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

\marginsize

2.6cm2.6cm2cm2cm

Relative pluripotential theory on compact Kähler manifolds

Tamás Darvas (University of Maryland)
Eleonora Di Nezza (Sorbonne Université)
Chinh H. Lu (Université d’Angers)
(To the memory of Jean–Pierre Demailly (1957-2022))

Abstract

Given a compact Kähler manifold, we survey the study of complex Monge–Ampère type equations with prescribed singularity type, developed by the authors in a series of papers. In addition, we give a general answer to a question of Guedj–Zeriahi about the finite energy range of the complex Monge–Ampère operator.

Chapter 1 Main results

Let (X,ω)(X,\omega) be a compact Kähler manifold of complex dimension nn, and let θ\theta be another smooth Kähler form. We consider the complex Monge–Ampère equation

(θ+ddcu)n=fωn.(\theta+dd^{c}u)^{n}=f\omega^{n}. (1.1)

When ff is smooth and positive, Yau [Yau78], proved that this equation admits a unique smooth solution, resolving the famous Calabi conjecture. This landmark result is one of the cornerstones for studying canonical Kähler metrics.

One can consider conical/edge behaviour along a divisor for solutions of (1.1). This is relevant in birational geometry and in the study of K-stability. It has been explored in many works, including [Don11, JMR16, GP16].
When f0f\geq 0, fLp(ωn),p>1,f\in L^{p}(\omega^{n}),\ p>1, and uu is allowed to be bounded θ\theta-plurisubharmonic, a solution to (1.1) exists by deep estimates of Kołodziej [Koł98, Koł03]. Furthermore, if one replaces fωnf\omega^{n} by a suitable Radon measure, it was shown in [GZ07] that solutions to (1.1) exist in a suitable finite energy space, mirroring results of Cegrell [Ceg98] in the local case.

When the cohomology class {θ}\{\theta\} is allowed to be degenerate, solutions to (1.1) and related equations have been explored in [EGZ09, BEGZ10, BBGZ13].

There have been numerous other extensions and generalizations in recent years. In this paper, we survey recent work done in [DDL18a, DDL21], where we consider solutions of (1.1) that have a prescribed singularity profile. We also give a general answer to a question raised by Guedj–Zeriahi [GZ07].

To fix notation and terminology, we assume that θ\theta is a closed smooth real (1,1)(1,1)-form such that {θ}\{\theta\} is big, and let ϕPSH(X,θ)\phi\in\textup{PSH}(X,\theta). Let f0f\geq 0, fLp(ωn),p>1f\in L^{p}(\omega^{n}),\ p>1.

We are looking for solutions uPSH(X,θ)u\in\textup{PSH}(X,\theta) of (1.1) satisfying

ϕCuϕ+C, for some C.\phi-C\leq u\leq\phi+C,\textup{ for some }C\in\mathbb{R}. (1.2)

If φ\varphi and φ\varphi^{\prime} are two θ\theta-plurisubharmonic functions on XX, then φ\varphi^{\prime} is said to be less singular than φ\varphi, i.e. φφ\varphi\preceq\varphi^{\prime}, if they satisfy φφ+C\varphi\leq\varphi^{\prime}+C for some CC\in\mathbb{R}. We say that φ\varphi has the same singularity as φ\varphi^{\prime}, i.e. φφ\varphi\simeq\varphi^{\prime}, if φφ\varphi\preceq\varphi^{\prime} and φφ\varphi^{\prime}\preceq\varphi. The latter condition is easily seen to yield an equivalence relation, whose equivalence classes are denoted by [φ][\varphi], φPSH(X,θ)\varphi\in{\rm PSH}(X,\theta). Using this terminology (1.2) simply says that [u]=[ϕ][u]=[\phi].

Typically ϕ\phi is unbounded, hence so is uu, and the left-hand-side of (1.1) has to be interpreted as the non-pluripolar complex Monge-Ampère measure of uu, considered in [BEGZ10] (see (2.2) below):

θun:=(θ+ddcu)n.\theta_{u}^{n}:=\langle(\theta+dd^{c}u)^{n}\rangle.

As it turns out, this problem is well posed only for potentials ϕ\phi having “model” singularity type [ϕ][\phi], that includes the case of analytic singularities, treated in [PS14].

We need to consider the following notion of envelope, only dependent on the singularity type [ϕ][\phi]:

Pθ[ϕ]=sup{vPSH(X,θ) s.t. v0,[ϕ]=[v]}.P_{\theta}[\phi]=\sup\{v\in\textup{PSH}(X,\theta)\textup{ s.t. }v\leq 0,\ [\phi]=[v]\}.

Since ϕsupXϕPθ[ϕ]\phi-\sup_{X}\phi\leq P_{\theta}[\phi], we have that [ϕ][Pθ[ϕ]][\phi]\leq[P_{\theta}[\phi]] and typically equality does not happen. When [ϕ]=[Pθ[ϕ]][\phi]=[P_{\theta}[\phi]], we say that ϕ\phi has model singularity type.

We state our first main result, initially proved in [DDL21]:

Theorem 1.1.

Suppose ϕPSH(X,θ)\phi\in\textup{PSH}(X,\theta) and [ϕ]=[Pθ[ϕ]][\phi]=[P_{\theta}[\phi]]. Let fLp(ωn)f\in L^{p}(\omega^{n}), p>1p>1 such that f0f\geq 0 and Xfωn=Xθϕn>0\int_{X}f\omega^{n}=\int_{X}\theta_{\phi}^{n}>0. Then the following hold:
(i) There exists uPSH(X,θ)u\in\textup{PSH}(X,\theta), unique up to a constant, such that [u]=[ϕ][u]=[\phi] and

θun=fωn.\theta_{u}^{n}=f\omega^{n}. (1.3)

(ii) For any λ>0\lambda>0 there exists a unique vPSH(X,θ)v\in\textup{PSH}(X,\theta), such that [v]=[ϕ][v]=[\phi] and

θvn=eλvfωn.\theta_{v}^{n}=e^{\lambda v}f\omega^{n}. (1.4)

As it turns out, the condition [ϕ]=[P[ϕ]][\phi]=[P[\phi]] is very natural and necessary for well posedness in this context. Due to Theorem 5.22, if (1.3) has a unique solution for any choice of fLf\in L^{\infty}, then the condition [ϕ]=[P[ϕ]][\phi]=[P[\phi]] must hold.

To prove this theorem, one needs to build up the variational approach of [BEGZ10] relative to the model singularity type [ϕ][\phi], motivating the title of this survey.

When KX>0K_{X}>0 and λ=1\lambda=1 (or KXK_{X} trivial and λ=0\lambda=0) solutions of (1.4) can be interpreted as Kähler–Einstein metrics with prescribed singularity.

Remark 1.2.

Potentials ϕ\phi with analytic singularity type can be locally written as

ϕ=c2log(j|fj|2)+g,\phi=\frac{c}{2}\log\big{(}\sum_{j}|f_{j}|^{2}\big{)}+g,

where fjf_{j} are holomorphic, c+c\in\mathbb{Q}_{+} and gg is bounded. Such potentials are always of model type ([RW14, Remark 4.6], [RS05], see also Proposition 5.24). In particular, discrete logarithmic singularity types are of model type, making connection with pluricomplex Green currents [CG09, PS14, RS05].

Our reader may wonder if there are other interesting enough potentials with model singularity type. We believe this to be the case:

\bullet By Theorem 3.14 below, Pθ[ψ]=Pθ[Pθ[ψ]]P_{\theta}[\psi]=P_{\theta}[P_{\theta}[\psi]] for any ψPSH(X,θ)\psi\in\textup{PSH}(X,\theta) with Xθψn>0\int_{X}\theta_{\psi}^{n}>0. In particular, Pθ[ψ]P_{\theta}[\psi] is a model potential, giving an abundance of potentials with model type singularity.

\bullet By Proposition 5.23 below, if ψPSH(X,θ)\psi\in\textup{PSH}(X,\theta) has small unbounded locus, and θψn/ωnLp(ωn),p>1\theta_{\psi}^{n}/\omega^{n}\in L^{p}(\omega^{n}),\ p>1 with Xθψn>0\int_{X}\theta_{\psi}^{n}>0, then ψ\psi has model type singularity.

\bullet Due to [RW14, DDL18], potentials with model type singularity naturally arise as degenerations along geodesic rays and in particular along test configurations.

As an application to Theorem 1.1 we prove the log-concavity property for non-pluripolar products, initially conjectured in [BEGZ10], proved in [DDL21].

Theorem 1.3.

Let T1,,TnT_{1},...,T_{n} be positive closed (1,1)(1,1)-currents on a compact Kähler manifold XX. Then

XT1Tn(XT1n)1n(XTnn)1n.\int_{X}\langle T_{1}\wedge\ldots\wedge T_{n}\rangle\geq\left(\int_{X}\langle T_{1}^{n}\rangle\right)^{\frac{1}{n}}\ldots\left(\int_{X}\langle T_{n}^{n}\rangle\right)^{\frac{1}{n}}.

In particular, TlogXTnT\to\log\int_{X}\langle T^{n}\rangle is concave function on the space of positive (1,1)(1,1)-currents.

Next we consider the situation when the right hand side of (1.1) is a non-pluripolar measure (not necessarily of the type fωnf\omega^{n}), and answer a question of Guedj–Zeriahi in our relative context.

For this we need to introduce relative finite energy classes. Let ϕPSH(X,θ)\phi\in\textup{PSH}(X,\theta) such that ϕ=P[ϕ]\phi=P[\phi] and Xθϕn>0\int_{X}\theta_{\phi}^{n}>0.

By a foundational result of Witt Nyström [Wit19], if [u][ϕ][u]\leq[\phi] then XθunXθϕn\int_{X}\theta_{u}^{n}\leq\int_{X}\theta_{\phi}^{n} (see Theorem 3.3 below). We say that uu has relative full mass with respect to ϕ\phi (notation: u(X,θ,ϕ)u\in\mathcal{E}(X,\theta,\phi)) if this inequality is extremized, i.e., [u][ϕ][u]\leq[\phi] and Xθun=Xθϕn\int_{X}\theta_{u}^{n}=\int_{X}\theta_{\phi}^{n}.

It is argued in Theorem 5.17 that for any Radon measure μ\mu not charging pluripolar sets such that μ(X)=Xθϕn\mu(X)=\int_{X}\theta_{\phi}^{n}, there exists a unique solution u(X,θ,ϕ))u\in\mathcal{E}(X,\theta,\phi)) to the equation

θun=μ.\theta_{u}^{n}=\mu. (1.5)

As pointed out in [GZ07], for potential theoretic reasons, it is natural to consider weighted subspaces of (X,θ,ϕ)\mathcal{E}(X,\theta,\phi). A weight is a continuous increasing function χ:[0,)[0,)\chi:[0,\infty)\rightarrow[0,\infty) such that χ(0)=0\chi(0)=0 and limtχ(t)=\lim_{t\to\infty}\chi(t)=\infty.

