uniform on $\tilde{\Omega}$ by [Reference Guedj, Guenancia and ZeriahiGGZ23, Proposition 1.8].

We can thus let $\varepsilon$ tend to zero in previous inequalities. Now $\|\varphi_{j,\varepsilon}\|_{L^{\infty}(\tilde{\Omega})} \rightarrow \|\varphi_j\|_{L^{\infty}(\Omega)}$ , and the latter is uniformly bounded in j by Proposition 6.1. For $\varepsilon=0$ , (6.3), (6.4), (6.13) and (6.14) thus provide uniform bounds in j, and conclude the proof of (6.1). The proof of Proposition 6.2 is thus complete.

6.3 Higher-order estimates and convergence

Once the uniform ${\mathcal C}^2$ -estimate is established (Proposition 6.2), one can then linearize the complex Monge–Ampère equation and apply standard elliptic theory (Evans–Krylov method and Schauder bootstrapping) to derive higher-order estimates.

Proposition 6.3. Given K a compact subset of ${\Omega} \setminus \{p\}$ and $\alpha>0,\ell \in \mathbb N$ , there exists $C(K,\ell,\alpha)>0$ such that for all $j \in \mathbb N$ , $ \|\varphi_j\|_{{\mathcal C}^{\ell,\alpha}(K)} \leq C(K,\ell,\alpha)$ .

It follows that the sequence $(\varphi_j)$ is relatively compact in ${\mathcal C}^{\infty}({\Omega}\setminus \{p\})$ . We let ${\mathcal K}$ denote the set of cluster values of the sequence $(\varphi_j)$ . Any function $\psi \in {\mathcal K}$ is:

  • psh in $\Omega$ and smooth in ${\Omega} \setminus \{p\}$ , with $\psi_{|\partial \Omega}=\phi$ ;

  • uniformly bounded in $\overline{\Omega}$ (Proposition 6.1);

  • continuous on $\overline{\Omega}$ , as the uniform limit of $(\varphi_{j_k})$ (see [Reference Guedj, Guenancia and ZeriahiGGZ23, Proposition 1.8]);

The set ${\mathcal K}$ is invariant under the action of $T_{\gamma} : \varphi \in {\mathcal T}_{\phi}(\Omega) \mapsto \psi \in {\mathcal T}_{\phi}(\Omega)$ , which associates, to a given $\varphi \in{\mathcal T}_{\phi}(\Omega)$ , the unique solution $\psi \in {\mathcal T}_{\phi}(\Omega)$ to the complex Monge–Ampère equation

\[(dd^c \psi)^n =\frac{e^{-\gamma\varphi} \mu_p}{\int_{\Omega} e^{-\gamma \varphi} \,d\mu_p}.\]

It follows from [Reference Guedj, Kolev and YeganefarGKY13, Proposition 12] that the functional $F_{\gamma}$ is constant on ${\mathcal K}$ and that ${\mathcal K}$ is pointwise invariant under the action of $T_{\gamma}$ . Thus, a cluster value of $(\varphi_j)$ provides a desired solution to Theorem C.

Appendix A

Sèbastien Boucksom

The purpose of this appendix is to provide an alternative approach to Proposition 5.8, emphasizing the role of b-divisors. We use [Reference Boucksom, de Fernex and FavreBdFF12] as a main reference for what follows.

A.1 Nef b-divisors over a point

Consider for the moment any normal singularity (X,p), and set $n:=\dim X$ .

