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

Weighted cscK metrics on Kähler varieties

Chung-Ming Pan Centre interuniversitaire de recherches en géométrie et topologies (CIRGET); Université du Québec à Montréal; Case postale 8888, Succursale centre-ville, Montréal, Québec, H3C 3P8, Canada pan.chung_ming@uqam.ca                               and  Tat Dat Tô Institut de Mathématiques de Jussieu-Paris Rive Gauche; Sorbonne Université - Campus Pierre et Marie Curie, 4 place Jussieu, 75252 Paris Cedex 05, France                        tat-dat.to@imj-prg.fr                       
(September 6, 2025)
Abstract.

We study the weighted constant scalar curvature Kähler equations on mildly singular Kähler varieties. Assuming the existence of a suitable resolution of singularities, we establish the existence of singular weighted cscK metrics when the weighted Mabuchi functional is coercive for an extremal weight. This extends the works of Chen-Cheng and He to the singular weighted setting. Moreover, we provide a method for constructing examples of singular cscK metrics inspired by the work of Arezzo–Pacard. In contrast to the usual gluing techniques, our approach does not require a precise understanding about of the metric behavior near the singular locus.

Key words and phrases:
Weighted cscK metric, Log terminal singularities, A priori estimates
1991 Mathematics Subject Classification:
53C55, 32J27, 32Q20, 32W20, 35A23

Introduction

The constant scalar curvature Kähler (cscK) metric problem has become one of the central focus in Kähler geometry during the last decades. The Yau–Tian–Donaldson conjecture asserts that given a compact Kähler manifold with a fixed Kähler class, the existence of cscK metrics in the Kähler class is equivalent to an algebro-geometric notion called ”K-stability”.

Several progresses in the literature [DR17, BDL20, CC21a, CC21b] have shown that the existence of a unique cscK metric in a Kähler class is equivalent to the coercivity of the Mabuchi functional, whose minimizers are cscK metrics. Boucksom–Hisamoto–Jonsson [BHJ19] demonstrated that the coercivity of the Mabuchi functional implies uniform K-stability (see [DR17, Der18, SD18, SD20] for a transcendental setup). Conversely, C. Li [Li22b] (and the recent transcendental version by Mesquita-Piccione [MP24]) showed that the strong uniform K-stability implies the coercivity of Mabuchi functional. The remaining challenge in proving the uniform Yau–Tian–Donaldson conjecture lies in establishing strong uniform K-stability from uniform K-stability.

Typical examples of cscK metrics are Kähler–Einstein metrics. Motivated by Minimal Model Program and moduli theory, Kähler–Einstein metrics have been well studied on smooth and mildly singular Kähler varieties [Aub78, Yau78, EGZ09, CDS15, BBE+19, BBJ21, LTW21, LTW22, Li22a] and their families [Koi83, RZ11a, RZ11b, SSY16, LWX19, DGG23, PT25] etc.

However, there are very few results regarding cscK metrics in the singular setting. We shall focus on the analytic part of the Yau–Tian–Donaldson conjecture on mildly singular varieties, particularly the relation between the existence of singular cscK metrics and the coercivity of the Mabuchi functional and explore under the weighted formalism. In a recent joint work with Trusiani [PTT23], when the Mabuchi functional is coercive, we establish the existence of singular cscK metrics on \mathbb{Q}-Gorenstein smoothable Kähler varieties with log terminal singularities. One of the key ingredients is the stability of the coercivity of the Mabuchi functional [PTT23, Thm. A]. A similar strategy for establishing openness of coercivity has been applied to the resolution setting by Boucksom–Jonsson–Trusiani [BJT24] under an appropriate condition on the resolution.

This article aims to remove the additional smoothable assumption and to investigate existence results in a more general weighted setting. The weighted framework introduced by Lahdili [Lah19] (see also [Ino22]) includes various notions of canonical Kähler metrics, for example, extremal Kähler metrics and Kähler–Ricci solitons. For further results on the existence of weighted cscK metrics, we refer to [AJL23, Lah23, DJL24, DJL25, HL25] and the references therein.

We quickly review the basic setup and notations for the weighted cscK metrics below (see Section 1 for more details).

Setting (GS).

Let XX be an nn-dimensional compact Kähler variety with log terminal singularities, let TAutred(X)T\subset\operatorname{Aut}_{\operatorname{red}}(X) be a maximal real torus in the reduced automorphism group, and let ω\omega be a TT-invariant Kähler metric on XX. Denote by TT_{\mathbb{C}} the complexification of TT, 𝔱\mathfrak{t} the Lie algebra of TT and mω:X𝔱m_{\omega}:X\to\mathfrak{t}^{\vee} a moment map associated to ω\omega. Consider functions v,w𝒞(𝔱,)v,w\in\mathcal{C}^{\infty}(\mathfrak{t}^{\vee},\mathbb{R}) with v>0v>0 on mω(X)m_{\omega}(X).

A metric ω\omega^{\prime} is a singular (v,w)(v,w)-cscK metric if it has locally bounded potentials on XX and it solves the weighted cscK equation Sv(ω)=w(mω)S_{v}(\omega^{\prime})=w(m_{\omega^{\prime}}) on XregX^{\mathrm{reg}}, where SvS_{v} is the weighted scalar curvature. Such an equation is the Euler–Lagrange equation of the (v,w)(v,w)-weighted Mabuchi functional 𝐌v,w\mathbf{M}_{v,w}. When w=w0w=\ell w_{0}, where w0>0w_{0}>0 on mω(X)m_{\omega}(X) and \ell is an affine function such that the Mabuchi functional 𝐌v,w\mathbf{M}_{v,w} is TT_{\mathbb{C}}-invariant, the corresponding (v,w)(v,w)-cscK metrics are called (v,w0)(v,w_{0})-extremal metrics. In this context, ww is referred to as an extremal weight. The cscK metrics and extremal Kähler metrics correspond to the special case of (1,1)(1,1)-extremal metrics.

Before presenting our results, we introduce the following additional condition:

Condition (A).

There exists a TT-equivariant resolution of singularities π:YX\pi:Y\to X such that YY is Kähler and π\pi is an isomorphism over π1(Xreg)\pi^{-1}(X^{\mathrm{reg}}). Also, there exist a Kähler metric ωY\omega_{Y} on YY, a positive constant K1>0K_{1}>0, and a function ρ1QPSH(Y)\rho_{1}\in\operatorname{QPSH}(Y) such that

Ric(ωY)K1(πω+ddcρ1)andYeK1ρ1ωYn<+.\operatorname{Ric}(\omega_{Y})\geq-K_{1}(\pi^{\ast}\omega+dd^{c}\rho_{1})\quad\text{and}\quad\int_{Y}e^{-K_{1}\rho_{1}}\omega^{n}_{Y}<+\infty.

In the above notation, QPSH(Y)\operatorname{QPSH}(Y) is the set of all quasi-plurisubharmonic functions on YY. A resolution of singularities described in Condition (A) is referred to as a resolution of Fano type in [BJT24] (see Section 3.3.1 and [BJT24, Sec. 4.1] for further discussions and examples).

Under Condition (A), we establish the following existence theorem for singular weighted cscK metrics:

Theorem A.

Let (X,ω)(X,\omega) be a compact Kähler variety with log terminal singularities that satisfies Setting (GS). Assume that (X,ω)(X,\omega) satisfies Condition (A), vv is log\log-concave and ww is an extremal weight. If the weighted Mabuchi functional 𝐌v,w\mathbf{M}_{v,w} is TT_{\mathbb{C}}-coercive, then XX admits a singular (v,w)(v,w)-cscK metric in {ω}\{\omega\}, which also minimizes 𝐌v,w\mathbf{M}_{v,w}.

For cscK and extremal metrics, the weight v1v\equiv 1 on 𝔱\mathfrak{t}^{\vee} (hence log\log-concave). In the smooth setup, existence results under the coercivity of the Mabuchi functional are obtained by [CC21b] and [He19] for cscK and extremal metrics, respectively. Our Theorem A, in particular, extends their results to singular settings. We state a direct corollary for the case of cscK metrics, corresponding v=1v=1 and w=4πnc1(X)[ω]n1[ω]nw=\frac{4\pi nc_{1}(X)\cdot[\omega]^{n-1}}{[\omega]^{n}} in the following.

Corollary B.

Let XX be a compact Kähler variety with log terminal singularities and let TAutred(X)T\subset\operatorname{Aut}_{\operatorname{red}}(X) be a maximal real torus. Assume that ω\omega is a TT-invariant Kähler metric and that (X,ω)(X,\omega) satisfies Condition (A). If the Mabuchi functional 𝐌ω\mathbf{M}_{\omega} on XX is TT_{\mathbb{C}}-coercive, then XX admits a singular cscK metric in {ω}\{\omega\} which also minimizes 𝐌ω\mathbf{M}_{\omega}. In particular, if Aut(X)\operatorname{Aut}(X) is discrete, the coercivity of the Mabuchi functional implies the existence of singular cscK metrics in {ω}\{\omega\}.

The strategy for proving Theorem A is to establish uniform a priori estimates for the solutions to a family of weighted cscK equations on the resolution of singularities π:YX\pi:Y\to X. Denote by EE the exceptional set of π\pi. Consider a family of perturbed Kähler metrics ωε:=πω+εωY\omega_{\varepsilon}:=\pi^{\ast}\omega+\varepsilon\omega_{Y} on the resolution for ε(0,1]\varepsilon\in(0,1]. Under Condition (A), and assuming ww is an extremal weight, [BJT24, Thm. A] has proved the openness of uniform coercivity on YY. By [DJL25, HL25], there exists a Kähler metric ωε,φε:=ωε+ddcφε\omega_{\varepsilon,\varphi_{\varepsilon}}:=\omega_{\varepsilon}+dd^{c}\varphi_{\varepsilon} solving the weighted cscK equation on YY. Then the uniform coercivity of weighted Mabuchi functional yields a uniform control on the entropy 𝐇ε(φε)\mathbf{H}_{\varepsilon}(\varphi_{\varepsilon}). Under Condition (A) and the bound on the entropy we establish priori estimates for the weighted cscK equation extending the estimates of Chen and Cheng to the degenerate weighted setting. Consequently, one can then extract a subsequence φε\varphi_{\varepsilon} converging in 𝒞oc(YE)\mathcal{C}^{\infty}_{\mathrm{\ell oc}}(Y\setminus E) to a bounded πω\pi^{\ast}\omega-psh function φ0\varphi_{0}, which is smooth on YEXregY\setminus E\simeq X^{\mathrm{reg}} and solves the weighted cscK equation there.

In particular, within the context of Corollary B, these estimates shows that singular Kähler–Einstein metrics on a Kähler varieties can be approximated by extremal metrics, provided Condition (A) holds, and this generalizes a recent result by Székelyhidi [Szé24, Thm. 3] for non-discrete automorphism group.

We highlight below the main difficulties and contributions made in the article:

  • LL^{\infty}-estimate and Condition (A). The main difficulty to obtain uniform LL^{\infty}-estimate is the lack of uniform bounds for the Ricci curvature of ωY\omega_{Y} with respect to the reference metrics (ωε)ε(\omega_{\varepsilon})_{\varepsilon} on the resolution of singularities. That is the reason why the estimates of Chen–Cheng [CC21a] and Guo–Phong [GP24] cannot be applied directly. To overcome this difficulty, we follow and generalize the Guo-Phong’s approach by incorporating Condition (A), which provides a weak version of lower bound for Ric(ωY)\operatorname{Ric}(\omega_{Y}), together with the strong openness [Ber13, GZ15], and Demailly–Kollar’s theorem [DK01]. Moreover, our LL^{\infty}-estimate, Theorem 2.1, does not require that vv be log-concave and ww be an extremal weight.

  • Constructing examples. In Section 3.3, we present a method for constructing singular cscK metrics inspired by Arezzo–Pacard [AP06] under Condition (A). Additionally, we propose a mixed construction that integrates the result on the smoothable setting [PTT23]. An important ingredient in the construction is the stability of the coercivity of the Mabuchi functional for the blow-up along compact submanifolds in the smooth locus of singular varieties (cf. Lemma 3.6) instead of desingularizations as in [BJT24]. Comparing to the usual gluing technique (cf. [AP06]), our method does not require a very precise understanding of how the metric behaves near the singular locus.

We provide further detailed comments on the proof and the technical assumption on the log\log-concavity of vv. For higher order estimates, we adapt the strategy of Chen–Cheng [CC21b] and its generalization for the weighted setting [HL25, DJL24]. Due to the degeneration of ωϵ\omega_{\epsilon}, we modify the test function for the Laplacian by adding a strictly πω\pi^{*}\omega-psh function ρ\rho with analytic singularities, ensuring ωε+ddcρCωY\omega_{\varepsilon}+dd^{c}\rho\geq C\omega_{Y} to absorb problematic terms. We obtain an integral Laplacian estimate with respect to μ=eCρωYn\mu=e^{C^{\prime}\rho}\omega_{Y}^{n}, giving a local integral Laplacian estimate away from the exceptional locus. The log\log-concavity of vv is used in the uniform integral Laplacian estimate in order to eliminate a problematic term in the weighted Aubin–Yau inequality. Local Laplacian and higher order estimates then follow from a generalization of Chen–Cheng’s local estimates for the weighted setting (cf. Appendix B). We remark that our estimates apply to general ww without requiring it to be an extremal weight.

Acknowledgements.

The authors are grateful to C. Arezzo, S. Boucksom, E. Di Nezza, S. Jubert, A. Lahdili, Y. Odaka, S. Sun, G. Székelyhidi, and A. Trusiani for helpful and inspiring discussions. The authors would like to thank S. Boucksom, M. Jonsson, A. Trusiani, and G. Székelyhidi for kindly sharing their articles. The authors would also like to thank V. Guedj and H. Guenancia for their suggestions that helped improve the exposition.

Part of this article is based upon work supported by the National Science Foundation under Grant No. DMS-1928930, while the first named author was in residence at the Simons Laufer Mathematical Sciences Institute (formerly MSRI) in Berkeley, California, during the Fall 2024 semester. The second named author is partially supported by ANR-21-CE40-0011-01 (research project MARGE), PEPS-JCJC-2024 (CRNS) and Tremplins-2024 (Sorbonne University).

1. Preliminaries

In this section, we recall several basic notions of pluripotential theory on singular spaces and weighted cscK metrics. We define dc:=i(¯)d^{c}:=\mathrm{i}(\bar{\partial}-\partial) and then we have ddc=2i¯dd^{c}=2\mathrm{i}\partial\bar{\partial}. By variety, we always mean an irreducible reduced complex analytic space.

1.1. Pluripotential theory on normal Kähler varieties

Let XX be an nn-dimensional compact normal complex variety. A smooth Kähler metric on XX is defined by a Kähler metric ω\omega on XregX_{\mathrm{reg}}, and it is locally a restriction of a Kähler metric defined near the image of a local embedding j:Xoc.Nj:X\underset{\mathrm{\ell oc}.}{\hookrightarrow}\mathbb{C}^{N}. By Kähler variety, we mean a complex variety XX equipped with a smooth Kähler metric ω\omega. More generally, a smooth form α\alpha on XX is defined as a smooth form on XregX^{\mathrm{reg}} such that α\alpha extends smoothly under any local embedding Xoc.NX\underset{\mathrm{\ell oc}.}{\hookrightarrow}\mathbb{C}^{N}.

Definition 1.1.

A function ϕ:X{}\phi:X\to\mathbb{R}\cup\{-\infty\} is ω\omega-plurisubharmonic (ω\omega-psh for short) if ϕ+u\phi+u is plurisubharmonic where uu is a local potential of ω\omega; i.e. ϕ+u\phi+u is the restriction of an psh function defined near an open neighborhood of im(j)\operatorname{im}(j). Denote by PSH(X,ω)\operatorname{PSH}(X,\omega) the set of all integrable ω\omega-psh functions.

By Bedford–Taylor’s theory [BT82], the complex Monge–Ampère operator can be extended to bounded ω\omega-psh functions on smooth complex manifolds. In the singular setting, the complex Monge–Ampère operator of locally bounded psh functions can also be defined by taking zero through singular locus (cf. [Dem85] for more details).

1.1.1. Finite energy class

Set V:=XωnV:=\int_{X}\omega^{n}. For all φPSH(X,ω)L(X)\varphi\in\operatorname{PSH}(X,\omega)\cap L^{\infty}(X), the Monge–Ampère energy is defined by

𝐄(φ):=1(n+1)j=0nXφωφjωnj,\mathbf{E}(\varphi):=\frac{1}{(n+1)}\sum_{j=0}^{n}\int_{X}\varphi\,\omega_{\varphi}^{j}\wedge\omega^{n-j},

where ωφ:=ω+ddcφ\omega_{\varphi}:=\omega+dd^{c}\varphi. The energy satisfies 𝐄(φ+c)=𝐄(φ)+cV\mathbf{E}(\varphi+c)=\mathbf{E}(\varphi)+cV for all cc\in\mathbb{R} and for φ,ψPSH(X,ω)L(X)\varphi,\psi\in\operatorname{PSH}(X,\omega)\cap L^{\infty}(X), if φψ\varphi\leq\psi then 𝐄(φ)𝐄(ψ)\mathbf{E}(\varphi)\leq\mathbf{E}(\psi) with equality iff φ=ψ\varphi=\psi. By the later property, 𝐄\mathbf{E} extends uniquely to PSH(X,ω)\operatorname{PSH}(X,\omega) by

𝐄(φ):=inf{𝐄(ψ)|φψPSH(X,ω)L(X)}.\mathbf{E}(\varphi):=\inf\left\{{\mathbf{E}(\psi)}\,\middle|\,{\varphi\leq\psi\in\operatorname{PSH}(X,\omega)\cap L^{\infty}(X)}\right\}.

The finite energy class is then given as

1(X,ω)={φPSH(X,ω)|𝐄(φ)>}.\mathcal{E}^{1}(X,\omega)=\left\{{\varphi\in\operatorname{PSH}(X,\omega)}\,\middle|\,{\mathbf{E}(\varphi)>-\infty}\right\}.

The space of finite energy potentials 1(X,ω)\mathcal{E}^{1}(X,\omega) admits a metric topology induced by the so-called d1d_{1}-distance, which can be expressed as follows (cf. [Dar17, Thm. 2.1], [DG18, Thm. B])

d1(u,v)=𝐄(u)+𝐄(v)2𝐄(Pω(u,v)),d_{1}(u,v)=\mathbf{E}(u)+\mathbf{E}(v)-2\mathbf{E}(P_{\omega}(u,v)),

where Pω(u,v):=(sup{wPSH(X,ω)|wmin(u,v)})P_{\omega}(u,v):=\left(\sup\left\{{w\in\operatorname{PSH}(X,\omega)}\,\middle|\,{w\leq\min(u,v)}\right\}\right)^{*}.

1.2. Weighted cscK metrics

In this section, we review some basic concepts related to weighted cscK metrics.

1.2.1. Automorphism group and holomorphic vector fields

We recall here certain well-known properties of the Lie algebra of Aut(X)\operatorname{Aut}(X) and some of its subgroups (cf. [LS94, Gau20]).

Let (X,ω)(X,\omega) be a compact Kähler manifold and denote JJ to be its complex structure. The automorphism group Aut(X)\operatorname{Aut}(X) is a complex Lie group, whose Lie algebra 𝔥={ξΓ(TX)|ξJ=0}\mathfrak{h}=\left\{{\xi\in\Gamma(TX)}\,\middle|\,{\mathcal{L}_{\xi}J=0}\right\} consists of real holomorphic vector fields on XX. A vector field ξ\xi is real holomorphic if and only if ξ:=12(ξiJξ)\xi^{\prime}:=\frac{1}{2}(\xi-iJ\xi) is a holomorphic, so 𝔥\mathfrak{h} can be identified with H0(X,TX)H^{0}(X,T_{X}) the space of holomorphic vector fields. Denote by Aut0(X)\operatorname{Aut}_{0}(X) the identity component of Aut(X)\operatorname{Aut}(X).

The Albanese torus is defined by (cf. [Uen75, Thm. 9.7])

Alb(X)=H0(X,ΩX1)/H1(X,),\text{Alb}(X)=H^{0}(X,\Omega_{X}^{1})^{\vee}/H_{1}(X,\mathbb{Z}),

where the inclusion H1(X,)H0(X,ΩX1)H_{1}(X,\mathbb{Z})\xhookrightarrow{}H^{0}(X,\Omega_{X}^{1})^{\vee} is induced via the map αcα\alpha\mapsto\int_{c}\alpha for any loop cc and holomorphic 11-form α\alpha. Then one can define a Lie group homomorphism τ:Aut0(X)Alb(X)\tau:\operatorname{Aut}_{0}(X)\rightarrow\text{Alb}(X) as follows. Fix a x0Xx_{0}\in X, for any σAut0(X)\sigma\in\operatorname{Aut}_{0}(X), we define τ(σ):αx0σx0α\tau(\sigma):\alpha\mapsto\int_{x_{0}}^{\sigma\cdot x_{0}}\alpha, which does not depend on x0x_{0} as α\alpha is harmonic. The homomorphism τ\tau induces an action of Aut0(X)\operatorname{Aut}_{0}(X) on Alb(X)\text{Alb}(X) by translation. Moreover, the derivative of τ\tau at the identity is τ:𝔥H0(X,ΩX1)\tau^{\prime}:\mathfrak{h}\rightarrow H^{0}(X,\Omega_{X}^{1})^{\vee}, ξ(τ(ξ):αα(ξ))\xi\mapsto(\tau^{\prime}(\xi):\alpha\mapsto\alpha(\xi)). Define 𝔥red:=kerτ\mathfrak{h}_{\operatorname{red}}:=\ker\tau^{\prime}, which consists of all holomorphic vector field ξH0(X,TX)\xi\in H^{0}(X,T_{X}) such that α(ξ)=0\alpha(\xi)=0 for all αH0(X,ΩX1)\alpha\in H^{0}(X,\Omega_{X}^{1}). By [LS94, Thm. 1], we have

𝔥red={ξH0(X,TX)ξ=1,0f=gjk¯k¯fzj,f𝒞(X,)}.\mathfrak{h}_{\operatorname{red}}=\{\xi\in H^{0}(X,T_{X})\mid\xi=\nabla^{1,0}f=g^{j\bar{k}}\partial_{\bar{k}}f\frac{\partial}{\partial z^{j}},\,f\in\mathcal{C}^{\infty}(X,\mathbb{C})\}.

Then 𝔥red\mathfrak{h}_{\operatorname{red}} generates a Lie subgroup Autred(X)Aut0(X)\operatorname{Aut}_{\operatorname{red}}(X)\subset\operatorname{Aut}_{0}(X) which acts trivially on the Albanese torus. We also identify 𝔥red𝔥\mathfrak{h}_{\operatorname{red}}\subset\mathfrak{h} with the Lie algebra of real holomorphic vector fields with zeros (cf. [LS94, Thm. 1]).

1.2.2. Weighted setting

Fix a real torus TAutred(X)T\subset\operatorname{Aut}_{\operatorname{red}}(X) with Lie algebra 𝔱𝔥red\mathfrak{t}\subset\mathfrak{h}_{\operatorname{red}}. Any closed, real, TT-invariant (1,1)(1,1)-form ω\omega admits a moment map mω:X𝔱m_{\omega}:X\rightarrow\mathfrak{t}^{\vee}, that is a unique (up to additive constant) TT-invariant smooth map such that for each ξ𝔱\xi\in\mathfrak{t}, mωξ:=mω,ξ:Xm_{\omega}^{\xi}:=\langle m_{\omega},\xi\rangle:X\rightarrow\mathbb{R} satisfies dmωξ=iξω=ω(ξ,)-dm_{\omega}^{\xi}=i_{\xi}\omega=\omega(\xi,\cdot); in other words, mωξm^{\xi}_{\omega} is a Hamiltonian function of ξ\xi with respect to ω\omega.

Denote by Pω:=im(mω)P_{\omega}:=\operatorname{im}(m_{\omega}) a compact set in 𝔱\mathfrak{t}^{\vee} and ωT:=T(X,ω)\mathcal{H}^{T}_{\omega}:=\mathcal{H}^{T}(X,\omega) the set of TT-invariant, strictly positive, smooth, ω\omega-psh functions. For any ϕωT\phi\in\mathcal{H}^{T}_{\omega}, we normalize mϕ:=mωϕm_{\phi}:=m_{\omega_{\phi}} by mϕξ=mωξ+dcϕ(ξ)m_{\phi}^{\xi}=m_{\omega}^{\xi}+d^{c}\phi(\xi) for all ξ𝔱\xi\in\mathfrak{t}. Then from [Lah19, Lem. 1], under this normalization, for any ϕωT\phi\in\mathcal{H}^{T}_{\omega}, one has Pωϕ=PωP_{\omega_{\phi}}=P_{\omega} and

Xmϕξωϕn=Xmωξωn,ξ𝔱.\int_{X}m^{\xi}_{\phi}\omega^{n}_{\phi}=\int_{X}m^{\xi}_{\omega}\omega^{n},\quad\forall\xi\in\mathfrak{t}. (1.1)

Let v,w𝒞(𝔱,)v,w\in\mathcal{C}^{\infty}(\mathfrak{t}^{\vee},\mathbb{R}) with v>0v>0 on P:=PωP:=P_{\omega}. We recall the following notations and definitions from [BJT24, Sec. 3],

Definition 1.2.

Let (θ,mθ)(\theta,m_{\theta}) be a TT-invariant pair and ff be a TT distribution.

  1. (1)

    A moment map mddcfm_{dd^{c}f} for the TT-invariant (1,1)(1,1)-current ddcfdd^{c}f is defined by

    mddcfξ:=dcf(ξ)=iξdcf;m_{dd^{c}f}^{\xi}:=d^{c}f(\xi)=i_{\xi}d^{c}f;
  2. (2)

    For a smooth function g:>0g:\mathbb{R}_{>0}\to\mathbb{R}, and another TT-invariant pair (ω,mω)(\omega,m_{\omega}),

    (g(v))(mω),mθ:=g(v(mω))αvα(mω)mθξα\langle(g(v))^{\prime}(m_{\omega}),m_{\theta}\rangle:=g^{\prime}(v(m_{\omega}))\sum_{\alpha}v_{\alpha}(m_{\omega})m_{\theta}^{\xi_{\alpha}}

    where (ξα)α(\xi_{\alpha})_{\alpha} is a basis of 𝔱\mathfrak{t}, and (vα)α(v_{\alpha})_{\alpha} are the partial derivatives of vv with respect to the dual basis (ξα)α(\xi^{\alpha})_{\alpha} in 𝔱\mathfrak{t}^{\vee};

  3. (3)

    The vv-weighted trace is given by

    trω,vθ:=trωθ+(logv)(mω),mθ;\operatorname{tr}_{\omega,v}\theta:=\operatorname{tr}_{\omega}\theta+\langle(\log v)^{\prime}(m_{\omega}),m_{\theta}\rangle;
  4. (4)

    The vv-weighted Laplacian is defined as

    Δω,vf:=trω,v(ddcf)=Δωf+(logv)(mω),mddcf;\Delta_{\omega,v}f:=\operatorname{tr}_{\omega,v}(dd^{c}f)=\Delta_{\omega}f+\langle(\log v)^{\prime}(m_{\omega}),m_{dd^{c}f}\rangle;
  5. (5)

    The vv-weighted Ricci curvature is

    Ricv(ω):=Ric(v(mω)ωn)=ddclog(v(mω)ωn)=Ric(ω)ddclogv(mω);\operatorname{Ric}_{v}(\omega):=\operatorname{Ric}(v(m_{\omega})\omega^{n})=-dd^{c}\log(v(m_{\omega})\omega^{n})=\operatorname{Ric}(\omega)-dd^{c}\log v(m_{\omega});
  6. (6)

    The vv-weighted scalar curvature is

    Sv(ω):=trω,vRicv(ω)=S(ω)v(mω),Δωmω+Δωv(mω)v(mω).S_{v}(\omega):=\operatorname{tr}_{\omega,v}\operatorname{Ric}_{v}(\omega)=S(\omega)-\frac{\langle v^{\prime}(m_{\omega}),\Delta_{\omega}m_{\omega}\rangle+\Delta_{\omega}v(m_{\omega})}{v(m_{\omega})}.

As [BJT24, (3.4)], by applying the interior product iξi_{\xi} to the trivial relation dcfωn=0d^{c}f\wedge\omega^{n}=0, one can obtain mddcfξωn=ndcfiξωωn1m_{dd^{c}f}^{\xi}\omega^{n}=nd^{c}f\wedge i_{\xi}\omega\wedge\omega^{n-1}. Combining this formula with

dv(mω)=αvα(mω)dmωξα=αvα(mω)iξαω,andv(mω),mddcf:=αvα(mω)mddcfξα,dv(m_{\omega})=\sum_{\alpha}v_{\alpha}(m_{\omega})dm_{\omega}^{\xi_{\alpha}}=-\sum_{\alpha}v_{\alpha}(m_{\omega})i_{\xi_{\alpha}}\omega,\quad\text{and}\quad\langle v^{\prime}(m_{\omega}),m_{dd^{c}f}\rangle:=\sum_{\alpha}v_{\alpha}(m_{\omega})m_{dd^{c}f}^{\xi_{\alpha}},

one can derive that

v(mω),mddcf=ndv(mω)dcfωn1ωn=trω(dv(mω)dcf).\langle v^{\prime}(m_{\omega}),m_{dd^{c}f}\rangle=n\frac{dv(m_{\omega})\wedge d^{c}f\wedge\omega^{n-1}}{\omega^{n}}=\operatorname{tr}_{\omega}(dv(m_{\omega})\wedge d^{c}f).

Therefore, we have the following formula for the weighted Laplacian

Δω,vf=Δωf+trω(dlogv(mω)dcf).\Delta_{\omega,v}f=\Delta_{\omega}f+\operatorname{tr}_{\omega}(d\log v(m_{\omega})\wedge d^{c}f). (1.2)

The vv-weighted Monge–Ampère operator is defined as

MAv(ϕ):=MAω,v(ϕ):=v(mϕ)(ω+ddcϕ)n\operatorname{MA}_{v}(\phi):=\operatorname{MA}_{\omega,v}(\phi):=v(m_{\phi})(\omega+dd^{c}\phi)^{n}

for any ϕωT\phi\in\mathcal{H}^{T}_{\omega}. From [BJT24, (3.32)], one has the following integration-by-parts formula

Xg(Δωϕ,vf)MAv(ϕ)=Xf(Δωϕ,vg)MAv(ϕ)=Xdf,dgωϕMAv(ϕ),\int_{X}g(\Delta_{\omega_{\phi},v}f)\operatorname{MA}_{v}(\phi)=\int_{X}f(\Delta_{\omega_{\phi},v}g)\operatorname{MA}_{v}(\phi)=-\int_{X}\langle df,dg\rangle_{\omega_{\phi}}\operatorname{MA}_{v}(\phi), (1.3)

for all TT-invariant distributions f,gf,g such that at least one of which is smooth, where

df,dgω:=trω(dfdcg).\langle df,dg\rangle_{\omega}:=\operatorname{tr}_{\omega}(df\wedge d^{c}g).

1.2.3. Weighted cscK equations

Consider the weighted cscK problem

Sv(ωϕ)=w(mϕ).S_{v}(\omega_{\phi})=w(m_{\phi}).