Let us assume that the weight χ\chi satisfies the following condition

t0,λ1,χ(λt)λMχ(t),\forall t\geq 0,\;\forall\lambda\geq 1,\;\chi(\lambda t)\leq\lambda^{M}\chi(t), (1.6)

where M1M\geq 1 is a fixed constant. Let χ(X,θ,ϕ)\mathcal{E}_{\chi}(X,\theta,\phi) denote the set of all u(X,θ,ϕ)u\in\mathcal{E}(X,\theta,\phi) such that

Eχ(u,ϕ):=Xχ(|uϕ|)θun<.E_{\chi}(u,\phi):=\int_{X}\chi(|u-\phi|)\theta_{u}^{n}<\infty.

When ϕ=Vθ\phi=V_{\theta}, we denote (X,θ)=(X,θ,Vθ)\mathcal{E}(X,\theta)=\mathcal{E}(X,\theta,V_{\theta}), χ(X,θ)=χ(X,θ,Vθ\mathcal{E}_{\chi}(X,\theta)=\mathcal{E}_{\chi}(X,\theta,V_{\theta}) and Eχ(u)=Eχ(u,Vθ)E_{\chi}(u)=E_{\chi}(u,V_{\theta}). Compared to [GZ07], we have changed the sign of the weight, but the weighted classes are the same. Also, notice that both low and high energy classes of [GZ07] satisfy the condition (1.6), allowing our treatment to be a bit more universal.

One may ask, under what condition is the solution u(X,θ,ϕ)u\in\mathcal{E}(X,\theta,\phi) of (1.5) an element of χ(X,θ,ϕ)\mathcal{E}_{\chi}(X,\theta,\phi). The theorem below provides precise answers to this question, containing perhaps the only novel result of this work:

Theorem 1.4.

Le μ\mu be a Radon measure with μ({ϕ=})=0\mu(\{\phi=-\infty\})=0 and Xθϕn=μ(X)>0\int_{X}\theta_{\phi}^{n}=\mu(X)>0. For χ\chi satisfying (1.6) the following conditions are equivalent:

  1. (i)

    There exists a constant C>0C>0 such that, for all uχ(X,θ,ϕ)u\in\mathcal{E}_{\chi}(X,\theta,\phi) with supXu=0\sup_{X}u=0, we have

    Xχ(ϕu)𝑑μCEχ(u,ϕ)M/(M+1)+C.\int_{X}\chi(\phi-u)d\mu\leq CE_{\chi}(u,\phi)^{M/(M+1)}+C.
  2. (ii)

    χ(|ϕu|)L1(μ)\chi(|\phi-u|)\in L^{1}(\mu), for all uχ(X,θ,ϕ)u\in\mathcal{E}_{\chi}(X,\theta,\phi).

  3. (iii)

    μ=θφn\mu=\theta_{\varphi}^{n}, for some φχ(X,θ,ϕ)\varphi\in\mathcal{E}_{\chi}(X,\theta,\phi), with supXφ=0\sup_{X}\varphi=0.

For slightly less general χ\chi, the equivalence between (i) and (iii) was obtained in [DV21]. The condition μ({ϕ=})=0\mu(\{\phi=-\infty\})=0 is necessary. Without it, the μ\mu-integrability of χ(ϕu)\chi(\phi-u) can not be discussed, as the values of this function are not defined on the set {ϕ=}\{\phi=-\infty\}. Of course this condition is vacuous if ϕ=0\phi=0.

In the particular case when θ\theta is Kähler and ϕ=0\phi=0, the equivalence between (ii) and (iii) answers a question asked by Guedj–Zeriahi in the comments following [GZ07, Theorem 4.1]. In [GZ07, Theorem C] the authors obtain the above result for χ(t)=tp,p1\chi(t)=t^{p},\ p\geq 1.

Prerequisites.

An effort has been made to keep prerequisites at a minimum. However due to size constraints, such requirements on part of the reader are inevitable. We assume that our reader is familiar with the basics of Bedfor-Taylor theory [BT76, BT82], and finite energy pluripotential theory in the big case, as elaborated in [BEGZ10]. The recent book [GZ17] is a comprehensive source that can initiate novice readers into this subject.

Organization.

In Chapter 2 we recall terminology of finite energy pluripotential theory from [BEGZ10] and prove some preliminary results, touching [DDL18b, DT21] in the process. In Chapter 3 we prove the monotonicity theorem due to Witt Nyström [Wit19] and the authors [DDL18a] for non-pluripolar product masses. Here we also introduce and study the basic concepts of relative pluripotential theory.

In Chapter 4 we prove a result about comparison of capacities and prove an integrating by parts formula, making contact with [Xia19, Lu21, Vu21].

In Chapter 5, we finally solve the complex Monge–Ampère equations (1.3) and (1.4) using the variational method of [BBGZ13] as in [DDL18a]. Due to the availability of integration parts, the technical assumption of small unbounded locus from [DDL18a] can be avoided, making our treatment here more transparent, compared to the original arguments in [DDL18a, DDL21].

The proof of Theorem 1.4 is given in Chapter 6, containing the novel results of this survey.

Relation to other works.

Since [DDL18b, DDL18a, DDL18, DDL21] appeared, a number of works have taken up the topics of this survey, and developed it further. Due to size constraints we can not treat these here, but let us mention a few exciting directions.

The work [DDL21a] introduced a pseudo-metric on the space of singularity types and studied the stability of solutions to (1.3), as the singularity type is varied. More precise results have been recently obtained in [DV22]. Connections with Lelong numbers have been studied in [Vu20].

The works [DX21, DZ22, DXZ23] used relative pluripotential theory to study partial Bergman kernels, K-stability and the Ross–Witt Nyström correspondence.

The works [Xia19a, Tru19, Gup22] have explored the metric geometry of the relative finite energy classes surveyed in this work.

Finally, A. Trusiani started an elaborate study of Kahler–Einstein metrics with prescribed singularity type in the Fano case [Tru20, Tru23].

Acknowledgments.

The first named author was partially supported by an Alfred P. Sloan Fellowship and National Science Foundation grant DMS–1846942. The second author is supported by the project SiGMA ANR-22-ERCS-0004-02. The third named author is partially supported by Centre Henri Lebesgue ANR-11-LABX-0020-01. The second and third named authors are partially supported by the project PARAPLUI ANR-20-CE40-0019.

Chapter 2 Preliminaries

We recall results from pluripotential theory of big cohomology classes, particularly results about non-pluripolar complex Monge–Ampere measures. We borrow notation and terminology from [BEGZ10], and we refer to this work for further details.

Let (X,ω)(X,\omega) be a compact Kähler manifold of dimension nn. Let θ\theta be a smooth closed (1,1)(1,1)-form on XX such that {θ}\{\theta\} is big, i.e., there exists ψPSH(X,θ)\psi\in{\rm PSH}(X,\theta) such that θ+ddcψεω\theta+dd^{c}\psi\geq\varepsilon\omega for some small constant ε>0\varepsilon>0. Here, dd and dcd^{c} are real differential operators defined as d:=+¯,dc:=i2π(¯).d:=\partial+\bar{\partial},\,d^{c}:=\frac{i}{2\pi}\left(\bar{\partial}-\partial\right). A function φ:X{}\varphi:X\rightarrow\mathbb{R}\cup\{-\infty\} is quasi-plurisubharmonic (qpsh) if it can be locally written as the sum of a plurisubharmonic function and a smooth function. φ\varphi is called θ\theta-plurisubharmonic (θ\theta-psh) if it is qpsh and θ+ddcφ0\theta+dd^{c}\varphi\geq 0 in the sense of currents. We let PSH(X,θ){\rm PSH}(X,\theta) denote the set of θ\theta-psh functions that are not identically -\infty.

A θ\theta-psh function φ\varphi is said to have analytic singularity type if there exists a constant c>0c>0 such that locally on XX,

φ=c2logj=1N|fj|2+g,\varphi=\frac{c}{2}\log\sum_{j=1}^{N}|f_{j}|^{2}+g,

where gg is bounded and f1,,fNf_{1},\dots,f_{N} are local holomorphic functions. The ample locus Amp({θ}){\rm Amp}(\{\theta\}) of {θ}\{\theta\} is the set of points xXx\in X such that there exists a Kähler current T{θ}T\in\{\theta\} with analytic singularity type and smooth in a neighbourhood of xx. The ample locus Amp({θ}){\rm Amp}(\{\theta\}) is a Zariski open subset, and it is nonempty [Bou04].

Let xXx\in X. Fixing a holomorphic chart xUXx\in U\subset X, the Lelong number ν(φ,x)\nu(\varphi,x) of φPSH(X,θ)\varphi\in\textup{PSH}(X,\theta) is defined as follows:

ν(φ,x)=sup{γ0 s.t. φ(z)γlogzx+O(1) on U}.\nu(\varphi,x)=\sup\{\gamma\geq 0\textup{ s.t. }\varphi(z)\leq\gamma\log\|z-x\|+O(1)\textup{ on }U\}. (2.1)

One can also associate to φ\varphi a multiplier ideal sheaf (φ)\mathcal{I}(\varphi) whose germs are holomorphic functions ff for which |f|2eφ|f|^{2}e^{-\varphi} is integrable.

If φ\varphi and φ\varphi^{\prime} are two θ\theta-psh functions on XX, then φ\varphi^{\prime} is said to be less singular than φ\varphi, i.e. φφ\varphi\preceq\varphi^{\prime}, if they satisfy φφ+C\varphi\leq\varphi^{\prime}+C for some CC\in\mathbb{R}. We say that φ\varphi has the same singularity as φ\varphi^{\prime}, i.e. φφ\varphi\simeq\varphi^{\prime}, if φφ\varphi\preceq\varphi^{\prime} and φφ\varphi^{\prime}\preceq\varphi. The latter condition is easily seen to yield an equivalence relation, whose equivalence classes are denoted by [φ][\varphi], φPSH(X,θ)\varphi\in{\rm PSH}(X,\theta).

A θ\theta-psh function φ\varphi is said to have minimal singularity type if it is less singular than any other θ\theta-psh function. Such θ\theta-psh functions with minimal singularity type always exist, one can consider for example

Vθ:=sup{φθ-psh,φ0 on X}.V_{\theta}:=\sup\left\{\varphi\,\,\theta\text{-psh},\varphi\leq 0\text{ on }X\right\}.

Trivially, a θ\theta-psh function with minimal singularity type is locally bounded in Amp({θ}){\rm Amp}(\{\theta\}). It follows from [DT23, Theorem 1.1] that VθV_{\theta} is C1,1¯C^{1,\bar{1}} in the ample locus Amp({θ}){\rm Amp}(\{\theta\}).