In what follows, a birational model means a projective birational morphism $\pi\colon X_\pi\to X$ with $X_\pi$ normal. A b-divisor over p is defined as a collection $B=(B_\pi)_\pi$ of $\mathbb R$ -divisors $B_\pi$ on $X_\pi$ for all birational models $\pi$ , compatible under push-forward, and such that each $B_\pi$ has support in $\pi^{-1}(p)$ . The $\mathbb R$ -vector space of b-divisors over p can thus be written as the projective limit

\[\mathrm{Div}_{\mathrm{b}}(X,p):=\varprojlim_\pi\mathrm{Div}_p(X_\pi),\]

where $\mathrm{Div}_p(X_\pi)$ denotes the (finite-dimensional) $\mathbb R$ -vector space of divisors on $X_\pi$ with support in $\pi^{-1}(p)$ , and we endow $\mathrm{Div}_{\mathrm{b}}(X,p)$ with the projective limit topology.

A b-divisor $B\in\mathrm{Div}_{\mathrm{b}}(X,p)$ is said to be Cartier if it is determined by some birational model $\pi$ , in the sense that $B_{\pi'}$ is the pullback of $B_\pi$ for any higher birational model $\pi'$ . There is a symmetric, multilinear intersection pairing

(A.1) \begin{equation}(B_1,\ldots,B_n)\mapsto(B_1\cdot\cdots\cdot B_n)\in\mathbb R\end{equation}

for Cartier b-divisors $B_i$ , defined as the intersection number $(B_{1,\pi}\cdot\cdots\cdot B_{n,\pi})$ computed on $X_\pi$ for any choice of common determination $\pi$ of the $B_i$ (the result being independent of the choice of $\pi$ , by the projection formula).

A valuation centered at p is a valuation $v \colon \mathcal{O}_{X,p}\to \mathbb R_{\ge 0}$ such that $v(\mathfrak{m}_p)>0$ on the maximal ideal $\mathfrak{m}_p \subset \mathcal{O}_{X,p}$ . It is further divisorial if it can be written as $v=c\mathrm{ord}_E$ for a prime divisor $E\subset\pi^{-1}(p)$ on some birational model $X_\pi$ and $c\in\mathbb Q_{>0}$ . Given a b-divisor B over $p\in X$ , we then set $v(B):=c \, \mathrm{ord}_E(B_\pi)$ . The function $v\mapsto v(B)$ so defined on the space $\mathrm{DivVal}(X,p)$ of divisorial valuations centered at p is homogeneous with respect to the scaling of $\mathbb Q_{>0}$ , and this yields a topological vector space isomorphism between $\mathrm{Div}_{\mathrm{b}}(X,p)$ and the space of homogeneous functions on $\mathrm{DivVal}(X,p)$ , endowed with the topology of pointwise convergence.

Pick a b-divisor B over p. If B is Cartier, we say that B is (relatively) nef if $B_\pi$ is $\pi$ -nef for some (hence, any) determination $\pi$ . In the general case, we say that B is nef if it can be written as a limit of nef Cartier b-divisors. By the negativity lemma, any nef b-divisor $B\in\mathrm{Div}_{\mathrm{b}}(X,p)$ is automatically antieffective, i.e. $v(B)\le 0$ for all $v\in\mathrm{DivVal}(X,p)$ . By [Reference Boucksom, de Fernex and FavreBdFF12, Lemma 2.10], we further have the following result.

Lemma A.1. A b-divisor B over p is nef if and only if, for each birational model $\pi$ , the numerical class of $B_\pi$ in $\mathrm{N}^1(X_\pi/X)$ is nef in codimension 1 (also known as movable).

Example A.2. Consider an ideal $\mathfrak{a}\subset\mathcal{O}_{X,p}$ , and assume that $\mathfrak{a}$ is primary, i.e. containing some power of the maximal ideal. Then $\mathfrak{a}$ determines a nef Cartier b-divisor $Z(\mathfrak{a})$ , defined by $v(Z(\mathfrak{a}))=-v(\mathfrak{a})$ for each $v\in\mathrm{DivVal}(X,p)$ , and determined on the normalized blow-up of $\mathfrak{a}$ . For any tuple of primary ideals $\mathfrak{a}_1,\ldots,\mathfrak{a}_n$ ,

\[e(\mathfrak{a}_1,\ldots,\mathfrak{a}_n)=-(Z(\mathfrak{a}_1)\cdot\cdots\cdot Z(\mathfrak{a}_n))\]

further coincides with the mixed multiplicity of the $\mathfrak{a}_i$ .