The above equation can be rewritten as

{v(mϕ)(ω+ddcϕ)n=eFωn,supXϕ=0,Δϕ,vF=w(mϕ)+trϕ,v(Ric(ω)),\begin{cases}v(m_{\phi})(\omega+dd^{c}\phi)^{n}=e^{F}\omega^{n},\quad\sup_{X}\phi=0,\\ \Delta_{\phi,v}F=-w(m_{\phi})+\operatorname{tr}_{\phi,v}(\operatorname{Ric}(\omega)),\end{cases}

where trϕ,v:=trωϕ,v\operatorname{tr}_{\phi,v}:=\operatorname{tr}_{\omega_{\phi},v} and Δϕ,v:=Δωϕ,v\Delta_{\phi,v}:=\Delta_{\omega_{\phi},v}. For any volume form μX\mu_{X} on XX, taking F~:=F+logωnμX\widetilde{F}:=F+\log\frac{\omega^{n}}{\mu_{X}}, the original coupled equations are equivalent to the following coupled equations

{v(mϕ)(ω+ddcϕ)n=eF~μX,supXϕ=0,Δϕ,vF~=w(mϕ)+trϕ,v(Ric(μX)).\begin{cases}v(m_{\phi})(\omega+dd^{c}\phi)^{n}=e^{\widetilde{F}}\mu_{X},\quad\sup_{X}\phi=0,\\ \Delta_{\phi,v}\widetilde{F}=-w(m_{\phi})+\operatorname{tr}_{\phi,v}(\operatorname{Ric}(\mu_{X})).\end{cases}

Since PP is compact, there are positive constants CvC_{v} and CwC_{w} such that, on PP,

Cv1v+α|vα|+α,β|vαβ|Cv, and Cw1|w|+α|wα|+α,β|wαβ|Cw.C_{v}^{-1}\leq v+\sum_{\alpha}|v_{\alpha}|+\sum_{\alpha,\beta}|v_{\alpha\beta}|\leq C_{v},\text{ and }\,C_{w}^{-1}\leq|w|+\sum_{\alpha}|w_{\alpha}|+\sum_{\alpha,\beta}|w_{\alpha\beta}|\leq C_{w}.

This yields that

|trωϕ,v(η)trωϕ(η))|=|(logv)(mϕ),mη|Cv,η,|\operatorname{tr}_{\omega_{\phi},v}(\eta)-\operatorname{tr}_{\omega_{\phi}}(\eta))|=|\langle(\log v)^{\prime}(m_{\phi}),m_{\eta}\rangle|\leq C_{v,\eta},

where Cv,ηC_{v,\eta} depending only on CvC_{v} and η\eta. In particular, for any φ,ψT(X,ω)\varphi,\psi\in\mathcal{H}^{T}(X,\omega) and η:=ωψ\eta:=\omega_{\psi},

|trωφ,v(ωψ)trωφ(ωψ)|=|(logv)(mφ),mψ|Cv,|\operatorname{tr}_{\omega_{\varphi},v}(\omega_{\psi})-\operatorname{tr}_{\omega_{\varphi}}(\omega_{\psi})|=|\langle(\log v)^{\prime}(m_{\varphi}),m_{\psi}\rangle|\leq C_{v}, (1.4)

since mφ(X)=mω(X)=Pm_{\varphi}(X)=m_{\omega}(X)=P is compact.

1.2.4. Weighted Mabuchi functional

Let μX\mu_{X} be a fixed TT-invariant volume form on XX. The vv-weighted relative entropy is defined by

𝐇v(ϕ):=𝐇v,μX(ϕ):=Xlog(MAv(ϕ)μX)MAv(ϕ).\mathbf{H}_{v}(\phi):=\mathbf{H}_{v,\mu_{X}}(\phi):=\int_{X}\log\left(\frac{\operatorname{MA}_{v}(\phi)}{\mu_{X}}\right)\operatorname{MA}_{v}(\phi).

The ww-weighted Monge–Ampère energy 𝐄w:=𝐄ω,w:ωT\mathbf{E}_{w}:=\mathbf{E}_{\omega,w}:\mathcal{H}_{\omega}^{T}\rightarrow\mathbb{R} is given by

(d𝐄w)ϕ(f):=Xfw(mωϕ)ωϕn,𝐄w(0)=0(d\mathbf{E}_{w})_{\phi}(f):=\int_{X}fw(m_{\omega_{\phi}})\omega^{n}_{\phi},\quad\mathbf{E}_{w}(0)=0

for any f𝒞(X)Tf\in\mathcal{C}^{\infty}(X)^{T} (cf. [Lah19, Lem. 3]).

Fix θ\theta a TT-invariant closed (1,1)(1,1)-form and mθm_{\theta} is its moment map. Then the twisted weighted Monge–Ampère energy 𝐄vθ:=𝐄ω,vθ:ωT\mathbf{E}_{v}^{\theta}:=\mathbf{E}_{\omega,v}^{\theta}:\mathcal{H}^{T}_{\omega}\rightarrow\mathbb{R} is the primitive of (v(mϕ)nθωϕn+v(mϕ),mθωϕn)\left(v(m_{\phi})n\theta\wedge\omega_{\phi}^{n}+\langle v^{\prime}(m_{\phi}),m_{\theta}\rangle\omega_{\phi}^{n}\right) with 𝐄ω,vθ(0)=0\mathbf{E}_{\omega,v}^{\theta}(0)=0; in other words,

(d𝐄ω,vθ)ϕ(f)=Xf(v(mϕ)nθωϕn+v(mϕ),mθωϕn),𝐄ω,vθ(0)=0.(d\mathbf{E}_{\omega,v}^{\theta})_{\phi}(f)=\int_{X}f\left(v(m_{\phi})n\theta\wedge\omega_{\phi}^{n}+\langle v^{\prime}(m_{\phi}),m_{\theta}\rangle\omega_{\phi}^{n}\right),\quad\mathbf{E}_{\omega,v}^{\theta}(0)=0.

for any f𝒞(X)Tf\in\mathcal{C}^{\infty}(X)^{T} (see [Lah19, Lem. 4] for the well-definedness).

Definition 1.3.

The weighted Mabuchi energy 𝐌v,w\mathbf{M}_{v,w} is an Euler–Lagrange functional of ωTϕ𝐌v,w(ϕ):=(w(mϕ)Sv(ωϕ))MAv(ϕ)\mathcal{H}^{T}_{\omega}\ni\phi\mapsto\mathbf{M}^{\prime}_{v,w}(\phi):=(w(m_{\phi})-S_{v}(\omega_{\phi}))\operatorname{MA}_{v}(\phi).

From [Lah19, BJT24], we have the following lemma.

Lemma 1.4.

The following formula holds on ωT\mathcal{H}_{\omega}^{T}

𝐌v,w:=𝐌v,w,μX=𝐇v,μX+𝐄vRic(μX)+𝐄vw.\mathbf{M}_{v,w}:=\mathbf{M}_{v,w,\mu_{X}}=\mathbf{H}_{v,\mu_{X}}+\mathbf{E}_{v}^{-\operatorname{Ric}(\mu_{X})}+\mathbf{E}_{vw}.

The above lemma follows from the fact that the functional ωTϕ𝐇v+𝐄vRic(μX)\mathcal{H}^{T}_{\omega}\ni\phi\mapsto\mathbf{H}_{v}+\mathbf{E}_{v}^{-\operatorname{Ric}(\mu_{X})} is an Euler–Lagrange functional for ϕSv(ωϕ)MAv(ϕ)\phi\mapsto-S_{v}(\omega_{\phi})\operatorname{MA}_{v}(\phi) (cf. [BJT24, Lem. 3.29]).

Lemma 1.5.

Any (v,w)(v,w)-cscK metric in {ω}\{\omega\} is a minimizer of 𝐌v,w,μX\mathbf{M}_{v,w,\mu_{X}} for any choice of μX\mu_{X}.

Proof.

It follows from [Lah23, Cor. 1] that any (v,w)(v,w)-cscK metric in {ω}\{\omega\} is a minimizer of 𝐌v,w,ωn\mathbf{M}_{v,w,\omega^{n}}. On the other hand, by [BJT24, Lem. 3.48] there is a constant CC such that 𝐌v,w,μX=𝐌v,w,ωn+C\mathbf{M}_{v,w,\mu_{X}}=\mathbf{M}_{v,w,\omega^{n}}+C, hence any (v,w)(v,w)-cscK metric in {ω}\{\omega\} is also a minimizer of 𝐌v,w,μX\mathbf{M}_{v,w,\mu_{X}}. ∎

1.2.5. Weighted extremal metrics

We recall here the definition of weighted extremal metrics and relative weighted Mabuchi energy, which is a modification of Mabuchi energy such that it is TT_{\mathbb{C}}-invariant. We refer to [BJT24, Sec. 3.6] for more details.

Lemma 1.6 ([BJT24, Lem. 3.34]).

The weighted Mabuchi functional 𝐌v,w\mathbf{M}_{v,w} is translation invariant and TT_{\mathbb{C}}-invariant iff for all affine function =ξ+c𝔱\ell=\xi+c\in\mathfrak{t}\bigoplus\mathbb{R} on 𝔱\mathfrak{t}^{\vee} we have X(mω)𝐌v,w(0)=0\int_{X}\ell(m_{\omega})\mathbf{M}^{\prime}_{v,w}(0)=0, i.e.

X(mω)w(mω)MAv(0)=X(mω)Sv(ω)MAv(0).\int_{X}\ell(m_{\omega})w(m_{\omega})\operatorname{MA}_{v}(0)=\int_{X}\ell(m_{\omega})S_{v}(\omega)\operatorname{MA}_{v}(0).

Now, let w0𝒞(P,>0)w_{0}\in\mathcal{C}^{\infty}(P,\mathbb{R}_{>0}). The weighted Futaki–Mabuchi pairing on the space 𝔱\mathfrak{t}\bigoplus\mathbb{R} of affine functions on 𝔱\mathfrak{t}^{\vee} is defined by

,:=X(mω)(mω)w0(mω)MAv(0)=X(mω)(mω)MAvw0(0).\langle\ell,\ell^{\prime}\rangle:=\int_{X}\ell(m_{\omega})\ell^{\prime}(m_{\omega})w_{0}(m_{\omega})\operatorname{MA}_{v}(0)=\int_{X}\ell(m_{\omega})\ell^{\prime}(m_{\omega})\operatorname{MA}_{vw_{0}}(0).

Then this is positive definite, and there exists the unique affine function ext=ω,v,w0ext\ell^{\operatorname{ext}}=\ell^{\operatorname{ext}}_{\omega,v,w_{0}} on 𝔱\mathfrak{t}^{\vee} such that

,ext=X(mω)Sv(ω)MAv(0).\langle\ell,\ell^{\operatorname{ext}}\rangle=\int_{X}\ell(m_{\omega})S_{v}(\omega)\operatorname{MA}_{v}(0). (1.5)

Then the weighted Mabuchi energy 𝐌v,w\mathbf{M}_{v,w} with w=w0extw=w_{0}\ell^{\operatorname{ext}} is TT_{\mathbb{C}}-invariant. We define 𝐌v,w0re:=𝐌v,w0ext\mathbf{M}^{\rm re\ell}_{v,w_{0}}:=\mathbf{M}_{v,w_{0}\ell^{\operatorname{ext}}} is the relative weighted Mabuchi energy.

Definition 1.7.

Let v,w0𝒞(P,>0)v,w_{0}\in\mathcal{C}^{\infty}(P,\mathbb{R}_{>0}). A metric ω\omega is called a (v,w0)(v,w_{0})-extremal metric of it is (v,w)(v,w)-cscK with w=w0extw=w_{0}\ell^{\operatorname{ext}}. In this case, ww is called an extremal weight.

1.2.6. Extremal Kähler metrics

We explain here how the problem of finding extremal Kähler metrics is a special case of the one for weighted setting. This corresponds to the (1,1)(1,1)-extremal metric, i.e. v=1v=1 and w0=1w_{0}=1.

Let (X,ω)(X,\omega) be a compact Kähler manifold, and let gg be the Riemannian metric defined by ω\omega. The metric ω\omega is said to be extremal if the Hamiltonian vector field ξω=JS(ω)\xi_{\omega}=J\nabla\operatorname{S}(\omega) is a Killing vector field for gg, i.e ξωg=0\mathcal{L}_{\xi_{\omega}}g=0.

Let TT be a maximal compact torus of Autred(X)\operatorname{Aut}_{\operatorname{red}}(X) and α\alpha be a Kähler class. From [FM95] (see also [Lah19, Sec. 3.1]), the projection of S(ω)\operatorname{S}(\omega) with respect to L2(ωn)L^{2}(\omega^{n})-inner product, to the sub-space {mωξ+c:ξ𝔱,c}\{m_{\omega}^{\xi}+c:\xi\in\mathfrak{t},c\in\mathbb{R}\} is written as ΠωT(S(ω))=mωξext+cext\Pi_{\omega}^{T}(\operatorname{S}(\omega))=m_{\omega}^{\xi_{\operatorname{ext}}}+c_{\operatorname{ext}} for some ξext𝔱\xi_{\operatorname{ext}}\in\mathfrak{t} where ξext\xi_{\operatorname{ext}} only depends on TT and {ω}\{\omega\}. In particular, it implies that cextc_{\operatorname{ext}} also depend only on TT and {ω}\{\omega\}, since

nc1(X){ω}n1=XS(ω)ωn=XΠωT(S(ω))ωn=Xmωξextωn+cext{ω}n,nc_{1}(X)\cdot\{\omega\}^{n-1}=\int_{X}\operatorname{S}(\omega)\omega^{n}=\int_{X}\Pi^{T}_{\omega}(\operatorname{S}(\omega))\omega^{n}=\int_{X}m_{\omega}^{\xi_{\operatorname{ext}}}\omega^{n}+c_{\operatorname{ext}}\{\omega\}^{n},

where the last integral only depends {ω}\{\omega\} and ξext\xi_{\operatorname{ext}} by (1.1). One can also normalize Xmωξextωn=0\int_{X}m_{\omega}^{\xi_{\operatorname{ext}}}\omega^{n}=0 to get cext=s¯:=nc1(X){ω}n1{ω}nc_{\operatorname{ext}}=\overline{s}:=n\frac{c_{1}(X)\cdot\{\omega\}^{n-1}}{\{\omega\}^{n}}. Therefore, we obtain the unique affine function ext(p)=ξext,p+cext\ell^{\operatorname{ext}}(p)=\langle\xi_{\operatorname{ext}},p\rangle+c_{\operatorname{ext}} defined in (1.5) and the problem of finding extremal metric is equivalent to the one for (1,ext)(1,\ell^{\operatorname{ext}})-cscK metric.

1.2.7. Extension on 1,T\mathcal{E}^{1,T} and coercivity

Denote by ω1,T:=1(X,ω)T1(X,ω)\mathcal{E}^{1,T}_{\omega}:=\mathcal{E}^{1}(X,\omega)^{T}\subset\mathcal{E}^{1}(X,\omega) the space of TT-invariant finite energy potentials and norm1,T(X,ω):={uω1,T𝐄ω(u)=0}\mathcal{E}^{1,T}_{\rm norm}(X,\omega):=\{u\in\mathcal{E}^{1,T}_{\omega}\mid\mathbf{E}_{\omega}(u)=0\}. From [BJT24, Prop. 3.41], one can extend all functionals above on ω1,T\mathcal{E}^{1,T}_{\omega}.

Since Aut0(X)\operatorname{Aut}_{0}(X) acts trivially on cohomologies, for each σAut0(X)\sigma\in\operatorname{Aut}_{0}(X), one can find a unique function τσPSH(X,ω)𝒞(X)\tau_{\sigma}\in\operatorname{PSH}(X,\omega)\cap\mathcal{C}^{\infty}(X) such that

σω=ω+ddcτσ,𝐄(τσ)=0.\sigma^{\ast}\omega=\omega+dd^{c}\tau_{\sigma},\quad\mathbf{E}(\tau_{\sigma})=0.

For uPSH(X,ω)u\in\operatorname{PSH}(X,\omega) and σAut0(X)\sigma\in\operatorname{Aut}_{0}(X), define σu:=σu+τσ\sigma\cdot u:=\sigma^{\ast}u+\tau_{\sigma}. Set

d1,T(u,0):=infσTd1(σu,0).d_{1,T}(u,0):=\inf_{\sigma\in T_{\mathbb{C}}}d_{1}(\sigma\cdot u,0). (1.6)
Definition 1.8.

The weighted Mabuchi functional 𝐌v,w\mathbf{M}_{v,w} is TT_{\mathbb{C}}-coercive if there exist constants A>0A>0 and B>0B>0 such that

𝐌v,w(ϕ)Ad1,T(ϕ,0)B\mathbf{M}_{v,w}(\phi)\geq Ad_{1,T}(\phi,0)-B

for any ϕnorm1,T(X,ω)\phi\in\mathcal{E}^{1,T}_{\rm norm}(X,\omega).

By definition, if 𝐌v,w\mathbf{M}_{v,w} is TT_{\mathbb{C}}-coercive, then it is TT_{\mathbb{C}}-invariant since it is bounded from below (cf. [BJT24, Section 1.1]).

1.3. Weighted variational formalism in the singular setting

We recall here the weighted formalism for the variational problem of singular weighted cscK metrics on Kähler varieties with log terminal singularities as introduced in [BJT24, Sec. 3.8, 4.1].

1.3.1. Reduced automorphism group and moment maps

Let (X,ω)(X,\omega) be a normal compact Kähler variety. The automorphism group Aut(X)\operatorname{Aut}(X) is a complex Lie group, whose Lie algebra is H0(X,TX)H0(Xreg,TXreg)H^{0}(X,T_{X})\simeq H^{0}(X^{\mathrm{reg}},T_{X^{\mathrm{reg}}}) the space of holomorphic vector fields, that are global section of the tangent sheaf TX:=Hom(ΩX1,𝒪X)T_{X}:=\operatorname{Hom}(\Omega_{X}^{1},\mathcal{O}_{X}). There exists an Aut0(X)\operatorname{Aut}_{0}(X)-equivariant resolution of singularities π:X~X\pi:\tilde{X}\to X, i.e. π\pi is an isomorphism over XregX^{\mathrm{reg}} and any σAut0(X)\sigma\in\operatorname{Aut}_{0}(X) can be extended to a unique σAut0(X~)\sigma^{\prime}\in\operatorname{Aut}_{0}(\tilde{X}). Hence, we get the inclusion Aut0(X)Aut0(X~)\operatorname{Aut}_{0}(X)\subset\operatorname{Aut}_{0}(\tilde{X}). Moreover, any holomorphic vector field on X~\tilde{X} descends to an element in H0(Xreg,TXreg)H^{0}(X^{\mathrm{reg}},T_{X^{\mathrm{reg}}}), so we get Aut0(X)Aut0(X~)\operatorname{Aut}_{0}(X)\simeq\operatorname{Aut}_{0}(\tilde{X}).

Given an Aut0(X)\operatorname{Aut}_{0}(X)-equivariant resolution of singularities π:X~X\pi:{\widetilde{X}}\to X, the reduced automorphism group Autred(X)Aut0(X)\operatorname{Aut}_{\operatorname{red}}(X)\subset\operatorname{Aut}_{0}(X) is defined as the subgroup of Aut0(X)Aut0(X~)\operatorname{Aut}_{0}(X)\simeq\operatorname{Aut}_{0}({\widetilde{X}}) acting trivially on the Albanese torus of X~{\widetilde{X}}. This definition is independent of the choice of Aut0(X)\operatorname{Aut}_{0}(X)-equivariant resolution of singularities π:X~X\pi:\tilde{X}\rightarrow X, by the bimeromorphic invariance of Albanese torus (cf. [Uen75, Prop. 9.12]). Since Aut0(X)Aut0(X~)\operatorname{Aut}_{0}(X)\simeq\operatorname{Aut}_{0}(\tilde{X}), we get Autred(X)Autred(X~)\operatorname{Aut}_{\operatorname{red}}(X)\simeq\operatorname{Aut}_{\operatorname{red}}(\tilde{X}).

Denote by 𝒵\mathcal{Z} (reps. 𝒵\mathcal{Z}^{\prime}) the space of locally ddcdd^{c}-exact real (1,1)(1,1)-form (resp. currents) on XX. In particular, any current θ𝒵\theta\in\mathcal{Z}^{\prime} can be written as θ=ω+ddcu\theta=\omega+dd^{c}u for some ω𝒵\omega\in\mathcal{Z} and uu is a distribution [BG13, Sec. 4.6.1]. Then the group Aut(X)\operatorname{Aut}(X) acts on 𝒵\mathcal{Z} and 𝒵\mathcal{Z}^{\prime}. In particular, Aut0(X)\operatorname{Aut}_{0}(X) acts trivially on the the classes of 𝒵\mathcal{Z} (cf. [BJT24, Lem. 3.54]): for any σAut0(X)\sigma\in\operatorname{Aut}_{0}(X), there exist f𝒞(X)f\in\mathcal{C}^{\infty}(X) such that σω=ω+ddcf\sigma^{*}\omega=\omega+dd^{c}f.

Fix a compact torus TAutred(X)T\subset\operatorname{Aut}_{\operatorname{red}}(X) with Lie algebra 𝔱\mathfrak{t}. Any TT-invariant (1,1)(1,1)-form (resp. current) ω𝒵\omega\in\mathcal{Z} admits a moment map mω:X𝔱m_{\omega}:X\rightarrow\mathfrak{t}^{\vee} which is a TT-invariant smooth map (resp. a distribution), unique up to an additive constant, such that for each ξ𝔱\xi\in\mathfrak{t}, mωξ:=mω,ξ:Xm_{\omega}^{\xi}:=\langle m_{\omega},\xi\rangle:X\rightarrow\mathbb{R} satisfies dmωξ=iξω=ω(ξ,)-dm_{\omega}^{\xi}=i_{\xi}\omega=\omega(\xi,\cdot). Denote by P:=im(mω)P:=\operatorname{im}(m_{\omega}) which is a compact subset in 𝔱\mathfrak{t}^{\vee} and normalize such that P=im(mωϕ)P=\operatorname{im}(m_{\omega_{\phi}}) for all TT-invariant smooth ω\omega-psh function ϕ\phi. Take v,w𝒞(t,)v,w\in\mathcal{C}^{\infty}(t^{\vee},\mathbb{R}) with infPv>0\inf_{P}v>0.

1.3.2. Singular weighted cscK metrics

Suppose that XX is \mathbb{Q}-Gorenstein, meaning that XX is normal and KXK_{X} is mm-Cartier for some mm\in\mathbb{N}^{\ast}. Below, we recall the definitions of adapted measures and log terminal singularities from [EGZ09, Sec. 5-6].

Definition 1.9.

Let hmh^{m} be a smooth hermitian metric on mKXmK_{X}. Taking Ω\Omega a local generator of mKXmK_{X}, the adapted measure associated with hmh^{m} is defined by

μh:=in2(ΩΩ¯|Ω|hm2)1/m.\mu_{h}:=\mathrm{i}^{n^{2}}\left(\frac{\Omega\wedge\overline{\Omega}}{\left\lvert{\Omega}\right\rvert_{h^{m}}^{2}}\right)^{1/m}.

This definition does not depend on the choice of Ω\Omega, and two adapted measures differ by a smooth positive density. The Ricci form of the adapted measure μh\mu_{h} is given as

Ric(μh):=ddclog|Ω|hm2/m\operatorname{Ric}(\mu_{h}):=dd^{c}\log|\Omega|^{2/m}_{h^{m}}

which belongs to 𝒵\mathcal{Z}. The \mathbb{Q}-Gorenstein variety XX has log terminal singularities if the measure μh\mu_{h} has finite masses near XsingX^{\mathrm{sing}}.

We now further assume XX is log terminal. As in [EGZ09, Sec. 5], Ric(μh)\operatorname{Ric}(\mu_{h}) is canonically attached to an element in H0(X,𝒞X/PHX)H^{0}(X,\mathcal{C}^{\infty}_{X}/\operatorname{PH}_{X}) where 𝒞X\mathcal{C}^{\infty}_{X} (resp. PHX\operatorname{PH}_{X}) is the subsheaf of continuous functions on XX that are local restrictions of smooth functions (resp. pluriharmonic functions) under local embeddings. The first Chern class of Ric(μh)\operatorname{Ric}(\mu_{h}), denoted {Ric(μh)}\{\operatorname{Ric}(\mu_{h})\}, is the image of Ric(μh)\operatorname{Ric}(\mu_{h}) in H1(X,PHX)H^{1}(X,\operatorname{PH}_{X}) via the connecting homomorphism {}\{\bullet\} in the following exact sequence

H0(X,𝒞X)H0(X,𝒞X/PHX){}H1(X,PHX)0H^{0}(X,\mathcal{C}^{\infty}_{X})\to H^{0}(X,\mathcal{C}^{\infty}_{X}/\operatorname{PH}_{X})\xrightarrow[]{\{\bullet\}}H^{1}(X,\operatorname{PH}_{X})\to 0

which is induced by the short exact sequence 0PHX𝒞X𝒞X/PHX00\to\operatorname{PH}_{X}\to\mathcal{C}^{\infty}_{X}\to\mathcal{C}^{\infty}_{X}/\operatorname{PH}_{X}\to 0.

Assume that hmh^{m} is a TT-invariant metric on mKXmK_{X} so that Ric(μh)\operatorname{Ric}(\mu_{h}) is an equivariant curvature form. Fix a compact torus TAutred(X)T\subset\operatorname{Aut}_{\operatorname{red}}(X) with Lie algebra 𝔱\mathfrak{t}. Set P:=Pω:=mω(X)𝔱P:=P_{\omega}:=m_{\omega}(X)\subset\mathfrak{t}^{\vee} and take v𝒞(P,>0)v\in\mathcal{C}^{\infty}(P,\mathbb{R}_{>0}). For any ϕωT\phi\in\mathcal{H}^{T}_{\omega}, the weighted Ricci current of ωϕ:=ω+ddcϕ\omega_{\phi}:=\omega+dd^{c}\phi is defined as

RicvT(ωϕ)=RicT(MAv(ϕ)):=ddclog(v(mϕ)ωϕnμh)+Ric(μh).\operatorname{Ric}_{v}^{T}(\omega_{\phi})=\operatorname{Ric}^{T}(\operatorname{MA}_{v}(\phi)):=-dd^{c}\log\left(\frac{v(m_{\phi})\omega_{\phi}^{n}}{\mu_{h}}\right)+\operatorname{Ric}(\mu_{h}).

and the vv-weighted scalar curvature is the distribution expressed by

Sv(ωϕ):=trωϕ(RicT(ωϕ)).S_{v}({\omega_{\phi}}):=\operatorname{tr}_{\omega_{\phi}}(\operatorname{Ric}^{T}(\omega_{\phi})).

One can extend the weighted energy and the weighted Ricci energy on ω1,T\mathcal{E}^{1,T}_{\omega} (cf. [PTT23, Sec. 4.1.2], [BJT24, Sec. 3.8]). Fix a TT-invariant adapted measure μX\mu_{X}, for any w𝒞(P,)w\in\mathcal{C}^{\infty}(P,\mathbb{R}), the weighted Mabuchi energy 𝐌v,w:ω1,T{+}\mathbf{M}_{v,w}:\mathcal{E}^{1,T}_{\omega}\rightarrow\mathbb{R}\cup\{+\infty\} can be expressed as

𝐌v,w:=𝐌v,w,μX:=𝐇v+𝐄vRic(μX)+𝐄vw.\mathbf{M}_{v,w}:=\mathbf{M}_{v,w,\mu_{X}}:=\mathbf{H}_{v}+\mathbf{E}^{-\operatorname{Ric}(\mu_{X})}_{v}+\mathbf{E}_{vw}.
Definition 1.10.

Let v𝒞(P,>0)v\in\mathcal{C}^{\infty}(P,\mathbb{R}_{>0}) and w𝒞(P,)w\in\mathcal{C}^{\infty}(P,\mathbb{R}). Then ωϕ{ω}\omega_{\phi}\in\{\omega\} is a singular (v,w)(v,w)-cscK metric if ϕPSH(X,ω)L(X)\phi\in\operatorname{PSH}(X,\omega)\cap L^{\infty}(X) which is also smooth on XregX^{\mathrm{reg}}, and

Sv(ωϕ)=w(mϕ)on Xreg.S_{v}(\omega_{\phi})=w(m_{\phi})\quad\text{on }X^{\mathrm{reg}}.

Let w0𝒞(P,>0)w_{0}\in\mathcal{C}^{\infty}(P,\mathbb{R}_{>0}) and ext\ell^{\operatorname{ext}} be the unique affine function ext=ω,v,w0ext\ell^{\operatorname{ext}}=\ell^{\operatorname{ext}}_{\omega,v,w_{0}} on 𝔱\mathfrak{t}^{\vee} such that

,ext=X(mω)Sv(ω)MAv(0).\langle\ell,\ell^{\operatorname{ext}}\rangle=\int_{X}\ell(m_{\omega})S_{v}(\omega)\operatorname{MA}_{v}(0).

Then ωϕ\omega_{\phi} is called a singular (v,w0)(v,w_{0})-extremal metric if it is a singular (v,w)(v,w)-cscK metric with w=w0extw=w_{0}\ell^{\operatorname{ext}}.

Let π:YX\pi:Y\rightarrow X be an TT-equivariant resolution of singularities. We can also define 𝐌πωre:πωT\mathbf{M}^{\rm re\ell}_{\pi^{*}\omega}:\pi^{*}\mathcal{H}^{T}_{\omega}\rightarrow\mathbb{R} by

𝐌πωre(πu):=𝐇ωYn(πu)+𝐄πωRic(ωY)(πu)+𝐄πω,vwext(πu).\mathbf{M}^{\rm re\ell}_{\pi^{*}\omega}(\pi^{*}u):=\mathbf{H}_{\omega_{Y}^{n}}(\pi^{*}u)+\mathbf{E}_{\pi^{*}\omega}^{-\operatorname{Ric}(\omega_{Y})}(\pi^{*}u)+\mathbf{E}_{\pi^{\ast}\omega,vw\ell^{\operatorname{ext}}}(\pi^{\ast}u).

Under Condition (A), following [BJT24, Lem. 4.21], one can extend the weighted energy, the weighted Ricci energy, and 𝐌πωre\mathbf{M}^{\rm re\ell}_{\pi^{*}\omega} to πω1,T\mathcal{E}^{1,T}_{\pi^{*}\omega}.

2. A priori estimates

In this section, we shall establish a priori estimates for weighted cscK equations on compact Kähler manifolds when the reference metrics are degenerating. These estimates are crucial for obtaining the existence of singular weighted cscK metrics in the next section.

Once and for all, we fix XX to be an nn-dimensional compact Kähler manifold, TAutred(X)T\subset\operatorname{Aut}_{\operatorname{red}}(X) a compact torus with Lie algebra 𝔱\mathfrak{t}, and ω\omega a TT-invariant Kähler metric on XX with the moment polytope PP of ω\omega and total volume VV. Take v𝒞(𝔱,)v\in\mathcal{C}^{\infty}(\mathfrak{t}^{\vee},\mathbb{R}) such that v>0v>0 on PP.

2.1. LL^{\infty}-estimates

In the following, we shall establish a uniform LL^{\infty}-estimate following and generalizing the approach of Guo and Phong [GP24, Thm. 3] with certain modifications (cf. [PTT23, Thm. 5.4]).

Theorem 2.1.