Given θ1+ddcφ1,,\theta^{1}+dd^{c}\varphi_{1},..., θp+ddcφp\theta^{p}+dd^{c}\varphi_{p} positive (1,1)(1,1)-currents, where θj\theta^{j} are closed smooth real (1,1)(1,1)-forms, following the construction of Bedford-Taylor [BT87] in the local setting, it has been shown in [BEGZ10] that the sequence of currents

𝟏j{φj>Vθjk}(θ1+ddcmax(φ1,Vθ1k))(θp+ddcmax(φp,Vθpk)){\bf 1}_{\bigcap_{j}\{\varphi_{j}>V_{\theta_{j}}-k\}}(\theta^{1}+dd^{c}\max(\varphi_{1},V_{\theta_{1}}-k))\wedge...\wedge(\theta^{p}+dd^{c}\max(\varphi_{p},V_{\theta_{p}}-k))

is non-decreasing in kk and converges weakly to the so called non-pluripolar product

θφ11θφpp.\langle\theta^{1}_{\varphi_{1}}\wedge\ldots\wedge\theta^{p}_{\varphi_{p}}\rangle. (2.2)

In the following, with a slight abuse of notation, we will denote the non-pluripolar product simply by θφ11θφpp\theta^{1}_{\varphi_{1}}\wedge\ldots\wedge\theta^{p}_{\varphi_{p}}. When p=np=n, the resulting positive (n,n)(n,n)-current is a Borel measure that does not charge pluripolar sets. Pluripolar sets are Borel measurable sets that are contained in some set {ψ=}\{\psi=-\infty\}, where ψPSH(X,θ)\psi\in\textup{PSH}(X,\theta).

For a θ\theta-psh function φ\varphi, the non-pluripolar complex Monge-Ampère measure of φ\varphi is

θφn:=(θ+ddcφ)n.\theta_{\varphi}^{n}:=\langle(\theta+dd^{c}\varphi)^{n}\rangle.

The volume of a big class {θ}\{\theta\} is defined by

Vol({θ}):=Amp({θ})θVθn.{\rm Vol}(\{\theta\}):=\int_{{\rm Amp}(\{\theta\})}\theta_{V_{\theta}}^{n}.

Alternatively, by [BEGZ10, Theorem 1.16], in the above expression one can replace VθV_{\theta} with any θ\theta-psh function with minimal singularity type. A θ\theta-psh function φ\varphi is said to have full Monge–Ampère mass if

Xθφn=Vol({θ}),\int_{X}\theta_{\varphi}^{n}={\rm Vol}(\{\theta\}),

and we then write φ(X,θ)\varphi\in\mathcal{E}(X,\theta).

An important property of the non-pluripolar product is that it is local with respect to the plurifine topology (see [BT87, Corollary 4.3],[BEGZ10, Section 1.2]). This topology is the coarsest such that all qpsh functions with values in \mathbb{R} are continuous. For convenience we record the following version of this result for later use.

Lemma 2.1.

Fix closed smooth big (1,1)(1,1)-forms θ1,,θn\theta^{1},...,\theta^{n}. Assume that φj,ψj,j=1,,n\varphi_{j},\psi_{j},j=1,...,n are θj\theta^{j}-psh functions such that φj=ψj\varphi_{j}=\psi_{j} on UU an open set in the plurifine topology. Then

𝟏Uθφ11θφnn=𝟏Uθψ11θψnn.{\bf 1}_{U}\theta^{1}_{\varphi_{1}}\wedge...\wedge\theta^{n}_{\varphi_{n}}={\bf 1}_{U}\theta^{1}_{\psi_{1}}\wedge...\wedge\theta^{n}_{\psi_{n}}.

Lemma 2.1 will be referred to as the plurifine locality property. We will often work with sets of the form {u<v}\{u<v\}, where u,vu,v are quasi-psh functions. These are always open in the plurifine topology.

Convergence theorems.

The Monge–Ampère capacity of a Borel set EXE\subset X is defined as

Capω(E):=sup{E(ω+ddcu)n:uPSH(X,ω),1u0}.{\rm Cap}_{\omega}(E):=\sup\left\{\int_{E}(\omega+dd^{c}u)^{n}\;:\;u\in{\rm PSH}(X,\omega),\;-1\leq u\leq 0\right\}.

A function uu is called quasi-continuous if for each ε>0\varepsilon>0, there exists an open set UU such that Capω(U)<ε{\rm Cap}_{\omega}(U)<\varepsilon and the restriction of uu on XUX\setminus U is continuous.

A sequence of functions uju_{j} converges in capacity to uu if, for any δ>0\delta>0,

limjCapω({xX:|uj(x)u(x)|>δ})=0.\lim_{j\to\infty}{\rm Cap}_{\omega}(\{x\in X\;:\;|u_{j}(x)-u(x)|>\delta\})=0.

We recall a classical convergence theorem from Bedford-Taylor theory. We refer to [GZ17, Theorem 4.26] for a proof of this result, which is a slight generalization of [Xin96, Theorem 1].

Proposition 2.2.

Let UnU\subset\mathbb{C}^{n} be an open set. Suppose {fj}j\{f_{j}\}_{j} are uniformly bounded quasi-continuous functions which converge in capacity to another quasi-continuous function ff on UU. Let {u1j}j,{u2j}j,,{unj}j\{u^{j}_{1}\}_{j},\{u^{j}_{2}\}_{j},\ldots,\{u^{j}_{n}\}_{j} be uniformly bounded plurisubharmonic functions on Ω\Omega, converging in capacity to u1,u2,,unu_{1},u_{2},\ldots,u_{n} respectively. Then we have the following weak convergence of measures:

fji¯u1ji¯u2ji¯unjfi¯u1i¯u2i¯un.f_{j}i\partial\bar{\partial}u_{1}^{j}\wedge i\partial\bar{\partial}u_{2}^{j}\wedge\ldots\wedge i\partial\bar{\partial}u_{n}^{j}\to fi\partial\bar{\partial}u_{1}\wedge i\partial\bar{\partial}u_{2}\wedge\ldots\wedge i\partial\bar{\partial}u_{n}.
Definition 2.3.

A Borel set EE is called quasi-open if for each ε>0\varepsilon>0, there exists an open set UU such that

Capω((UE)(EU))ε.{\rm Cap}_{\omega}((U\setminus E)\cup(E\setminus U))\leq\varepsilon.

Quasi-closed sets are defined similarly.

As mentioned above, we will often work with sets of the form {u<v}\{u<v\} where uu and vv are quasi-psh functions. In general these sets are not open but merely quasi-open.

If a sequence of positive measures μj\mu_{j} converges weakly in UU to a positive measure μ\mu then a elementary argument shows that if EE is open and VV is closed then

lim infjμj(E)μ(E),lim supjμj(V)μ(V).\liminf_{j\to\infty}\mu_{j}(E)\geq\mu(E),\qquad\limsup_{j\to\infty}\mu_{j}(V)\leq\mu(V).

Using the above facts one can easily argue the following result:

Lemma 2.4.

Assume uju_{j} is a sequence of uniformly bounded θ\theta-psh functions in UXU\subset X converging in capacity to a θ\theta-psh function uu. Suppose {fj}j\{f_{j}\}_{j} are uniformly bounded non-negative quasi-continuous functions which converge in capacity to another quasi-continuous function f0f\geq 0 on UU.
If EUE\subset U is a quasi-open set then

lim infjEfj(θ+ddcuj)nEf(θ+ddcu)n.\liminf_{j\to\infty}\int_{E}f_{j}(\theta+dd^{c}u_{j})^{n}\geq\int_{E}f(\theta+dd^{c}u)^{n}.

If VUV\subset U is a quasi-closed set then

lim supjVfj(θ+ddcuj)nVf(θ+ddcu)n.\limsup_{j\to\infty}\int_{V}f_{j}(\theta+dd^{c}u_{j})^{n}\leq\int_{V}f(\theta+dd^{c}u)^{n}.

We will also need the following basic convergence result:

Lemma 2.5.

Assume that μj\mu_{j} is a sequence of positive Borel measures converging weakly to μ\mu. Assume that there exists a continuous function f:[0,)[0,)f:[0,\infty)\rightarrow[0,\infty) with f(0)=0f(0)=0 such that, for any Borel set EE,

μj(E)+μ(E)f(Capω(E)).\mu_{j}(E)+\mu(E)\leq f\left({\rm Cap}_{\omega}(E)\right).

Let uju_{j} be a sequence of uniformly bounded quasi-continuous functions which converges in capacity to a bounded quasi-continuous function uu. Then ujμjuμu_{j}\mu_{j}\rightarrow u\mu in the sense of Radon measures on XX.

Proof.

Fixing ε>0\varepsilon>0 there exists a continuous function vv on XX such that

Capω({xX:u(x)v(x)})<ε.{\rm Cap}_{\omega}(\{x\in X\;:u(x)\neq v(x)\})<\varepsilon.

Let A>0A>0 be a constant such that |uj|+|u|+|v|A|u_{j}|+|u|+|v|\leq A on XX. Fix δ>0\delta>0. For j>Nj>N large enough we have, by the assumption that uju_{j} converges in capacity to uu, that

Capω({xX:|uj(x)u(x)|>δ)<ε.{\rm Cap}_{\omega}(\{x\in X\;:\;|u_{j}(x)-u(x)|>\delta)<\varepsilon.

Fixing a continuous function χ\chi and j>Nj>N, it follows from the above that

|X(χujμjχudμ)|X|χ(uju)|μj+|Xχu(μjμ)|\displaystyle|\int_{X}(\chi u_{j}\mu_{j}-\chi ud\mu)|\leq\int_{X}|\chi(u_{j}-u)|\mu_{j}+|\int_{X}\chi u(\mu_{j}-\mu)|
δX|χ|μj+AsupX|χ|μj({|uju|>δ})\displaystyle\leq\delta\int_{X}|\chi|\mu_{j}+A\sup_{X}|\chi|\mu_{j}(\{|u_{j}-u|>\delta\})
+|Xχ(uv)(μjμ)|+|Xχv(μjμ)|\displaystyle+|\int_{X}\chi(u-v)(\mu_{j}-\mu)|+|\int_{X}\chi v(\mu_{j}-\mu)|
δX|χ|μj+AsupX|χ|f(ε)+|Xχ(uv)(μjμ)|+|Xχv(μjμ)|\displaystyle\leq\delta\int_{X}|\chi|\mu_{j}+A\sup_{X}|\chi|f(\varepsilon)+|\int_{X}\chi(u-v)(\mu_{j}-\mu)|+|\int_{X}\chi v(\mu_{j}-\mu)|
δX|χ|μj+2AsupX|χ|f(ε)+|Xχv(μjμ)|.\displaystyle\leq\delta\int_{X}|\chi|\mu_{j}+2A\sup_{X}|\chi|f(\varepsilon)+|\int_{X}\chi v(\mu_{j}-\mu)|.

Since vv is continuous on XX the last term converges to 0 as jj\to\infty. This completes the proof. ∎

The following lower-semicontinuity property of non-pluripolar products from [DDL18a] will be key in the sequel:

Theorem 2.6.