Example A.3. For any valuation v centered at p, the valuation ideals

\[\mathfrak{a}_m(v):=\{f\in\mathcal{O}_{X,p}\mid v(f)\ge m\}\]

define a graded sequence of primary ideals $\mathfrak{a}_\bullet(v)$ , and hence a nef b-divisor over p

\[Z(v):=Z(\mathfrak{a}_\bullet(v))=\lim_m m^{-1} Z(\mathfrak{a}_m(v)),\]

(see [Reference Boucksom, de Fernex and FavreBdFF12, Lemma 2.11]), which is not Cartier in general.

Lemma A.4. If $B\in\mathrm{Div}_{\mathrm{b}}(X,p)$ is nef, then $B\le - v(B) Z(v)$ for all $v\in\mathrm{DivVal}(X,p)$ ,

Proof. Write $v=c \, \mathrm{ord}_E$ for a prime divisor E on $X_\pi$ and $c\in\mathbb Q_{>0}$ . Then Z(v) coincides with $\mathrm{Env}_\pi(-c^{-1} E)$ (see [Reference Boucksom, de Fernex and FavreBdFF12, Definition 2.3]), and the result thus follows from [Reference Boucksom, de Fernex and FavreBdFF12, Proposition 2.12].

A.2 Normalized volume and b-divisors

From now on, we assume that the normal singularity $p\in X$ is further isolated.

By [Reference Boucksom, de Fernex and FavreBdFF12, Theorem 4.14], the intersection pairing (A.1) then extends to arbitrary tuples of nef b-divisors over p. This extended pairing takes values in $\mathbb R\cup\{-\infty\}$ , and is symmetric, additive and non-decreasing in each variable, and continuous along decreasing nets.

Definition A.5. For any nef b-divisor B over p, we define the Hilbert–Samuel multiplicity of B as

\[e(B):=-B^n\in [0,+\infty].\]

When (X,p) is further klt, we define the log canonical threshold of B as

\[\mathrm{lct}(B):=\inf_{v\in\mathrm{DivVal}(X,p)}\frac{\mathrm{A}_X(v)}{-v(B)}\in [0,+\infty),\]

where $\mathrm{A}_X(v)\ge 0$ denotes the log discrepancy of v.

Example A.6. For any primary ideal $\mathfrak{a}\subset\mathcal{O}_{X,p}$ , the associated nef Cartier b-divisor $B:=Z(\mathfrak{a})$ (see Example A.2) satisfies $e(B)=e(\mathfrak{a})$ , and $\mathrm{lct}(B)=\mathrm{lct}(\mathfrak{a})$ when (X,p) is klt.

Example A.7. Pick any valuation v centered at p, with associated nef b-divisor Z(v) (see Example A.3). Then it follows from [Reference Boucksom, de Fernex and FavreBdFF12, Remark 4.17] that the volume $\mathrm{Vol}(v):=\lim_{m\to\infty}({n!}/{m^n})\dim\mathcal{O}_{X,p}/\mathfrak{a}_m(v) $ satisfies

(A.2) \begin{equation}\mathrm{Vol}(v)=e(Z(v)).\end{equation}

Lemma A.8. For each nef b-divisor B over p, we have $e(B)=\sup_{C\ge B} e(C)$ , where C ranges over all nef Cartier b-divisors of the form $C=m^{-1}Z(\mathfrak{a})$ for a primary ideal $\mathfrak{a}\subset\mathcal{O}_{X,p}$ and $m\in\mathbb Z_{>0}$ , and such that $C\ge B$ .