Let μ\mu be a TT-invariant smooth volume form such that ωn=gμ\omega^{n}=g\mu with g1/nLK0\|g^{1/n}\|_{L^{\infty}}\leq K_{0} for some constant K0>0K_{0}>0. Suppose that (φ,F)T(X,ω)×𝒞(X)T(\varphi,F)\in\mathcal{H}^{T}(X,\omega)\times\mathcal{C}^{\infty}(X)^{T} is a solution to the coupled equations

{v(mφ)(ω+ddcφ)n=eFμ,supXφ=0,Δφ,vF=S+trφ,v(Ric(μ)),\begin{cases}v(m_{\varphi})(\omega+dd^{c}\varphi)^{n}=e^{F}\mu,\quad\sup_{X}\varphi=0,\\ \Delta_{\varphi,v}F=-S+\operatorname{tr}_{\varphi,v}(\operatorname{Ric}(\mu)),\end{cases}

for some S𝒞(X)TS\in\mathcal{C}^{\infty}(X)^{T}. In addition, assume that there are positive constants K1,K3,K4K_{1},K_{3},K_{4} such that

  1. (1)

    K1(ω+ddcρ1)Ric(μ)-K_{1}(\omega+dd^{c}\rho_{1})\leq{\rm Ric}(\mu), where ρ1\rho_{1} is a TT-invariant quasi-psh function, supXρ1=0\sup_{X}\rho_{1}=0 and XeK1ρ1𝑑μ<+\int_{X}e^{-K_{1}\rho_{1}}d\mu<+\infty;

  2. (2)

    𝐇μ(φ)=Xlog(ωφnμ)ωφnK3{\bf H}_{\mu}(\varphi)=\int_{X}\log\left(\frac{\omega_{\varphi}^{n}}{\mu}\right)\omega_{\varphi}^{n}\leqslant K_{3};

  3. (3)

    there exists α>0\alpha>0 such that Xeα(ϕsupXϕ)μK4\int_{X}e^{-\alpha(\phi-\sup_{X}\phi)}\mu\leqslant K_{4} for all ϕPSH(X,ω)\phi\in{\rm PSH}(X,\omega).

Then there is a uniform constant C>0C>0 depending only on n,maxXS,V,α,K0,K1,K3,K4,Cv,μ,ρ1n,\max_{X}S,V,\alpha,K_{0},K_{1},K_{3},K_{4},C_{v},\mu,\rho_{1} such that

φLC and FC.\|\varphi\|_{L^{\infty}}\leq C\,\text{ and }\,F\leq C.

Furthermore, if we have that

Ric(μ)K2(ω+ddcρ2),\operatorname{Ric}(\mu)\leq K_{2}(\omega+dd^{c}\rho_{2}),

for a TT-invariant ω\omega-psh function ρ2𝒞(XZ)\rho_{2}\in\mathcal{C}^{\infty}(X\setminus Z), with Z:={ρ2=}Z:=\{\rho_{2}=-\infty\} and supXρ2=0\sup_{X}\rho_{2}=0, there is a constant C2>0C_{2}>0 depending only on n,maxX|S|,V,α,K0,K1,K2,K3,K4,Cv,μ,ρ1n,\max_{X}|S|,V,\alpha,K_{0},K_{1},K_{2},K_{3},K_{4},C_{v},\mu,\rho_{1} such that

FK2ρ2C2.F\geq K_{2}\rho_{2}-C_{2}.
Proof.

Consider τk:>0\tau_{k}:\mathbb{R}\to\mathbb{R}_{>0} a sequence of positive smooth functions decreasing towards the function xx𝟙>0(x).x\mapsto x\cdot\mathds{1}_{\mathbb{R}_{>0}}(x). Let ϕk\phi_{k} be a solution to the following auxiliary complex Monge–Ampère equation

V1(ω+ddcϕk)n=τk(φ+λF)+1AkeFμ,supXϕk=0,V^{-1}(\omega+dd^{c}\phi_{k})^{n}=\frac{\tau_{k}(-\varphi+\lambda F)+1}{A_{k}}e^{F}\mu,\quad\sup_{X}\phi_{k}=0, (2.1)

where

Ak=X(τk(φ+λF)+1)eFμk+X(φ+λF)+eFμ+eFμ=A.A_{k}=\int_{X}(\tau_{k}(-\varphi+\lambda F)+1)e^{F}\mu\xrightarrow[k\to+\infty]{}\int_{X}(-\varphi+\lambda F)_{+}e^{F}\mu+\int e^{F}\mu=A_{\infty}.

We remark that ϕk\phi_{k} is also TT-invariant since the RHS of (2.1) is TT-invariant. Applying Young’s inequality with χ(s)=(s+1)log(s+1)s\chi(s)=(s+1)\log(s+1)-s and χ(s)=ess1\chi^{\ast}(s)=e^{s}-s-1,

X(φ)eFωnXχ(α1eF)ωn+Xχ(αφ)ωn\int_{X}(-\varphi)e^{F}\omega^{n}\leq\int_{X}\chi(\alpha^{-1}e^{F})\omega^{n}+\int_{X}\chi^{*}(-\alpha\varphi)\omega^{n}

where α>0\alpha>0 is the constant in (3). It follows from (2) and (3) that Cv1VAC(K3,K4,V)C_{v}^{-1}V\leq A_{\infty}\leq C(K_{3},K_{4},V). Thus Cv1VAkC1=C(K3,K4,V)C_{v}^{-1}V\leq A_{k}\leq C_{1}=C(K_{3},K_{4},V) for kk sufficiently large.

By strong openness [Ber13, GZ15], we have a constant δ>0\delta>0 such that I:=Xe(K1+δ)ρ1<+I:=\int_{X}e^{-(K_{1}+\delta)\rho_{1}}<+\infty. Using Demailly’s approximation theorem [Dem92] (see also [Dem12, Sec. 14B]) and Demailly–Kollár’s convergence result [DK01, Main Thm. 0.2 (2)], there exists a sequence quasi-plurisubharmonic functions (ρ1,k)k(\rho_{1,k})_{k} with analytic singularities and smooth away from their singular locus, such that K1(ω+ddcρ1,k)1kωRic(μ)-K_{1}(\omega+dd^{c}\rho_{1,k})-\frac{1}{k}\omega\leq\operatorname{Ric}(\mu), ρ1,kρ1\rho_{1,k}\to\rho_{1} in L1L^{1} as k+k\to+\infty, and Xe(K1+δ)ρ1,k2I\int_{X}e^{-(K_{1}+\delta)\rho_{1,k}}\leq 2I for all k1k\gg 1. Hence, K1(ω+ddcρ1,k)Ric(μ)-K_{1}^{\prime}(\omega+dd^{c}\rho^{\prime}_{1,k})\leq\operatorname{Ric}(\mu) where K1=K1+1kK_{1}^{\prime}=K_{1}+\frac{1}{k} with k>1/δk>1/\delta and ρ1,k=K1K1+1/kρ1,k\rho^{\prime}_{1,k}=\frac{K_{1}}{K_{1}+1/k}\rho_{1,k}. Replacing ρ1\rho_{1} by ρ1,k\rho^{\prime}_{1,k} and K1K_{1} by K1K_{1}^{\prime}, one can assume ρ1\rho_{1} has analytics singularities.

Consider the function

Φ=ϵ(ϕk+Λ)nn+1φ+λ(F+K1ρ1)\Phi=-\epsilon(-\phi_{k}+\Lambda)^{\frac{n}{n+1}}-\varphi+\lambda(F+K_{1}\rho_{1})

with Λ=(2nn+1ϵ)n+1\Lambda=\left(\frac{2n}{n+1}\epsilon\right)^{n+1} and ϵ=((n+1)(n+λs¯+L)n2)nn+1Ak1n+1\epsilon=\left(\frac{(n+1)(n+\lambda\bar{s}+L)}{n^{2}}\right)^{\frac{n}{n+1}}A_{k}^{\frac{1}{n+1}}, where s¯:=maxXS\bar{s}:=\max_{X}S, L:=λ(Cv,μ+K1Cv)+2CvL:=\lambda(C_{v,\mu}+K_{1}C_{v})+2C_{v} with Cv,μ>0C_{v,\mu}>0 a constant independent of φ\varphi so that (logv)(mφ),mRic(μ)Cv,μ\langle(\log v)^{\prime}(m_{\varphi}),m_{\operatorname{Ric}(\mu)}\rangle\geq-C_{v,\mu}, and λ>0\lambda>0 is a constant such that n+λs¯>0n+\lambda\bar{s}>0 and λK1<1/2\lambda K_{1}<1/2. Since ρ1(x)\rho_{1}(x)\to-\infty as xZ1:={ρ1=}x\to Z_{1}:=\{\rho_{1}=-\infty\}, the maximal points of Φ\Phi only occur in XZ1X\setminus Z_{1}. Fix x0XZ1x_{0}\in X\setminus Z_{1} a maximal point of Φ\Phi. At x0x_{0}, we have

0\displaystyle 0 Δφ,vΦ=ΔωφΦ+(logv)(mφ),mddcΦ\displaystyle\geq\Delta_{\varphi,v}\Phi=\Delta_{\omega_{\varphi}}\Phi+\langle(\log v)^{\prime}(m_{\varphi}),m_{dd^{c}\Phi}\rangle
ϵnn+1(ϕk+Λ)1n+1ΔωφϕkΔωφφ+λΔφ,v(F+K1ρ1)\displaystyle\geq\frac{\epsilon n}{n+1}(-\phi_{k}+\Lambda)^{-\frac{1}{n+1}}\Delta_{\omega_{\varphi}}\phi_{k}-\Delta_{\omega_{\varphi}}\varphi+\lambda\Delta_{\varphi,v}(F+K_{1}\rho_{1})
+ϵnn+1(ϕk+Λ)1n+1(logv)(mφ),mddcϕk(logv)(mφ),mddcφ.\displaystyle\quad+\frac{\epsilon n}{n+1}(-\phi_{k}+\Lambda)^{\frac{-1}{n+1}}\langle(\log v)^{\prime}(m_{\varphi}),m_{dd^{c}\phi_{k}}\rangle-\langle(\log v)^{\prime}(m_{\varphi}),m_{dd^{c}\varphi}\rangle.

Since Cv(logv)(mϕ),mddcψCv-C_{v}\leq\langle(\log v)^{\prime}(m_{\phi}),m_{dd^{c}\psi}\rangle\leq C_{v} for any ϕ,ψPSH(X,ω)T\phi,\psi\in\operatorname{PSH}(X,\omega)^{T} (see (1.4)), we infer that

0\displaystyle 0 ϵnn+1(ϕk+Λ)1n+1ΔωφϕkΔωφφ+λΔφ,v(F+K1ρ1)Cv(ϵnn+1Λ1n+1+1)\displaystyle\geq\frac{\epsilon n}{n+1}(-\phi_{k}+\Lambda)^{-\frac{1}{n+1}}\Delta_{\omega_{\varphi}}\phi_{k}-\Delta_{\omega_{\varphi}}\varphi+\lambda\Delta_{\varphi,v}(F+K_{1}\rho_{1})-C_{v}\left(\frac{\epsilon n}{n+1}\Lambda^{\frac{-1}{n+1}}+1\right)
=ϵnn+1(ϕk+Λ)1n+1(trωφωϕktrωφω)trωφ(ωφω)+λ(S+trωφ(Ric(μ)+K1ddcρ1))\displaystyle=\frac{\epsilon n}{n+1}(-\phi_{k}+\Lambda)^{-\frac{1}{n+1}}(\operatorname{tr}_{\omega_{\varphi}}\omega_{\phi_{k}}-\operatorname{tr}_{\omega_{\varphi}}\omega)-\operatorname{tr}_{\omega_{\varphi}}(\omega_{\varphi}-\omega)+\lambda(-S+\operatorname{tr}_{\omega_{\varphi}}({\rm Ric}(\mu)+K_{1}dd^{c}\rho_{1}))
+λ((logv)(mφ),mRic(μ)+K1(logv)(mφ),mddcρ1)2Cv.\displaystyle\quad+\lambda\left(\langle(\log v)^{\prime}(m_{\varphi}),m_{\operatorname{Ric}(\mu)}\rangle+K_{1}\langle(\log v)^{\prime}(m_{\varphi}),m_{dd^{c}\rho_{1}}\rangle\right)-2C_{v}.

Then

0\displaystyle 0 n2ϵn+1(ϕk+Λ)1n+1(τk(φ+λF)+1Ak)1/nnλs¯L+(1nϵn+1Λ1n+1λK1)trωφω\displaystyle\geq\frac{n^{2}\epsilon}{n+1}(-\phi_{k}+\Lambda)^{-\frac{1}{n+1}}\left(\frac{\tau_{k}(-\varphi+\lambda F)+1}{A_{k}}\right)^{1/n}-n-\lambda\bar{s}-L+\left(1-\frac{n\epsilon}{n+1}\Lambda^{-\frac{1}{n+1}}-\lambda K_{1}\ \right)\operatorname{tr}_{\omega_{\varphi}}\omega
n2ϵn+1(ϕk+Λ)1n+1(τk(φ+λF)+1Ak)1/nnλs¯L.\displaystyle\geq\frac{n^{2}\epsilon}{n+1}(-\phi_{k}+\Lambda)^{-\frac{1}{n+1}}\left(\frac{\tau_{k}(-\varphi+\lambda F)+1}{A_{k}}\right)^{1/n}-n-\lambda\bar{s}-L.

Since n+λs¯>0n+\lambda\bar{s}>0 and λK1<1/2\lambda K_{1}<1/2, at x0x_{0}, we obtain

φ+λ(F+K1ρ1)φ+λF((n+λs¯+L)(n+1)n2ϵ)nAk(ϕk+Λ)nn+1;-\varphi+\lambda(F+K_{1}\rho_{1})\leq-\varphi+\lambda F\leq\left(\frac{(n+\lambda\bar{s}+L)(n+1)}{n^{2}\epsilon}\right)^{n}A_{k}(-\phi_{k}+\Lambda)^{\frac{n}{n+1}}; (2.2)

therefore, Φ(x0)0\Phi(x_{0})\leq 0 and Φ0\Phi\leq 0 on XX. By the choice of ϵ,Λ\epsilon,\Lambda and VAkC(K3,K4,V)V\leq A_{k}\leq C(K_{3},K_{4},V), and Young’s inequality, we derive that for any δ>0\delta>0

λ(F+K1ρ1)φ+λ(F+K1ρ1)C(V,K1,K3,K4)(ϕk+Λ)nn+1δϕk+C2,\lambda(F+K_{1}\rho_{1})\leq-\varphi+\lambda(F+K_{1}\rho_{1})\leq C(V,K_{1},K_{3},K_{4})(-\phi_{k}+\Lambda)^{\frac{n}{n+1}}\leq-\delta\phi_{k}+C_{2}, (2.3)

with C2=C(δ,V,K2,K3,K4)C_{2}=C(\delta,V,K_{2},K_{3},K_{4}).

From Condition (A), we have YeK1ρ1μ<+\int_{Y}e^{-K_{1}\rho_{1}}\mu<+\infty. The strong openness [Ber13, GZ15] yields a constant 0<a10<a\ll 1 such that Xe(1+a)K1ρ1μCa\int_{X}e^{-(1+a)K_{1}\rho_{1}}\mu\leq C_{a} for some constant Ca>0C_{a}>0. By (2.3), Hölder inequality and (3), for β=1+a/2λ\beta=\frac{1+a/2}{\lambda} and δ>0\delta>0 such that δβ<α/γ\delta\beta<\alpha/\gamma^{*}, with γ=1+a1+a/2\gamma=\frac{1+a}{1+a/2} and 1γ+1γ=1\frac{1}{\gamma}+\frac{1}{\gamma^{\ast}}=1, we obtain

XeβλFμeβC2Xeαγϕk(1+a2)K1ρ1μ=eβC2Xeαγϕk(1+a)K1ρ1/γμC(a,Ca,α).\int_{X}e^{\beta\lambda F}\mu\leq e^{\beta C_{2}}\int_{X}e^{-\frac{\alpha}{\gamma^{*}}\phi_{k}-(1+\frac{a}{2})K_{1}\rho_{1}}\mu=e^{\beta C_{2}}\int_{X}e^{-\frac{\alpha}{\gamma^{*}}\phi_{k}-(1+a)K_{1}\rho_{1}/\gamma}\mu\leq C(a,C_{a},\alpha). (2.4)

By a refined version of Kołodziej’s LL^{\infty}-estimate [Koł98] (see [DGG23, Thm. A] for the version we referred), a uniform control Cv1v(mφ)CvC_{v}^{-1}\leq v(m_{\varphi})\leq C_{v} and (2.4), we obtain φLC(n,V,α,K1,K3,K4,Cv,μ,ρ1)\|\varphi\|_{L^{\infty}}\leq C(n,V,\alpha,K_{1},K_{3},K_{4},C_{v},\mu,\rho_{1}). Also, combining the LL^{\infty}-estimate of φ\varphi with (2.4), we infer that

(τk(φ+λF)+1)eFLp(X,ωn)C(n,α,V,K1,K3,K4,Cv,μ,ρ1)\|(\tau_{k}(-\varphi+\lambda F)+1)e^{F}\|_{L^{p^{\prime}}(X,\omega^{n})}\leq C(n,\alpha,V,K_{1},K_{3},K_{4},C_{v},\mu,\rho_{1})

for some p>1p^{\prime}>1 and for all k>0k>0 sufficiently large. Again, Kołodziej’s LL^{\infty}-estimate yields a uniform bound ϕkLC(n,V,α,K1,K3,K4,Cv,ρ1)\|\phi_{k}\|_{L^{\infty}}\leq C(n,V,\alpha,K_{1},K_{3},K_{4},C_{v},\rho_{1}). Then the inequality (2.2) provides a uniform upper bound for FF.

In the second part, we consider the function H:=F+(K2+1)φK2ρ2H:=F+(K_{2}+1)\varphi-K_{2}\rho_{2}. Since ρ2=\rho_{2}=-\infty along ZZ, one can assume that HH admits a minimum at x0XZx_{0}\in X\setminus Z. At x0x_{0}, we have

0Δφ,vH\displaystyle 0\leq\Delta_{\varphi,v}H =S+trωφ(Ric(μ))+(K2+1)n(K2+1)trωφωK2trωφ(ω+ddcρ2)+K2trωφω\displaystyle=-S+\operatorname{tr}_{\omega_{\varphi}}(\operatorname{Ric}(\mu))+(K_{2}+1)n-(K_{2}+1)\operatorname{tr}_{\omega_{\varphi}}\omega-K_{2}\operatorname{tr}_{\omega_{\varphi}}(\omega+dd^{c}\rho_{2})+K_{2}\operatorname{tr}_{\omega_{\varphi}}\omega
+(logv)(mφ),mRic(μ)+(K2+1)(logv)(mφ),mddcϕK2(logv)(mφ),mddcρ2\displaystyle\quad+\langle(\log v)^{\prime}(m_{\varphi}),m_{\operatorname{Ric}(\mu)}\rangle+(K_{2}+1)\langle(\log v)^{\prime}(m_{\varphi}),m_{dd^{c}\phi}\rangle-K_{2}\langle(\log v)^{\prime}(m_{\varphi}),m_{dd^{c}\rho_{2}}\rangle
S+trωφ(Ric(μ)K2(ω+ddcρ2))+(K2+1)ntrωφω+Cv,μ+2(K2+1)Cv\displaystyle\leq-S+\operatorname{tr}_{\omega_{\varphi}}(\operatorname{Ric}(\mu)-K_{2}(\omega+dd^{c}\rho_{2}))+(K_{2}+1)n-\operatorname{tr}_{\omega_{\varphi}}\omega+C_{v,\mu}+2(K_{2}+1)C_{v}
(K2+1)nStrωφω+Cv,μ+2(K2+1)Cv\displaystyle\leq(K_{2}+1)n-S-\operatorname{tr}_{\omega_{\varphi}}\omega+C_{v,\mu}+2(K_{2}+1)C_{v}
(K2+1)nSn(ωnωφn)1/n+Cv,μ+2(K2+1)Cv\displaystyle\leq(K_{2}+1)n-S-n\left(\frac{\omega^{n}}{\omega^{n}_{\varphi}}\right)^{1/n}+C_{v,\mu}+2(K_{2}+1)C_{v}
(K2+1)nSn(gv(mφ)eF)1/n+Cv,μ+2(K2+1)Cv\displaystyle\leq(K_{2}+1)n-S-n\left(\frac{gv(m_{\varphi})}{e^{F}}\right)^{1/n}+C_{v,\mu}+2(K_{2}+1)C_{v}
(K2+1)nminXSnK0Cv1/neF/n+Cv,μ+2(K2+1)Cv.\displaystyle\leq(K_{2}+1)n-\min_{X}S-nK_{0}C_{v}^{-1/n}e^{-F/n}+C_{v,\mu}+2(K_{2}+1)C_{v}.

Therefore, at x0x_{0}, F(x0)C(n,K0,K2,minXS,Cv,μ,Cv)F(x_{0})\geq-C(n,K_{0},K_{2},\min_{X}S,C_{v,\mu},C_{v}). We obtain

F\displaystyle F K2(ρ2ρ2(x0))C(n,K0,K2,minXS,Cv,μ,Cv)(K2+1)φL\displaystyle\geq K_{2}(\rho_{2}-\rho_{2}(x_{0}))-C(n,K_{0},K_{2},\min_{X}S,C_{v,\mu},C_{v})-(K_{2}+1)\|\varphi\|_{L^{\infty}}
K2ρ2C(n,K0,,K4,maxX|S|,Cv,μ,ρ1)\displaystyle\geq K_{2}\rho_{2}-C(n,K_{0},\ldots,K_{4},\max_{X}|S|,C_{v},\mu,\rho_{1})

as required. ∎

2.2. Local LpL^{p}-estimate for Laplacian

Let ωX\omega_{X} be another TT-invariant Kähler metric on XX. Since PωX=im(mωX)P_{\omega_{X}}=\operatorname{im}(m_{\omega_{X}}) is compact, one can further assume that

Cv1|v|+α|vα|+α,β|vαβ|Cvon PωX.C_{v}^{-1}\leq|v|+\sum_{\alpha}|v_{\alpha}|+\sum_{\alpha,\beta}|v_{\alpha\beta}|\leq C_{v}\quad\text{on }P_{\omega_{X}}.

Consider the following weighted cscK equations

{v(mφ)(ω+ddcφ)n=eFωXn,supXφ=0,Δφ,vF=S+trφ,v(Ric(ωX)).\begin{cases}v(m_{\varphi})(\omega+dd^{c}\varphi)^{n}=e^{F}\omega_{X}^{n},\quad\sup_{X}\varphi=0,\\ \Delta_{\varphi,v}F=-S+\operatorname{tr}_{\varphi,v}(\operatorname{Ric}(\omega_{X})).\end{cases} (2.5)

In this section, we shall further assume 𝔱ζlogv(ζ)\mathfrak{t}^{\vee}\ni\zeta\mapsto\log v(\zeta)\in\mathbb{R} to be concave. With the concavity condition on logv\log v, from Lemma A.1, we have the following weighted Aubin–Yau type inequality:

Δφ,vlogtrωXωφΔωXFtrωXωφtrωφωX𝒞\Delta_{\varphi,v}\log\operatorname{tr}_{\omega_{X}}\omega_{\varphi}\geq\frac{\Delta_{\omega_{X}}F}{\operatorname{tr}_{\omega_{X}}\omega_{\varphi}}-\mathcal{B}\operatorname{tr}_{\omega_{\varphi}}\omega_{X}-\mathcal{C}

for some uniform constant ,𝒞>0\mathcal{B},\mathcal{C}>0.

Proposition 2.2.

Suppose that (φ,F)T(X,ω)×𝒞(X)T(\varphi,F)\in\mathcal{H}^{T}(X,\omega)\times\mathcal{C}^{\infty}(X)^{T} is a solution to (2.5). Fix p>1p>1. Assume that ωCωωX\omega\leq C_{\omega}\omega_{X}, Ric(ωX)AωX\operatorname{Ric}(\omega_{X})\leq A\omega_{X}, Bisec(ωX)B\operatorname{Bisec}(\omega_{X})\geq-B, ω+ddcρCρωX\omega+dd^{c}\rho\geq C_{\rho}\omega_{X} with Cρ>0C_{\rho}>0, and

φLC0,FC0,andFK2ρC0,\|\varphi\|_{L^{\infty}}\leq C_{0},\quad F\leq C_{0},\quad\text{and}\quad F\geq K_{2}\rho-C_{0},

where ρ𝒞(X{ρ=})\rho\in\mathcal{C}^{\infty}(X\setminus\{\rho=-\infty\}) is a TT-invariant ω\omega-psh function. Then for any 𝒦\mathcal{K} compact set of X{ρ=}X\setminus\{\rho=-\infty\}, one has the following estimate

trωXωφL2p+2(𝒦,ωXn)C1,\|\operatorname{tr}_{\omega_{X}}\omega_{\varphi}\|_{L^{2p+2}(\mathcal{K},\omega_{X}^{n})}\leq C_{1},

where C1C_{1} only depends on 𝒦,n,p,A,B,C0,Cρ,minXS\mathcal{K},n,p,A,B,C_{0},C_{\rho},\min_{X}S.

Proof.

We shall adapt the approach in the smooth setting of Chen and Cheng [CC21a] for cscK metrics and also its generalization by [DJL24] and [HL25] for weighted cscK metrics. We highlight some differences in the following:

  • We shall work with the trace taken with respect to the reference metric ωX\omega_{X} instead of ω\omega as ω\omega is moving when we are going to apply the result. However, the metric ω\omega still plays a role and must be carefully merged during the computations, especially since we only have the upper bound ωCωωX\omega\leq C_{\omega}\omega_{X}.

  • In the last step, we need to use bounds on φ\varphi and FF. Special attention is required for FF since its lower bound is uniform only up to a term involving K2ρK_{2}\rho, which is not bounded from below.

Take u:=ea(F+bφbρ)trωXωφ0u:=e^{-a(F+b\varphi-b\rho)}\operatorname{tr}_{\omega_{X}}\omega_{\varphi}\geq 0, where a,ba,b are constants to be determined later. Recall that from (1.2),

Δω,vf=Δωf+trω(dlogv(mω)dcf);\Delta_{\omega,v}f=\Delta_{\omega}f+\operatorname{tr}_{\omega}(d\log v(m_{\omega})\wedge d^{c}f);

hence, we have

Δφ,vu=Δφ,velogu=|du|ωφ2u+uΔωφlogu+trωφ(dlogv(mωφ)dcu)=|du|ωφ2u+uΔωφlogu+utrωφ(dlogv(mωφ)dclogu)=|du|ωφ2u+uΔφ,vloguuΔφ,vlogu=auΔφ,v(F+bφcρ)+uΔφ,vlog(trωXωφ).\begin{split}\Delta_{\varphi,v}u&=\Delta_{\varphi,v}e^{\log u}=\frac{|du|^{2}_{\omega_{\varphi}}}{u}+u\Delta_{\omega_{\varphi}}\log u+\operatorname{tr}_{\omega_{\varphi}}(d\log v(m_{\omega_{\varphi}})\wedge d^{c}u)\\ &=\frac{|du|^{2}_{\omega_{\varphi}}}{u}+u\Delta_{\omega_{\varphi}}\log u+u\operatorname{tr}_{\omega_{\varphi}}(d\log v(m_{\omega_{\varphi}})\wedge d^{c}\log u)=\frac{|du|^{2}_{\omega_{\varphi}}}{u}+u\Delta_{\varphi,v}\log u\\ &\geq u\Delta_{\varphi,v}\log u=-au\Delta_{\varphi,v}(F+b\varphi-c\rho)+u\Delta_{\varphi,v}\log(\operatorname{tr}_{\omega_{X}}\omega_{\varphi}).\end{split} (2.6)

Using Ric(ωX)AωX\operatorname{Ric}(\omega_{X})\leq A\omega_{X}, |(logv)(mφ),mRic(ωX)|Cv,Ric(ωX)|\langle(\log v)^{\prime}(m_{\varphi}),m_{\operatorname{Ric}(\omega_{X})}\rangle|\leq C_{v,\operatorname{Ric}(\omega_{X})}, trφ,vωφn+Cv\operatorname{tr}_{\varphi,v}\omega_{\varphi}\leq n+C_{v}, and trωφωρCρtrφ,vωXCv\operatorname{tr}_{\omega_{\varphi}}{\omega_{\rho}}\geq C_{\rho}\operatorname{tr}_{\varphi,v}\omega_{X}-C_{v}, we derive

Δφ,v(F+bφbρ)=S+trφ,v(Ric(ωX))+btrφ,v(ddcφddcρ)minXS+AtrωφωX+(logv)(mφ),mRic(ωX)btrφ,vωρ+btrφ,vωφminXS+(AbCρ)trωφωX+b(2Cv+n)+Cv,Ric(ωX)=C1+(AbCρ)trωφωX\begin{split}\Delta_{\varphi,v}(F+b\varphi-b\rho)&=-S+\operatorname{tr}_{\varphi,v}(\operatorname{Ric}(\omega_{X}))+b\operatorname{tr}_{\varphi,v}(dd^{c}\varphi-dd^{c}\rho)\\ &\leq-\min_{X}S+A\operatorname{tr}_{\omega_{\varphi}}\omega_{X}+\langle(\log v)^{\prime}(m_{\varphi}),m_{\operatorname{Ric}(\omega_{X})}\rangle-b\operatorname{tr}_{\varphi,v}\omega_{\rho}+b\operatorname{tr}_{\varphi,v}\omega_{\varphi}\\ &\leq-\min_{X}S+(A-bC_{\rho})\operatorname{tr}_{\omega_{\varphi}}\omega_{X}+b(2C_{v}+n)+C_{v,\operatorname{Ric}(\omega_{X})}\\ &=C_{1}+(A-bC_{\rho})\operatorname{tr}_{\omega_{\varphi}}\omega_{X}\end{split} (2.7)

where C1>0C_{1}>0 only depends on minXS,b,Cv,Cv,Ric(ωX)-\min_{X}S,b,C_{v},C_{v,\operatorname{Ric}(\omega_{X})}.

By a weighted Aubin–Yau’s inequality for weighted Monge–Ampère equation obtained in [DJL24] (see also Lemma A.1 for the version we apply), we have

Δφ,vlogtrωXωφ1trωXωφΔωXFtrωφωX𝒞\Delta_{\varphi,v}\log\operatorname{tr}_{\omega_{X}}\omega_{\varphi}\geq\frac{1}{\operatorname{tr}_{\omega_{X}}\omega_{\varphi}}\Delta_{\omega_{X}}F-\mathcal{B}\operatorname{tr}_{\omega_{\varphi}}\omega_{X}-\mathcal{C} (2.8)

where ,𝒞\mathcal{B},\mathcal{C} depend on ωX,Cv\omega_{X},C_{v}. Combining (2.6), (2.7), and (2.8), we get

Δφ,vuea(F+bφbρ){(aC1+𝒞)trωXωφ+ΔωXF+(abCρAa)trωXωφtrωφωX}.\Delta_{\varphi,v}u\geq e^{-a(F+b\varphi-b\rho)}\left\{-(aC_{1}+\mathcal{C})\operatorname{tr}_{\omega_{X}}\omega_{\varphi}+\Delta_{\omega_{X}}F+(abC_{\rho}-Aa-\mathcal{B})\operatorname{tr}_{\omega_{X}}\omega_{\varphi}\operatorname{tr}_{\omega_{\varphi}}\omega_{X}\right\}.