Let θj,j{1,,n}\theta^{j},j\in\{1,\ldots,n\} be smooth closed real (1,1)(1,1)-forms on XX whose cohomology classes are big. Suppose that for all j{1,,n}j\in\{1,\ldots,n\} we have uj,ujkPSH(X,θj)u_{j},u_{j}^{k}\in\textup{PSH}(X,\theta^{j}) such that ujkuju^{k}_{j}\to u_{j} in capacity as kk\to\infty, and let χk,χ0\chi_{k},\chi\geq 0 be quasi-continuous and uniformly bounded such that χkχ\chi_{k}\to\chi in capacity. Then

lim infkXχkθu1k1θunknXχθu11θunn.\liminf_{k\to\infty}\int_{X}\chi_{k}\theta^{1}_{u^{k}_{1}}\wedge\ldots\wedge\theta^{n}_{u^{k}_{n}}\geq\int_{X}\chi\theta^{1}_{u_{1}}\wedge\ldots\wedge\theta^{n}_{u_{n}}. (2.3)

If additionally,

Xθu11θunnlim supkXθu1k1θunkn,\int_{X}\theta^{1}_{u_{1}}\wedge\ldots\wedge\theta^{n}_{u_{n}}\geq\limsup_{k\rightarrow\infty}\int_{X}\theta^{1}_{u^{k}_{1}}\wedge\ldots\wedge\theta^{n}_{u^{k}_{n}}, (2.4)

then χkθu1k1θunknχθu11θunn\chi_{k}\theta^{1}_{u^{k}_{1}}\wedge\ldots\wedge\theta^{n}_{u^{k}_{n}}\to\chi\theta^{1}_{u_{1}}\wedge\ldots\wedge\theta^{n}_{u_{n}} in the weak sense of measures on XX.

Proof.

Set Ω:=j=1nAmp(θj)\Omega:=\bigcap_{j=1}^{n}{\rm Amp}({\theta^{j}}) and fix an open relatively compact subset UU of Ω\Omega. Then the functions VθjV_{\theta^{j}} are bounded on UU. We now use a classical idea in pluripotential theory. Fix C>0,ε>0C>0,\varepsilon>0 and consider

fjk,C,ε:=max(ujkVθj+C,0)max(ujkVθj+C,0)+ε,j=1,,n,k,f_{j}^{k,C,\varepsilon}:=\frac{\max(u_{j}^{k}-V_{\theta^{j}}+C,0)}{\max(u_{j}^{k}-V_{\theta^{j}}+C,0)+\varepsilon},\ j=1,...,n,\ k\in\mathbb{N}^{*},

and

ujk,C:=max(ujk,VθjC).u_{j}^{k,C}:=\max(u_{j}^{k},V_{\theta^{j}}-C).

Observe that for C,jC,j fixed, the functions ujk,CVθjCu_{j}^{k,C}\geq V_{\theta^{j}}-C are uniformly bounded in UU (since VθjV_{\theta^{j}} is bounded in UU) and converge in capacity to ujCu_{j}^{C} as kk\to\infty. Moreover, fjk,C,ε=0f_{j}^{k,C,\varepsilon}=0 if ujkVθjCu_{j}^{k}\leq V_{\theta^{j}}-C. By locality of the non-pluripolar product we can write

fk,C,εχkθu1k1θunkn=fk,C,εχkθu1k,C1θunk,Cn,f^{k,C,\varepsilon}\chi_{k}\theta^{1}_{u^{k}_{1}}\wedge\ldots\wedge\theta^{n}_{u^{k}_{n}}=f^{k,C,\varepsilon}\chi_{k}\theta^{1}_{u^{k,C}_{1}}\wedge\ldots\wedge\theta^{n}_{u^{k,C}_{n}},

where fk,C,ε=f1k,C,εfnk,C,εf^{k,C,\varepsilon}=f_{1}^{k,C,\varepsilon}\cdots f_{n}^{k,C,\varepsilon}. For each C,εC,\varepsilon fixed the functions fk,C,εf^{k,C,\varepsilon} are quasi-continuous, uniformly bounded (with values in [0,1][0,1]) and converge in capacity to fC,ε:=f1C,εfnC,εf^{C,\varepsilon}:=f_{1}^{C,\varepsilon}\cdots f_{n}^{C,\varepsilon}, where fjC,εf_{j}^{C,\varepsilon} is defined by

fjC,ε:=max(ujVθj+C,0)max(ujVθj+C,0)+ε.f_{j}^{C,\varepsilon}:=\frac{\max(u_{j}-V_{\theta^{j}}+C,0)}{\max(u_{j}-V_{\theta^{j}}+C,0)+\varepsilon}.

With the information above we can apply Proposition 2.2 to get that

fk,C,εχkθu1k,C1θunk,CnfC,εχθu1C1θunCnask,f^{k,C,\varepsilon}\chi_{k}\theta^{1}_{u^{k,C}_{1}}\wedge\ldots\wedge\theta^{n}_{u^{k,C}_{n}}\rightarrow f^{C,\varepsilon}\chi\theta^{1}_{u^{C}_{1}}\wedge\ldots\wedge\theta^{n}_{u^{C}_{n}}\ \textrm{as}\ k\to\infty,

in the weak sense of measures on UU. In particular since 0fk,C,ε10\leq f^{k,C,\varepsilon}\leq 1 we have that

lim infkXχkθu1k1θunkn\displaystyle\liminf_{k\to\infty}\int_{X}\chi_{k}\theta^{1}_{u^{k}_{1}}\wedge\ldots\wedge\theta^{n}_{u^{k}_{n}} \displaystyle\geq lim infkUfk,C,εχθu1k,C1θunk,Cn\displaystyle\liminf_{k\to\infty}\int_{U}f^{k,C,\varepsilon}\chi\theta^{1}_{u^{k,C}_{1}}\wedge\ldots\wedge\theta^{n}_{u^{k,C}_{n}}
\displaystyle\geq UfC,εχθu1C1θunCn.\displaystyle\int_{U}f^{C,\varepsilon}\chi\theta^{1}_{u^{C}_{1}}\wedge\ldots\wedge\theta^{n}_{u^{C}_{n}}.

Now, letting ε0\varepsilon\to 0 and then CC\to\infty, by definition of the non-pluripolar product we obtain

lim infkXχkθu1k1θunknUχθu11θunn.\liminf_{k\to\infty}\int_{X}\chi_{k}\theta^{1}_{u^{k}_{1}}\wedge\ldots\wedge\theta^{n}_{u^{k}_{n}}\geq\int_{U}\chi\theta^{1}_{u_{1}}\wedge\ldots\wedge\theta^{n}_{u_{n}}.

Finally, letting UU increase to Ω\Omega and noting that the complement of Ω\Omega is pluripolar, we conclude the proof of the first statement of the theorem.

To prove the last statement, we first notice that we actually have equality in (2.4) and the limsup is a lim, as one can just plug χ=1\chi=1 in (2.3).

Now let BB\in\mathbb{R} such that χ,χkB\chi,\chi_{k}\leq B. By (2.3) we get that

lim infkX(Bχk)θu1k1θunknX(Bχ)θu11θunn.\liminf_{k\to\infty}\int_{X}(B-\chi_{k})\theta^{1}_{u^{k}_{1}}\wedge\ldots\wedge\theta^{n}_{u^{k}_{n}}\geq\int_{X}(B-\chi)\theta^{1}_{u_{1}}\wedge\ldots\wedge\theta^{n}_{u_{n}}.

Flipping the signs and using (equality in) (2.4), we conclude the following inequality, finishing the proof:

lim supkXχkθu1k1θunknXχθu11θunn.\limsup_{k\to\infty}\int_{X}\chi_{k}\theta^{1}_{u^{k}_{1}}\wedge\ldots\wedge\theta^{n}_{u^{k}_{n}}\leq\int_{X}\chi\theta^{1}_{u_{1}}\wedge\ldots\wedge\theta^{n}_{u_{n}}.

Envelopes.

If ff is a function on XX, we define the envelope of ff in the class PSH(X,θ){\rm PSH}(X,\theta) by

Pθ(f):=(sup{uPSH(X,θ):uf}),P_{\theta}(f):=\left(\sup\{u\in{\rm PSH}(X,\theta)\;:\;u\leq f\}\right)^{*},

with the convention that sup=\sup\emptyset=-\infty. Observe that Pθ(f)PSH(X,θ)P_{\theta}(f)\in{\rm PSH}(X,\theta) if and only if there exists some uPSH(X,θ)u\in{\rm PSH}(X,\theta) lying below ff. Note also that Vθ=Pθ(0)V_{\theta}=P_{\theta}(0), and that Pθ(f+C)=Pθ(f)+CP_{\theta}(f+C)=P_{\theta}(f)+C for any constant CC.

In the particular case f=min(ψ,ϕ)f=\min(\psi,\phi), we denote the envelope as Pθ(ψ,ϕ):=Pθ(min(ψ,ϕ))P_{\theta}(\psi,\phi):=P_{\theta}(\min(\psi,\phi)). We observe that Pθ(ψ,ϕ)=Pθ(Pθ(ψ),Pθ(ϕ))P_{\theta}(\psi,\phi)=P_{\theta}(P_{\theta}(\psi),P_{\theta}(\phi)), so w.l.o.g. we can assume ψ,ϕ\psi,\phi are two θ\theta-psh functions.

In our first technical result about envelopes, we show that the mass θPθ(f)n\theta_{P_{\theta}(f)}^{n} is concentrated on the contact set {Pθ(f)=f}\{P_{\theta}(f)=f\}:

Theorem 2.7.

Assume ff is quasi-continuous on XX and Pθ(f)PSH(X,θ)P_{\theta}(f)\in{\rm PSH}(X,\theta). Then

{Pθ(f)<f}(θ+ddcPθ(f))n=0.\int_{\{P_{\theta}(f)<f\}}(\theta+dd^{c}P_{\theta}(f))^{n}=0.
Proof.

We can assume that θω\theta\leq\omega. Since Pθ(f)CP_{\theta}(f)\leq C is bounded from above, by replacing ff by min(f,C)\min(f,C) we can assume that ff is bounded from above. Shifting ff by a constant we can also assume f0f\leq 0.

Step 1. We assume that ff is bounded from below fC0f\geq-C_{0}. For each j1j\geq 1, there is an open set UjXU_{j}\subset X such that Capω(Uj)2j1{\rm Cap}_{\omega}(U_{j})\leq 2^{-j-1} and the restriction of ff on XUjX\setminus U_{j} is continuous. By taking kjUk\cup_{k\geq j}U_{k} we can assume that the sequence UjU_{j} is decreasing. By the Tietze extension theorem, there is a function fjf_{j} continuous on XX such that fj=ff_{j}=f on Dj:=XUjD_{j}:=X\setminus U_{j}, moreover C0fj0-C_{0}\leq f_{j}\leq 0. For each jj we define

gj:=supkjfk.g_{j}:=\sup_{k\geq j}f_{k}.