Proof. Since B is the limit of the decreasing net $(\mathrm{Env}_\pi(B_\pi))$ (see [Reference Boucksom, de Fernex and FavreBdFF12, Remark 2.17]), it is enough to prove the result when $B=\mathrm{Env}_\pi(B_\pi)$ , by continuity of the intersection pairing along decreasing nets. By [Reference Boucksom, de Fernex and FavreBdFF12, Theorem 4.11], we can then write B as the limit of a decreasing sequence $(C_i)$ of nef Cartier b-divisors of the desired form, and we are done since $e(C_i)\to e(B)$ .

Consider now a psh function $\varphi$ on X. The collection of its Lelong numbers on all birational models defines a homogeneous function $v\mapsto v(\varphi)$ on $\mathrm{DivVal}(X,p)$ , and hence an antieffective b-divisor $Z(\varphi)$ over p, such that $v(Z(\varphi))=-v(\varphi)$ .

Proposition A.9. The b-divisor $Z(\varphi)$ is nef. Furthermore:

  1. (i) if $\varphi$ is locally bounded outside p, then

    \[e(Z(\varphi))\le e(\varphi):=(dd^c\varphi)^n(\{p\});\]
  2. (ii) if (X,p) is klt, then $\mathrm{lct}(Z(\varphi))=\mathrm{lct}(\varphi)$ .

Proof. Consider the closed positive (1,1)-current $T:=dd^c\varphi$ , and pick a log resolution $\pi\colon X_\pi\to X$ of (X,p). The Siu decomposition of $\pi^\star T=dd^c\pi^\star\varphi$ shows that $\pi^\star T+[Z(\varphi)_\pi]$ is a positive current with zero generic Lelong numbers along each component of $\pi^{-1}(p)$ . By Demailly regularization, it follows that the class of $Z(\varphi)_\pi$ in $\mathrm{N}^1(X_\pi/X)$ is nef in codimension 1, and hence that $Z(\varphi)$ is nef (see Lemma A.1).

Assume next that $\varphi$ is locally bounded outside p, and pick a primary ideal $\mathfrak{a}\subset\mathcal{O}_{X,p}$ and $m\in\mathbb Z_{>0}$ such that $C:=m^{-1} Z(\mathfrak{a})\ge Z(\varphi)$ . Choose a finite set of local generators $(f_i)$ of $\mathfrak{a}$ , and consider the psh function $\psi:=m^{-1}\log\sum_i|f_i|$ . Then $Z(\varphi)\le C=Z(\psi)$ and, hence, $\varphi\le\psi+O(1)$ (to see this, pull back $\varphi$ and $\psi$ to a log resolution of $\mathfrak{a}$ , and use the Siu decomposition). By Demailly’s comparison theorem, it follows that $e(C)=e(\psi)\le e(\varphi)$ , and taking the supremum over C yields part (i), by Lemma A.8.

Finally, part (ii) is a rather simple consequence of [Reference Berman, Boucksom and JonssonBBJ21, Theorem B.5] applied to the pullback of $\varphi$ to a log resolution of (X,p).

We can now state the following variant of Proposition 5.8.

Theorem A.10. Let (X,p) be an isolated klt singularity. Then

\[\widehat{\mathrm{vol}}(X,p)=\inf_B e(B)\mathrm{lct}(B)^n=\inf_\varphi e(\varphi)\mathrm{lct}(\varphi)^n,\]

where B runs over all nef b-divisors over p, and $\varphi$ runs over all psh functions on X that are locally bounded outside p.