Using the fact that

trωXωφtrωφωX(trωXωφ)nn1(ωXnωφn)1n1=(trωXωφ)nn1(veF)1n1(trωXωφ)nn1(CveF)1n1,\operatorname{tr}_{\omega_{X}}\omega_{\varphi}\operatorname{tr}_{\omega_{\varphi}}\omega_{X}\geq(\operatorname{tr}_{\omega_{X}}\omega_{\varphi})^{\frac{n}{n-1}}\left(\frac{\omega_{X}^{n}}{\omega_{\varphi}^{n}}\right)^{\frac{1}{n-1}}=(\operatorname{tr}_{\omega_{X}}\omega_{\varphi})^{\frac{n}{n-1}}\left(ve^{-F}\right)^{\frac{1}{n-1}}\geq(\operatorname{tr}_{\omega_{X}}\omega_{\varphi})^{\frac{n}{n-1}}\left(C_{v}e^{F}\right)^{\frac{-1}{n-1}},

and choosing bb large enough such that abCρAaabCρ/2abC_{\rho}-Aa-\mathcal{B}\geq abC_{\rho}/2, we get

Δφ,vu(aC1+𝒞)u+abCρ2C1n1eFn1(trωXωφ)1n1u+ea(F+bφbρ)ΔωXF.\displaystyle\Delta_{\varphi,v}u\geq-(aC_{1}+\mathcal{C})u+\frac{abC_{\rho}}{2}C^{-\frac{1}{n-1}}e^{\frac{-F}{n-1}}(\operatorname{tr}_{\omega_{X}}\omega_{\varphi})^{\frac{1}{n-1}}u+e^{-a(F+b\varphi-b\rho)}\Delta_{\omega_{X}}F. (2.9)

Since |u|ωφ2trωXωφ|u|ωX2|\nabla u|_{\omega_{\varphi}}^{2}\operatorname{tr}_{\omega_{X}}\omega_{\varphi}\geq|\nabla u|^{2}_{\omega_{X}}, we infer that

12p+1Δφ,vu2p+1=u2pΔφ,vu+2pu2p1|u|ωφ2u2pΔφ,vu+2pu2p2ea(F+bφbρ)|u|ωX2\frac{1}{2p+1}\Delta_{\varphi,v}u^{2p+1}=u^{2p}\Delta_{\varphi,v}u+2pu^{2p-1}|\nabla u|^{2}_{\omega_{\varphi}}\geq u^{2p}\Delta_{\varphi,v}u+2pu^{2p-2}e^{-a(F+b\varphi-b\rho)}|\nabla u|^{2}_{\omega_{X}} (2.10)

By (2.9), (2.10), and the weighted integration-by-parts formula (1.3), we obtain

0=X12p+1Δφ,vu2p+1v(mφ)ωφnX2pu2p2ea(F+bφbρ)+F|u|ωX2ωXn(aC1+𝒞)Xu2p+1eFωXn+abCρ2Cv1n1Xu2p+1en2n1F(trωXωφ)1n1ωXn+Xu2pea(F+bφbρ)+FΔωXFωXn=:I.\begin{split}0&=\int_{X}\frac{1}{2p+1}\Delta_{\varphi,v}u^{2p+1}v(m_{\varphi})\omega_{\varphi}^{n}\\ &\geq\int_{X}2pu^{2p-2}e^{-a(F+b\varphi-b\rho)+F}|\nabla u|^{2}_{\omega_{X}}\omega_{X}^{n}-(aC_{1}+\mathcal{C})\int_{X}u^{2p+1}e^{F}\omega_{X}^{n}\\ &\quad\quad+\frac{abC_{\rho}}{2}C_{v}^{-\frac{1}{n-1}}\int_{X}u^{2p+1}e^{\frac{n-2}{n-1}F}(\operatorname{tr}_{\omega_{X}}\omega_{\varphi})^{\frac{1}{n-1}}\omega_{X}^{n}+\underset{=:I}{\underbrace{\int_{X}u^{2p}e^{-a(F+b\varphi-b\rho)+F}\Delta_{\omega_{X}}F\omega_{X}^{n}}}.\end{split} (2.11)

Set G=(1a)Fab(φρ)G=(1-a)F-ab(\varphi-\rho) and consider a>1a>1. The last term II can be written as the following two parts

I=Xu2pΔωXFeGωXn=11aXu2pΔωX(G)eGωXn=:I1+ab1aXu2pΔωX(φρ)eGωXn=:I2.\displaystyle I=\int_{X}u^{2p}\Delta_{\omega_{X}}Fe^{G}\omega_{X}^{n}=\underset{=:I_{1}}{\underbrace{\frac{1}{1-a}\int_{X}u^{2p}\Delta_{\omega_{X}}(G)e^{G}\omega_{X}^{n}}}+\underset{=:I_{2}}{\underbrace{\frac{ab}{1-a}\int_{X}u^{2p}\Delta_{\omega_{X}}(\varphi-\rho)e^{G}\omega_{X}^{n}}}.

For I1I_{1}, the weighted integration-by-parts formula (1.3) implies

I1\displaystyle I_{1} =11aXu2pΔωX(G)eGωXn\displaystyle=\frac{1}{1-a}\int_{X}u^{2p}\Delta_{\omega_{X}}(G)e^{G}\omega_{X}^{n}
=1a1Xu2p|G|ωX2eGωXn+1a1Xn2pu2p1eG𝑑udcGωXn1.\displaystyle=\frac{1}{a-1}\int_{X}u^{2p}|\nabla G|^{2}_{\omega_{X}}e^{G}\omega_{X}^{n}+\frac{1}{a-1}\int_{X}n2pu^{2p-1}e^{G}du\wedge d^{c}G\wedge\omega_{X}^{n-1}.

By Cauchy–Schwarz inequality,

|2pu2p1dudcGωXn1ωXn|4p22u2p2|u|ωX2+12u2p|G|ωX2.\left|2pu^{2p-1}\frac{du\wedge d^{c}G\wedge\omega_{X}^{n-1}}{\omega_{X}^{n}}\right|\leq\frac{4p^{2}}{2}u^{2p-2}|\nabla u|^{2}_{\omega_{X}}+\frac{1}{2}u^{2p}|\nabla G|^{2}_{\omega_{X}}.

The above two inequalities yield

I12p2a1Xu2p2|u|ωX2eGωXn.\displaystyle I_{1}\geq-\frac{2p^{2}}{a-1}\int_{X}u^{2p-2}|\nabla u|^{2}_{\omega_{X}}e^{G}\omega_{X}^{n}. (2.12)

We next control the term I2I_{2}. We compute

ΔωX,v(φρ)=trωX,v(ωφωρ)\displaystyle\Delta_{\omega_{X},v}(\varphi-\rho)=\operatorname{tr}_{\omega_{X},v}(\omega_{\varphi}-\omega_{\rho}) =trωX(ωφωρ)+(logv)(mωX),(mφmρ)\displaystyle=\operatorname{tr}_{\omega_{X}}(\omega_{\varphi}-\omega_{\rho})+\langle(\log v)^{\prime}(m_{\omega_{X}}),(m_{\varphi}-m_{\rho})\rangle
trωXωφ+C3\displaystyle\leq\operatorname{tr}_{\omega_{X}}\omega_{\varphi}+C_{3}

where C3C_{3} only depends on P,CvP,C_{v}. Since a>1a>1, we get

I2ab1aXu2ptrωXωφeGωXn+abC31aXu2peGωXn.\displaystyle I_{2}\geq\frac{ab}{1-a}\int_{X}u^{2p}\operatorname{tr}_{\omega_{X}}\omega_{\varphi}e^{G}\omega_{X}^{n}+\frac{abC_{3}}{1-a}\int_{X}u^{2p}e^{G}\omega_{X}^{n}. (2.13)

Combining (2.11), (2.12) and (2.13), one can derive

0\displaystyle 0 2(pp2a1)Xu2p2eG|u|ωX2ωXn(aC1+𝒞)Xu2p+1eFωXn\displaystyle\geq 2\left(p-\frac{p^{2}}{a-1}\right)\int_{X}u^{2p-2}e^{G}|\nabla u|^{2}_{\omega_{X}}\omega_{X}^{n}-(aC_{1}+\mathcal{C})\int_{X}u^{2p+1}e^{F}\omega_{X}^{n}
+abCρ2Cv1n1Xu2p+1en2n1F(trωXωφ)1n1ωXn+ab1aXu2ptrωXωφeGωXn+abC31aXu2peGωXn.\displaystyle\qquad+\frac{abC_{\rho}}{2}C_{v}^{-\frac{1}{n-1}}\int_{X}u^{2p+1}e^{\frac{n-2}{n-1}F}(\operatorname{tr}_{\omega_{X}}\omega_{\varphi})^{\frac{1}{n-1}}\omega_{X}^{n}+\frac{ab}{1-a}\int_{X}u^{2p}\operatorname{tr}_{\omega_{X}}\omega_{\varphi}e^{G}\omega_{X}^{n}+\frac{abC_{3}}{1-a}\int_{X}u^{2p}e^{G}\omega_{X}^{n}.

Taking a>1a>1 sufficiently large so that p>p2/(a1)p>p^{2}/(a-1), and using the upper bound FC0F\leq C_{0}, one gains n2n1FFC0/(n1)\frac{n-2}{n-1}F\geq F-C_{0}/(n-1) and thus,

0\displaystyle 0 C4X(trωXωφ)2p+1e(2p+1)(GF)+FωXnC5Xu2peGωXn+C6X(trωXωφ)2p+1+1n1e(2p+1)(GF)+FωXn,\displaystyle\geq-C_{4}\int_{X}(\operatorname{tr}_{\omega_{X}}\omega_{\varphi})^{2p+1}e^{(2p+1)(G-F)+F}\omega_{X}^{n}-C_{5}\int_{X}u^{2p}e^{G}\omega_{X}^{n}+C_{6}\int_{X}(\operatorname{tr}_{\omega_{X}}\omega_{\varphi})^{2p+1+\frac{1}{n-1}}e^{(2p+1)(G-F)+F}\omega_{X}^{n},

where C4,C5,C6C_{4},C_{5},C_{6} only depend on a,b,p,C0,Cρ,Cv,na,b,p,C_{0},C_{\rho},C_{v},n and |(v(mωX))|ωX2|\nabla(v(m_{\omega_{X}}))|^{2}_{\omega_{X}}. Set μ:=e(2p+1)(GF)+FωXn=e[a(2p+1)1]Fab(2p+1)φ+ab(2p+1)ρ)ωXn\mu:=e^{(2p+1)(G-F)+F}\omega_{X}^{n}=e^{-[a(2p+1)-1]F-ab(2p+1)\varphi+ab(2p+1)\rho)}\omega_{X}^{n}. Choose b1b\gg 1 such that ab(2p+1)>K2(a(2p+1)1)ab(2p+1)>K_{2}(a(2p+1)-1). By FK2ρC0F\geq K_{2}\rho-C_{0} in the assumption,

C71eab(2p+1)ρμωXnC7,C_{7}^{-1}e^{ab(2p+1)\rho}\leq\frac{\mu}{\omega_{X}^{n}}\leq C_{7},

and it implies that C81XμC8C_{8}^{-1}\leq\int_{X}\mu\leq C_{8}. Therefore, we obtain

0\displaystyle 0 C4X(trωXωφ)2p+1𝑑μC5X(trωXωφ)2p𝑑μ+C6X(trωXωφ)2p+1+1n1𝑑μ,\displaystyle\geq-C_{4}\int_{X}(\operatorname{tr}_{\omega_{X}}\omega_{\varphi})^{2p+1}d\mu-C_{5}\int_{X}(\operatorname{tr}_{\omega_{X}}\omega_{\varphi})^{2p}d\mu+C_{6}\int_{X}(\operatorname{tr}_{\omega_{X}}\omega_{\varphi})^{2p+1+\frac{1}{n-1}}d\mu,

and Hölder’s inequality shows trωXωφL2p+1(X,μ)C\|\operatorname{tr}_{\omega_{X}}\omega_{\varphi}\|_{L^{2p+1}(X,\mu)}\leq C. This concludes the proof. ∎

2.3. Local higher-order estimates

We next provide higher-order estimates that are away from the degenerating locus. We shall use the following local estimate, which generalizes [CC21a] for cscK equations (see also [DJL24, HL25]). Its proof follows a similar argument in [CC21a] with further analysis of the weighted terms. For full details, the reader is referred to Appendix B.

Theorem 2.3.

Assume that 𝔱ζlogv(ζ)\mathfrak{t}^{\vee}\ni\zeta\mapsto\log v(\zeta)\in\mathbb{R} is concave and w𝒞(𝔱,)w\in\mathcal{C}^{\infty}(\mathfrak{t}^{\vee},\mathbb{R}). Let ϕ\phi be a smooth solution of

{v(mddcϕ)det(ϕkj¯)=eGΔϕ,vG=w(mddcϕ)\begin{cases}v(m_{dd^{c}\phi})\det{(\phi_{k\bar{j}})}=e^{G}\\ \Delta_{\phi,v}G=-w(m_{dd^{c}\phi})\end{cases}

in B1(0)nB_{1}(0)\subset\mathbb{C}^{n}, where Δϕ,vf:=ϕkj¯kj¯f+trddcϕ(dlogv(mddcϕ)dcf)\Delta_{\phi,v}f:=\phi^{k\bar{j}}\partial_{k}\partial_{\bar{j}}f+\operatorname{tr}_{dd^{c}\phi}(d\log v(m_{dd^{c}\phi})\wedge d^{c}f), such that ΔϕLp(B1(0))\Delta\phi\in L^{p}(B_{1}(0)) and trϕωeucLp(B1(0))\operatorname{tr}_{\phi}\omega_{\mathrm{euc\ell}}\in L^{p}(B_{1}(0)) for some p>3np>3n. Then there exists a constant A>0A>0, depending only on CvC_{v}, CwC_{w}, pp, ΔϕLp(B1(0))\|\Delta\phi\|_{L^{p}(B_{1}(0))}, trϕωeucLp(B1(0))\|\operatorname{tr}_{\phi}\omega_{\mathrm{euc\ell}}\|_{L^{p}(B_{1}(0))}, such that |ϕkj¯|A|\phi_{k\bar{j}}|\leq A, |G|A|\nabla G|\leq A in B12(0))B_{\frac{1}{2}}(0)) and for any k2k\geq 2

DkϕL(B1/2(0))C(k,A,ϕL(B1(0))).\|D^{k}\phi\|_{L^{\infty}(B_{1/2}(0))}\leq C(k,A,\|\phi\|_{L^{\infty}(B_{1}(0))}).

Combining Theorem 2.1, Proposition 2.2, and Theorem 2.3, we obtain the following result:

Theorem 2.4.

Under the same assumption of Theorem 2.1 and assuming that logv\log v is concave and S=w(mϕ),w𝒞(𝔱,)S=w(m_{\phi}),w\in\mathcal{C}^{\infty}(\mathfrak{t}^{\vee},\mathbb{R}), there is a uniform constant C>0C>0 depending only on n,V,α,K0,K1,K2,K3,K4,Cv,Cw,μ,ρ1n,V,\alpha,K_{0},K_{1},K_{2},K_{3},K_{4},C_{v},C_{w},\mu,\rho_{1} such that

φLC and FC and FK2ρ2C,\|\varphi\|_{L^{\infty}}\leq C\,\text{ and }\,F\leq C\,\text{ and }F\geq K_{2}\rho_{2}-C,

where ρ2\rho_{2} is a quasi-plurisubharmonic function with analytic singularities, such that ω+ddcρ2\omega+dd^{c}\rho_{2} is a Kähler current with Ric(μ)K2(ω+ddcρ2)\operatorname{Ric}(\mu)\leq K_{2}(\omega+dd^{c}\rho_{2}).

Moreover, for any compact set KX{ρ2=}K\subset X\setminus\{\rho_{2}=-\infty\} and \ell\in\mathbb{N}, we have φε𝒞(K,ωX)CK,\|\varphi_{\varepsilon}\|_{\mathcal{C}^{\ell}(K,\omega_{X})}\leq C_{K,\ell} for some uniform constant CK,>0C_{K,\ell}>0, where ωX\omega_{X} is a fixed Kähler metric on XX.

3. Existence of singular weighted cscK metrics

We shall combine the a priori estimates obtained in the previous section and the uniform coercivity established in [BJT24] to construct singular weighted cscK metrics.

3.1. Setup

In the sequel, we always assume the following setting:

Setting 3.1.

Let XX be an nn-dimensional compact Kähler variety with log terminal singularities. Fix a compact torus TAutred(X)T\subset\operatorname{Aut}_{\operatorname{red}}(X) and denote by 𝔱=Lie(T)\mathfrak{t}=\operatorname{Lie}(T). Assume that XX admits a TT-equivariant resolution of singularities π:YX\pi:Y\to X with KY=πKX+iaiEiK_{Y}=\pi^{*}K_{X}+\sum_{i}a_{i}E_{i}, ai>1a_{i}>-1, YY is Kähler, and π\pi is an isomorphism over XregX^{\mathrm{reg}}. Let ωY\omega_{Y} be a TT-invariant Kähler metric on YY. Given a TT-invariant Kähler metric ω\omega on XX, by [Bou04, Thm. 3.17], there exists a TT-invariant quasi-psh function ρ\rho that is smooth on YEY\setminus E and has analytic singularities along EE such that πω+ddcρ\pi^{\ast}\omega+dd^{c}\rho is a Kähler current, i.e. πω+ddcρCρωY\pi^{\ast}\omega+dd^{c}\rho\geq C_{\rho}\omega_{Y} for some Cρ>0C_{\rho}>0. Fix a TT-invariant adapted measure μX\mu_{X} on XX, which is normalized by Ylog(πμXωYn)πωn=0\int_{Y}\log\left(\frac{\pi^{*}\mu_{X}}{\omega_{Y}^{n}}\right)\pi^{*}\omega^{n}=0.

For ε(0,1]\varepsilon\in(0,1], we denote by ωε:=πω+εωY\omega_{\varepsilon}:=\pi^{\ast}\omega+\varepsilon\omega_{Y} and Pε:=im(mωε)=im(mπω)+εim(mωY)P_{\varepsilon}:=\operatorname{im}(m_{\omega_{\varepsilon}})=\operatorname{im}(m_{\pi^{*}\omega})+\varepsilon\operatorname{im}(m_{\omega_{Y}}) which contained in a compact set of 𝔱\mathfrak{t}^{\vee} for all ε\varepsilon small. Let v,w𝒞(𝔱,)v,w\in\mathcal{C}^{\infty}(\mathfrak{t}^{\vee},\mathbb{R}) with v,w>0v,w>0 on PωP_{\omega}. Since PωP_{\omega} is compact, there are constants Cv,Cw1C_{v},C_{w}\geq 1 such that for any ε\varepsilon sufficiently small, on PεP_{\varepsilon},

Cv1v+α|vα|+α,β|vαβ|Cv,andCw1|w|+α|wα|+α,β|wαβ|Cw.C_{v}^{-1}\leq v+\sum_{\alpha}|v_{\alpha}|+\sum_{\alpha,\beta}|v_{\alpha\beta}|\leq C_{v},\quad\text{and}\quad C_{w}^{-1}\leq|w|+\sum_{\alpha}|w_{\alpha}|+\sum_{\alpha,\beta}|w_{\alpha\beta}|\leq C_{w}.

Denote by εext\ell^{\operatorname{ext}}_{\varepsilon} (reps. ext\ell^{\operatorname{ext}}) the extremal affine function on 𝔱\mathfrak{t}^{\vee} associated to ωε\omega_{\varepsilon} (resp. ω\omega). Since ωε\omega_{\varepsilon} converges smoothly to πω\pi^{\ast}\omega, the moment map mωεm_{\omega_{\varepsilon}} converges smoothly to mπωm_{\pi^{\ast}\omega} and εextext\ell^{\operatorname{ext}}_{\varepsilon}\rightarrow\ell^{\operatorname{ext}} in 𝔱\mathfrak{t}\oplus\mathbb{R} (cf. [BJT24, Lem. 4.18]).

Assume that a smooth pair (φε,Fε)(\varphi_{\varepsilon},F_{\varepsilon}) solves

{v(mφε)(ωε+ddcφε)n=eFεωYn,supYφε=0,Δφε,vFε=(wεext)(mφε)+trφε,v(Ric(ωY)).\begin{cases}v(m_{\varphi_{\varepsilon}})(\omega_{\varepsilon}+dd^{c}\varphi_{\varepsilon})^{n}=e^{F_{\varepsilon}}\omega_{Y}^{n},\quad\sup_{Y}\varphi_{\varepsilon}=0,\\ \Delta_{\varphi_{\varepsilon},v}F_{\varepsilon}=-(w\ell^{\operatorname{ext}}_{\varepsilon})(m_{\varphi_{\varepsilon}})+\operatorname{tr}_{\varphi_{\varepsilon},v}(\operatorname{Ric}(\omega_{Y})).\end{cases} (3.1)

Under Setting 3.1, one can conclude the following Theorem  3.2 by applying Theorem 2.4.

Theorem 3.2.

Under Setting 3.1, suppose that Condition (A) holds and logv\log v is concave. Let (φε,Fε)(\varphi_{\varepsilon},F_{\varepsilon}) be a solution to (3.1). If there is a constant C>0C>0 independent of sufficient small ε>0\varepsilon>0 such that

𝐇ε(φε)Ylog(ωε,φεnωYn)ωε,φεnC,\mathbf{H}_{\varepsilon}(\varphi_{\varepsilon})\leq\int_{Y}\log\left(\frac{\omega_{\varepsilon,\varphi_{\varepsilon}}^{n}}{\omega_{Y}^{n}}\right)\omega_{\varepsilon,\varphi_{\varepsilon}}^{n}\leq C,

then there exists a uniform constant C0>0C_{0}>0 such that

φεLC0,C0FεK2ρC0,\|\varphi_{\varepsilon}\|_{L^{\infty}}\leq C_{0},\quad C_{0}\geq F_{\varepsilon}\geq K_{2}\rho-C_{0},

where K2K_{2} is a positive constant such that Ric(ωY)K2(πω+ddcρ)\operatorname{Ric}(\omega_{Y})\leq K_{2}(\pi^{\ast}\omega+dd^{c}\rho). Moreover, for any compact subset KYExc(π)K\subset Y\setminus\operatorname{Exc}(\pi) and \ell\in\mathbb{N}, we have φε𝒞(K,ωY)CK,\|\varphi_{\varepsilon}\|_{\mathcal{C}^{\ell}(K,\omega_{Y})}\leq C_{K,\ell} for some uniform constant CK,>0C_{K,\ell}>0.

3.2. Proof of Theorem A

We now prove the following result on the existence of singular weighted cscK metrics.

Theorem 3.3.

Under Setting 3.1, suppose that Condition (A) is fulfilled and logv\log v is concave. If the weighted Mabuchi functional 𝐌ω,v,wext\mathbf{M}_{\omega,v,w\ell^{\operatorname{ext}}} is TT_{\mathbb{C}}-coercive, then XX admits a singular (v,w)(v,w)-extremal metric (i.e. (v,wext)(v,w\ell^{\operatorname{ext}})-cscK metric) in {ω}\{\omega\}, and it is also a minimizer of 𝐌ω,v,wext\mathbf{M}_{\omega,v,w\ell^{\operatorname{ext}}}.

Proof.

Part 1: existence. Since 𝐌Xre:=𝐌ω,v,wext\mathbf{M}^{\rm re\ell}_{X}:=\mathbf{M}_{\omega,v,w\ell^{\operatorname{ext}}} is TT_{\mathbb{C}}-coercive, there are positive constants A0A_{0} and B0B_{0} such that

𝐌XreA0d1,TB0on norm1(X,ω)T.\mathbf{M}^{\rm re\ell}_{X}\geq A_{0}d_{1,T}-B_{0}\quad\text{on }\mathcal{E}^{1}_{{\mathrm{norm}}}(X,\omega)^{T}.

By [BJT24, Thm. A], for any A(0,A0)A\in(0,A_{0}), there exists B>0B>0 such that for all sufficiently small ε>0\varepsilon>0,

𝐌εreAd1,ωε,TBonnorm1(Y,ωε)T,\mathbf{M}_{\varepsilon}^{\rm re\ell}\geq Ad_{1,\omega_{\varepsilon},T}-B\quad\text{on}\ \,\mathcal{E}^{1}_{\mathrm{norm}}(Y,\omega_{\varepsilon})^{T}, (3.2)

where d1,ωε,Td_{1,\omega_{\varepsilon},T} is d1,Td_{1,T} with respect to ωε\omega_{\varepsilon} in (1.6) and

𝐌εre:=𝐌ωε,v,wεext=𝐇ε,v,ωYn+𝐄ε,vRic(ωY)+𝐄ε,vwεext\mathbf{M}_{\varepsilon}^{\rm re\ell}:=\mathbf{M}_{\omega_{\varepsilon},v,w\ell^{\operatorname{ext}}_{\varepsilon}}=\mathbf{H}_{\varepsilon,v,\omega_{Y}^{n}}+\mathbf{E}_{\varepsilon,v}^{-\operatorname{Ric}(\omega_{Y})}+\mathbf{E}_{\varepsilon,vw\ell^{\operatorname{ext}}_{\varepsilon}}

is the weighted Mabuchi functional on (Y,ωε)(Y,\omega_{\varepsilon}). By the existence result obtained in [DJL25, HL25], there is a smooth pair (φε,Fε)(\varphi_{\varepsilon},F_{\varepsilon}) solving

{v(mφε)(ωε+ddcφε)n=eFεωYn,𝐄ωε(φε)=0Δφε,vFε=(wεext)(mφε)+trφε,v(Ric(ωY))\begin{cases}v(m_{\varphi_{\varepsilon}})(\omega_{\varepsilon}+dd^{c}\varphi_{\varepsilon})^{n}=e^{F_{\varepsilon}}\omega_{Y}^{n},\quad\mathbf{E}_{\omega_{\varepsilon}}(\varphi_{\varepsilon})=0\\ \Delta_{\varphi_{\varepsilon},v}F_{\varepsilon}=-(w\ell^{\operatorname{ext}}_{\varepsilon})(m_{\varphi_{\varepsilon}})+\operatorname{tr}_{\varphi_{\varepsilon},v}(\operatorname{Ric}(\omega_{Y}))\end{cases} (3.3)

Namely, ωε,φε\omega_{\varepsilon,\varphi_{\varepsilon}} is a (v,wεext)(v,w\ell^{\operatorname{ext}}_{\varepsilon})-cscK metric and φε\varphi_{\varepsilon} minimizes 𝐌εre\mathbf{M}_{\varepsilon}^{\rm re\ell}.

From Lemma 1.5, we know that φε\varphi_{\varepsilon} minimizes the relative Mabuchi functional 𝐌εre\mathbf{M}_{\varepsilon}^{\rm re\ell}, so 𝐌εre(φε)𝐌εre(0)C1\mathbf{M}_{\varepsilon}^{\rm re\ell}(\varphi_{\varepsilon})\leq\mathbf{M}_{\varepsilon}^{\rm re\ell}(0)\leq C_{1}, which gives

d1,ωε,T(φε,0)C1+BAd_{1,\omega_{\varepsilon},T}(\varphi_{\varepsilon},0)\leq\frac{C_{1}+B}{A}

using the coercivity condition. By the definition of d1,ωε,Td_{1,\omega_{\varepsilon},T}, for each ε\varepsilon there exists σεT\sigma_{\varepsilon}\in T_{\mathbb{C}} such that d1,ωε(σεφε,0)C1+BAd_{1,\omega_{\varepsilon}}(\sigma_{\varepsilon}\cdot\varphi_{\varepsilon},0)\leq\frac{C_{1}+B}{A}. Set φ~ε:=σεφε\tilde{\varphi}_{\varepsilon}:=\sigma_{\varepsilon}\cdot\varphi_{\varepsilon}. By the TT_{\mathbb{C}}-action, we have

ωε,φ~ε=σεωε,φε,mφ~ε=σεmφε\omega_{\varepsilon,\tilde{\varphi}_{\varepsilon}}=\sigma_{\varepsilon}^{*}\omega_{\varepsilon,\varphi_{\varepsilon}},\quad m_{\tilde{\varphi}_{\varepsilon}}=\sigma_{\varepsilon}^{*}m_{\varphi_{\varepsilon}}

(see e.g. [BJT24, Lem. 3.19]). Pulling back the first equation in (3.3) by σε\sigma_{\varepsilon}, we get

v(mφ~ε)(ωε+ddcφ~ε)n=σε(v(mφε)(ωε+ddcφε)n)=eσεFε(σωY)n=eF~εωYnv(m_{\tilde{\varphi}_{\varepsilon}})(\omega_{\varepsilon}+dd^{c}\tilde{\varphi}_{\varepsilon})^{n}=\sigma_{\varepsilon}^{\ast}\left(v(m_{\varphi_{\varepsilon}})(\omega_{\varepsilon}+dd^{c}\varphi_{\varepsilon})^{n}\right)=e^{\sigma^{*}_{\varepsilon}F_{\varepsilon}}(\sigma^{*}\omega_{Y})^{n}=e^{\tilde{F}_{\varepsilon}}\omega_{Y}^{n}

where F~ε=σεFε+log(σεωY)nωYn\tilde{F}_{\varepsilon}=\sigma_{\varepsilon}^{*}F_{\varepsilon}+\log\frac{(\sigma_{\varepsilon}^{*}\omega_{Y})^{n}}{\omega_{Y}^{n}}. For any TT-invariant Kähler metric ω\omega on YY and σT\sigma\in T_{\mathbb{C}}, we have

σ(trω,v)=trσω,v(σ)andσ(Δω,v)=Δσω,v(σ).\sigma^{\ast}\left(\operatorname{tr}_{\omega,v}\bullet\right)=\operatorname{tr}_{\sigma^{\ast}\omega,v}(\sigma^{\ast}\bullet)\quad\text{and}\quad\sigma^{\ast}\left(\Delta_{\omega,v}\bullet\right)=\Delta_{\sigma^{\ast}\omega,v}(\sigma^{\ast}\bullet).

Combining this with the second equation in (3.3) yields

Δφ~ε,vF~ε\displaystyle\Delta_{\tilde{\varphi}_{\varepsilon},v}\tilde{F}_{\varepsilon} =Δσεωε,φε,vσεFε+Δφ~ε,vlog(σεωY)nωYn=σεΔφε,vFε+Δφ~ε,vlog(σεωY)nωYn\displaystyle=\Delta_{\sigma_{\varepsilon}^{\ast}\omega_{\varepsilon,\varphi_{\varepsilon}},v}\sigma_{\varepsilon}^{\ast}F_{\varepsilon}+\Delta_{\tilde{\varphi}_{\varepsilon},v}\log\frac{(\sigma_{\varepsilon}^{*}\omega_{Y})^{n}}{\omega_{Y}^{n}}=\sigma_{\varepsilon}^{*}\Delta_{\varphi_{\varepsilon},v}F_{\varepsilon}+\Delta_{\tilde{\varphi}_{\varepsilon},v}\log\frac{(\sigma_{\varepsilon}^{*}\omega_{Y})^{n}}{\omega_{Y}^{n}}
=(wεext)(mφ~ε)+trφ~ε,v(Ric(σεωY))+trφ~ε,v(Ric(ωY)Ric(σεωY))\displaystyle=-(w\ell^{\operatorname{ext}}_{\varepsilon})(m_{\tilde{\varphi}_{\varepsilon}})+\operatorname{tr}_{\tilde{\varphi}_{\varepsilon},v}(\operatorname{Ric}(\sigma_{\varepsilon}^{*}\omega_{Y}))+\operatorname{tr}_{\tilde{\varphi}_{\varepsilon},v}(\operatorname{Ric}(\omega_{Y})-\operatorname{Ric}(\sigma_{\varepsilon}^{*}\omega_{Y}))
=(wεext)(mφ~ε)+trφ~ε,v(Ric(ωY)).\displaystyle=-(w\ell^{\operatorname{ext}}_{\varepsilon})(m_{\tilde{\varphi}_{\varepsilon}})+\operatorname{tr}_{\tilde{\varphi}_{\varepsilon},v}(\operatorname{Ric}(\omega_{Y})).