We observe that gjg_{j} is lower-semicontinuous on XX, gj=fg_{j}=f on DjD_{j}, and gjg_{j} decreases to some function gg on XX. Since the sequence (Dj)(D_{j}) is increasing, it follows that gj=fg_{j}=f on DkD_{k} for all kjk\leq j. Thus letting jj\to\infty gives g=fg=f on DkD_{k} for all kk. This implies g=fg=f in XX except for a set of capacity equal to zero. Hence g=fg=f quasi-everywhere in XX, and Pθ(f)=Pθ(g)P_{\theta}(f)=P_{\theta}(g). Since gjg_{j} is lower-semicontinuous in XX, by the balayage method, [BT82, Corollary 9.2], we have

Ω(1ePθ(gj)gj)(θ+ddcPθ(gj))n=0,\int_{\Omega}(1-e^{P_{\theta}(g_{j})-g_{j}})(\theta+dd^{c}P_{\theta}(g_{j}))^{n}=0,

where Ω\Omega is the ample locus of {θ}\{\theta\}. Fix an open set GΩG\Subset\Omega. Since fjf_{j} is uniformly bounded on XX, we infer that Pθ(gj)VθP_{\theta}(g_{j})-V_{\theta} is uniformly bounded, hence there is a constant B>0B>0 such that BPθ(gj)0-B\leq P_{\theta}(g_{j})\leq 0 in GG. It follows from the plurifine locality that, for all Borel set EE,

𝟏GE(θ+ddcPθ(gj))n𝟏GE(ω+ddcmax(Pθ(gj),B))nBnCapω(E).{\bf 1}_{G\cap E}(\theta+dd^{c}P_{\theta}(g_{j}))^{n}\leq{\bf 1}_{G\cap E}(\omega+dd^{c}\max(P_{\theta}(g_{j}),-B))^{n}\leq B^{n}{\rm Cap}_{\omega}(E).

It follows from all the above that

G\displaystyle\int_{G} |1ePθ(gj)f|(θ+ddcPθ(gj))n\displaystyle|1-e^{P_{\theta}(g_{j})-f}|(\theta+dd^{c}P_{\theta}(g_{j}))^{n}
=DjG|1ePθ(gj)f|(θ+ddcPθ(gj))n+UjG|1ePθ(gj)f|(θ+ddcPθ(gj))n\displaystyle=\int_{D_{j}\cap G}|1-e^{P_{\theta}(g_{j})-f}|(\theta+dd^{c}P_{\theta}(g_{j}))^{n}+\int_{U_{j}\cap G}|1-e^{P_{\theta}(g_{j})-f}|(\theta+dd^{c}P_{\theta}(g_{j}))^{n}
=DjG(1ePθ(gj)gj)(θ+ddcPθ(gj))n+UjG|1ePθ(gj)f|(θ+ddcPθ(gj))n\displaystyle=\int_{D_{j}\cap G}(1-e^{P_{\theta}(g_{j})-g_{j}})(\theta+dd^{c}P_{\theta}(g_{j}))^{n}+\int_{U_{j}\cap G}|1-e^{P_{\theta}(g_{j})-f}|(\theta+dd^{c}P_{\theta}(g_{j}))^{n}
BnsupX|1ePθ(gj)f|Capω(Uj)C2j1.\displaystyle\leq B^{n}\sup_{X}|1-e^{P_{\theta}(g_{j})-f}|{\rm Cap}_{\omega}(U_{j})\leq C2^{-j-1}.

The functions |1ePθ(gj)f||1-e^{P_{\theta}(g_{j})-f}| are uniformly bounded and (by construction) converge in capacity to the quasi-continuous function |1ePθ(f)f||1-e^{P_{\theta}(f)-f}|. It thus follows from Proposition 2.2 that

|1ePθ(gj)f|(θ+ddcPθ(gj))n weakly converges to |1ePθ(f)f|(θ+ddcPθ(f))n,|1-e^{P_{\theta}(g_{j})-f}|(\theta+dd^{c}P_{\theta}(g_{j}))^{n}\;\text{ weakly converges to }\;|1-e^{P_{\theta}(f)-f}|(\theta+dd^{c}P_{\theta}(f))^{n},

hence

lim infj(C2j1)\displaystyle\liminf_{j\to\infty}(C2^{-j-1}) lim infjG|1eP(gj)f|(θ+ddcPθ(gj))n\displaystyle\geq\liminf_{j\to\infty}\int_{G}|1-e^{P(g_{j})-f}|(\theta+dd^{c}P_{\theta}(g_{j}))^{n}
G|1ePθ(f)f|(θ+ddcPθ(f))n0.\displaystyle\geq\int_{G}|1-e^{P_{\theta}(f)-f}|(\theta+dd^{c}P_{\theta}(f))^{n}\geq 0.

We then infer that

G|1ePθ(f)f|(θ+ddcPθ(f))n=0.\int_{G}|1-e^{P_{\theta}(f)-f}|(\theta+dd^{c}P_{\theta}(f))^{n}=0.

Letting GG increase to Ω\Omega, we can conclude that (θ+ddcPθ(f))n(\theta+dd^{c}P_{\theta}(f))^{n} is concentrated on the contact set {Pθ(f)=f}\{P_{\theta}(f)=f\}.

Step 2. For the general case we approximate ff by fj:=max(f,j)f_{j}:=\max(f,-j). By the first step, we have

{Pθ(fj)<fj}(θ+ddcPθ(fj))n=0.\int_{\{P_{\theta}(f_{j})<f_{j}\}}(\theta+dd^{c}P_{\theta}(f_{j}))^{n}=0.

Fix C>0C>0 and an open set GΩG\Subset\Omega. Consider the quasi-open set U:=G{Pθ(f)>VθC}U:=G\cap\{P_{\theta}(f)>V_{\theta}-C\}. Since Pθ(fj)Pθ(f)P_{\theta}(f_{j})\geq P_{\theta}(f), thanks to the plurifine property we have

𝟏U(θ+ddcmax(Pθ(fj),VθC))n=𝟏U(θ+ddcPθ(fj))n,{\bf 1}_{U}(\theta+dd^{c}\max(P_{\theta}(f_{j}),V_{\theta}-C))^{n}={\bf 1}_{U}(\theta+dd^{c}P_{\theta}(f_{j}))^{n},

hence

U(1ePθ(fj)fj))(θ+ddcmax(Pθ(fj),VθC))n=0.\int_{U}(1-e^{P_{\theta}(f_{j})-f_{j})})(\theta+dd^{c}\max(P_{\theta}(f_{j}),V_{\theta}-C))^{n}=0.

Observe that Pθ(fj)P_{\theta}(f_{j}) decreases to Pθ(f)P_{\theta}(f), hence it converges in capacity. Letting jj\to\infty and using Lemma 2.4 we obtain

U(1ePθ(f)f))(θ+ddcmax(Pθ(f),VθC))n=0,\int_{U}(1-e^{P_{\theta}(f)-f)})(\theta+dd^{c}\max(P_{\theta}(f),V_{\theta}-C))^{n}=0,

hence

U(1ePθ(f)f))(θ+ddcPθ(f))n=0.\int_{U}(1-e^{P_{\theta}(f)-f)})(\theta+dd^{c}P_{\theta}(f))^{n}=0.

From this, letting CC\to\infty and UΩU\nearrow\Omega we obtain the result. ∎

In order to prove a regularity result for θVθn\theta_{V_{\theta}}^{n}, we will need two preliminaries lemmas. We recall that C1,1¯(X)C^{1,\overline{1}}(X) denotes the space of continuous function with bounded distributional Laplacian w.r.t. ω\omega.

Lemma 2.8.

If f1,f2C1,1¯(X)f_{1},f_{2}\in C^{1,\bar{1}}(X), then Pω(f1,f2)C1,1¯(X)P_{\omega}(f_{1},f_{2})\in C^{1,\bar{1}}(X) and for i=1,2i=1,2 the functions fif_{i} and Pω(f1,f2)P_{\omega}(f_{1},f_{2}) are equal up to second order at almost every point on the set {Pω(f1,f2)=fi}\{P_{\omega}(f_{1},f_{2})=f_{i}\}. In particular, the measures

𝟏{Pω(f1,f2)=f1}ωf1n,𝟏{Pω(f1,f2)=f2}ωf2n,(j=1,2),{\bf 1}_{\{P_{\omega}(f_{1},f_{2})=f_{1}\}}\omega_{f_{1}}^{n},\quad{\bf 1}_{\{P_{\omega}(f_{1},f_{2})=f_{2}\}}\omega_{f_{2}}^{n},\quad(j=1,2),

are positive and

ωPω(f1,f2)n\displaystyle{\omega}_{P_{\omega}(f_{1},f_{2})}^{n} =\displaystyle= 𝟏{Pω(f1,f2)=f1}ωf1n+𝟏{Pω(f1,f2)=f2}ωf2n𝟏{Pω(f1,f2)=f1=f2}ωf1n\displaystyle{\bf 1}_{\{P_{\omega}(f_{1},f_{2})=f_{1}\}}\omega_{f_{1}}^{n}+{\bf 1}_{\{P_{\omega}(f_{1},f_{2})=f_{2}\}}\omega_{f_{2}}^{n}-{\bf 1}_{\{P_{\omega}(f_{1},f_{2})=f_{1}=f_{2}\}}\omega_{f_{1}}^{n}
=\displaystyle= 𝟏{Pω(f1,f2)=f1}ωf1n+𝟏{Pω(f1,f2)=f2}ωf2n𝟏{Pω(f1,f2)=f1=f2}ωf2n.\displaystyle{\bf 1}_{\{P_{\omega}(f_{1},f_{2})=f_{1}\}}\omega_{f_{1}}^{n}+{\bf 1}_{\{P_{\omega}(f_{1},f_{2})=f_{2}\}}\omega_{f_{2}}^{n}-{\bf 1}_{\{P_{\omega}(f_{1},f_{2})=f_{1}=f_{2}\}}\omega_{f_{2}}^{n}.

The C1,1¯C^{1,\bar{1}} regularity of the envelope Pω(f1)P_{\omega}(f_{1}) is due to Berman [Ber19]. For the C1,1¯C^{1,\bar{1}} regularity of the envelope Pθ(f1)P_{\theta}(f_{1}) in the case {θ}\{\theta\} is big we refer to [DT23]. The C1,1¯C^{1,\bar{1}} regularity of the rooftop envelope Pω(f1,f2)P_{\omega}(f_{1},f_{2}) was proved in [DR16, Theorem 2.5(ii)]. For a detailed presentation of these results, we refer to [Dar19, Appendix A.1]. For Hessian estimates which give the optimal C1,1C^{1,1} regularity, see [Tos18, Theorem 3.1]. Using the C1,1¯C^{1,\bar{1}} estimates, one can reason the same way as in the short argument of [Dar17, Proposition 2.2] to conclude (2.8).

Lemma 2.9.

Let φ,ψPSH(X,θ)\varphi,\psi\in{\rm PSH}(X,\theta). Then

θmax(φ,ψ)n𝟏{ψφ}θφn+𝟏{φ<ψ}θψn.\theta_{\max(\varphi,\psi)}^{n}\geq{\bf 1}_{\{\psi\leq\varphi\}}\theta_{\varphi}^{n}+{\bf 1}_{\{\varphi<\psi\}}\theta_{\psi}^{n}. (2.6)

In particular, if φψ\varphi\leq\psi then 𝟏{φ=ψ}θφn𝟏{φ=ψ}θψn.{\bf 1}_{\{\varphi=\psi\}}\theta_{\varphi}^{n}\leq{\bf 1}_{\{\varphi=\psi\}}\theta_{\psi}^{n}.

Proof.