Proof. By Theorem 2.16 we have $\widehat{\mathrm{vol}}(X,p)=\inf_\mathfrak{a} e(\mathfrak{a})\mathrm{lct}(\mathfrak{a})^n$ , where $\mathfrak{a}\subset\mathcal{O}_{X,p}$ runs over all primary divisors, and hence $\widehat{\mathrm{vol}}(X,p)\ge\inf_Be(B)\mathrm{lct}(B)^n$ , by Example A.6. Conversely, pick a nef b-divisor B over p. For any $v\in\mathrm{DivVal}(X,p)$ , Lemma A.4 yields $B\le -v(B) Z(v)$ . By monotonicity and homogeneity of the intersection pairing, this yields $B^n\le (-v(B))^n Z(v)^n$ , i.e. $e(B)\ge(-v(B))^n \mathrm{Vol}(v)$ , by (A.2). Thus,

\[e(B)\bigg(\frac{\mathrm{A}_X(v)}{-v(B)}\bigg)^n\ge\mathrm{A}_X(v)^n\mathrm{Vol}(v)\ge\widehat{\mathrm{vol}}(X,p).\]

Taking the infimum over v yields $e(B)\mathrm{lct}(B)^n\ge\widehat{\mathrm{vol}}(X,p)$ for any nef b-divisor B over p, and hence also $e(\varphi)\mathrm{lct}(\varphi)^n\ge\widehat{\mathrm{vol}}(X,p)$ for any psh function $\varphi$ locally bounded outside p, by Proposition A.9.

Acknowledgements

We thank Sébastien Boucksom for writing Appendix A, which enhances Proposition 5.8. We are grateful to the referee for suggesting many improvements.

Conflicts of interest

None.

Financial support

The authors are supported by the project Hermetic (ANR-11-LABX-0040) and the ANR project Paraplui. The second author is supported by a grant from the Knut and Alice Wallenberg foundation.

Journal information

Compositio Mathematica is owned by the Foundation Compositio Mathematica and published by the London Mathematical Society in partnership with Cambridge University Press. All surplus income from the publication of Compositio Mathematica is returned to mathematics and higher education through the charitable activities of the Foundation, the London Mathematical Society and Cambridge University Press.

Footnotes

With an appendix by Sébastien Boucksom.