Therefore, (σεφε,F~ε)(\sigma_{\varepsilon}\cdot\varphi_{\varepsilon},\tilde{F}_{\varepsilon}) also solves (3.3). Replacing φε\varphi_{\varepsilon} by σεφε\sigma_{\varepsilon}\cdot\varphi_{\varepsilon} and FεF_{\varepsilon} by F~ε\tilde{F}_{\varepsilon}, we thus obtain

d1,ωε(φε,0)C1+BA.d_{1,\omega_{\varepsilon}}(\varphi_{\varepsilon},0)\leq\frac{C_{1}+B}{A}. (3.4)

Since 𝐌εre\mathbf{M}_{\varepsilon}^{\rm re\ell} is TT_{\mathbb{C}}-invariant, this still minimize 𝐌εre\mathbf{M}_{\varepsilon}^{\rm re\ell}.

Recall that πω+ddcρCρωY\pi^{\ast}\omega+dd^{c}\rho\geq C_{\rho}\omega_{Y}. We shall verify the assumptions in Theorem 2.1 with ρ2=ρ\rho_{2}=\rho and μ=ωYn\mu=\omega_{Y}^{n}. It is obvious that the condition (3) holds. We now show that the uniform control on the entropies 𝐇ε(φε)\mathbf{H}_{\varepsilon}(\varphi_{\varepsilon}), which will confirm the condition (2).

If a closed (1,1)(1,1)-current ΘAπω\Theta\geq-A\pi^{\ast}\omega for some constant A>0A>0, then

𝐄ε,vΘ(ψ)C(A,v,Θ)(d1,ε(ψ,0)+1).\mathbf{E}_{\varepsilon,v}^{\Theta}(\psi)\leq C(A,v,\Theta)(d_{1,\varepsilon}(\psi,0)+1). (3.5)

for any ε(0,1]\varepsilon\in(0,1] and for all ψ1(Y,ωε)\psi\in\mathcal{E}^{1}(Y,\omega_{\varepsilon}) (c.f [BJT24, Lem. 4.22]). By Condition (A) and the strong openness [Ber13, GZ15], there is a constant a>1a>1 such that YeaK1ρ1ωYn<+\int_{Y}e^{-aK_{1}\rho_{1}}\omega_{Y}^{n}<+\infty. Set μY:=eaK1ρ1ωYn\mu_{Y}:=e^{-aK_{1}\rho_{1}}\omega_{Y}^{n}. Similar to [BJT24, Sec. 4.4], we consider

𝐇^ε,v(ψ):=𝐇μY(MAε,v(ψ))andRic^:=Ric(ωY)+K1ddcρ1,\widehat{\mathbf{H}}_{\varepsilon,v}(\psi):=\mathbf{H}_{\mu_{Y}}(\operatorname{MA}_{\varepsilon,v}(\psi))\quad\text{and}\quad\widehat{\operatorname{Ric}}:=\operatorname{Ric}(\omega_{Y})+K_{1}dd^{c}\rho_{1},

then Ric^K1πωK1ωε\widehat{\operatorname{Ric}}\geq-K_{1}\pi^{\ast}\omega\geq-K_{1}\omega_{\varepsilon} by assumption. It follows from [BJT24, Lem. A.2] and [BJT24, Prop. 3.44] that

1a𝐇^ε,v(ψ)=1a𝐇ε,v(ψ)+YK1ρ1MAε,v(ψ)\frac{1}{a}\widehat{\mathbf{H}}_{\varepsilon,v}(\psi)=\frac{1}{a}\mathbf{H}_{\varepsilon,v}(\psi)+\int_{Y}K_{1}\rho_{1}\operatorname{MA}_{\varepsilon,v}(\psi) (3.6)

and

𝐄ε,vRic^(ψ)=𝐄ε,vRic(ωY)(ψ)YK1ρ1MAε,v(ψ)+YK1ρ1MAε,v(0).\mathbf{E}^{-\widehat{\operatorname{Ric}}}_{\varepsilon,v}(\psi)=\mathbf{E}^{-{\operatorname{Ric}(\omega_{Y})}}_{\varepsilon,v}(\psi)-\int_{Y}K_{1}\rho_{1}\operatorname{MA}_{\varepsilon,v}(\psi)+\int_{Y}K_{1}\rho_{1}\operatorname{MA}_{\varepsilon,v}(0). (3.7)

Therefore, we get

1a𝐇ε,v,ωYn(ψ)+𝐄ε,vRic(ωY)(ψ)=1a𝐇^ε,v(ψ)+𝐄ε,vRic^(ψ)YK1ρ1MAε,v(0),\frac{1}{a}\mathbf{H}_{\varepsilon,v,\omega_{Y}^{n}}(\psi)+\mathbf{E}_{\varepsilon,v}^{-\operatorname{Ric}(\omega_{Y})}(\psi)=\frac{1}{a}\widehat{\mathbf{H}}_{\varepsilon,v}(\psi)+\mathbf{E}_{\varepsilon,v}^{-\widehat{\operatorname{Ric}}}(\psi)-\int_{Y}K_{1}\rho_{1}\operatorname{MA}_{\varepsilon,v}(0), (3.8)

with YK1ρ1MAε,v(0)YK1ρ1v(mπω)(πω)n\int_{Y}K_{1}\rho_{1}\operatorname{MA}_{\varepsilon,v}(0)\rightarrow\int_{Y}K_{1}\rho_{1}v(m_{\pi^{*}\omega})(\pi^{*}\omega)^{n} as ε0\varepsilon\to 0. Here without loss of generality, we assume that Yρ1v(mπω)(πω)n=0\int_{Y}\rho_{1}v(m_{\pi^{*}\omega})(\pi^{*}\omega)^{n}=0. Then one can derive

𝐌εre(ψ)=𝐇ε,v,ωYn(ψ)+𝐄ε,vRic(ωY)(ψ)+𝐄ε,vwext(ψ)(11a)𝐇ε,v,ωYn(ψ)+𝐄ε,vRic^(ψ)+𝐄ε,vwext(ψ)C.\begin{split}\mathbf{M}_{\varepsilon}^{\rm re\ell}(\psi)&=\mathbf{H}_{\varepsilon,v,\omega_{Y}^{n}}(\psi)+\mathbf{E}_{\varepsilon,v}^{-\operatorname{Ric}(\omega_{Y})}(\psi)+\mathbf{E}_{\varepsilon,vw\ell^{\operatorname{ext}}}(\psi)\\ &\geq\left(1-\frac{1}{a}\right)\mathbf{H}_{\varepsilon,v,\omega_{Y}^{n}}(\psi)+\mathbf{E}_{\varepsilon,v}^{-\widehat{\operatorname{Ric}}}(\psi)+\mathbf{E}_{\varepsilon,vw\ell^{\operatorname{ext}}}(\psi)-C.\end{split} (3.9)

By [BJT24, Lem. 4.20], |𝐄ε(ψ)|C2d1,ωε(ψ,0)|\mathbf{E}_{\varepsilon}(\psi)|\leq C_{2}d_{1,\omega_{\varepsilon}}(\psi,0) for all ψ1(Y,ωε)\psi\in\mathcal{E}^{1}(Y,\omega_{\varepsilon}), hence (3.4) implies that

|𝐄ε,vwext(φε)|C3=C(A,B,C1,C2,Cv,Cw),|\mathbf{E}_{\varepsilon,vw\ell^{\operatorname{ext}}}(\varphi_{\varepsilon})|\leq C_{3}=C(A,B,C_{1},C_{2},C_{v},C_{w}),

for all ε\varepsilon small, and then (3.5) and (3.9) imply that

(11a)𝐇ε,v,ωYn(φε)C𝐄ε,vwext(φε)+𝐄ε,vRic^(φε)C1+C3+C(K1,v,Ric^)C4.\left(1-\frac{1}{a}\right)\mathbf{H}_{\varepsilon,v,\omega_{Y}^{n}}(\varphi_{\varepsilon})\leq C-\mathbf{E}_{\varepsilon,vw\ell^{\operatorname{ext}}}(\varphi_{\varepsilon})+\mathbf{E}^{\widehat{\operatorname{Ric}}}_{\varepsilon,v}(\varphi_{\varepsilon})\leq C_{1}+C_{3}+C(K_{1},v,\widehat{\operatorname{Ric}})C_{4}.

Therefore, 𝐇ε,v,ωYn(φε)C\mathbf{H}_{\varepsilon,v,\omega_{Y}^{n}}(\varphi_{\varepsilon})\leq C^{\prime} for all sufficiently small ε>0\varepsilon>0 and for some uniform constant C>0C^{\prime}>0. We have

|𝐇ε,v,ωYn(φε)𝐇ε,ωYn(φε)|=|Ylog(v(mφε)ωε,φεnωYn)v(mφε)ωε,φεnYlog(ωε,φεnωYn)ωε,φεn|\displaystyle|\mathbf{H}_{\varepsilon,v,\omega_{Y}^{n}}(\varphi_{\varepsilon})-\mathbf{H}_{\varepsilon,\omega_{Y}^{n}}(\varphi_{\varepsilon})|=\left|\int_{Y}\log\left(\frac{v(m_{\varphi_{\varepsilon}})\omega_{\varepsilon,\varphi_{\varepsilon}}^{n}}{\omega_{Y}^{n}}\right)v(m_{\varphi_{\varepsilon}})\omega_{\varepsilon,\varphi_{\varepsilon}}^{n}-\int_{Y}\log\left(\frac{\omega_{\varepsilon,\varphi_{\varepsilon}}^{n}}{\omega_{Y}^{n}}\right)\omega_{\varepsilon,\varphi_{\varepsilon}}^{n}\right|
=|Y(11v(mφε))log(v(mφε)ωε,φεnωYn)v(mφε)ωε,φεn+Ylog(v(mφε))ωε,φεn|CvC+Cv,\displaystyle=\left|\int_{Y}\left(1-\frac{1}{v(m_{\varphi_{\varepsilon}})}\right)\log\left(\frac{v(m_{\varphi_{\varepsilon}})\omega_{\varepsilon,\varphi_{\varepsilon}}^{n}}{\omega_{Y}^{n}}\right)v(m_{\varphi_{\varepsilon}})\omega_{\varepsilon,\varphi_{\varepsilon}}^{n}+\int_{Y}\log(v(m_{\varphi_{\varepsilon}}))\omega_{\varepsilon,\varphi_{\varepsilon}}^{n}\right|\leq C_{v}^{\prime}C^{\prime}+C_{v}^{\prime},

where Cv>0C^{\prime}_{v}>0 depend only on CvC_{v}; hence, we obtain a uniform upper bound for 𝐇ε,ωYn(φε)\mathbf{H}_{\varepsilon,\omega_{Y}^{n}}(\varphi_{\varepsilon}) as required.

By Theorem 2.4, we obtained uniform LL^{\infty} and local 𝒞\mathcal{C}^{\ell}-estimates for φεsupYφε\varphi_{\varepsilon}-\sup_{Y}\varphi_{\varepsilon}. The Arzelà–Ascoli theorem shows that there is a subsequence φεsupYφε\varphi_{\varepsilon}-\sup_{Y}\varphi_{\varepsilon} converging in 𝒞oc(YE)\mathcal{C}_{\mathrm{\ell oc}}^{\infty}(Y\setminus E) to φ0PSH(Y,πω)L(Y)𝒞(YE)\varphi_{0}\in\operatorname{PSH}(Y,\pi^{*}\omega)\cap L^{\infty}(Y)\cap\mathcal{C}^{\infty}(Y\setminus E) which satisfies (v,wext)(v,w\ell^{\operatorname{ext}})-cscK equations on YEY\setminus E. This deduces the existence of a singular weighted cscK metric ω+ddcψ0\omega+dd^{c}\psi_{0} in {ω}\{\omega\} on XX, where ψ0PSH(X,ω)L(X)𝒞(Xreg)\psi_{0}\in\operatorname{PSH}(X,\omega)\cap L^{\infty}(X)\cap\mathcal{C}^{\infty}(X^{\mathrm{reg}}) and φ0=πψ0\varphi_{0}=\pi^{*}\psi_{0}.

Part 2: Minimizer. It remains to show that ψ0\psi_{0} is a minimizer of 𝐌Xre\mathbf{M}_{X}^{\rm re\ell} on ω1,T\mathcal{E}^{1,T}_{\omega}. Note that φε\varphi_{\varepsilon} are uniformly bounded and converging locally smoothly to πψ0\pi^{\ast}\psi_{0} on YEY\setminus E. Once can derive 𝐄ε,vwext(φε)𝐄vwext(πψ0)\mathbf{E}_{\varepsilon,vw\ell^{\operatorname{ext}}}(\varphi_{\varepsilon})\to\mathbf{E}_{vw\ell^{\operatorname{ext}}}(\pi^{\ast}\psi_{0}) as ε0\varepsilon\to 0. Thus, by (3.8) with a=1a=1 and by the semi-continuity with respect to strong convergence of 𝐇^ε,v\widehat{\mathbf{H}}_{\varepsilon,v} and 𝐄ε,vRic^\mathbf{E}_{\varepsilon,v}^{-\widehat{\operatorname{Ric}}} (cf. [BJT24, Lem. 4.22]) , one gets

lim infε0𝐌εre(φε)𝐇^v(πψ0)+𝐄vRic^(πψ0)+𝐄vwext(πψ0),\liminf_{\varepsilon\to 0}\mathbf{M}_{\varepsilon}^{\rm re\ell}(\varphi_{\varepsilon})\geq\widehat{\mathbf{H}}_{v}(\pi^{*}\psi_{0})+\mathbf{E}^{-\widehat{\operatorname{Ric}}}_{v}(\pi^{*}\psi_{0})+\mathbf{E}_{vw\ell^{\operatorname{ext}}}(\pi^{*}\psi_{0}),

where we used the fact that YK1ρ1MAε,v(0)YK1ρ1v(mπω)(πω)n=0\int_{Y}K_{1}\rho_{1}\operatorname{MA}_{\varepsilon,v}(0)\rightarrow\int_{Y}K_{1}\rho_{1}v(m_{\pi^{*}\omega})(\pi^{*}\omega)^{n}=0. It follows from [BJT24, Lem. 4.26] that

𝐇^v(πψ0)+𝐄vRic^(πψ0)𝐇μX(ψ0)+𝐄vRic(μX)(ψ0).\widehat{\mathbf{H}}_{v}(\pi^{*}\psi_{0})+\mathbf{E}^{-\widehat{\operatorname{Ric}}}_{v}(\pi^{*}\psi_{0})\geq\mathbf{H}_{\mu_{X}}(\psi_{0})+\mathbf{E}^{-{\operatorname{Ric}(\mu_{X})}}_{v}(\psi_{0}).

This shows that

lim infε0𝐌εre(φε)𝐌Xre(ψ0).\liminf_{\varepsilon\to 0}\mathbf{M}_{\varepsilon}^{\rm re\ell}(\varphi_{\varepsilon})\geq\mathbf{M}_{X}^{\rm re\ell}(\psi_{0}). (3.10)

Fix an arbitrary uω1,Tu\in\mathcal{E}^{1,T}_{\omega}. Without loss of generality, assume that 𝐇v,μX(u)<+\mathbf{H}_{v,\mu_{X}}(u)<+\infty and denote by f=ωun/ωnf=\omega^{n}_{u}/\omega^{n} which is TT-invariant. By [PTT23, Lem. 3.4], there exists 0fj𝒞(X)0\leq f^{j}\in\mathcal{C}^{\infty}(X) converging to ff in LχL^{\chi} as j0j\rightarrow 0 with χ(s):=(s+1)log(s+1)s\chi(s):=(s+1)\log(s+1)-s. Consider fT,j(x):=Tfj(σx)𝑑μ(σ)f^{T,j}(x):=\int_{T}f^{j}(\sigma\cdot x)d\mu(\sigma), where μ\mu is the Haar measure on TT normalized by T𝑑μ(σ)=1\int_{T}d\mu(\sigma)=1. We claim that fT,j𝒞(X)f^{T,j}\in\mathcal{C}^{\infty}(X) also converges to ff in LχL^{\chi}. Indeed, since χ\chi is convex,

χ(|fT,jf|(x))=χ(|T(fT,jf)(σx)𝑑μ(σ)|)Tχ(|fjf|(σx))𝑑μ(σ).\chi(|f^{T,j}-f|(x))=\chi\left(\left|\int_{T}(f^{T,j}-f)(\sigma\cdot x)d\mu(\sigma)\right|\right)\leq\int_{T}\chi(|f^{j}-f|(\sigma\cdot x))d\mu(\sigma).

By Fubini’s theorem and TT-invariance of ω\omega, we obtain that

xXχ(|fj,Tf|(x))ωn(x)xX(σTχ(|fjf|(σx))𝑑μ(σ))ωn(x)\displaystyle\int_{x\in X}\chi(|f^{j,T}-f|(x))\omega^{n}(x)\leq\int_{x\in X}\left(\int_{\sigma\in T}\chi(|f^{j}-f|(\sigma\cdot x))d\mu(\sigma)\right)\omega^{n}(x)
=σT(xXχ(|fjf|(σx))ωn(x))𝑑μ(σ)=σT(xXχ(|fjf|(σx))ωn(σx))𝑑μ(σ).\displaystyle=\int_{\sigma\in T}\left(\int_{x\in X}\chi(|f^{j}-f|(\sigma\cdot x))\omega^{n}(x)\right)d\mu(\sigma)=\int_{\sigma\in T}\left(\int_{x\in X}\chi(|f^{j}-f|(\sigma\cdot x))\omega^{n}(\sigma\cdot x)\right)d\mu(\sigma).

Hence, fT,jf^{T,j} converges to ff in LχL^{\chi}. Consider uj𝒞(Xreg)L(X)u_{j}\in\mathcal{C}^{\infty}(X^{\mathrm{reg}})\cap L^{\infty}(X) solving

(ω+ddcuj)n=cjfT,jωnwithsupXuj=0(\omega+dd^{c}u_{j})^{n}=c_{j}f^{T,j}\omega^{n}\quad\text{with}\quad\sup_{X}u_{j}=0

where cj>0c_{j}>0 is a normalizing constant. Moreover, it follows from [CC24, Thm. 1.3] that uju_{j} is continuous on XX. By [PTT23, Lem. 3.4], uju_{j} converges strongly to uu, and 𝐇μX(uj)𝐇μX(u)\mathbf{H}_{\mu_{X}}(u_{j})\rightarrow\mathbf{H}_{\mu_{X}}(u); thus, 𝐌v,wext(uj)𝐌v,wext(u)\mathbf{M}_{v,w\ell^{\operatorname{ext}}}(u_{j})\to\mathbf{M}_{v,w\ell^{\operatorname{ext}}}(u).

Then for each jj fixed, one can find a family of functions uj,εPSH(Y,ωε)u_{j,\varepsilon}\in\operatorname{PSH}(Y,\omega_{\varepsilon}) such that

  • (uj,ε)ε(0,1)(u_{j,\varepsilon})_{\varepsilon\in(0,1)} are uniformly bounded and continuous;

  • uj,εu_{j,\varepsilon} converges locally smoothly on YEY\setminus E as ε0\varepsilon\to 0;

  • uj,εu_{j,\varepsilon} decreases to πuj\pi^{\ast}u_{j} as ε0\varepsilon\to 0.

Since πuj\pi^{*}u_{j} is continuous on YY, Dini’s theorem implies uj,εu_{j,\varepsilon} converges uniformly to πuj\pi^{*}u_{j} as ε0\varepsilon\rightarrow 0. Indeed, for a fixed jj, consider uj,εu_{j,\varepsilon} the unique solution to the following equation

(ωε+ddcuj,ε)n=euj,επujπ(cjfT,jωn).(\omega_{\varepsilon}+dd^{c}u_{j,\varepsilon})^{n}=e^{u_{j,\varepsilon}-\pi^{\ast}u_{j}}\pi^{*}(c_{j}f^{T,j}\omega^{n}).

Then by [EGZ09], one has a uniform 𝒞0\mathcal{C}^{0}-estimate for uj,εu_{j,\varepsilon} on YY, local 𝒞2\mathcal{C}^{2}-estimate for uj,εu_{j,\varepsilon} away from EE, and uj,εu_{j,\varepsilon} converges locally smoothly to πuj\pi^{*}u_{j} in YEY\setminus E. We now check that uj,εu_{j,\varepsilon} is decreasing as ε0\varepsilon\to 0. For 0<ε1<ε20<\varepsilon_{1}<\varepsilon_{2}, we have

(ωε2+ddcuj,ε1)n(ωε1+ddcuj,ε1)n=euj,ε1πujπ(cjfT,jωn),(\omega_{\varepsilon_{2}}+dd^{c}u_{j,\varepsilon_{1}})^{n}\geq(\omega_{\varepsilon_{1}}+dd^{c}u_{j,\varepsilon_{1}})^{n}=e^{u_{j,\varepsilon_{1}}-\pi^{\ast}u_{j}}\pi^{*}(c_{j}f^{T,j}\omega^{n}),

so uj,ε1u_{j,\varepsilon_{1}} is a subsolution to the equation

(ωε2+ddcφ)=eφeπujπ(cjfT,jωn).(\omega_{\varepsilon_{2}}+dd^{c}\varphi)=e^{\varphi}e^{-\pi^{\ast}u_{j}}\pi^{*}(c_{j}f^{T,j}\omega^{n}).

Thus, we obtain that uj,ε1uj,ε2u_{j,\varepsilon_{1}}\leq u_{j,\varepsilon_{2}} for any 0<ε1<ε20<\varepsilon_{1}<\varepsilon_{2}.

Combining with (3.10), we get

𝐌Xre(ψ0)lim infε0𝐌εre(φε)lim infε0𝐌εre(uε,j)=𝐌Xre(uj),\mathbf{M}_{X}^{\rm re\ell}(\psi_{0})\leq\liminf_{\varepsilon\to 0}\mathbf{M}_{\varepsilon}^{\rm re\ell}(\varphi_{\varepsilon})\leq\liminf_{\varepsilon\to 0}\mathbf{M}_{\varepsilon}^{\rm re\ell}(u_{\varepsilon,j})=\mathbf{M}_{X}^{\rm re\ell}(u_{j}),

where the second inequality follows from the fact that φε\varphi_{\varepsilon} is a minimizer for 𝐌ε,v,wext\mathbf{M}_{\varepsilon,v,w\ell^{\operatorname{ext}}} and the last equality follows from Lemma 3.4 below. Letting j+j\rightarrow+\infty, we obtain that 𝐌v,wext(ψ0)𝐌v,wext(u)\mathbf{M}_{v,w\ell^{\operatorname{ext}}}(\psi_{0})\leq\mathbf{M}_{v,w\ell^{\operatorname{ext}}}(u) for arbitrary uω1,Tu\in\mathcal{E}^{1,T}_{\omega} and this finishes the proof. ∎

Lemma 3.4.

Consider uPSH(X,ω)T𝒞0(X)u\in\operatorname{PSH}(X,\omega)^{T}\cap\mathcal{C}^{0}(X) such that uu is smooth on XregX^{\mathrm{reg}} and 𝐇μX(u)<+\mathbf{H}_{\mu_{X}}(u)<+\infty. Assume that uεPSH(Y,ωε)T𝒞0(Y)𝒞(YE)u_{\varepsilon}\in\operatorname{PSH}(Y,\omega_{\varepsilon})^{T}\cap\mathcal{C}^{0}(Y)\cap\mathcal{C}^{\infty}(Y\setminus E) converges smoothly to πu\pi^{*}u in YEY\setminus E, uniformly on YY and 𝐇ε,ωYn(uε)𝐇ωYn(πu)\mathbf{H}_{\varepsilon,\omega^{n}_{Y}}(u_{\varepsilon})\rightarrow\mathbf{H}_{\omega^{n}_{Y}}(\pi^{*}u) as ε0\varepsilon\to 0. Then 𝐌εre(uε)𝐌Xre(u)\mathbf{M}_{\varepsilon}^{\rm re\ell}(u_{\varepsilon})\rightarrow\mathbf{M}^{\rm re\ell}_{X}(u) as ε0\varepsilon\rightarrow 0.

Proof.

The proof follows a similar approach to that of [BJT24, Prop. 4.16] where the authors assume uu to be smooth on XX instead of XregX^{\mathrm{reg}}. Here, we only provide a brief outline and indicate the necessary modifications.

Set D=KYπKX=iaiEiD=K_{Y}-\pi^{*}K_{X}=\sum_{i}a_{i}E_{i} and define ρY/X=logπμXωYn\rho_{Y/X}=\log\frac{\pi^{\ast}\mu_{X}}{\omega_{Y}^{n}}. Then we have

[D]ddcρY/X=πRic(μX)Ric(ωYn).[D]-dd^{c}\rho_{Y/X}=\pi^{*}\operatorname{Ric}(\mu_{X})-\operatorname{Ric}(\omega_{Y}^{n}).

For uPSH(X,ω)L(X)u\in\operatorname{PSH}(X,\omega)\cap L^{\infty}(X) and any smooth (1,1)(1,1)-form θ\theta (or positive closed (1,1)(1,1)-current),

𝐄πωθ(πu):=j=0n1Y(πu)θ(πω+ddcπu)j(πω)n1j.\mathbf{E}^{\theta}_{\pi^{*}\omega}(\pi^{*}u):=\sum_{j=0}^{n-1}\int_{Y}(\pi^{*}u)\theta\wedge(\pi^{*}\omega+dd^{c}\pi^{*}u)^{j}\wedge(\pi^{*}\omega)^{n-1-j}.

Note that 𝐄πω[D](πu)=0\mathbf{E}^{[D]}_{\pi^{*}\omega}(\pi^{*}u)=0 for any uPSH(X,ω)L(X)u\in\operatorname{PSH}(X,\omega)\cap L^{\infty}(X). Following the same argument in [BJT24, Lem. 4.15], one can infer that for any uPSH(X,ω)L(X)u\in\operatorname{PSH}(X,\omega)\cap L^{\infty}(X) and smooth on XregX^{\mathrm{reg}} we have

(𝐇ωYn,v(πu)+𝐄πωRic(ωY)(πu))(𝐇μX,v(u)+𝐄ωRic(μX)(u))=𝐄πω[D](πu)=0;(\mathbf{H}_{\omega_{Y}^{n},v}(\pi^{*}u)+\mathbf{E}^{-\operatorname{Ric}(\omega_{Y})}_{\pi^{*}\omega}(\pi^{*}u))-(\mathbf{H}_{\mu_{X},v}(u)+\mathbf{E}^{-\operatorname{Ric}(\mu_{X})}_{\omega}(u))=\mathbf{E}^{[D]}_{\pi^{*}\omega}(\pi^{*}u)=0;

therefore,

𝐇ωYn,v(πu)+𝐄πωRic(ωY)(πu)=𝐇μX,v(u)+𝐄ωRic(μX)(u).\mathbf{H}_{\omega_{Y}^{n},v}(\pi^{*}u)+\mathbf{E}^{-\operatorname{Ric}(\omega_{Y})}_{\pi^{*}\omega}(\pi^{*}u)=\mathbf{H}_{\mu_{X},v}(u)+\mathbf{E}^{-\operatorname{Ric}(\mu_{X})}_{\omega}(u). (3.11)

Since uεu_{\varepsilon} converges uniformly to πu\pi^{*}u, we have 𝐄ωεRic(ωY)(uε)𝐄πωRic(ωY)(πu)\mathbf{E}_{\omega_{\varepsilon}}^{-\operatorname{Ric}(\omega_{Y})}(u_{\varepsilon})\to\mathbf{E}_{\pi^{*}\omega}^{-\operatorname{Ric}(\omega_{Y})}(\pi^{*}u) as ε0\varepsilon\to 0. Moreover, since εextext\ell^{\operatorname{ext}}_{\varepsilon}\rightarrow\ell^{\operatorname{ext}} in 𝔱\mathfrak{t}\oplus\mathbb{R}, we have 𝐄ε,vwext(uε)𝐄ω,vwext(u)\mathbf{E}_{\varepsilon,vw\ell^{\operatorname{ext}}}(u_{\varepsilon})\to\mathbf{E}_{\omega,vw\ell^{\operatorname{ext}}}(u). The hypothesis 𝐇ε,ωYn(uε)𝐇ωYn(πu)\mathbf{H}_{\varepsilon,\omega_{Y}^{n}}(u_{\varepsilon})\rightarrow\mathbf{H}_{\omega_{Y}^{n}}(\pi^{*}u) implies 𝐇ε,ωYn,v(uε)𝐇ωYn,v(πu)\mathbf{H}_{\varepsilon,\omega_{Y}^{n},v}(u_{\varepsilon})\rightarrow\mathbf{H}_{\omega_{Y}^{n},v}(\pi^{*}u) by generalized dominated convergent theorem and uεπuu_{\varepsilon}\to\pi^{\ast}u locally smoothly on YEY\setminus E. All in all, these yield

limε0𝐌εre(uε)\displaystyle\lim_{\varepsilon\to 0}\mathbf{M}_{\varepsilon}^{\rm re\ell}(u_{\varepsilon}) =𝐇ωYn,v(πu)+𝐄πωRic(ωY)(πu)+𝐄ω,vwext(u)\displaystyle=\mathbf{H}_{\omega_{Y}^{n},v}(\pi^{*}u)+\mathbf{E}_{\pi^{*}\omega}^{-\operatorname{Ric}(\omega_{Y})}(\pi^{*}u)+\mathbf{E}_{\omega,vw\ell^{\operatorname{ext}}}(u)
=𝐇μX,v(u)+𝐄ωRic(μX)(u)+𝐄ω,vwext(u)=𝐌Xre(u),\displaystyle=\mathbf{H}_{\mu_{X},v}(u)+\mathbf{E}^{-\operatorname{Ric}(\mu_{X})}_{\omega}(u)+\mathbf{E}_{\omega,vw\ell^{\operatorname{ext}}}(u)=\mathbf{M}_{X}^{\rm re\ell}(u),

as required, where the second equality follows from (3.11). ∎

3.2.1. Singular cscK and extremal metrics

We begin by considering the problem of finding a singular cscK metric.

Proof of Corollary B.

Take TAutred(X)T\subset\operatorname{Aut}_{\operatorname{red}}(X) to be a maximal torus and π:YX\pi:Y\rightarrow X to be the TT-equivariant resolution of singularities in Condition (A). For cscK metrics, we have v=1v=1, w=1w=1, and ext=s¯\ell^{\operatorname{ext}}=\bar{s} where s¯:=nVωc1(X){ω}n1\bar{s}:=\frac{n}{V_{\omega}}c_{1}(X)\cdot\{\omega\}^{n-1}, so 𝐌ω=𝐌v,ext\mathbf{M}_{\omega}=\mathbf{M}_{v,\ell^{\operatorname{ext}}}. By the hypothesis, 𝐌v,ext\mathbf{M}_{v,\ell^{\operatorname{ext}}} is TT_{\mathbb{C}}-coercive, the existence of a singular cscK metric on XX now follows from Theorem 3.3. ∎

Let XX be a normal compact Kähler variety with log terminal singularities. Fix a Kähler class α\alpha and take a maximal torus TAutred(X)T\subset\operatorname{Aut}_{\operatorname{red}}(X). Let π:YX\pi:Y\rightarrow X be a TT-equivariant resolution of singularities. We now consider the problem of the existence of singular extremal metrics in the Kähler class α\alpha.

Let ξext\xi_{\operatorname{ext}} be the extremal vector field defined by TT and α\alpha (cf. Section 1.2.6) and let ωα\omega\in\alpha be a TT-equivariant Kähler metric. A singular extremal metric in this setting is defined as a positive current of the form πω+ddcφ\pi^{*}\omega+dd^{c}\varphi where φPSH(Y,πω)L(Y)\varphi\in\operatorname{PSH}(Y,\pi^{\ast}\omega)\cap L^{\infty}(Y) and φ\varphi is smooth away from Exc(π)\operatorname{Exc}(\pi). Additionally, πω+ddcφ\pi^{*}\omega+dd^{c}\varphi is a genuine (1,ext)(1,\ell^{\operatorname{ext}})-cscK metric on YExc(π)Y\setminus\operatorname{Exc}(\pi) where ext(p)=ξext,p+s¯\ell^{\operatorname{ext}}(p)=\langle\xi_{\operatorname{ext}},p\rangle+\bar{s} , where s¯:=nVωc1(X){ω}n1\bar{s}:=\frac{n}{V_{\omega}}c_{1}(X)\cdot\{\omega\}^{n-1}.