Let ψk:=max(ψ,Vθk)\psi_{k}:=\max(\psi,V_{\theta}-k) and φk:=max(φ,Vθk)\varphi_{k}:=\max(\varphi,V_{\theta}-k).

By the locality of the Monge–Ampère measure with respect to the plurifine topology it follows that

𝟏{ψk>φk}θmax(ψk,φk)n=𝟏{ψk>φk}θψkn.{\bf 1}_{\{\psi_{k}>\varphi_{k}\}}\theta_{\max(\psi_{k},\varphi_{k})}^{n}={\bf 1}_{\{\psi_{k}>\varphi_{k}\}}\theta_{\psi_{k}}^{n}.

and

𝟏{ψk<φk}θmax(ψk,φk)n=𝟏{ψk<φk}θφkn{\bf 1}_{\{\psi_{k}<\varphi_{k}\}}\theta_{\max(\psi_{k},\varphi_{k})}^{n}={\bf 1}_{\{\psi_{k}<\varphi_{k}\}}\theta_{\varphi_{k}}^{n}

holds in the ample locus of {θ}\{\theta\} where all the functions above are locally bounded. As the non-pluripolar products are extended trivially over XX, we see that the above inequality holds over XX in the sense of measures. Considering max(ψk,φk+t)\max(\psi_{k},\varphi_{k}+t) and letting t0t\searrow 0 we obtain

θmax(φk,ψk)n𝟏{ψkφk}θφkn+𝟏{φk<ψk}θψkn.\theta_{\max(\varphi_{k},\psi_{k})}^{n}\geq{\bf 1}_{\{\psi_{k}\leq\varphi_{k}\}}\theta_{\varphi_{k}}^{n}+{\bf 1}_{\{\varphi_{k}<\psi_{k}\}}\theta_{\psi_{k}}^{n}.

Multiplying with 𝟏{min(φ,ψ)>Vθk}{\bf 1}_{\{\min(\varphi,\psi)>V_{\theta}-k\}}, and using plurifine locality we arrive at

𝟏{min(φ,ψ)>Vθk}θmax(φ,ψ)n𝟏{min(φ,ψ)>Vθk}{ψφ}θφn+𝟏{min(φ,ψ)>Vθk}{φ<ψ}θψn.{\bf 1}_{\{\min(\varphi,\psi)>V_{\theta}-k\}}\theta_{\max(\varphi,\psi)}^{n}\geq{\bf 1}_{\{\min(\varphi,\psi)>V_{\theta}-k\}\cap\{\psi\leq\varphi\}}\theta_{\varphi}^{n}+{\bf 1}_{\{\min(\varphi,\psi)>V_{\theta}-k\}\cap\{\varphi<\psi\}}\theta_{\psi}^{n}.

Letting kk\to\infty, (2.6) follows. ∎

We are now ready to prove a regularity result about θVθn\theta_{V_{\theta}}^{n}, following [DT21].

Theorem 2.10.

If ϕPSH(X,θ)\phi\in{\rm PSH}(X,\theta) and ϕ0\phi\leq 0, then

𝟏{ϕ=0}(θ+ddcϕ)n=𝟏{ϕ=0}θn.{\bf 1}_{\{\phi=0\}}(\theta+dd^{c}\phi)^{n}={\bf 1}_{\{\phi=0\}}\theta^{n}.

In particular, one has

(θ+ddcVθ)n=𝟏{Vθ=0}θn.(\theta+dd^{c}V_{\theta})^{n}={\bf 1}_{\{V_{\theta}=0\}}\theta^{n}.
Proof.

It suffices to prove the first statement as the second follows from this and Theorem 2.7. Without loss of generality, we can assume θω\theta\leq\omega. Note that ϕ\phi is a ω\omega-psh function as well. We proceed in two steps.

Step 1. We want to prove in this step that

𝟏{ϕ=0}(ω+ddcϕ)n=𝟏{ϕ=0}ωn.{\bf 1}_{\{\phi=0\}}(\omega+dd^{c}\phi)^{n}={\bf 1}_{\{\phi=0\}}\omega^{n}.

Let fjf_{j} be a sequence of smooth ω\omega-psh functions on XX decreasing to ϕ\phi. Set ϕj=Pω(fj,0)\phi_{j}=P_{\omega}(f_{j},0) and observe that ϕϕj\phi\leq\phi_{j}. Using

{ϕ=0}{ϕj=0}.\{\phi=0\}\subseteq\{\phi_{j}=0\}. (2.7)

and Lemma 2.8 we then get

ωϕjn=𝟏{ϕj=0}ωn𝟏{ϕ=0}ωn.\omega_{\phi_{j}}^{n}={\bf 1}_{\{\phi_{j}=0\}}\omega^{n}\geq{\bf 1}_{\{\phi=0\}}\omega^{n}. (2.8)

The functions max(ϕj,1)\max(\phi_{j},-1) are uniformly bounded and decrease to max(ϕ,1)\max(\phi,-1), hence

(ω+ddcmax(ϕj,1))n(ω+ddcmax(ϕ,1))n(\omega+dd^{c}\max(\phi_{j},-1))^{n}\to(\omega+dd^{c}\max(\phi,-1))^{n}

weakly. Since {ϕ=0}=X{ϕ<0}\{\phi=0\}=X\setminus\{\phi<0\} is a closed subset of XX, we get

lim supj{ϕ=0}(ω+ddcmax(ϕj,1))n{ϕ=0}(ω+ddcmax(ϕ,1))n.\limsup_{j\to\infty}\int_{\{\phi=0\}}(\omega+dd^{c}\max(\phi_{j},-1))^{n}\leq\int_{\{\phi=0\}}(\omega+dd^{c}\max(\phi,-1))^{n}.

Observe also that {ϕ=0}{ϕj>1}\{\phi=0\}\subset\{\phi_{j}>-1\} and {ϕ=0}{ϕ>1}\{\phi=0\}\subset\{\phi>-1\}. By plurifine locality we thus have

lim supj{ϕ=0}(ω+ddcϕj)n{ϕ=0}(ω+ddcϕ)n.\limsup_{j\to\infty}\int_{\{\phi=0\}}(\omega+dd^{c}\phi_{j})^{n}\leq\int_{\{\phi=0\}}(\omega+dd^{c}\phi)^{n}. (2.9)

By (2.8) we then get

{ϕ=0}ωn{ϕ=0}(ω+ddcϕ)n.\int_{\{\phi=0\}}\omega^{n}\leq\int_{\{\phi=0\}}(\omega+dd^{c}\phi)^{n}.

Then, by Lemma 2.9 we get the identity

𝟏{ϕ=0}ωϕn=𝟏{ϕ=0}ωn.{\bf 1}_{\{\phi=0\}}\omega_{\phi}^{n}={\bf 1}_{\{\phi=0\}}\omega^{n}. (2.10)

Step 2. Now, fix A>0A>0 such that θ+Aω\theta+A\omega is a Kähler form. Since ϕ\phi is θ\theta-psh function, it follows that ϕ\phi is (θ+tω)(\theta+t\omega)-psh, for t0.t\geq 0. Let gC0(X,)g\in C^{0}(X,\mathbb{R}) and consider the function

Q(t):={ϕ=0}g(θ+tω+ddcϕ)n{ϕ=0}g(θ+tω)nQ(t):=\int_{\{\phi=0\}}g(\theta+t\omega+dd^{c}\phi)^{n}-\int_{\{\phi=0\}}g(\theta+t\omega)^{n}

defined for t0.t\geq 0. Then by multilinearity of the non-pluripolar product and the multilinearity of the product of forms, it is clear that Q(t)Q(t) is a polynomial in tt. Thanks to (2.10) we can infer that for any t>At>A

𝟏{ϕ=0}(θ+tω+ddcϕ)n=𝟏{ϕ=0}(θ+tω)n.{\bf 1}_{\{\phi=0\}}(\theta+t\omega+dd^{c}\phi)^{n}={\bf 1}_{\{\phi=0\}}(\theta+t\omega)^{n}.

This implies that the polynomial Q(t)Q(t) is identically zero for t>A,t>A, hence Q(t)0Q(t)\equiv 0. It then follows that Q(0)=0.Q(0)=0. Since gC0(X,)g\in C^{0}(X,\mathbb{R}) is arbitrary we have the desired equality between measures. ∎

As it turns out, (2.9) follows from a more general result. This was proved in [DDL21a, Proposition 4.6], and the proof we give below fixes an imprecision in the original argument.

Lemma 2.11.

Suppose that uj,uPSH(X,θ)u_{j},u\in\textup{PSH}(X,\theta) with uj,uu_{j},u\leq0. If ujuL10\|u_{j}-u\|_{L^{1}}\to 0 then

lim supj+{uj=0}θujn{u=0}θun.\limsup_{j\rightarrow+\infty}\int_{\{u_{j}=0\}}\theta_{u_{j}}^{n}\leq\int_{\{u=0\}}\theta_{u}^{n}.
Proof.

Let vk=usc(supjkuj)v_{k}=\textup{usc}(\sup_{j\geq k}u_{j}). It is well known that vkuv_{k}\searrow u. Also ukvk0u_{k}\leq v_{k}\leq 0, so {uk=0}{vk=0}\{u_{k}=0\}\subset\{v_{k}=0\}. As a result, due to Lemma 2.9, we have 𝟏{uk=0}θukn𝟏{uk=0}θvkn.{\bf 1}_{\{u_{k}=0\}}\theta_{u_{k}}^{n}\leq{\bf 1}_{\{u_{k}=0\}}\theta_{v_{k}}^{n}.

Let b,c>0b,c>0. Using plurifine locality and the above we get that

{uj=0}θujn{uj=0}θvjn{vj=0}θvjn={vj=0}θmax(vj,Vθc)nXebvjθmax(vj,Vθc)n.\int_{\{u_{j}=0\}}\theta_{u_{j}}^{n}\leq\int_{\{u_{j}=0\}}\theta_{v_{j}}^{n}\leq\int_{\{v_{j}=0\}}\theta_{v_{j}}^{n}=\int_{\{v_{j}=0\}}\theta_{\max(v_{j},V_{\theta}-c)}^{n}\leq\int_{X}e^{bv_{j}}\theta_{\max(v_{j},V_{\theta}-c)}^{n}.

Since ebvj,ebue^{bv_{j}},e^{bu} are uniformly bounded and non-negative function, taking the limit Theorem 2.6 gives that

lim supj+{uj=0}θujnXebuθmax(u,Vθc)n.\limsup_{j\rightarrow+\infty}\int_{\{u_{j}=0\}}\theta_{u_{j}}^{n}\leq\int_{X}e^{bu}\theta_{\max(u,V_{\theta}-c)}^{n}.

Now letting bb\to\infty, and subsequently cc\to\infty, the conclusion follows. ∎

In our investigation of relative pluripotential theory, the following envelope construction will be essential: given ϕPSH(X,θ)\phi\in\textup{PSH}(X,\theta) we consider

PSH(X,θ)ψPθ[ψ](ϕ)PSH(X,θ).\textup{PSH}(X,\theta)\ni\psi\to\ P_{\theta}[\psi](\phi)\in\textup{PSH}(X,\theta).