References

Åhag, P., Cegrell, U. and Czyz, R. , On Dirichlet’s principle and problem , Math. Scand. 110 (2012), 235250.CrossRefGoogle Scholar
Åhag, P., Cegrell, U., Kołodziej, S., Pham, H. H. and Zeriahi, A., Partial pluricomplex energy and integrability exponents of PSH functions , Adv. Math. 222 (2009), 20362058.CrossRefGoogle Scholar
Bakker, B., Guenancia, H. and Lehn, C., Algebraic approximation and decomposition theorem for Kähler Calabi-Yau varieties , Invent. Math. 228 (2022), 12551308.CrossRefGoogle Scholar
Bedford, E. and Taylor, B. A., A new capacity for plurisubharmonic functions, Acta Math. 149 (1982), 140.Google Scholar
Berman, R. and Berndtsson, B., The volume of Kähler–Einstein Fano varieties and convex bodies, J. Reine Angew. Math. 723 (2017), 127152.Google Scholar
Berman, R. and Berndtsson, B., Moser–Trudinger type inequalities for complex Monge–Ampère operators and Aubin’s “hypothèse fondamentale”, Ann. Fac. Sci. Toulouse Math. (6) 31 (2022), 595645.CrossRefGoogle Scholar
Berman, R. J. and Boucksom, S., Growth of balls of holomorphic sections and energy at equilibrium , Invent. Math. 181 (2010), 337394.CrossRefGoogle Scholar
Berman, R. J., Boucksom, S., Eyssidieux, P., Guedj, V. and Zeriahi, A., Kähler–Einstein metrics and Kähler–Ricci flow on log Fano varieties, J. Reine Angew. Math. 751 (2019), 2789.Google Scholar
Berman, R. J., Boucksom, S., Guedj, V. and Zeriahi, A., A variational approach to complex Monge–Ampère equations, Publ. Math. Inst. Hautes Études Sci. 117 (2013), 179245.Google Scholar
Berman, R. J., Boucksom, S. and Jonsson, M., A variational approach to the Yau–Tian–Donaldson conjecture, J. Amer. Math. Soc. 34 (2021), 605652.Google Scholar
Błocki, Z., The domain of definition of the complex Monge–Ampère operator, Amer. J. Math. 128 (2006), 519530.CrossRefGoogle Scholar
Blum, H., Existence of valuations with smallest normalized volume, Compositio Math. 154 (2018), 820849.Google Scholar
Boucksom, S., de Fernex, T. and Favre, C., The volume of an isolated singularity, Duke Math. J. 161 (2012), 14551520.Google Scholar
Boucksom, S., Favre, C. and Jonsson, M., Valuations and plurisubharmonic singularities, Publ. Res. Inst. Math. Sci. 44 (2008), 449494.Google Scholar
Caffarelli, L., Kohn, J. J., Nirenberg, L. and Spruck, J., The Dirichlet problem for nonlinear second-order elliptic equations. II. Complex Monge–Ampère, and uniformly elliptic, equations , Comm. Pure Appl. Math. 38 (1985), 209252.CrossRefGoogle Scholar
Cegrell, U., Pluricomplex energy, Acta Math. 180 (1998), 187217.Google Scholar
Cegrell, U., The general definition of the complex Monge–Ampère operator , Ann. Inst. Fourier (Grenoble) 54 (2004), 159179.CrossRefGoogle Scholar
Chen, X. X., Donaldson, S. and Sun, S., Kähler–Einstein metrics on Fano manifolds, I, II & III, J. Amer. Math. Soc. 28 (2015), 183197, 199–234 & 235–278.Google Scholar
Datar, V., Fu, X. and Song, J., Kähler–Einstein metric near an isolated log canonical singularity, J. Reine Angew. Math. 797 (2023), 79116.Google Scholar
Demailly, J.-P., Mesures de Monge–Ampère et caractérisation géométrique des variétés algébriques affines , Mém. Soc. Math. Fr. (N.S.) 19 (1985).Google Scholar
Demailly, J.-P., Estimates on Monge–Ampère operators derived from a local algebra inequality, in Complex analysis and digital geometry (Uppsala Universitet, Uppsala, 2009), 131143.Google Scholar
Demailly, J.-P., Analytic methods in algebraic geometry, Surveys of Modern Mathematics, vol. 1 (International Press–Higher Education Press, Beijing, 2012).Google Scholar
Demailly, J.-P. and Kollàr, J., Semi-continuity of complex singularity exponents and Kähler–Einstein metrics on Fano orbifolds , Ann. Sci. Éc. Norm. Supéer. (4) 34 (2001), 525556.CrossRefGoogle Scholar
Demailly, J.-P. and Pham, H. H., A sharp lower bound for the log canonical threshold, Acta Math. 212 (2014), 19.Google Scholar