Under Condition (A), Theorem 3.3 implies the following existence result of singular extremal metrics:

Theorem 3.5.

Under the above setting, moreover, assume that XX satisfies Condition (A). If the weighted Mabuchi functional 𝐌ω,v,ext\mathbf{M}_{\omega,v,\ell^{\operatorname{ext}}} on XX is TT_{\mathbb{C}}-coercive, then XX admits a singular extremal metric in α\alpha.

3.3. Constructing examples of singular cscK metrics

We shall give a way to construct examples of singular cscK metric in the spirit of Arezzo–Pacard [AP06] (see also [AP09, APS11, Szé15]) and use variational argument and our existence result. Before illustrating the process, we need the following lemma:

Lemma 3.6.

Let (X,ω)(X,\omega) be a compact Kähler variety with log terminal singularities and let f:YXf:Y\to X be a blowup along a compact submanifold SXregS\subset X^{\mathrm{reg}} of codimension 2\geq 2. Consider ωY\omega_{Y} a Kähler metric on YY and ωε=fω+εωY\omega_{\varepsilon}=f^{\ast}\omega+\varepsilon\omega_{Y} for ε[0,1]\varepsilon\in[0,1]. If 𝐌ω\mathbf{M}_{\omega} is coercive, then 𝐌ωε\mathbf{M}_{\omega_{\varepsilon}} is coercive for all sufficiently small ε>0\varepsilon>0.

Proof.

The proof follows the same strategy in [PTT23, Thm. 4.11] and [BJT24, Thm. A]. Since ff is a blowup of SS, for mm\in\mathbb{N}^{\ast} so that mKYmK_{Y} and mKXmK_{X} are both \mathbb{Q}-Cartier, there is a smooth hermitain metric hYh_{Y} of mKYmK_{Y} such that

1mΘ(mKY,hY)Dfω-\frac{1}{m}\Theta(mK_{Y},h_{Y})\geq-Df^{\ast}\omega (3.12)

for some constant D0D\geq 0. Denote by ΘY=1mΘ(mKY,hY)\Theta_{Y}=\frac{1}{m}\Theta(mK_{Y},h_{Y}) and ΘX=1mΘ(mKX,hX)\Theta_{X}=\frac{1}{m}\Theta(mK_{X},h_{X}) for some hermitian metric hXh_{X} on mKXmK_{X}. We also set μY\mu_{Y} (resp. μX\mu_{X}) to be the corresponding probability measure of hYh_{Y} (resp. hXh_{X}).

Recall that from [PTT23, Sec. 4],

𝐌ωε=𝐇μY,ωε+s¯ε𝐄ωεn𝐄ΘY,ωεCεon 1(Y,ωε)\mathbf{M}_{\omega_{\varepsilon}}=\mathbf{H}_{\mu_{Y},\omega_{\varepsilon}}+\bar{s}_{\varepsilon}\mathbf{E}_{\omega_{\varepsilon}}-n\mathbf{E}_{\Theta_{Y},\omega_{\varepsilon}}-C_{\varepsilon}\quad\text{on }\mathcal{E}^{1}(Y,\omega_{\varepsilon})

and

𝐌ω=𝐇μX,ω+s¯𝐄ωn𝐄ΘX,ωCon 1(X,ω)\mathbf{M}_{\omega}=\mathbf{H}_{\mu_{X},\omega}+\bar{s}\mathbf{E}_{\omega}-n\mathbf{E}_{\Theta_{X},\omega}-C\quad\text{on }\mathcal{E}^{1}(X,\omega)

where s¯ε=nc1(Y)[ωε]n1[ωε]n\bar{s}_{\varepsilon}=\frac{nc_{1}(Y)\cdot[\omega_{\varepsilon}]^{n-1}}{[\omega_{\varepsilon}]^{n}}, s¯=nc1(X)[ω]n1[ω]n\bar{s}=\frac{nc_{1}(X)\cdot[\omega]^{n-1}}{[\omega]^{n}}, Cε=𝐇μY,ωε(0)=Ylog(ωεnμY)μYC_{\varepsilon}=\mathbf{H}_{\mu_{Y},\omega_{\varepsilon}}(0)=\int_{Y}\log\left(\frac{\omega_{\varepsilon}^{n}}{\mu_{Y}}\right)\mu_{Y} and C=𝐇μX,ωε(0)=Xlog(ωnμX)μXC=\mathbf{H}_{\mu_{X},\omega_{\varepsilon}}(0)=\int_{X}\log\left(\frac{\omega^{n}}{\mu_{X}}\right)\mu_{X}. Let A0,B0>0A_{0},B_{0}>0 be two constant such that 𝐌ωA0(𝐄ω)B0\mathbf{M}_{\omega}\geq A_{0}(-\mathbf{E}_{\omega})-B_{0} on norm1(X,ω)\mathcal{E}^{1}_{{\mathrm{norm}}}(X,\omega). We claim that for all 0<A<A00<A<A_{0} there are ε0>0\varepsilon_{0}>0 and B>0B>0 such that for all ε(0,ε0)\varepsilon\in(0,\varepsilon_{0}),

𝐌ωεA(𝐄ωε)B\mathbf{M}_{\omega_{\varepsilon}}\geq A(-\mathbf{E}_{\omega_{\varepsilon}})-B

s on norm1(Y,ωε)\mathcal{E}^{1}_{{\mathrm{norm}}}(Y,\omega_{\varepsilon}).

Suppose otherwise, for an A(0,A0)A\in(0,A_{0}), there are εk0\varepsilon_{k}\to 0, Bk+B_{k}\to+\infty, as k+k\to+\infty, and uknorm1(Y,ωεk)u_{k}\in\mathcal{E}^{1}_{{\mathrm{norm}}}(Y,\omega_{\varepsilon_{k}}) such that

𝐌ωεk(uk)<A(𝐄ωεk(uk))Bk.\mathbf{M}_{\omega_{\varepsilon_{k}}}(u_{k})<A(-\mathbf{E}_{\omega_{\varepsilon_{k}}}(u_{k}))-B_{k}.

Without loss of generality, one can assume that uku_{k} is bounded. Otherwise, by [PTT23, Lem. 3.4], one can find a sequence bounded ωk\omega_{k}-psh functions (vk,j)j(v_{k,j})_{j} converging strongly to uku_{k} and their entropies also converge to 𝐇μY,ωεk(uk)\mathbf{H}_{\mu_{Y},\omega_{\varepsilon_{k}}}(u_{k}). Then for sufficiently large jj, we have 𝐌ωεk(vk,j)<A(𝐄ωεk(vk,j))Bk\mathbf{M}_{\omega_{\varepsilon_{k}}}(v_{k,j})<A(-\mathbf{E}_{\omega_{\varepsilon_{k}}}(v_{k,j}))-B_{k}.

Note that from (3.12), for all ψnorm1(Y,ωε)\psi\in\mathcal{E}_{{\mathrm{norm}}}^{1}(Y,\omega_{\varepsilon}),

n𝐄ΘY,ωε(ψ)\displaystyle-n\mathbf{E}_{\Theta_{Y},\omega_{\varepsilon}}(\psi) =1Vεj=0n1YψΘYωε,ψjωεn1j\displaystyle=-\frac{1}{V_{\varepsilon}}\sum_{j=0}^{n-1}\int_{Y}\psi\Theta_{Y}\wedge\omega_{\varepsilon,\psi}^{j}\wedge\omega_{\varepsilon}^{n-1-j}
1Vεj=0n1YψDfωωε,ψjωεn1j(n+1)D𝐄ωε(ψ).\displaystyle\geq\frac{1}{V_{\varepsilon}}\sum_{j=0}^{n-1}\int_{Y}\psi Df^{\ast}\omega\wedge\omega_{\varepsilon,\psi}^{j}\wedge\omega_{\varepsilon}^{n-1-j}\geq(n+1)D\mathbf{E}_{\omega_{\varepsilon}}(\psi).

Hence,

𝐇μY,ωε(uk)+(s¯εk+(n+1)D)𝐄ωεk(uk)𝐌ωεk(uk)<A(𝐄ωεk(uk))Bk.\mathbf{H}_{\mu_{Y},\omega_{\varepsilon}}(u_{k})+(\bar{s}_{\varepsilon_{k}}+(n+1)D)\mathbf{E}_{\omega_{\varepsilon_{k}}}(u_{k})\leq\mathbf{M}_{\omega_{\varepsilon_{k}}}(u_{k})<A(-\mathbf{E}_{\omega_{\varepsilon_{k}}}(u_{k}))-B_{k}.

After enlarging DD, one may assume that for all ε[0,1]\varepsilon\in[0,1], (n+1)D+As¯ε>δ(n+1)D+A-\bar{s}_{\varepsilon}>\delta for a uniform δ>0\delta>0. Then we have Bk<((n+1)D+As¯ε)(𝐄ωεk(uk))B_{k}<((n+1)D+A-\bar{s}_{\varepsilon})(-\mathbf{E}_{\omega_{\varepsilon_{k}}}(u_{k})) and this implies that dk:=𝐄ωεk(uk)+d_{k}:=-\mathbf{E}_{\omega_{\varepsilon_{k}}}(u_{k})\to+\infty as k+k\to+\infty. Let gk(s)g_{k}(s) be the d1d_{1}-geodesic connecting 0 and uku_{k} in norm1(Y,ωεk)\mathcal{E}^{1}_{{\mathrm{norm}}}(Y,\omega_{\varepsilon_{k}}). Fix a constant R>0R>0 and define vk:=gk(R)v_{k}:=g_{k}(R). By the convexity of Mabuchi functional [PTT23, Prop. 4.7],

𝐌ωεk(vk)dkRdk𝐌ωεk(0)+Rdk𝐌ωεk(uk)Rdk(AdkBk)AR.\mathbf{M}_{\omega_{\varepsilon_{k}}}(v_{k})\leq\frac{d_{k}-R}{d_{k}}\mathbf{M}_{\omega_{\varepsilon_{k}}}(0)+\frac{R}{d_{k}}\mathbf{M}_{\omega_{\varepsilon_{k}}}(u_{k})\leq\frac{R}{d_{k}}(Ad_{k}-B_{k})\leq AR.

From the expression of Mabuchi functional, we have 𝐇μY,ωεk(vk)((n+1)D+As¯εk)R\mathbf{H}_{\mu_{Y},\omega_{\varepsilon_{k}}}(v_{k})\leq((n+1)D+A-\bar{s}_{\varepsilon_{k}})R.

In the argument below, although YY is singular, corresponding proofs in [BJT24] proceed exactly the same. By the strong compactness [BJT24, Thm. 2.10], up to a subsequence, vkv_{k} converges in v0norm1(Y,fω)v_{0}\in\mathcal{E}^{1}_{{\mathrm{norm}}}(Y,f^{\ast}\omega) and 𝐄ωεk(vk)𝐄fω(v0)\mathbf{E}_{\omega_{\varepsilon_{k}}}(v_{k})\to\mathbf{E}_{f^{\ast}\omega}(v_{0}) as k+k\to+\infty. Note that vv can descend to a function in norm1(X,ω)\mathcal{E}^{1}_{{\mathrm{norm}}}(X,\omega), which we still denote by vv. By [BJT24, Lem. 4.16], 𝐇μX,ω(v)n𝐄ΘX,ω(v)𝐇μY,fω(v)n𝐄ΘY,fω(v)\mathbf{H}_{\mu_{X},\omega}(v)-n\mathbf{E}_{\Theta_{X},\omega}(v)\leq\mathbf{H}_{\mu_{Y},f^{\ast}\omega}(v)-n\mathbf{E}_{\Theta_{Y},f^{\ast}\omega}(v) and [BJT24, Lem. 4.6 and Lem. 4.9] shows that 𝐌ω0(v0)lim infk+𝐌ωεk(vk)\mathbf{M}_{\omega_{0}}(v_{0})\leq\liminf_{k\to+\infty}\mathbf{M}_{\omega_{\varepsilon_{k}}}(v_{k}). All in all, we obtain

A0RB0𝐌ω(v0)lim infk+𝐌ωεk(vk)AR.A_{0}R-B_{0}\leq\mathbf{M}_{\omega}(v_{0})\leq\liminf_{k\to+\infty}\mathbf{M}_{\omega_{\varepsilon_{k}}}(v_{k})\leq AR.

Letting R=B0A0A+1R=\frac{B_{0}}{A_{0}-A}+1, this yields a contradiction. ∎

3.3.1. Construct singular cscK on blowups of singular KEs

Let (X,ω)(X,\omega) be a compact Kähler variety with log-terminal singularity. Suppose that KXK_{X} is mm-Cartier for some mm\in\mathbb{N}^{\ast}, and pick a hermitian metric hh on mKXmK_{X}. Assume that either

  • KXK_{X} is ample and ω=imΘ(mKX,h)\omega=\frac{\mathrm{i}}{m}\Theta(mK_{X},h); or

  • KXK_{X} numerically trivial and imΘ(mKX,h)=0\frac{\mathrm{i}}{m}\Theta(mK_{X},h)=0; or else

  • KXK_{X} is anti-ample, ω=imΘ(mKX,h)\omega=-\frac{\mathrm{i}}{m}\Theta(mK_{X},h), and (X,KX)(X,K_{X}) is K-stable.

Note that in the above cases, {ω}\{\omega\} contains a unique singular Kähler–Einstein metric and the Mabuchi function with respect to ω\omega is coercive. We further assume that XX admits a resolution of Fano type π:X^X\pi:\widehat{X}\to X.

Let f:YXf:Y\to X be a blowup of NN distinct points in the smooth locus of XX. Denote the irreducible components of the exceptional divisor of ff by (Di)i(D_{i})_{i}. Since ff is an isomorphism near the singularities of YY and XX, one can verify that YY also admits a resolution of Fano type.

For any a:=(a1,,aN)>0na:=(a_{1},\cdots,a_{N})\in\mathbb{R}_{>0}^{n}, there exists a constant δa>0\delta_{a}>0, such that for all δ(0,δa)\delta\in(0,\delta_{a}), the class {fω}δi=1Naic1(𝒪(Di))\{f^{\ast}\omega\}-\delta\sum_{i=1}^{N}a_{i}c_{1}(\mathcal{O}(D_{i})) contains a Kähler metric ωY,a,δ\omega_{Y,a,\delta}. By Lemma 3.6, the Mabuchi functional with respect to fω+εωY,a,δf^{\ast}\omega+\varepsilon\omega_{Y,a,\delta} is coercive for all sufficiently small ε\varepsilon. Corollary B then ensures the existence of a singular cscK metric in the class (1+ε){fω}εδi=1Naic1(𝒪(Di))(1+\varepsilon)\{f^{\ast}\omega\}-\varepsilon\delta\sum_{i=1}^{N}a_{i}c_{1}(\mathcal{O}(D_{i})) on YY.

If XX is a Kähler variety with log terminal singularities, Aut(X)\operatorname{Aut}(X) is discrete and it admits a crepant resolution π:YX\pi:Y\to X, it is not difficult to check that π:YX\pi:Y\to X satisfies Condition (A). In the case of Kähler–Einstein varieties with log terminal singularities and discrete automorphism groups that admit crepant resolutions, the above construction provides a method to produce numerous singular cscK metrics on their blowups at points within the smooth locus.

In dimension two, all surfaces with canonical singularities admit crepant resolutions. In dimension three, the famous result of [BKR01, Thm. 1.2] establishes that singular varieties locally modeled on 3/G\mathbb{C}^{3}/G, where G<SL(3,)G<\mathrm{SL}(3,\mathbb{C}) is finite, also admit crepant resolutions. Three-dimensional ODP singularity also admits a crepant resolution, and it is not a quotient singularity.

We also extract the following example from [Szé24, Rmk. 34] and [BJT24, Example 4.10] that is not a crepant resolution.

Example 3.7 (Isolated cone singularities).

Let VV be a smooth projective variety and let LL be an ample line bundle on VV. Set Ca(V,L):=Specm0H0(V,Lm)C_{a}(V,L):=\operatorname{Spec}\sum_{m\geq 0}H^{0}(V,L^{m}) the corresponding affine cone. Assuming KVrLK_{V}\sim_{\mathbb{Q}}r\cdot L for some rr\in\mathbb{Q}, by [Kol13, Lem. 3.1], Ca(V,L)C_{a}(V,L) is klt if and only if r<0r<0. Moreover, Ca(V,L)C_{a}(V,L) is canonical if and only if r1r\leq-1. Therefore, one can choose r(1,0)r\in(-1,0) to get a klt isolated singularity which is not canonical.

Assume that XX has klt isolated singularities, and each singular point is locally isomorphic to an affine cone Ca(V,L)C_{a}(V,L) where VV is a Fano manifold. Then blowing up the singularities yields a resolution of singularities such that π:YX\pi:Y\rightarrow X is projective. In this case, KY=πKXiaiEi-K_{Y}=-\pi^{*}K_{X}-\sum_{i}a_{i}E_{i} is π\pi-nef if and only if XX has canonical singularities, i.e. ai0a_{i}\geq 0. When KY-K_{Y} is π\pi-nef, Condition (A) holds by choosing ρ1=0\rho_{1}=0.

We now consider a slightly more general situation that KY-K_{Y} is not π\pi-nef and explain Condition (A) in this setting. Let μX\mu_{X} and μY\mu_{Y} be two smooth TT-invariant volume forms on XX and YY, respectively. Then define a TT-invariant current Ric(μY)+ddcψ:=πRic(μX)iai[Ei],\operatorname{Ric}(\mu_{Y})+dd^{c}\psi:=\pi^{\ast}\operatorname{Ric}(\mu_{X})-\sum_{i}a_{i}[E_{i}], where ψ=iailog|si|hi2\psi=-\sum_{i}a_{i}\log|s_{i}|^{2}_{h_{i}} for some smooth Hermitian metric hih_{i} on 𝒪X(Ei)\mathcal{O}_{X}(E_{i}). Since EiE_{i} are disjoint in our construction, one has {πω}ai>0aic1(𝒪X(Ei))>0\{\pi^{\ast}\omega\}-\sum_{a_{i}>0}a_{i}c_{1}(\mathcal{O}_{X}(E_{i}))>0. Therefore, there exists a constant K>0K>0 such that ai>0ai[Ei]K{πω}ai>0ddcailog|si|hi2-\sum_{a_{i}>0}a_{i}[E_{i}]\geq-K\{\pi^{\ast}\omega\}-\sum_{a_{i}>0}dd^{c}a_{i}\log|s_{i}|^{2}_{h^{\prime}_{i}} for some smooth hermitian metrics hih^{\prime}_{i} on 𝒪X(Ei)\mathcal{O}_{X}(E_{i}). We then have Ric(μY)+ddc(ψ+ai>0ailog|si|hi2)Kπω\operatorname{Ric}(\mu_{Y})+dd^{c}(\psi+\sum_{a_{i}>0}a_{i}\log|s_{i}|^{2}_{h^{\prime}_{i}})\geq-K\pi^{*}\omega. Set ϕ:=ψ+ai>0ailog|si|hi2=uaj<0ajlog|sj|hj2\phi:=\psi+\sum_{a_{i}>0}a_{i}\log|s_{i}|^{2}_{h_{i}^{\prime}}=u-\sum_{a_{j}<0}a_{j}\log|s_{j}|^{2}_{h_{j}} for some u𝒞(Y)u\in\mathcal{C}^{\infty}(Y). One can obtain XeϕμYn<+\int_{X}e^{-\phi}\mu_{Y}^{n}<+\infty as aj>1a_{j}>-1. Hence, Condition (A) holds by taking ρ1=ϕ/K\rho_{1}=\phi/K.

3.3.2. Mixing construction with smoothing

Let us stress that this construction also works on the \mathbb{Q}-Gorenstein smoothable setting. Consider a \mathbb{Q}-Gorenstein smoothing f:(𝒳,ω𝒳)𝔻f:(\mathcal{X},\omega_{\mathcal{X}})\to\mathbb{D} of (X0,ω𝒳|X0)(X_{0},{\omega_{\mathcal{X}}}_{|X_{0}}) where X0X_{0} is a Kähler–Einstein variety with log terminal singularities and Aut(X)\operatorname{Aut}(X) is discrete, and {ω𝒳|X0}\{{\omega_{\mathcal{X}}}_{|X_{0}}\} is a Kähler–Einstein class. Denote by 𝒵\mathcal{Z} the singular set of 𝒵\mathcal{Z}. Take a finite set of points {p1,,pN}\{p_{1},\cdots,p_{N}\} in X0regX_{0}^{\mathrm{reg}}. There exists smooth curves C1,,CNC_{1},\cdots,C_{N} in 𝒳\mathcal{X} such that these curves are disjoint, each CiC_{i} intersects X0X_{0} transversely at the single point pip_{i}, Ci𝒳𝒵C_{i}\subset\mathcal{X}\setminus\mathcal{Z}, and the restriction f|Ci:Ci𝔻f_{|C_{i}}:C_{i}\to\mathbb{D} is an isomorphism. Now consider the blowup map μ:𝒴𝒳\mu:\mathcal{Y}\to\mathcal{X} along all the curves (Ci)i(C_{i})_{i} and let Y0=B{p1,,pN}XY_{0}=\operatorname{B\ell}_{\{p_{1},\cdots,p_{N}\}}X. Then π=fμ:𝒴𝔻\pi=f\circ\mu:\mathcal{Y}\to\mathbb{D} is a \mathbb{Q}-Gorenstein smoothing of (Y0,ω𝒴,ε|Y0)(Y_{0},{\omega_{\mathcal{Y},\varepsilon}}_{|Y_{0}}) where ω𝒴,ε=μω𝒳+εω𝒴\omega_{\mathcal{Y},\varepsilon}=\mu^{\ast}\omega_{\mathcal{X}}+\varepsilon\omega_{\mathcal{Y}}. Here ω𝒴\omega_{\mathcal{Y}} is a hermitian metric on 𝒴\mathcal{Y} and relatively Kähler that defined by

ω𝒴=μω𝒳i=1NaiΘhi(𝒪(Ei))\omega_{\mathcal{Y}}=\mu^{\ast}\omega_{\mathcal{X}}-\sum_{i=1}^{N}a_{i}\Theta_{h_{i}}(\mathcal{O}(E_{i}))

where EiE_{i} is the exceptional divisor over CiC_{i}, ai>0a_{i}>0 and hih_{i} is some hermitian metric on 𝒪(Ei)\mathcal{O}(E_{i}) such that ω\omega. Since the Mabuchi functional 𝐌ω𝒳|X0\mathbf{M}_{{\omega_{\mathcal{X}}}_{|X_{0}}} on X0X_{0} is coercive, it follows from Lemma 3.6 𝐌ω𝒴,ε|Y0\mathbf{M}_{{\omega_{\mathcal{Y},\varepsilon}}_{|Y_{0}}} on Y0Y_{0} as well. By [PTT23, Thm. C], this implies the existence of a cscK metric in the class {ω𝒴,ε|Y0}\{{\omega_{\mathcal{Y},\varepsilon}}|Y_{0}\}.

For examples of smoothable Calabi–Yau varieties, we refer the reader to [DG18, Sec. 8]. For the smoothable Fano case, [LX19, Liu22] prove that mildly singular cubic varieties in dimensions three and four are K-stable. Additionally, explicit examples of K-stable singular cubic threefolds can be found, for instance, in [CTZ25, Sec. 3, 4] and [CMTZ24, Sec. 5].

Appendix A Weighted Aubin–Yau inequality

In this section, for the reader’s convenience, we provide detailed proof of a Laplacian inequality (see also [DJL24, Lem. 5.6]), which generalizes [Siu87, p. 98–99] to the weighted setting.

Lemma A.1.

Let ω\omega and ωX\omega_{X} be two TT-invariant Kähler metrics. Assume that φωT\varphi\in\mathcal{H}^{T}_{\omega} satisfies

v(mφ)(ω+ddcφ)n=eFωXn.v(m_{\varphi})(\omega+dd^{c}\varphi)^{n}=e^{F}\omega_{X}^{n}.

Then there exist positive constants 0,𝒞0\mathcal{B}_{0},\mathcal{C}_{0} such that

Δωφ,vlogtrωXωφ\displaystyle\Delta_{\omega_{\varphi},v}\log\operatorname{tr}_{\omega_{X}}\omega_{\varphi} ΔωXFtrωXωφ0trωφωX𝒞01trωXωφα,β(logv)αβ(mφ)dmφξα,dmφξβωX.\displaystyle\geq\frac{\Delta_{\omega_{X}}F}{\operatorname{tr}_{\omega_{X}}\omega_{\varphi}}-\mathcal{B}_{0}\operatorname{tr}_{\omega_{\varphi}}\omega_{X}-\mathcal{C}_{0}-\frac{1}{\operatorname{tr}_{\omega_{X}}\omega_{\varphi}}\sum_{\alpha,\beta}(\log v)_{\alpha\beta}(m_{\varphi})\langle dm^{\xi_{\alpha}}_{\varphi},dm^{\xi_{\beta}}_{\varphi}\rangle_{\omega_{X}}.

In particular, if 𝔱plogv(p)\mathfrak{t}^{\vee}\ni p\mapsto\log v(p)\in\mathbb{R} is concave, then

Δωφ,vlogtrωXωφΔωXFtrωXωφ0trωφωX𝒞0.\Delta_{\omega_{\varphi},v}\log\operatorname{tr}_{\omega_{X}}\omega_{\varphi}\geq\frac{\Delta_{\omega_{X}}F}{\operatorname{tr}_{\omega_{X}}\omega_{\varphi}}-\mathcal{B}_{0}\operatorname{tr}_{\omega_{\varphi}}\omega_{X}-\mathcal{C}_{0}.

We note that the constants 0=B+2Cv2CRic\mathcal{B}_{0}=B+2C_{v}^{2}C_{\operatorname{Ric}} and 𝒞0=Cv2(C2+Cv2Cξ)\mathcal{C}_{0}=C_{v}^{2}(C_{2}+C_{v}^{2}C_{\xi}) where

  • Bisec(ωX)B\operatorname{Bisec}(\omega_{X})\geq-B is a negative lower bound for the bisectional curvature of ωX\omega_{X};

  • Cv>0C_{v}>0 is a constant such that

    Cv1v+α|vα|+α,β|vαβ|Cvon P=im(mω),Cv1|v|+α|vα|+α,β|vαβ|Cvon PωX=im(mωX);\begin{split}C_{v}^{-1}&\leq v+\sum_{\alpha}|v_{\alpha}|+\sum_{\alpha,\beta}|v_{\alpha\beta}|\leq C_{v}\quad{\text{on $P=\operatorname{im}(m_{\omega})$}},\\ C_{v}^{-1}&\leq|v|+\sum_{\alpha}|v_{\alpha}|+\sum_{\alpha,\beta}|v_{\alpha\beta}|\leq C_{v}\quad{\text{on $P_{\omega_{X}}=\operatorname{im}(m_{\omega_{X}})$}};\end{split} (A.1)
  • C2C_{2} depending only on nn and C1C_{1} where C1>0C_{1}>0 is a constant so that for any α\alpha,

    C1ωXJξαωXC1ωX;-C_{1}\omega_{X}\leq\mathcal{L}_{J\xi_{\alpha}}\omega_{X}\leq C_{1}\omega_{X}; (A.2)
  • CRic>0C_{\operatorname{Ric}}>0 and Cξ>0C_{\xi}>0 are constants so that mRic(ωX)ξα(X)CRic\|m_{\operatorname{Ric}(\omega_{X})}^{\xi_{\alpha}}(X)\|\leq C_{\operatorname{Ric}} and |ξα|ωX2Cξ|\xi_{\alpha}|_{\omega_{X}}^{2}\leq C_{\xi}, for all α\alpha, respectively, where \|\cdot\| is a norm on the dual Lie algebra of TT.

Remark A.2.

The concavity condition on logv(p)\log v(p) holds in many interested cases. Here we extract some examples from [Lah19, Sec. 3]:

  • cscK and extremal metrics: v(p)=1v(p)=1,

  • Kähler–Ricci solitons: v(p)=eξ,pv(p)=e^{\langle\xi,p\rangle} for some fixed ξ𝔱\xi\in\mathfrak{t},

  • Kähler metrics given by the generalized Calabi construction: v(p)=Πj(ξj,p+aj)djv(p)=\Pi_{j}(\langle\xi_{j},p\rangle+a_{j})^{d_{j}} with dj>0d_{j}>0.

Proof.

Before entering the proof, we recall a basic equality of Lie derivative about the quotient of volume forms. Suppose that α\alpha is an (n,n)(n,n)-form and β\beta is a volume form on XX. For all vector fields VV, one can derive the following

V(αβ)=VαβαβVββ.\mathcal{L}_{V}\left(\frac{\alpha}{\beta}\right)=\frac{\mathcal{L}_{V}\alpha}{\beta}-\frac{\alpha}{\beta}\cdot\frac{\mathcal{L}_{V}\beta}{\beta}.