This was introduced by Ross and Witt Nyström [RW14] in their construction of geodesic rays, building on ideas of Rashkovskii and Sigurdsson [RS05] in the local setting. Starting from the “rooftop envelope” Pθ(ψ,ϕ)P_{\theta}(\psi,\phi) we introduce

Pθ[ψ](ϕ):=(limCPθ(ψ+C,ϕ)).P_{\theta}[\psi](\phi):=\Big{(}\lim_{C\to\infty}P_{\theta}(\psi+C,\phi)\Big{)}^{*}.

It is easy to see that Pθ[ψ](ϕ)P_{\theta}[\psi](\phi) only depends on the singularity type of ψ\psi. When ϕ=Vθ\phi=V_{\theta}, we will simply write Pθ[ψ]:=Pθ[ψ](Vθ)P_{\theta}[\psi]:=P_{\theta}[\psi](V_{\theta}), and we refer to this potential as the envelope of the singularity type [ψ][\psi]. We note the following simple concavity result about the operator Pθ[](ϕ)P_{\theta}[\cdot](\phi).

Lemma 2.12.

The operator Pθ[](ϕ)P_{\theta}[\cdot](\phi) is concave: if u,v,ϕPSH(X,θ)u,v,\phi\in{\rm PSH}(X,\theta) and t(0,1)t\in(0,1) then

Pθ[tu+(1t)v](ϕ)tPθ[u](ϕ)+(1t)Pθ[v](ϕ).P_{\theta}[tu+(1-t)v](\phi)\geq tP_{\theta}[u](\phi)+(1-t)P_{\theta}[v](\phi).
Proof.

Assume u,vPSH(X,θ)u,v\in{\rm PSH}(X,\theta) and t(0,1)t\in(0,1). Then, for all C>0C>0,

Pθ(min(tu+(1t)v+C,ϕ))tPθ(min(u+C,ϕ))+(1t)Pθ(min(v+C,ϕ)),P_{\theta}(\min(tu+(1-t)v+C,\phi))\geq tP_{\theta}(\min(u+C,\phi))+(1-t)P_{\theta}(\min(v+C,\phi)),

because the right-hand side is a θ\theta-psh function lying below min(tu+(1t)v+C,0)\min(tu+(1-t)v+C,0). Letting CC\nearrow\infty we arrive at the result. ∎

Chapter 3 The basics of relative pluripotential theory

3.1 Monotonicity of non-pluripolar product masses

In what follows, unless otherwise stated, we work with θ\theta a smooth closed real (1,1)(1,1)-form whose cohomology class is big. We start with the following result, saying that potentials with the same singularity type have also the same global mass.

Lemma 3.1.

Let u,vPSH(X,θ)u,v\in{\rm PSH}(X,\theta). If uu and vv have the same singularity type, then Xθun=Xθvn\int_{X}\theta_{u}^{n}=\int_{X}\theta_{v}^{n}.

The above result is due to Witt Nyström [Wit19]. A different proof has been given in [LN22] using the monotonicity of the energy functional. We give below a direct proof using a standard approximation process. Another different proof has been recently given in [Vu21], where generalized non-pluripolar products of positive currents are studied.

Proof.

Step 1. Assume that θ\theta is a Kähler form.

We first prove the following claim: if there exists a constant C>0C>0 such that u=vu=v on the open set U:={min(u,v)<C}U:=\{\min(u,v)<-C\} then Xθun=Xθvn\int_{X}\theta_{u}^{n}=\int_{X}\theta_{v}^{n}.

Fix t>Ct>C. Since u=vu=v on UU, we have max(u,t)=max(v,t)\max(u,-t)=\max(v,-t) on UU. Since UU is open, we have

𝟏U(θ+ddcmax(u,t))n=𝟏U(θ+ddcmax(v,t))n.{\bf 1}_{U}(\theta+dd^{c}\max(u,-t))^{n}={\bf 1}_{U}(\theta+dd^{c}\max(v,-t))^{n}.

We also have {ut}={vt}U\{u\leq-t\}=\{v\leq-t\}\subset U. Indeed, if u(x)tu(x)\leq-t then xUx\in U because t<C-t<-C. But on UU we have u=vu=v, hence v(x)=u(x)tv(x)=u(x)\leq-t.

By plurifine locality we thus have

θmax(u,t)n\displaystyle\theta_{\max(u,-t)}^{n} =𝟏{u>t}θmax(u,t)n+𝟏{ut}θmax(u,t)n\displaystyle={\bf 1}_{\{u>-t\}}\theta_{\max(u,-t)}^{n}+{\bf 1}_{\{u\leq-t\}}\theta_{\max(u,-t)}^{n}
=𝟏{u>t}θun+𝟏{vt}θmax(v,t)n,\displaystyle={\bf 1}_{\{u>-t\}}\theta_{u}^{n}+{\bf 1}_{\{v\leq-t\}}\theta_{\max(v,-t)}^{n},

and

θmax(v,t)n=𝟏{v>t}θvn+𝟏{vt}θmax(v,t)n.\theta_{\max(v,-t)}^{n}={\bf 1}_{\{v>-t\}}\theta_{v}^{n}+{\bf 1}_{\{v\leq-t\}}\theta_{\max(v,-t)}^{n}.

Integrating over XX, since Xθmax(u,t)n=Xθmax(v,t)n=Vol(θ)\int_{X}\theta_{\max(u,-t)}^{n}=\int_{X}\theta_{\max(v,-t)}^{n}={\rm Vol}({\theta}) (recall that max(u,t)\max(u,-t) and max(u,t)\max(u,-t) are bounded functions), we get

{u>t}θun={v>t}θvn.\int_{\{u>-t\}}\theta_{u}^{n}=\int_{\{v>-t\}}\theta_{v}^{n}.

Letting tt\to\infty, the claim follows.

Now we prove the general case when θ\theta is Kähler. Since uu and vv have the same singularity type, we can assume that vuv+B0v\leq u\leq v+B\leq 0, for some positive constant BB. For each a(0,1)a\in(0,1) we set va:=avv_{a}:=av, ua:=max(u,va)u_{a}:=\max(u,v_{a}) and C:=Ba(1a)1C:=Ba(1-a)^{-1}. To use the claim above we need to check that ua=vau_{a}=v_{a} on the open set Ua:={min(ua,va)<C}U_{a}:=\{\min(u_{a},v_{a})<-C\}. Observe that min(ua,va)=va\min(u_{a},v_{a})=v_{a} because vauav_{a}\leq u_{a}. If xUax\in U_{a} then av(x)<Cav(x)<-C hence (1a)v(x)<B(1-a)v(x)<-B, which implies (recall that v+Buv+B\geq u)

av(x)v(x)+Bu(x).av(x)\geq v(x)+B\geq u(x).

We infer that va(x)u(x)v_{a}(x)\geq u(x), hence ua(x)=va(x)u_{a}(x)=v_{a}(x). We can thus apply the claim above to get Xθuan=Xθvan\int_{X}\theta_{u_{a}}^{n}=\int_{X}\theta_{v_{a}}^{n}. Since non-pluripolar products are multilinear, see [BEGZ10, Proposition 1.4], we have that

Xθvan=anXθvn+k=0n1ak(1a)nkXθvkθnkXθvn\int_{X}\theta_{v_{a}}^{n}=a^{n}\int_{X}\theta_{v}^{n}+\sum_{k=0}^{n-1}a^{k}(1-a)^{n-k}\int_{X}\theta_{v}^{k}\wedge\theta^{n-k}\to\int_{X}\theta_{v}^{n}

as a1a\nearrow 1. Since uauu_{a}\searrow u as a1a\nearrow 1, by Theorem 2.6 we have

lim infa1XθuanXθun.\liminf_{a\to 1^{-}}\int_{X}\theta_{u_{a}}^{n}\geq\int_{X}\theta_{u}^{n}.

We thus have XθunXθvn\int_{X}\theta_{u}^{n}\leq\int_{X}\theta_{v}^{n}. By symmetry we get equality, finishing the proof of Step 1.

Step 2. We treat the general case when {θ}\{\theta\} is merely big. We use an idea from [DT21]. Fix s>0s>0 so large that θ+sω\theta+s\omega is Kähler. For t>st>s we have, by the first step,

X(θ+tω+ddcu)n=X(θ+tω+ddcv)n.\int_{X}(\theta+t\omega+dd^{c}u)^{n}=\int_{X}(\theta+t\omega+dd^{c}v)^{n}.

Since non-pluripolar products are multilinear ([BEGZ10, Proposition 1.4]) we have for all t>st>s,

k=0n(nk)Xθukωnktnk=k=0n(nk)Xθvkωnktnk.\sum_{k=0}^{n}\binom{n}{k}\int_{X}\theta_{u}^{k}\wedge\omega^{n-k}t^{n-k}=\sum_{k=0}^{n}\binom{n}{k}\int_{X}\theta_{v}^{k}\wedge\omega^{n-k}t^{n-k}.

We thus obtain an equality between two polynomials for all t0t\geq 0, and identifying the coefficients we infer the desired equality. ∎

We continue with the following immediate generalization of the above result:

Proposition 3.2.

Let θj,j{1,,n}\theta^{j},j\in\{1,\ldots,n\} be smooth closed real (1,1)(1,1)-forms on XX whose cohomology classes are pseudoeffective. Let uj,vjPSH(X,θj)u_{j},v_{j}\in\textup{PSH}(X,\theta^{j}) such that uju_{j} has the same singularity type as vj,j{1,,n}v_{j},\ j\in\{1,\ldots,n\}. Then

Xθu11θunn=Xθv11θvnn.\int_{X}\theta^{1}_{u_{1}}\wedge\ldots\wedge\theta^{n}_{u_{n}}=\int_{X}\theta^{1}_{v_{1}}\wedge\ldots\wedge\theta^{n}_{v_{n}}.

The proof of this result uses the ideas from [BEGZ10, Corollary 2.15].

Proof.

First we note that we can assume that the classes {θj}\{\theta^{j}\} are in fact big. Indeed, if this is not the case we can just replace each θj\theta^{j} with θj+εω\theta^{j}+\varepsilon\omega, and using the multi-linearity of the non-pluripolar product ([BEGZ10, Proposition 1.4]) we can let ε0\varepsilon\to 0 at the end of our argument to conclude the statement for pseudoeffective classes.