Druel, S., A decomposition theorem for singular spaces with trivial canonical class of dimension at most five , Invent. Math. 211 (2018), 245296.CrossRefGoogle Scholar
Ein, L., Lazarsfeld, R. and Smith, K. E., Uniform approximation of Abhyankar valuation ideals in smooth function fields, Amer. J. Math. 125 (2003), 409440.CrossRefGoogle Scholar
Eyssidieux, P., Guedj, V. and Zeriahi, A., Singular Kähler–Einstein metrics, J. Amer. Math. Soc. 22 (2009), 607639.Google Scholar
Fornæss, J.-E. and Narasimhan, R., The Levi problem on complex spaces with singularities , Math. Ann. 248 (1980), 4772.CrossRefGoogle Scholar
Fu, X., Dirichlet problem of complex Monge–Ampère equation near an isolated KLT singularity, Commun. Contemp. Math. 25 (2023), 2250019.Google Scholar
Fujita, K., Uniform K-stability and plt blowups of log Fano pairs, Kyoto J. Math. 59 (2019), 399418.Google Scholar
Guan, B. and Li, Q., Complex Monge–Ampère equations and totally real submanifolds , Adv. Math. 225 (2010), 11851223.CrossRefGoogle Scholar
Guan, Q. and Zhou, X., A proof of Demailly’s strong openness conjecture , Ann. of Math. (2) 182 (2015), 605616.CrossRefGoogle Scholar
Guedj, V., Guenancia, H. and Zeriahi, A., Continuity of singular Kähler–Einstein potentials , Int. Math. Res. Not. IMRN 2 (2023), 13551377.CrossRefGoogle Scholar
Guedj, V., Kolev, B. and Yeganefar, N., Kähler–Einstein fillings , J. Lond. Math. Soc. (2) 88 (2013), 737760.CrossRefGoogle Scholar
Guedj, V. and Zeriahi, A., Degenerate complex Monge–Ampère equations, EMS Tracts in Mathematics, vol. 26 (European Mathematical Society (EMS), Zürich, 2017).CrossRefGoogle Scholar
Hartshorne, R., Algebraic geometry , Graduate Texts in Mathematics, vol. 52 (Springer, New York–Heidelberg, 1977).Google Scholar
Hein, H.-J. and Sun, S., Calabi-Yau manifolds with isolated conical singularities, Publ. Math. Inst. Hautes Études Sci. 126 (2017), 73130.Google Scholar
Höring, A. and Peternell, T., Algebraic integrability of foliations with numerically trivial canonical bundle , Invent. Math. 216 (2019), 395419.CrossRefGoogle Scholar
Kołodziej, S., The complex Monge–Ampère equation , Acta Math. 180 (1998), 69117.CrossRefGoogle Scholar
Li, C., Minimizing normalized volumes of valuations, Math. Z. 289 (2018), 491513.Google Scholar
Li, C., G-uniform stability and Kähler–Einstein metrics on Fano varieties , Invent. Math. 227 (2022), 661744.CrossRefGoogle Scholar
Li, C., Tian, G. and Wang, F., On Yau–Tian–Donaldson conjecture for singular Fano varieties, Comm. Pure Appl. Math. 74 (2021), 17481800.Google Scholar
Li, C., Wang, X. and Xu, C., Algebraicity of the metric tangent cones and equivariant K-stability, J. Amer. Math. Soc. 34 (2021), 11751214.Google Scholar
Liu, Y., The volume of singular Kähler–Einstein Fano varieties, Compositio Math. 154 (2018), 11311158.Google Scholar
Liu, Y., Xu, C. and Zhang, Z., Finite generation for valuations computing stability thresholds and applications to K-stability, Ann. of Math. (2) 196 (2022), 507566.CrossRefGoogle Scholar
Pham, H. H., The weighted log canonical threshold , C. R. Math. Acad. Sci. Paris 352 (2014), 283288.Google Scholar
Pham, H. H., Log canonical thresholds and Monge–Ampère masses , Math. Ann. 370 (2018), 555566.CrossRefGoogle Scholar
Phong, D. H., Song, J. and Sturm, K., Complex Monge–Ampère equations , Surv. Differ. Geom. 17 (2012), 327410.CrossRefGoogle Scholar
Siu, Y.-T., Lectures on Hermitian–Einstein metrics for stable bundles and Kähler–Einstein metrics , DMV Seminar, vol. 8 (Birkhäuser, Basel, 1987).Google Scholar
Tian, G., On Kähler–Einstein metrics on certain Kähler manifolds with $C_1(M)>0$ , Invent. Math. 89 (1987), 225246.CrossRefGoogle Scholar
Yau, S. T., On the Ricci curvature of a compact Kähler manifold and the complex Monge-Ampère equation . I, Comm. Pure Appl. Math. 31 (1978), 339411.CrossRefGoogle Scholar
Zeriahi, A., A stronger version of Demailly’s estimate on Monge–Ampère operators, in Complex analysis and digital geometry (Uppsala Universitet, Uppsala, 2009), 144146.Google Scholar