As a consequence, we have

V(ωφnωXn)=n(Vωφ)ωφn1ωφnn(VωX)ωXn1ωXn\mathcal{L}_{V}\left(\frac{\omega_{\varphi}^{n}}{\omega_{X}^{n}}\right)=\frac{n(\mathcal{L}_{V}\omega_{\varphi})\wedge\omega_{\varphi}^{n-1}}{\omega_{\varphi}^{n}}-\frac{n(\mathcal{L}_{V}\omega_{X})\wedge\omega_{X}^{n-1}}{\omega_{X}^{n}} (A.3)

and

V(trωXωφ)=n(Vωφ)ωXn1+n(n1)(VωX)ωφωXn2ωXn(trωXωφ)n(VωX)ωXn1ωXn\mathcal{L}_{V}\left(\operatorname{tr}_{\omega_{X}}\omega_{\varphi}\right)=\frac{n(\mathcal{L}_{V}\omega_{\varphi})\wedge\omega_{X}^{n-1}+n(n-1)(\mathcal{L}_{V}\omega_{X})\wedge\omega_{\varphi}\wedge\omega_{X}^{n-2}}{\omega_{X}^{n}}-(\operatorname{tr}_{\omega_{X}}\omega_{\varphi})\frac{n(\mathcal{L}_{V}\omega_{X})\wedge\omega_{X}^{n-1}}{\omega_{X}^{n}} (A.4)

By the standard Aubin–Yau’s inequality, we have

ΔωφlogtrωXωφΔωX(Flogv(mφ))trωXωφBtrωφωX,\Delta_{\omega_{\varphi}}\log\operatorname{tr}_{\omega_{X}}\omega_{\varphi}\geq\frac{\Delta_{\omega_{X}}(F-\log v(m_{\varphi}))}{\operatorname{tr}_{\omega_{X}}\omega_{\varphi}}-B\operatorname{tr}_{\omega_{\varphi}}\omega_{X},

and thus,

Δωφ,vlogtrωXωφΔωX(Flogv(mφ))trωXωφBtrωφωX+trωφ[dlogv(mφ)dclogtrωXωφ]=ΔωXFtrωXωφBtrωφωX+1trωXωφ{trωφ[dlogv(mφ)dctrωXωφ]ΔωXlogv(mφ)}.\begin{split}&\Delta_{\omega_{\varphi},v}\log\operatorname{tr}_{\omega_{X}}\omega_{\varphi}\\ &\geq\frac{\Delta_{\omega_{X}}(F-\log v(m_{\varphi}))}{\operatorname{tr}_{\omega_{X}}\omega_{\varphi}}-B\operatorname{tr}_{\omega_{\varphi}}\omega_{X}+\operatorname{tr}_{\omega_{\varphi}}[d\log v(m_{\varphi})\wedge d^{c}\log\operatorname{tr}_{\omega_{X}}\omega_{\varphi}]\\ &=\frac{\Delta_{\omega_{X}}F}{\operatorname{tr}_{\omega_{X}}\omega_{\varphi}}-B\operatorname{tr}_{\omega_{\varphi}}\omega_{X}+\frac{1}{\operatorname{tr}_{\omega_{X}}\omega_{\varphi}}\bigg{\{}\operatorname{tr}_{\omega_{\varphi}}[d\log v(m_{\varphi})\wedge d^{c}\operatorname{tr}_{\omega_{X}}\omega_{\varphi}]-\Delta_{\omega_{X}}\log v(m_{\varphi})\bigg{\}}.\end{split} (A.5)

It suffices to establish a suitable lower bound for the terms in {}\{\cdots\} in (A.5). Note that PP does not depend on φ\varphi and dc=J1dJd^{c}=J^{-1}\circ d\circ J. By Cartan’s formula, iVdcf=iJVdf=JVfi_{V}d^{c}f=-i_{JV}df=-\mathcal{L}_{JV}f. Using (A.4), (A.2) and (A.1), we obtain the following:

trωφ(dlogv(mφ)dctrωXωφ)=α(logv)α(mφ)iξαdctrωXωφ=α(logv)α(mφ)(JξαtrωXωφ)=α(logv)α(mφ)(n(Jξαωφ)ωXn1+n(n1)(JξαωX)ωφωXn2ωXntrωXωφn(JξαωX)ωXn1ωXn)α(logv)α(mφ)(n(Jξαωφ)ωXn1ωXn)C2Cv2trωXωφ=α(logv)α(mφ)ΔωXmφξαC2Cv2trωXωφ=1v(mφ)(ΔωXv(mφ)α,βvαβ(mφ)dmφξα,dmφξβωX)C2Cv2trωXωφ=ΔωXlogv(mφ)α,β(logv)αβ(mφ)dmφξα,dmφξβωXC2Cv2trωXωφ\begin{split}&\operatorname{tr}_{\omega_{\varphi}}(d\log v(m_{\varphi})\wedge d^{c}\operatorname{tr}_{\omega_{X}}\omega_{\varphi})=\sum_{\alpha}(\log v)_{\alpha}(m_{\varphi})i_{\xi_{\alpha}}d^{c}\operatorname{tr}_{\omega_{X}}\omega_{\varphi}=-\sum_{\alpha}(\log v)_{\alpha}(m_{\varphi})(\mathcal{L}_{J\xi_{\alpha}}\operatorname{tr}_{\omega_{X}}\omega_{\varphi})\\ &=-\sum_{\alpha}(\log v)_{\alpha}(m_{\varphi})\left(\frac{n(\mathcal{L}_{J\xi_{\alpha}}\omega_{\varphi})\wedge\omega_{X}^{n-1}+n(n-1)(\mathcal{L}_{J\xi_{\alpha}}\omega_{X})\wedge\omega_{\varphi}\wedge\omega_{X}^{n-2}}{\omega_{X}^{n}}-\operatorname{tr}_{\omega_{X}}\omega_{\varphi}\cdot\frac{n(\mathcal{L}_{J\xi_{\alpha}}\omega_{X})\wedge\omega_{X}^{n-1}}{\omega_{X}^{n}}\right)\\ &\geq-\sum_{\alpha}(\log v)_{\alpha}(m_{\varphi})\left(\frac{n(\mathcal{L}_{J\xi_{\alpha}}\omega_{\varphi})\wedge\omega_{X}^{n-1}}{\omega_{X}^{n}}\right)-C_{2}C_{v}^{2}\operatorname{tr}_{\omega_{X}}\omega_{\varphi}\\ &=\sum_{\alpha}(\log v)_{\alpha}(m_{\varphi})\Delta_{\omega_{X}}m_{\varphi}^{\xi_{\alpha}}-C_{2}C_{v}^{2}\operatorname{tr}_{\omega_{X}}\omega_{\varphi}\\ &=\frac{1}{v(m_{\varphi})}\left(\Delta_{\omega_{X}}v(m_{\varphi})-\sum_{\alpha,\beta}v_{\alpha\beta}(m_{\varphi})\langle dm_{\varphi}^{\xi_{\alpha}},dm_{\varphi}^{\xi_{\beta}}\rangle_{\omega_{X}}\right)-C_{2}C_{v}^{2}\operatorname{tr}_{\omega_{X}}\omega_{\varphi}\\ &=\Delta_{\omega_{X}}\log v(m_{\varphi})-\sum_{\alpha,\beta}(\log v)_{\alpha\beta}(m_{\varphi})\langle dm_{\varphi}^{\xi_{\alpha}},dm_{\varphi}^{\xi_{\beta}}\rangle_{\omega_{X}}-C_{2}C_{v}^{2}\operatorname{tr}_{\omega_{X}}\omega_{\varphi}\end{split} (A.6)

where C2>0C_{2}>0 is a constant depending only on C1C_{1} and nn. To conclude, combining (A.5) and (A.6), we finally obtain

Δωφ,vlogtrωXωφΔωXFtrωXωφBtrωφωXC2Cv21trωXωφα,β(logv)αβ(mφ)dmφξα,dmφξβωX\displaystyle\Delta_{\omega_{\varphi},v}\log\operatorname{tr}_{\omega_{X}}\omega_{\varphi}\geq\frac{\Delta_{\omega_{X}}F}{\operatorname{tr}_{\omega_{X}}\omega_{\varphi}}-B\operatorname{tr}_{\omega_{\varphi}}\omega_{X}-C_{2}C_{v}^{2}-\frac{1}{\operatorname{tr}_{\omega_{X}}\omega_{\varphi}}\sum_{\alpha,\beta}(\log v)_{\alpha\beta}(m_{\varphi})\langle dm^{\xi_{\alpha}}_{\varphi},dm^{\xi_{\beta}}_{\varphi}\rangle_{\omega_{X}}

as required. ∎

Appendix B Weighted local Chen–Cheng’s 𝒞2\mathcal{C}^{2}-estimate

This section aims to give details for the 𝒞2\mathcal{C}^{2}-estimate in Theorem 2.3. The proof follows a similar argument in [CC21a, Prop. 4.1] with further analysis of the weighted terms. Recall that we have two local equations on B1(0)nB_{1}(0)\subset\mathbb{C}^{n},

v(mddcϕ)det(ϕij¯)=eGandΔϕ,vG=S=w(mddcϕ).v(m_{dd^{c}\phi})\det(\phi_{i\bar{j}})=e^{G}\quad\text{and}\quad\Delta_{\phi,v}G=-S=-w(m_{dd^{c}\phi}).

Here we denote Δf:=k=1nkk¯f\Delta f:=\sum_{k=1}^{n}\partial_{k}\partial_{\bar{k}}f and Δϕ,vf:=i,j=1nϕij¯ij¯f+12dlogv(mddcϕ),dfddcϕ\Delta_{\phi,v}f:=\sum_{i,j=1}^{n}\phi^{i\bar{j}}\partial_{i}\partial_{\bar{j}}f+\frac{1}{2}\langle d\log v(m_{dd^{c}\phi}),df\rangle_{dd^{c}\phi} which are different from the scaling we used before. We first prove the following lemma:

Lemma B.1.

There exists positive constants K,CK,C depending on v,w,GL(B1(0))v,w,\|G\|_{L^{\infty}(B_{1}(0))} and ϕL(B1(0))\|\phi\|_{L^{\infty}(B_{1}(0))} such that

Δϕ,vuC(Δϕ)uon B1(0)\Delta_{\phi,v}u\geq-C\cdot(\Delta\phi)\cdot u\quad\text{on $B_{1}(0)$}

where u=eG2|dG|ϕ2+KΔϕu=e^{\frac{G}{2}}|dG|_{\phi}^{2}+K\Delta\phi.

Proof.

We first remark that by arithmetic and geometric means inequality, one has

Δϕndet(ϕij¯)1/n=neG/nv(mddcϕ)1/ncandtrϕωndet(ϕij¯)1/n=neG/nv(mddcϕ)1/nc.\Delta\phi\geq n\det(\phi_{i\bar{j}})^{1/n}=ne^{G/n}v(m_{dd^{c}\phi})^{-1/n}\geq c\quad\text{and}\quad\operatorname{tr}_{\phi}\omega\geq n\det(\phi_{i\bar{j}})^{-1/n}=ne^{-G/n}v(m_{dd^{c}\phi})^{1/n}\geq c.

where ω\omega is the Euclidean metric and c>0c>0 is a constant depending only on nn, GL(B1(0))\|G\|_{L^{\infty}(B_{1}(0))} and CvC_{v}. We next review the following estimate by Chen and Cheng [CC21a, p. 16, (4.3)]. Under the normal coordinates with respect to ϕij¯\phi_{i\bar{j}}, we may assume that ωij¯(x0)=δij\omega_{i\bar{j}}(x_{0})=\delta_{ij}, ϕij¯(x0)=λiδij\phi_{i\bar{j}}(x_{0})=\lambda_{i}\delta_{ij} and kϕij¯(x0)=0\partial_{k}\phi_{i\bar{j}}(x_{0})=0 at a fixed point x0x_{0}, then, at x0x_{0}, we have

eG2Δϕ(eG2|dG|ϕ2)=Δϕ|ϕG|ϕ2+12ϕii¯(iGi¯|ϕG|ϕ2+i¯Gi|ϕG|ϕ2)+14|ϕG|ϕ4+12ΔϕG|ϕG|ϕ2=ϕii¯(iΔϕGi¯G+i¯ΔϕGiG)+12Ricϕ(ϕG,ϕG)+|ϕϕG|ϕ2+|ϕϕ¯G|ϕ2+12ϕii¯ϕjj¯(GiGjGi¯j¯+GiGj¯Gji¯+Gi¯Gj¯Gij+Gi¯GjGij¯)+14|ϕG|ϕ4+12ΔϕG|ϕG|ϕ2ϕii¯(iΔϕGi¯G+i¯ΔϕGiG)+12Ricϕ(ϕG,ϕG)+|ϕϕ¯G|ϕ2+12ϕii¯ϕjj¯(GiGj¯Gji¯+Gi¯GjGij¯)+12ΔϕG|ϕG|ϕ2\begin{split}e^{-\frac{G}{2}}\Delta_{\phi}(e^{\frac{G}{2}}|dG|_{\phi}^{2})&=\Delta_{\phi}|\nabla^{\phi}G|_{\phi}^{2}+\frac{1}{2}\phi^{i\bar{i}}(\partial_{i}G\cdot\partial_{\bar{i}}|\nabla^{\phi}G|_{\phi}^{2}+\partial_{\bar{i}}G\cdot\partial_{i}|\nabla^{\phi}G|_{\phi}^{2})+\frac{1}{4}|\nabla^{\phi}G|_{\phi}^{4}+\frac{1}{2}\Delta_{\phi}G\cdot|\nabla^{\phi}G|_{\phi}^{2}\\ &=\phi^{i\bar{i}}(\partial_{i}\Delta_{\phi}G\cdot\partial_{\bar{i}}G+\partial_{\bar{i}}\Delta_{\phi}G\cdot\partial_{i}G)+\frac{1}{2}\operatorname{Ric}_{\phi}(\nabla^{\phi}G,\nabla^{\phi}G)+|\nabla^{\phi}\nabla^{\phi}G|_{\phi}^{2}+|\nabla^{\phi}\overline{\nabla^{\phi}}G|_{\phi}^{2}\\ &\quad+\frac{1}{2}\phi^{i\bar{i}}\phi^{j\bar{j}}(G_{i}G_{j}G_{\bar{i}\bar{j}}+G_{i}G_{\bar{j}}G_{j\bar{i}}+G_{\bar{i}}G_{\bar{j}}G_{ij}+G_{\bar{i}}G_{j}G_{i\bar{j}})+\frac{1}{4}|\nabla^{\phi}G|_{\phi}^{4}+\frac{1}{2}\Delta_{\phi}G\cdot|\nabla^{\phi}G|_{\phi}^{2}\\ &\geq\phi^{i\bar{i}}(\partial_{i}\Delta_{\phi}G\cdot\partial_{\bar{i}}G+\partial_{\bar{i}}\Delta_{\phi}G\cdot\partial_{i}G)+\frac{1}{2}\operatorname{Ric}_{\phi}(\nabla^{\phi}G,\nabla^{\phi}G)+|\nabla^{\phi}\overline{\nabla^{\phi}}G|_{\phi}^{2}\\ &\quad+\frac{1}{2}\phi^{i\bar{i}}\phi^{j\bar{j}}(G_{i}G_{\bar{j}}G_{j\bar{i}}+G_{\bar{i}}G_{j}G_{i\bar{j}})+\frac{1}{2}\Delta_{\phi}G\cdot|\nabla^{\phi}G|_{\phi}^{2}\end{split} (B.1)

where Ricϕ:=Ric(ddcϕ)=ddclogdet(ϕij¯)\operatorname{Ric}_{\phi}:=\operatorname{Ric}(dd^{c}\phi)=-dd^{c}\log\det(\phi_{i\bar{j}}). The last inequality comes from the fact that

14|ϕG|ϕ4+12ϕii¯ϕjj¯(GiGjGi¯j¯+Gi¯Gj¯Gij)+|ϕϕG|ϕ2=|Gij+12λi1λj1GiGj|ϕ20.\frac{1}{4}|\nabla^{\phi}G|_{\phi}^{4}+\frac{1}{2}\phi^{i\bar{i}}\phi^{j\bar{j}}(G_{i}G_{j}G_{\bar{i}\bar{j}}+G_{\bar{i}}G_{\bar{j}}G_{ij})+|\nabla^{\phi}\nabla^{\phi}G|_{\phi}^{2}=\left|G_{ij}+\frac{1}{2}\sqrt{\lambda_{i}^{-1}\lambda_{j}^{-1}}G_{i}G_{j}\right|_{\phi}^{2}\geq 0.

From the first equation, we have Ric(ddcϕ)=ddc(Glogv(mddcϕ))\operatorname{Ric}(dd^{c}\phi)=-dd^{c}(G-\log v(m_{dd^{c}\phi})). Rewrite (B.1) as follows

eG2Δϕ(eG2|dG|ϕ2)\displaystyle e^{-\frac{G}{2}}\Delta_{\phi}(e^{\frac{G}{2}}|dG|_{\phi}^{2}) ϕii¯(iΔϕGi¯G+i¯ΔϕGiG)ϕii¯ϕjj¯Gij¯GiGj¯\displaystyle\geq\phi^{i\bar{i}}(\partial_{i}\Delta_{\phi}G\cdot\partial_{\bar{i}}G+\partial_{\bar{i}}\Delta_{\phi}G\cdot\partial_{i}G)-\phi^{i\bar{i}}\phi^{j\bar{j}}G_{i\bar{j}}G_{i}G_{\bar{j}}
+ϕii¯ϕjj¯(logv(mddcϕ))ij¯GiGj¯+ϕii¯ϕjj¯Gij¯GiGj¯+|ϕϕ¯G|ϕ2+12ΔϕG|dG|ϕ2\displaystyle\quad+\phi^{i\bar{i}}\phi^{j\bar{j}}(\log v(m_{dd^{c}\phi}))_{i\bar{j}}G_{i}G_{\bar{j}}+\phi^{i\bar{i}}\phi^{j\bar{j}}G_{i\bar{j}}G_{i}G_{\bar{j}}+|\nabla^{\phi}\overline{\nabla^{\phi}}G|_{\phi}^{2}+\frac{1}{2}\Delta_{\phi}G\cdot|dG|_{\phi}^{2}
=2trϕ(dΔϕGdcG)+ϕii¯ϕjj¯(logv(mddcϕ))ij¯GiGj¯+|ϕϕ¯G|ϕ2+12ΔϕG|dG|ϕ2.\displaystyle=2\operatorname{tr}_{\phi}(d\Delta_{\phi}G\wedge d^{c}G)+\phi^{i\bar{i}}\phi^{j\bar{j}}(\log v(m_{dd^{c}\phi}))_{i\bar{j}}G_{i}G_{\bar{j}}+|\nabla^{\phi}\overline{\nabla^{\phi}}G|_{\phi}^{2}+\frac{1}{2}\Delta_{\phi}G\cdot|dG|_{\phi}^{2}.

Considering the weighted version of the above inequality, one can infer

eG2Δϕ,v(eG2|dG|ϕ2)2trϕ(dΔϕGdcG)+ϕii¯ϕjj¯(logv(mddcϕ))ij¯GiGj¯+|ϕϕ¯G|ϕ2+12ΔϕG|dG|ϕ2+12trϕ(dlogv(mddcϕ)dcG)|dG|ϕ2+trϕ(dlogv(mddcϕ)dc|dG|ϕ2)=2trϕ(dΔϕGdcG)+ϕii¯ϕjj¯(logv(mddcϕ))ij¯GiGj¯+|ϕϕ¯G|ϕ2+12Δϕ,vG|dG|ϕ2+trϕ(dlogv(mddcϕ)dc|dG|ϕ2)\begin{split}e^{-\frac{G}{2}}\Delta_{\phi,v}(e^{\frac{G}{2}}|dG|_{\phi}^{2})&\geq 2\operatorname{tr}_{\phi}(d\Delta_{\phi}G\wedge d^{c}G)+\phi^{i\bar{i}}\phi^{j\bar{j}}(\log v(m_{dd^{c}\phi}))_{i\bar{j}}G_{i}G_{\bar{j}}+|\nabla^{\phi}\overline{\nabla^{\phi}}G|_{\phi}^{2}+\frac{1}{2}\Delta_{\phi}G\cdot|dG|_{\phi}^{2}\\ &\quad+\frac{1}{2}\operatorname{tr}_{\phi}(d\log v(m_{dd^{c}\phi})\wedge d^{c}G)|dG|_{\phi}^{2}+\operatorname{tr}_{\phi}(d\log v(m_{dd^{c}\phi})\wedge d^{c}|dG|_{\phi}^{2})\\ &=2\operatorname{tr}_{\phi}(d\Delta_{\phi}G\wedge d^{c}G)+\phi^{i\bar{i}}\phi^{j\bar{j}}(\log v(m_{dd^{c}\phi}))_{i\bar{j}}G_{i}G_{\bar{j}}+|\nabla^{\phi}\overline{\nabla^{\phi}}G|_{\phi}^{2}+\frac{1}{2}\Delta_{\phi,v}G\cdot|dG|_{\phi}^{2}\\ &\quad+\operatorname{tr}_{\phi}(d\log v(m_{dd^{c}\phi})\wedge d^{c}|dG|_{\phi}^{2})\end{split} (B.2)

Under normal coordinates with respect to ϕij¯\phi_{i\bar{j}} at x0x_{0}, we obtain

trϕ(dlogv(mddcϕ)dc|dG|ϕ2)=12ϕii¯((logv(mddcϕ))i[ϕjj¯GjGj¯]i¯+(logv(mddcϕ))i¯[ϕjj¯GjGj¯]i)\displaystyle\operatorname{tr}_{\phi}(d\log v(m_{dd^{c}\phi})\wedge d^{c}|dG|_{\phi}^{2})=\frac{1}{2}\phi^{i\bar{i}}\left((\log v(m_{dd^{c}\phi}))_{i}[\phi^{j\bar{j}}G_{j}G_{\bar{j}}]_{\bar{i}}+(\log v(m_{dd^{c}\phi}))_{\bar{i}}[\phi^{j\bar{j}}G_{j}G_{\bar{j}}]_{i}\right)
=12ϕjj¯(Gj¯ϕii¯(logv(mddcϕ))iGji¯+Gjϕii¯(logv(mddcϕ))iGj¯i¯+Gj¯ϕii¯(logv(mddcϕ))i¯Gji+Gjϕii¯(logv(mddcϕ))i¯Gj¯i)\displaystyle=\frac{1}{2}\phi^{j\bar{j}}\Big{(}G_{\bar{j}}\phi^{i\bar{i}}(\log v(m_{dd^{c}\phi}))_{i}G_{j\bar{i}}+G_{j}\phi^{i\bar{i}}(\log v(m_{dd^{c}\phi}))_{i}G_{\bar{j}\bar{i}}+G_{\bar{j}}\phi^{i\bar{i}}(\log v(m_{dd^{c}\phi}))_{\bar{i}}G_{ji}+G_{j}\phi^{i\bar{i}}(\log v(m_{dd^{c}\phi}))_{\bar{i}}G_{\bar{j}i}\Big{)}
=2trϕ(dGdc[trϕ(dlogv(mddcϕ)dcG)])(ϕii¯ϕjj¯(logv(mddcϕ))ijGi¯Gj¯)ϕii¯ϕjj¯(logv(mddcϕ))ij¯Gi¯Gj\displaystyle=2\operatorname{tr}_{\phi}(dG\wedge d^{c}[\operatorname{tr}_{\phi}(d\log v(m_{dd^{c}\phi})\wedge d^{c}G)])-\Re\left(\phi^{i\bar{i}}\phi^{j\bar{j}}(\log v(m_{dd^{c}\phi}))_{ij}G_{\bar{i}}G_{\bar{j}}\right)-\phi^{i\bar{i}}\phi^{j\bar{j}}(\log v(m_{dd^{c}\phi}))_{i\bar{j}}G_{\bar{i}}G_{j}

and the inequality (B.2) can be expressed as

eG2Δϕ,v(eG2|dG|ϕ2)2trϕ(dΔϕ,vGdcG)+|ϕϕ¯G|ϕ2+12Δϕ,vG|ϕG|ϕ2(ϕii¯ϕjj¯(logv(mddcϕ))ijGi¯Gj¯).\begin{split}e^{-\frac{G}{2}}\Delta_{\phi,v}(e^{\frac{G}{2}}|dG|_{\phi}^{2})&\geq 2\operatorname{tr}_{\phi}(d\Delta_{\phi,v}G\wedge d^{c}G)+|\nabla^{\phi}\overline{\nabla^{\phi}}G|_{\phi}^{2}+\frac{1}{2}\Delta_{\phi,v}G\cdot|\nabla^{\phi}G|_{\phi}^{2}\\ &\quad-\Re\left(\phi^{i\bar{i}}\phi^{j\bar{j}}(\log v(m_{dd^{c}\phi}))_{ij}G_{\bar{i}}G_{\bar{j}}\right).\end{split} (B.3)

Recall that

dmddcϕξα=iξαddcϕ=iξα(21ϕpq¯dzpdz¯q)=21ϕpq¯(ξα)pdz¯q+21ϕpq¯(ξα′′)q¯dzp.dm^{\xi_{\alpha}}_{dd^{c}\phi}=-i_{\xi_{\alpha}}dd^{c}\phi=-i_{\xi_{\alpha}}(2\sqrt{-1}\phi_{p\bar{q}}dz^{p}\wedge d\bar{z}^{q})=-2\sqrt{-1}\phi_{p\bar{q}}(\xi_{\alpha}^{\prime})^{p}d\bar{z}^{q}+2\sqrt{-1}\phi_{p\bar{q}}(\xi_{\alpha}^{\prime\prime})^{\bar{q}}dz^{p}.

Hence,

ij2logv(mddcϕ)=i{α(αlogv)(mddcϕ)21ϕj¯(ξα′′)¯}\displaystyle\partial_{ij}^{2}\log v(m_{dd^{c}\phi})=\partial_{i}\left\{\sum_{\alpha}(\partial_{\alpha}\log v)(m_{dd^{c}\phi})\cdot 2\sqrt{-1}\phi_{j\bar{\ell}}(\xi^{\prime\prime}_{\alpha})^{\bar{\ell}}\right\} (B.4)
=4α,β(αβ2logv)(mddcϕ)[ϕiq¯ϕj¯(ξβ′′)q¯(ξα′′)¯]+2α(αlogv)(mddcϕ)1[ϕj¯i(ξα′′)¯+ϕj¯(ξα′′)i¯].\displaystyle=4\sum_{\alpha,\beta}(\partial_{\alpha\beta}^{2}\log v)(m_{dd^{c}\phi})\cdot[-\phi_{i\bar{q}}\phi_{j\bar{\ell}}(\xi^{\prime\prime}_{\beta})^{\bar{q}}(\xi^{\prime\prime}_{\alpha})^{\bar{\ell}}]+2\sum_{\alpha}(\partial_{\alpha}\log v)(m_{dd^{c}\phi})\cdot\sqrt{-1}[\phi_{j\bar{\ell}i}(\xi^{\prime\prime}_{\alpha})^{\bar{\ell}}+\phi_{j\bar{\ell}}(\xi^{\prime\prime}_{\alpha})^{\bar{\ell}}_{i}].

Note that ξα\xi^{\prime}_{\alpha} is a holomorphic vector field for any α\alpha. Therefore, (ξα)i¯j=0=(ξα′′)ij¯(\xi^{\prime}_{\alpha})^{j}_{\bar{i}}=0=(\xi^{\prime\prime}_{\alpha})^{\bar{j}}_{i} for any i,j{1,,n}i,j\in\{1,\cdots,n\}.

Combining (B.3), (B.4), and the fact that (ξα′′)i¯=0(\xi^{\prime\prime}_{\alpha})^{\bar{\ell}}_{i}=0, under normal coordinates with respect to ϕ\phi at x0x_{0}, we get

eG2Δϕ,v(eG2|dG|ϕ2)\displaystyle e^{-\frac{G}{2}}\Delta_{\phi,v}(e^{\frac{G}{2}}|dG|_{\phi}^{2}) 2trϕ(dΔϕ,vGdcG)+|ϕϕ¯G|ϕ2+12Δϕ,vG|ϕG|ϕ2\displaystyle\geq 2\operatorname{tr}_{\phi}(d\Delta_{\phi,v}G\wedge d^{c}G)+|\nabla^{\phi}\overline{\nabla^{\phi}}G|_{\phi}^{2}+\frac{1}{2}\Delta_{\phi,v}G\cdot|\nabla^{\phi}G|_{\phi}^{2}
+4α,β(αβ2logv)(mddcϕ)(ξβ′′)i¯(ξα′′)j¯Gi¯Gj¯\displaystyle\quad+4\Re\sum_{\alpha,\beta}(\partial_{\alpha\beta}^{2}\log v)(m_{dd^{c}\phi})(\xi^{\prime\prime}_{\beta})^{\bar{i}}(\xi^{\prime\prime}_{\alpha})^{\bar{j}}G_{\bar{i}}G_{\bar{j}}
2trϕ(dΔϕ,vGdcG)+|ϕϕ¯G|ϕ2+12Δϕ,vG|ϕG|ϕ24Cv4Cξ|dG|ω2\displaystyle\geq 2\operatorname{tr}_{\phi}(d\Delta_{\phi,v}G\wedge d^{c}G)+|\nabla^{\phi}\overline{\nabla^{\phi}}G|_{\phi}^{2}+\frac{1}{2}\Delta_{\phi,v}G\cdot|\nabla^{\phi}G|_{\phi}^{2}-4C_{v}^{4}C_{\xi}|dG|_{\omega}^{2}

where Cξ>0C_{\xi}>0 is a constant such that |ξ|ω2Cξ|\xi|^{2}_{\omega}\leq C_{\xi}. Using the second equation Δϕ,vG=w(mddcϕ)\Delta_{\phi,v}G=-w(m_{dd^{c}\phi}), we then derive

trϕ(dΔϕ,vGdcG)\displaystyle\operatorname{tr}_{\phi}(d\Delta_{\phi,v}G\wedge d^{c}G) =trϕ(dw(mddcϕ)dcG)=αwα(mddcϕ)trϕ(dmddcϕξαdcG)\displaystyle=-\operatorname{tr}_{\phi}(dw(m_{dd^{c}\phi})\wedge d^{c}G)=-\sum_{\alpha}w_{\alpha}(m_{dd^{c}\phi})\operatorname{tr}_{\phi}(dm_{dd^{c}\phi}^{\xi_{\alpha}}\wedge d^{c}G)
=αwα(mddcϕ)(ξα)iGiCξCw|dG|ωC1(Δϕ)|dG|ϕ2\displaystyle=\sum_{\alpha}w_{\alpha}(m_{dd^{c}\phi})\Re(\xi^{\prime}_{\alpha})^{i}G_{i}\geq-C_{\xi}C_{w}|dG|_{\omega}\geq-C_{1}-(\Delta\phi)\cdot|dG|_{\phi}^{2}

where C1>0C_{1}>0 depending only on Cξ,CvC_{\xi},C_{v}. Then

eG2Δϕ,v(eG2|dG|ϕ2)2C1C2(Δϕ)|dG|ϕ2+|ϕϕ¯G|ϕ2e^{-\frac{G}{2}}\Delta_{\phi,v}(e^{\frac{G}{2}}|dG|_{\phi}^{2})\geq-2C_{1}-C_{2}(\Delta\phi)|dG|_{\phi}^{2}+|\nabla^{\phi}\overline{\nabla^{\phi}}G|_{\phi}^{2} (B.5)

where C2>0C_{2}>0 depend only on Cξ,Cv,CwC_{\xi},C_{v},C_{w}.

Next, we compute, under normal coordinates with respect to ϕij¯\phi_{i\bar{j}},

Δϕ,v(Δϕ)\displaystyle\Delta_{\phi,v}(\Delta\phi) =Δϕ(Δϕ)+trϕ(dlogv(mddcϕ)dcΔϕ)\displaystyle=\Delta_{\phi}(\Delta\phi)+\operatorname{tr}_{\phi}(d\log v(m_{dd^{c}\phi})\wedge d^{c}\Delta\phi)
=trω(1¯logdet(ϕij¯))+trϕ(dlogv(mddcϕ)dcΔϕ)\displaystyle=\operatorname{tr}_{\omega}(\sqrt{-1}\partial\bar{\partial}\log\det(\phi_{i\bar{j}}))+\operatorname{tr}_{\phi}(d\log v(m_{dd^{c}\phi})\wedge d^{c}\Delta\phi)
=ΔGΔlogv(mddcϕ)+trϕ(dlogv(mddcϕ)dcΔϕ)\displaystyle=\Delta G-\Delta\log v(m_{dd^{c}\phi})+\operatorname{tr}_{\phi}(d\log v(m_{dd^{c}\phi})\wedge d^{c}\Delta\phi)
ΔGα,β(αβ2logv)(mddcϕ)dmddcϕξα,dmddcϕξβC3Δϕ\displaystyle\geq\Delta G-\sum_{\alpha,\beta}(\partial_{\alpha\beta}^{2}\log v)(m_{dd^{c}\phi})\langle dm_{dd^{c}\phi}^{\xi_{\alpha}},dm_{dd^{c}\phi}^{\xi_{\beta}}\rangle-C_{3}\Delta\phi
ΔG2Cv2(Δϕ)2C3Δϕ,\displaystyle\geq\Delta G-2C_{v}^{2}(\Delta\phi)^{2}-C_{3}\Delta\phi,

where C3C_{3} depend only on nn and a constant C>0C>0 such that CωξαωCω-C\omega\leq\mathcal{L}_{\xi_{\alpha}}\omega\leq C\omega. The fourth line follows from the same estimate (A.6).

Then we obtain

Δϕ,v(eG2|dG|ϕ2+KΔϕ)\displaystyle\Delta_{\phi,v}(e^{\frac{G}{2}}|dG|_{\phi}^{2}+K\Delta\phi) C1eG2C2(Δϕ)eG2|dG|ϕ2+eG2|ϕϕ¯G|ϕ2+KΔG2Cv2(Δϕ)2C3Δϕ.\displaystyle\geq-C_{1}e^{\frac{G}{2}}-C_{2}(\Delta\phi)e^{\frac{G}{2}}|dG|_{\phi}^{2}+e^{\frac{G}{2}}|\nabla^{\phi}\overline{\nabla^{\phi}}G|_{\phi}^{2}+K\Delta G-2C_{v}^{2}(\Delta\phi)^{2}-C_{3}\Delta\phi.

Recall that |ϕϕ¯G|ϕ2=i,jϕii¯ϕjj¯|Gij¯|2|\nabla^{\phi}\overline{\nabla^{\phi}}G|_{\phi}^{2}=\sum_{i,j}\phi^{i\bar{i}}\phi^{j\bar{j}}|G_{i\bar{j}}|^{2} and GLC4\|G\|_{L^{\infty}}\leq C_{4}. By Cauchy–Schwarz inequality, we have the following estimate

eG2|ϕϕ¯G|ϕ2+eC4K24(Δϕ)2eC4i,jϕii¯ϕjj¯|Gij¯|2+eC4K24i,jϕii¯ϕjj¯i,jK|Gij¯|KiK|Gii¯|K|ΔG|.\displaystyle e^{\frac{G}{2}}|\nabla^{\phi}\overline{\nabla^{\phi}}G|_{\phi}^{2}+e^{C_{4}}\frac{K^{2}}{4}(\Delta\phi)^{2}\geq e^{-C_{4}}\sum_{i,j}\phi^{i\bar{i}}\phi^{j\bar{j}}|G_{i\bar{j}}|^{2}+e^{C_{4}}\frac{K^{2}}{4}\sum_{i,j}\phi_{i\bar{i}}\phi_{j\bar{j}}\geq\sum_{i,j}K|G_{i\bar{j}}|\geq K\sum_{i}K|G_{i\bar{i}}|\geq K|\Delta G|.

Moreover, since Δϕc\Delta\phi\geq c; hence, we derive the following estimate

Δϕ,vu=Δϕ,v(eG2|dG|ϕ2+KΔϕ)C5K(Δϕ)u,\Delta_{\phi,v}u=\Delta_{\phi,v}(e^{\frac{G}{2}}|dG|_{\phi}^{2}+K\Delta\phi)\geq-C_{5}\cdot K\cdot(\Delta\phi)\cdot u,

for a constant C5>0C_{5}>0 depending only on nn, CvC_{v}, CwC_{w} and some sufficient large K>0K>0 depending on Cv,C1,C3,C4,cC_{v},C_{1},C_{3},C_{4},c and GL(B1(0))\|G\|_{L^{\infty}(B_{1}(0))}. ∎

Using Lemma B.1, one can obtain Theorem 2.3 following an argument in Chen–Cheng’s article [CC21a]:

Proof of Theorem 2.3.

Let η\eta be a positive smooth function with compact support in B1(0)B_{1}(0), η1\eta\equiv 1 on B3/4(0)B_{3/4}(0) and 0η10\leq\eta\leq 1. We first claim that there is a positive constant

D1=D1(n,GL(B1(0)),v,w,ΔϕLn+1(B1(0)),Cξ,Cη)D_{1}=D_{1}(n,\|G\|_{L^{\infty}(B_{1}(0))},v,w,\|\Delta\phi\|_{L^{n+1}(B_{1}(0))},C_{\xi},C_{\eta})

such that |dG|ϕ2L1(B3/4(0))D1\||dG|_{\phi}^{2}\|_{L^{1}(B_{3/4}(0))}\leq D_{1}, where Cη>0C_{\eta}>0 is a constant such that |dη|ω,|ddcη|ωCη|d\eta|_{\omega},|dd^{c}\eta|_{\omega}\leq C_{\eta} and Cξ>0C_{\xi}>0 is a constant such that |ξα|ωCξ|\xi_{\alpha}|_{\omega}\leq C_{\xi}.

To see the claim, set CG>0C_{G}>0 a constant such that |G|CG|G|\leq C_{G} and eGCGe^{-G}\leq C_{G} on B1(0)B_{1}(0). Then we have

B3/4(0)|dG|ϕ2ωnB1(0)η|dG|ϕ2eGv(mddcϕ)(ddcϕ)nCGCvB1(0)η𝑑GdcG(ddcϕ)n1\displaystyle\int_{B_{3/4}(0)}|dG|_{\phi}^{2}\omega^{n}\leq\int_{B_{1}(0)}\eta|dG|_{\phi}^{2}e^{-G}v(m_{dd^{c}\phi})(dd^{c}\phi)^{n}\leq C_{G}C_{v}\int_{B_{1}(0)}\eta dG\wedge d^{c}G\wedge(dd^{c}\phi)^{n-1}
=CGCvB1(0)ηGddcG(ddcϕ)n1GdηdcG(ddcϕ)n1\displaystyle=C_{G}C_{v}\int_{B_{1}(0)}-\eta Gdd^{c}G\wedge(dd^{c}\phi)^{n-1}-Gd\eta\wedge d^{c}G\wedge(dd^{c}\phi)^{n-1}
=CGCvB1(0)ηGn[w(mddcϕ)(ddcϕ)n+dlogv(mddcϕ)dcG(ddcϕ)n1]\displaystyle=C_{G}C_{v}\int_{B_{1}(0)}\eta Gn[w(m_{dd^{c}\phi})(dd^{c}\phi)^{n}+d\log v(m_{dd^{c}\phi})\wedge d^{c}G\wedge(dd^{c}\phi)^{n-1}]
+12CGCvB1(0)G2𝑑dcη(ddcϕ)n1\displaystyle\qquad+\frac{1}{2}C_{G}C_{v}\int_{B_{1}(0)}G^{2}dd^{c}\eta\wedge(dd^{c}\phi)^{n-1}
neCGCG2Cv2CwB1(0)ωn+12CG3CvCηB1(0)(Δϕ)n1ωn\displaystyle\leq ne^{C_{G}}C_{G}^{2}C_{v}^{2}C_{w}\int_{B_{1}(0)}\omega^{n}+\frac{1}{2}C_{G}^{3}C_{v}C_{\eta}\int_{B_{1}(0)}(\Delta\phi)^{n-1}\omega^{n}
12nCGCv[B1(0)G2η𝑑dclogv(mddcϕ)(ddcϕ)n1+G2dlogv(mddcϕ)dcη(ddcϕ)n1]\displaystyle\qquad-\frac{1}{2}nC_{G}C_{v}\left[\int_{B_{1}(0)}G^{2}\eta dd^{c}\log v(m_{dd^{c}\phi})\wedge(dd^{c}\phi)^{n-1}+G^{2}d\log v(m_{dd^{c}\phi})\wedge d^{c}\eta\wedge(dd^{c}\phi)^{n-1}\right]
neCGCG2Cv2CwB1(0)ωn+12CG3CvCηB1(0)(Δϕ)n1ωn+12nCGCv[B1(0)eCGCG2Cv3Cξ2(Δϕ)ωn+eCGCG2Cv3CξCηωn]\displaystyle\leq ne^{C_{G}}C_{G}^{2}C_{v}^{2}C_{w}\int_{B_{1}(0)}\omega^{n}+\frac{1}{2}C_{G}^{3}C_{v}C_{\eta}\int_{B_{1}(0)}(\Delta\phi)^{n-1}\omega^{n}+\frac{1}{2}nC_{G}C_{v}\left[\int_{B_{1}(0)}e^{C_{G}}C_{G}^{2}C_{v}^{3}C_{\xi}^{2}(\Delta\phi)\omega^{n}+e^{C_{G}}C_{G}^{2}C_{v}^{3}C_{\xi}C_{\eta}\omega^{n}\right]
D1.\displaystyle\leq D_{1}.

Hence, the claim follows. We provide more details about the inequality next to the last one. Since

j¯logv(mddcϕ)=1v(mddcϕ)αvα(mddcϕ)(21ϕkj¯(ξα)k),\partial_{\bar{j}}\log v(m_{dd^{c}\phi})=\frac{1}{v(m_{dd^{c}\phi})}\sum_{\alpha}v_{\alpha}(m_{dd^{c}\phi})(-2\sqrt{-1}\phi_{k\bar{j}}(\xi_{\alpha}^{\prime})^{k}),

we have trϕ(dlogv(mddcϕ)dcη)Cv2CξCη\operatorname{tr}_{\phi}(d\log v(m_{dd^{c}\phi})\wedge d^{c}\eta)\leq C_{v}^{2}C_{\xi}C_{\eta}; thus,

dlogv(mddcϕ)dcη(ddcϕ)nCv2CξCη(ddcϕ)n=Cv2CξCηeGv(mddcϕ)ωneCGCv3CξCηωn.d\log v(m_{dd^{c}\phi})\wedge d^{c}\eta\wedge(dd^{c}\phi)^{n}\leq C_{v}^{2}C_{\xi}C_{\eta}(dd^{c}\phi)^{n}=C_{v}^{2}C_{\xi}C_{\eta}\frac{e^{G}}{v(m_{dd^{c}\phi})}\omega^{n}\leq e^{C_{G}}C_{v}^{3}C_{\xi}C_{\eta}\omega^{n}.

On the other hand, for the term involving ddclogv(mddcϕ)dd^{c}\log v(m_{dd^{c}\phi}),

ij¯2logv(mddcϕ)=i{α(αlogv)(mddcϕ)(21ϕkj¯(ξα)k)}\displaystyle\partial_{i\bar{j}}^{2}\log v(m_{dd^{c}\phi})=\partial_{i}\left\{\sum_{\alpha}(\partial_{\alpha}\log v)(m_{dd^{c}\phi})(-2\sqrt{-1}\phi_{k\bar{j}}(\xi_{\alpha}^{\prime})^{k})\right\}
=4α,β(αβ2logv)(mddcϕ)ϕiq¯ϕkj¯(ξβ)q¯(ξα)k+2α(αlogv)(mddcϕ)(1)[ϕkj¯i(ξα)k+ϕkj¯(ξα)ik].\displaystyle=4\sum_{\alpha,\beta}(\partial_{\alpha\beta}^{2}\log v)(m_{dd^{c}\phi})\phi_{i\bar{q}}\phi_{k\bar{j}}(\xi^{\prime}_{\beta})^{\bar{q}}(\xi_{\alpha}^{\prime})^{k}+2\sum_{\alpha}(\partial_{\alpha}\log v)(m_{dd^{c}\phi})(-\sqrt{-1})[\phi_{k\bar{j}i}(\xi^{\prime}_{\alpha})^{k}+\phi_{k\bar{j}}(\xi^{\prime}_{\alpha})^{k}_{i}].

Note that ξαddcϕ=0\mathcal{L}_{\xi_{\alpha}}dd^{c}\phi=0 from the GG-invariant assumption. Therefore,

0\displaystyle 0 =12ξddcϕ=12d(iξddcϕ)=d(1ϕk¯(ξ)kdz¯1ϕk¯(ξ′′)¯dzk)\displaystyle=\frac{1}{2}\mathcal{L}_{\xi}dd^{c}\phi=\frac{1}{2}d(i_{\xi}dd^{c}\phi)=d\left(\sqrt{-1}\phi_{k\bar{\ell}}(\xi^{\prime})^{k}d\bar{z}^{\ell}-\sqrt{-1}\phi_{k\bar{\ell}}(\xi^{\prime\prime})^{\bar{\ell}}dz^{k}\right)
=(1ϕk¯(ξ)ikdzidz¯1ϕk¯(ξ′′)j¯¯dz¯jdzk)1ϕk¯(ξ′′)i¯dzidzk+1ϕk¯(ξ)j¯kdz¯jdz¯\displaystyle=\left(\sqrt{-1}\phi_{k\bar{\ell}}(\xi^{\prime})^{k}_{i}dz^{i}\wedge d\bar{z}^{\ell}-\sqrt{-1}\phi_{k\bar{\ell}}(\xi^{\prime\prime})^{\bar{\ell}}_{\bar{j}}d\bar{z}^{j}\wedge dz^{k}\right)-\sqrt{-1}\phi_{k\bar{\ell}}(\xi^{\prime\prime})^{\bar{\ell}}_{i}dz^{i}\wedge dz^{k}+\sqrt{-1}\phi_{k\bar{\ell}}(\xi^{\prime})^{k}_{\bar{j}}d\bar{z}^{j}\wedge d\bar{z}^{\ell}
=1(ϕγβ¯(ξ)αγ+ϕαδ¯(ξ′′)β¯δ¯)dzαdz¯β1ϕk¯(ξ′′)i¯dzidzk+1ϕk¯(ξ)j¯kdz¯jdz¯\displaystyle=\sqrt{-1}\left(\phi_{\gamma\bar{\beta}}(\xi^{\prime})^{\gamma}_{\alpha}+\phi_{\alpha\bar{\delta}}(\xi^{\prime\prime})^{\bar{\delta}}_{\bar{\beta}}\right)dz^{\alpha}\wedge d\bar{z}^{\beta}-\sqrt{-1}\phi_{k\bar{\ell}}(\xi^{\prime\prime})^{\bar{\ell}}_{i}dz^{i}\wedge dz^{k}+\sqrt{-1}\phi_{k\bar{\ell}}(\xi^{\prime})^{k}_{\bar{j}}d\bar{z}^{j}\wedge d\bar{z}^{\ell}

and this implies ϕγβ¯(ξ)αγ+ϕαδ¯(ξ′′)β¯δ¯=0\phi_{\gamma\bar{\beta}}(\xi^{\prime})^{\gamma}_{\alpha}+\phi_{\alpha\bar{\delta}}(\xi^{\prime\prime})^{\bar{\delta}}_{\bar{\beta}}=0. In the normal coordinates of ddcϕdd^{c}\phi, we get

ϕij¯ϕkj¯(ξα)ik=12(ϕij¯ϕkj¯(ξα)ik+ϕij¯ϕj¯(ξα)i¯)=12ϕij¯(ϕkj¯(ξα)ik+ϕi¯(ξα′′)j¯¯)=0.\displaystyle\phi^{i\bar{j}}\phi_{k\bar{j}}(\xi^{\prime}_{\alpha})^{k}_{i}=\frac{1}{2}\left(\phi^{i\bar{j}}\phi_{k\bar{j}}(\xi^{\prime}_{\alpha})^{k}_{i}+\overline{\phi^{i\bar{j}}\phi_{\ell\bar{j}}(\xi^{\prime}_{\alpha})^{\ell}_{i}}\right)=\frac{1}{2}\phi^{i\bar{j}}\left(\phi_{k\bar{j}}(\xi^{\prime}_{\alpha})^{k}_{i}+\phi_{i\bar{\ell}}(\xi^{\prime\prime}_{\alpha})^{\bar{\ell}}_{\bar{j}}\right)=0.

This yields that trϕddclogv(mddcϕ)Cv2Cξ2(Δϕ)\operatorname{tr}_{\phi}dd^{c}\log v(m_{dd^{c}\phi})\leq C_{v}^{2}C_{\xi}^{2}(\Delta\phi) and

ddclogv(mddcϕ)(ddcϕ)n1Cv2Cξ2(Δϕ)(ddcϕ)neCGCv3Cξ2(Δϕ)ωn.dd^{c}\log v(m_{dd^{c}\phi})\wedge(dd^{c}\phi)^{n-1}\leq C_{v}^{2}C_{\xi}^{2}(\Delta\phi)(dd^{c}\phi)^{n}\leq e^{C_{G}}C_{v}^{3}C_{\xi}^{2}(\Delta\phi)\omega^{n}.

We also remark that Δϕ,vf=12v(mddcϕ)α(v(mddcϕ)gϕαββf)\Delta_{\phi,v}f=\frac{1}{2v(m_{dd^{c}\phi})}\partial_{\alpha}(v(m_{dd^{c}\phi})g_{\phi}^{\alpha\beta}\partial_{\beta}f) where α,β\alpha,\beta are real coordinates and [(gϕ)αβ]1α,β2n[(g_{\phi})_{\alpha\beta}]_{1\leq\alpha,\beta\leq 2n} is the Riemannian metric associate to ddcϕdd^{c}\phi. Therefore, it follows from Lemma B.1, that

α(v(mddcϕ)gϕαββu)2Cv(mddcϕ)(Δϕ)u.\partial_{\alpha}(v(m_{dd^{c}\phi})g_{\phi}^{\alpha\beta}\partial_{\beta}u)\geq-2Cv(m_{dd^{c}\phi})(\Delta\phi)u.

Using [CC21a, Lem. 6.3, arXiv version]), we derive that

uL(B1/2(0))D2(uL1(B3/4(0))+1)C,\|u\|_{L^{\infty}(B_{1/2}(0))}\leq D_{2}(\|u\|_{L^{1}(B_{3/4}(0))}+1)\leq C, (B.6)

where D2D_{2} depends on CvC_{v}, CwC_{w}, pp, SL\|S\|_{L^{\infty}}, ΔϕLp(B1(0))\|\Delta\phi\|_{L^{p}(B_{1}(0))}, trϕωLp(B1(0))\|\operatorname{tr}_{\phi}\omega\|_{L^{p}(B_{1}(0))}. Recall that u=eG2|dG|ϕ2+KΔϕu=e^{\frac{G}{2}}|dG|_{\phi}^{2}+K\Delta\phi; hence we get

uL1(B3/4(0))eCG/2|dG|ϕ2L1(B3/4(0))+KΔϕL1(B3/4(0))eCG/2D1+KΔϕL1(B3/4(0)).\|u\|_{L^{1}(B_{3/4}(0))}\leq e^{C_{G}/2}\||dG|^{2}_{\phi}\|_{L^{1}(B_{3/4}(0))}+K\|\Delta\phi\|_{L^{1}(B_{3/4}(0))}\leq e^{C_{G}/2}D_{1}+K\|\Delta\phi\|_{L^{1}(B_{3/4}(0))}.

Then combining this with (B.6) implies the uniform LL^{\infty}-estimates of |dG|ϕ|dG|_{\phi} and Δϕ\Delta\phi on B1/2(0)B_{1/2}(0) which only depend on CvC_{v}, CwC_{w}, pp, SL\|S\|_{L^{\infty}}, ΔϕLp(B1(0))\|\Delta\phi\|_{L^{p}(B_{1}(0))}, trϕωLp(B1(0))\|\operatorname{tr}_{\phi}\omega\|_{L^{p}(B_{1}(0))}. Then the standard Evans–Krylov estimate and bootstrapping argument imply higher order estimates on B1/2(0)B_{1/2}(0) for ϕ,G\phi,G which depend on CvC_{v}, CwC_{w}, pp, SL\|S\|_{L^{\infty}}, ΔϕLp(B1(0))\|\Delta\phi\|_{L^{p}(B_{1}(0))}, trϕωLp(B1(0))\|\operatorname{tr}_{\phi}\omega\|_{L^{p}(B_{1}(0))}. ∎

References

  • [AJL23] V. Apostolov, S. Jubert & A. Lahdili – “Weighted K-stability and coercivity with applications to extremal Kähler and Sasaki metrics”, Geom. Topol. 27 (2023), no. 8, p. 3229–3302.
  • [AP06] C. Arezzo & F. Pacard – “Blowing up and desingularizing constant scalar curvature Kähler manifolds”, Acta Math. 196 (2006), no. 2, p. 179–228.
  • [AP09] by same author, “Blowing up Kähler manifolds with constant scalar curvature. II”, Ann. of Math. (2) 170 (2009), no. 2, p. 685–738.
  • [APS11] C. Arezzo, F. Pacard & M. Singer – “Extremal metrics on blowups”, Duke Math. J. 157 (2011), no. 1, p. 1–51.
  • [Aub78] T. Aubin – “Équations du type Monge-Ampère sur les variétés kählériennes compactes”, Bull. Sci. Math. (2) 102 (1978), no. 1, p. 63–95.
  • [BT82] E. Bedford & B. A. Taylor – “A new capacity for plurisubharmonic functions”, Acta Math. 149 (1982), no. 1-2, p. 1–40.
  • [BBE+19] R. J. Berman, S. Boucksom, P. Eyssidieux, V. Guedj & A. Zeriahi – “Kähler-Einstein metrics and the Kähler-Ricci flow on log Fano varieties”, J. Reine Angew. Math. 751 (2019), p. 27–89.
  • [BBJ21] R. J. Berman, S. Boucksom & M. Jonsson – “A variational approach to the Yau-Tian-Donaldson conjecture”, J. Amer. Math. Soc. 34 (2021), no. 3, p. 605–652.
  • [BDL20] R. J. Berman, T. Darvas & C. H. Lu – “Regularity of weak minimizers of the KK-energy and applications to properness and KK-stability”, Ann. Sci. Éc. Norm. Supér. (4) 53 (2020), no. 2, p. 267–289.
  • [Ber13] B. Berndtsson – “The openness conjecture for plurisubharmonic functions”, arXiv:1305.5781 (2013).
  • [Bou04] S. Boucksom – “Divisorial Zariski decompositions on compact complex manifolds”, Ann. Sci. École Norm. Sup. (4) 37 (2004), no. 1, p. 45–76.
  • [BG13] S. Boucksom & V. Guedj – “Regularizing properties of the Kähler-Ricci flow”, in An introduction to the Kähler-Ricci flow, Lecture Notes in Math., vol. 2086, Springer, Cham, 2013, p. 189–237.
  • [BHJ19] S. Boucksom, T. Hisamoto & M. Jonsson – “Uniform K-stability and asymptotics of energy functionals in Kähler geometry”, J. Eur. Math. Soc. (JEMS) 21 (2019), no. 9, p. 2905–2944.
  • [BJT24] S. Boucksom, M. Jonsson & A. Trusiani – “Weighted extremal Kähler metrics on resolutions of singularities”, arXiv:2412.06096 (2024).
  • [BKR01] T. Bridgeland, A. King & M. Reid – “The McKay correspondence as an equivalence of derived categories”, J. Amer. Math. Soc. 14 (2001), no. 3, p. 535–554.
  • [CMTZ24] I. Cheltsov, L. Marquand, Y. Tschinkel & Z. Zhang – “Equivariant geometry of singular cubic threefolds, ii”, arXiv:2405.02744 (2024).
  • [CTZ25] I. Cheltsov, Y. Tschinkel & Z. Zhang – “Equivariant geometry of singular cubic threefolds”, Forum Math. Sigma 13 (2025), p. Paper No. e9.
  • [CC21a] X. Chen & J. Cheng – “On the constant scalar curvature Kähler metrics (I)—A priori estimates”, J. Amer. Math. Soc. 34 (2021), no. 4, p. 909–936.
  • [CC21b] by same author, “On the constant scalar curvature Kähler metrics (II)—Existence results”, J. Amer. Math. Soc. 34 (2021), no. 4, p. 937–1009.
  • [CDS15] X. Chen, S. Donaldson & S. Sun – “Kähler-Einstein metrics on Fano manifolds. I, II, III”, J. Amer. Math. Soc. 28 (2015), no. 1, p. 183–197, 199–234, 235–278.
  • [CC24] Y.-W. L. Cho & Y.-J. Choi – “Continuity of solutions to complex Monge-Ampère equations on compact Kähler spaces”, arXiv:2401.03935 (2024).
  • [Dar17] T. Darvas – “Metric geometry of normal Kähler spaces, energy properness, and existence of canonical metrics”, Int. Math. Res. Not. IMRN (2017), no. 22, p. 6752–6777.
  • [DR17] T. Darvas & Y. A. Rubinstein – “Tian’s properness conjectures and Finsler geometry of the space of Kähler metrics”, J. Amer. Math. Soc. 30 (2017), no. 2, p. 347–387.
  • [Dem85] J.-P. Demailly – “Mesures de Monge-Ampère et caractérisation géométrique des variétés algébriques affines”, Mém. Soc. Math. France (N.S.) (1985), no. 19, p. 124.
  • [Dem92] by same author, “Regularization of closed positive currents and intersection theory”, J. Algebraic Geom. 1 (1992), no. 3, p. 361–409.
  • [Dem12] by same author, Analytic methods in algebraic geometry, Surveys of Modern Mathematics, vol. 1, International Press, Somerville, MA; Higher Education Press, Beijing, 2012.
  • [DK01] J.-P. Demailly & J. Kollár – “Semi-continuity of complex singularity exponents and Kähler-Einstein metrics on Fano orbifolds”, Ann. Sci. École Norm. Sup. (4) 34 (2001), no. 4, p. 525–556.
  • [Der18] R. Dervan – “Relative K-stability for Kähler manifolds”, Math. Ann. 372 (2018), no. 3-4, p. 859–889.
  • [DR17] R. Dervan & J. Ross – “K-stability for Kähler manifolds”, Math. Res. Lett. 24 (2017), no. 3, p. 689–739.
  • [DG18] E. Di  Nezza & V. Guedj – “Geometry and topology of the space of Kähler metrics on singular varieties”, Compos. Math. 154 (2018), no. 8, p. 1593–1632.
  • [DGG23] E. Di  Nezza, V. Guedj & H. Guenancia – “Families of singular Kähler–Einstein metrics”, J. Eur. Math. Soc. (JEMS) 25 (2023), no. 7, p. 2697–2762.
  • [DJL24] E. Di  Nezza, S. Jubert & A. Lahdili – “Weighted cscK metrics (I): a priori estimates”, arXiv:2407.09929 (2024).
  • [DJL25] by same author, “Weighted cscK metric (II): the continuity method”, arXiv:2503.22183 (2025).
  • [DG18] S. Druel & H. Guenancia – “A decomposition theorem for smoothable varieties with trivial canonical class”, J. Éc. polytech. Math. 5 (2018), p. 117–147.
  • [EGZ09] P. Eyssidieux, V. Guedj & A. Zeriahi – “Singular Kähler-Einstein metrics”, J. Amer. Math. Soc. 22 (2009), no. 3, p. 607–639.
  • [FM95] A. Futaki & T. Mabuchi – “Bilinear forms and extremal Kähler vector fields associated with Kähler classes”, Math. Ann. 301 (1995), no. 2, p. 199–210.
  • [Gau20] P. Gauduchon – “ Calabi’s extremal Kähler metrics: an elementary introduction”, Lecture Notes (2020).
  • [GZ15] Q. Guan & X. Zhou – “A proof of Demailly’s strong openness conjecture”, Ann. of Math. (2) 182 (2015), no. 2, p. 605–616.
  • [GP24] B. Guo & D. H. Phong – “Uniform entropy and energy bounds for fully non-linear equations”, Comm. Anal. Geom. 32 (2024), no. 8, p. 2305–2325.
  • [HL25] J. Han & Y. Liu – “On the existence of weighted-cscK metrics”, Adv. Math. 463 (2025), p. Paper No. 110125, 54.
  • [He19] W. He – “On Calabi’s extremal metric and properness”, Trans. Amer. Math. Soc. 372 (2019), no. 8, p. 5595–5619.
  • [Ino22] E. Inoue – “Constant μ\mu-scalar curvature Kähler metric—formulation and foundational results”, J. Geom. Anal. 32 (2022), no. 5, p. Paper No. 145, 53.
  • [Koi83] N. Koiso – “Einstein metrics and complex structures”, Invent. Math. 73 (1983), no. 1, p. 71–106.
  • [Kol13] J. KollárSingularities of the minimal model program, Cambridge Tracts in Mathematics, vol. 200, Cambridge University Press, Cambridge, 2013, With a collaboration of Sándor Kovács.
  • [Koł98] S. Kołodziej – “The complex Monge-Ampère equation”, Acta Math. 180 (1998), no. 1, p. 69–117.
  • [Lah19] A. Lahdili – “Kähler metrics with constant weighted scalar curvature and weighted K-stability”, Proc. Lond. Math. Soc. (3) 119 (2019), no. 4, p. 1065–1114.
  • [Lah23] by same author, “Convexity of the weighted Mabuchi functional and the uniqueness of weighted extremal metrics”, Math. Res. Lett. 30 (2023), no. 2, p. 541–576.
  • [LS94] C. LeBrun & S. R. Simanca – “Extremal Kähler metrics and complex deformation theory”, Geom. Funct. Anal. 4 (1994), no. 3, p. 298–336.
  • [Li22a] C. Li – “GG-uniform stability and Kähler-Einstein metrics on Fano varieties”, Invent. Math. 227 (2022), no. 2, p. 661–744.
  • [Li22b] by same author, “Geodesic rays and stability in the cscK problem”, Ann. Sci. Éc. Norm. Supér. (4) 55 (2022), no. 6, p. 1529–1574.
  • [LTW21] C. Li, G. Tian & F. Wang – “On the Yau-Tian-Donaldson conjecture for singular Fano varieties”, Comm. Pure Appl. Math. 74 (2021), no. 8, p. 1748–1800.
  • [LTW22] by same author, “The uniform version of Yau-Tian-Donaldson conjecture for singular Fano varieties”, Peking Math. J. 5 (2022), no. 2, p. 383–426.
  • [LWX19] C. Li, X. Wang & C. Xu – “On the proper moduli spaces of smoothable Kähler-Einstein Fano varieties”, Duke Math. J. 168 (2019), no. 8, p. 1387–1459.
  • [Liu22] Y. Liu – “K-stability of cubic fourfolds”, J. Reine Angew. Math. 786 (2022), p. 55–77.
  • [LX19] Y. Liu & C. Xu – “K-stability of cubic threefolds”, Duke Math. J. 168 (2019), no. 11, p. 2029–2073.
  • [MP24] P. Mesquita-Piccione – “A non-Archimedean theory of complex spaces and the cscK problem”, arXiv:2409.06221 (2024).
  • [PTT23] C.-M. Pan, T. D. Tô & A. Trusiani – “Singular cscK metrics on smoothable varieties”, arXiv:2312.13653 (2023).
  • [PT25] C.-M. Pan & A. Trusiani – “Kähler–Einstein metrics on families of Fano varieties”, J. Reine Angew. Math. (2025), no. 819, p. 45–87.
  • [RZ11a] X. Rong & Y. Zhang – “Continuity of extremal transitions and flops for Calabi-Yau manifolds”, J. Differential Geom. 89 (2011), no. 2, p. 233–269, Appendix B by Mark Gross.
  • [RZ11b] W.-D. Ruan & Y. Zhang – “Convergence of Calabi-Yau manifolds”, Adv. Math. 228 (2011), no. 3, p. 1543–1589.
  • [Siu87] Y. T. SiuLectures on Hermitian-Einstein metrics for stable bundles and Kähler-Einstein metrics, DMV Seminar, vol. 8, Birkhäuser Verlag, Basel, 1987.
  • [SD18] Z. Sjöström Dyrefelt – “K-semistability of cscK manifolds with transcendental cohomology class”, J. Geom. Anal. 28 (2018), no. 4, p. 2927–2960.
  • [SD20] by same author, “On K-polystability of cscK manifolds with transcendental cohomology class”, Int. Math. Res. Not. IMRN (2020), no. 9, p. 2769–2817, With an appendix by Ruadhaí Dervan.
  • [SSY16] C. Spotti, S. Sun & C. Yao – “Existence and deformations of Kähler-Einstein metrics on smoothable \mathbb{Q}-Fano varieties”, Duke Math. J. 165 (2016), no. 16, p. 3043–3083.
  • [Szé15] G. Székelyhidi – “Blowing up extremal Kähler manifolds II”, Invent. Math. 200 (2015), no. 3, p. 925–977.
  • [Szé24] by same author, “Singular Kähler-Einstein metrics and RCD spaces”, arXiv:2408.10747 (2024).
  • [Uen75] K. UenoClassification theory of algebraic varieties and compact complex spaces, Lecture Notes in Mathematics, Vol. 439, Springer-Verlag, Berlin-New York, 1975, Notes written in collaboration with P. Cherenack.
  • [Yau78] S.-T. Yau – “On the Ricci curvature of a compact Kähler manifold and the complex Monge-Ampère equation. I”, Comm. Pure Appl. Math. 31 (1978), no. 3, p. 339–411.