For each tΔ={t=(t1,,tn)n|tj>0}t\in\Delta=\{t=(t_{1},...,t_{n})\in\mathbb{R}^{n}\ |\ t_{j}>0\} consider ut:=jtjuju_{t}:=\sum_{j}t_{j}u_{j}, vt:=jtjvjv_{t}:=\sum_{j}t_{j}v_{j} and θt:=jtjθj\theta^{t}:=\sum_{j}t_{j}\theta^{j}. Clearly, {θt}\{\theta^{t}\} is big, and utu_{t} has the same singularity type as vtv_{t}. Hence it follows from Lemma 3.1 that X(θutt)n=X(θvtt)n\int_{X}(\theta^{t}_{u_{t}})^{n}=\int_{X}(\theta^{t}_{v_{t}})^{n} for all tΔt\in\Delta. On the other hand, using multi-linearity of the non-pluripolar product again ([BEGZ10, Proposition 1.4]), we see that both tX(θutt)nt\to\int_{X}(\theta^{t}_{u_{t}})^{n} and tX(θvtt)nt\to\int_{X}(\theta^{t}_{v_{t}})^{n} are homogeneous polynomials of degree nn. Our last identity forces all the coefficients of these polynomials to be equal, giving the statement of our result. ∎

Using the above result we argue the monotonicity of non-pluripolar products, conjectured in [BEGZ10] and proved in [DDL18a] (see [Vu21, Theorem 1.1] for a more general result):

Theorem 3.3.

Let θj,j{1,,n}\theta^{j},j\in\{1,\ldots,n\} be smooth closed real (1,1)(1,1)-forms on XX whose cohomology classes are pseudoeffective. Let uj,vjPSH(X,θj)u_{j},v_{j}\in\textup{PSH}(X,\theta^{j}) be such that uju_{j} is less singular than vjv_{j} for all j{1,,n}j\in\{1,\ldots,n\}. Then

Xθu11θunnXθv11θvnn.\int_{X}\theta^{1}_{u_{1}}\wedge\ldots\wedge\theta^{n}_{u_{n}}\geq\int_{X}\theta^{1}_{v_{1}}\wedge\ldots\wedge\theta^{n}_{v_{n}}.
Proof.

By the same reason as in Proposition 3.2, we can assume that the classes {θj}\{\theta^{j}\} are in fact big. For each t>0t>0 we set vjt:=max(ujt,vj)v_{j}^{t}:=\max(u_{j}-t,v_{j}) for j=1,,nj=1,...,n. Observe that the vjtv_{j}^{t} converge decreasingly to vjv_{j} as tt\to\infty. In particular, by [GZ05, Proposition 3.7] the convergence holds in capacity. As vjtv_{j}^{t} and uju_{j} have the same singularity type, it follows from Proposition 3.2 that

Xθu11θunn=Xθv1t1θvntn.\int_{X}\theta^{1}_{u_{1}}\wedge\ldots\wedge\theta^{n}_{u_{n}}=\int_{X}\theta^{1}_{v_{1}^{t}}\wedge\ldots\wedge\theta^{n}_{v_{n}^{t}}.

Letting tt\to\infty, the first part of Theorem 2.6 allows to conclude the argument. ∎

Remark 3.4.

Condition (2.4) in Theorem 2.6 is automatically satisfied if ujkuju^{k}_{j}\nearrow u_{j} a.e. as kk\to\infty. Indeed, in this case ujkuju_{j}^{k}\to u_{j} in capacity (see [GZ17, Proposition 4.25]), and by Theorem 3.3 we have Xθu11θunnlim supkXθu1k1θunkn\int_{X}\theta^{1}_{u_{1}}\wedge\ldots\wedge\theta^{n}_{u_{n}}\geq\limsup_{k}\int_{X}\theta^{1}_{u^{k}_{1}}\wedge\ldots\wedge\theta^{n}_{u^{k}_{n}}.

Since for all uPSH(X,θ)u\in{\rm PSH}(X,\theta) we have P(u,Vθ+C)Pθ[u]P(u,V_{\theta}+C)\nearrow P_{\theta}[u] as CC\to\infty, we conclude that

Xθun=XθPθ[u]n.\int_{X}\theta_{u}^{n}=\int_{X}\theta_{P_{\theta}[u]}^{n}.

Similarly, for ujPSH(X,θj)u_{j}\in\textup{PSH}(X,\theta^{j}) the same ideas allow to conclude that

Xθu11θunn=XθPθ[u1]1θPθ[un]n.\int_{X}\theta^{1}_{u_{1}}\wedge\ldots\wedge\theta^{n}_{u_{n}}=\int_{X}\theta^{1}_{P_{\theta}[u_{1}]}\wedge\ldots\wedge\theta^{n}_{P_{\theta}[u_{n}]}.

3.2 Model potentials and relative full mass classes

Rooftop envelopes revisited.

We survey some results from [DDL18a]. We first start with the following simple observation.

Lemma 3.5.

Assume u,vPSH(X,θ)u,v\in{\rm PSH}(X,\theta) and uu is more singular than vv. Then

supX(uPθ[v])=supXu.\sup_{X}(u-P_{\theta}[v])=\sup_{X}u.
Proof.

Since Pθ[v]0P_{\theta}[v]\leq 0, we have supX(uPθ[v])supXu\sup_{X}(u-P_{\theta}[v])\geq\sup_{X}u. By the assumption that uu is more singular than vv, we have usupXumin(v+C,0)u-\sup_{X}u\leq\min(v+C,0) for some constant CC, hence usupXuPθ[v]u-\sup_{X}u\leq P_{\theta}[v]. This proves the other inequality, finishing the proof. ∎

We next prove the following result about the complex Monge–Ampère measure of Pθ(u,v)P_{\theta}(u,v) from [GLZ19, Lemma 4.1] (see [Dar17] for a particular result in the Kähler case).

Theorem 3.6.

Suppose φ,ψ,Pθ(φ,ψ)PSH(X,θ)\varphi,\psi,P_{\theta}(\varphi,\psi)\in{\rm PSH}(X,\theta). Then

θPθ(φ,ψ)n𝟏{Pθ(φ,ψ)=φ}θφn+𝟏{Pθ(φ,ψ)=ψ}θψn.\theta_{P_{\theta}(\varphi,\psi)}^{n}\leq{\bf 1}_{\{P_{\theta}(\varphi,\psi)=\varphi\}}\theta_{\varphi}^{n}+{\bf 1}_{\{P_{\theta}(\varphi,\psi)=\psi\}}\theta_{\psi}^{n}.

In particular,

θPθ[ψ](φ)n𝟏{Pθ[ψ](φ)=φ}θφnandθPθ[ψ]n𝟏{Pθ[ψ]=0}θn.\theta_{P_{\theta}[\psi](\varphi)}^{n}\leq{\bf 1}_{\{P_{\theta}[\psi](\varphi)=\varphi\}}\theta_{\varphi}^{n}\quad\text{and}\quad\theta_{P_{\theta}[\psi]}^{n}\leq{\bf 1}_{\{P_{\theta}[\psi]=0\}}\theta^{n}.
Proof.

Since the function min(φ,ψ)\min(\varphi,\psi) is quasi-continuous on XX, by Theorem 2.7, the Monge–Ampère measure θPθ(φ,ψ)n\theta_{P_{\theta}(\varphi,\psi)}^{n} is concentrated on the contact set {Pθ(φ,ψ)=min(φ,ψ)}\{P_{\theta}(\varphi,\psi)=\min(\varphi,\psi)\}. By Lemma 2.9 we have

𝟏{Pθ(φ,ψ)=φ}θPθ(φ,ψ)n𝟏{Pθ(φ,ψ)=φ}θφnand𝟏{Pθ(φ,ψ)=ψ}θPθ(φ,ψ)n𝟏{Pθ(φ,ψ)=ψ}θψn.{\bf 1}_{\{P_{\theta}(\varphi,\psi)=\varphi\}}\theta_{P_{\theta}(\varphi,\psi)}^{n}\leq{\bf 1}_{\{P_{\theta}(\varphi,\psi)=\varphi\}}\theta_{\varphi}^{n}\quad\text{and}\quad{\bf 1}_{\{P_{\theta}(\varphi,\psi)=\psi\}}\theta_{P_{\theta}(\varphi,\psi)}^{n}\leq{\bf 1}_{\{P_{\theta}(\varphi,\psi)=\psi\}}\theta_{\psi}^{n}.

From this the first inequality follows. Using this, for each t>0t>0 we have

θPθ(ψ+t,φ)n\displaystyle\theta_{P_{\theta}(\psi+t,\varphi)}^{n} 𝟏{Pθ(ψ+t,φ)=ψ+t}θψn+𝟏{Pθ(ψ+t,φ)=φ}θφn\displaystyle\leq{\bf 1}_{\{P_{\theta}(\psi+t,\varphi)=\psi+t\}}\theta_{\psi}^{n}+{\bf 1}_{\{P_{\theta}(\psi+t,\varphi)=\varphi\}}\theta_{\varphi}^{n}
𝟏{ψ+tφ}θψn+𝟏{Pθ(ψ+t,φ)=φ}θφn.\displaystyle\leq{\bf 1}_{\{\psi+t\leq\varphi\}}\theta_{\psi}^{n}+{\bf 1}_{\{P_{\theta}(\psi+t,\varphi)=\varphi\}}\theta_{\varphi}^{n}.

Since θψn\theta_{\psi}^{n} vanishes on the pluripolar set {ψ=}\{\psi=-\infty\}, we have limt𝟏{ψφt}θψn=0\lim_{t\to\infty}{\bf 1}_{\{\psi\leq\varphi-t\}}\theta_{\psi}^{n}=0. Observe also that Pθ(ψ+t,φ)Pθ[ψ](φ)P_{\theta}(\psi+t,\varphi)\nearrow P_{\theta}[\psi](\varphi) as tt\nearrow\infty, hence Theorem 2.6 and Remark 3.4 ensure that θPθ(ψ+t,φ)n\theta_{P_{\theta}(\psi+t,\varphi)}^{n} converges weakly to θPθ[ψ](φ)n\theta_{P_{\theta}[\psi](\varphi)}^{n}. We also have that {Pθ(ψ+t,φ)=φ}{Pθ[ψ](φ)=φ}\{P_{\theta}(\psi+t,\varphi)=\varphi\}\subset\{P_{\theta}[\psi](\varphi)=\varphi\}. Letting tt\to\infty thus gives

θPθ[ψ](φ)n𝟏{Pθ[ψ](φ)=φ}θφn.\theta_{P_{\theta}[\psi](\varphi)}^{n}\leq{\bf 1}_{\{P_{\theta}[\psi](\varphi)=\varphi\}}\theta_{\varphi}^{n}.

In particular, when φ=Vθ\varphi=V_{\theta}, this yields

θPθ[ψ]n=θPθ[ψ](Vθ)n𝟏{Pθ[ψ](Vθ)=Vθ}θVθn𝟏{Pθ[ψ]=0}θn,\theta_{P_{\theta}[\psi]}^{n}=\theta_{P_{\theta}[\psi](V_{\theta})}^{n}\leq{\bf 1}_{\{P_{\theta}[\psi](V_{\theta})=V_{\theta}\}}\theta_{V_{\theta}}^{n}\leq{\bf 1}_{\{P_{\theta}[\psi]=0\}}\theta^{n},

where the last inequality follows from Theorem 2.10 since {Pθ[ψ]=0}{Vθ=0}\{P_{\theta}[\psi]=0\}\subset\{V_{\theta}=0\}. ∎

The domination principle and uniqueness.

Definition 3.7.

Given a potential ϕPSH(X,θ)\phi\in{\rm PSH}(X,\theta), the relative full mass class (X,θ,ϕ)\mathcal{E}(X,\theta,\phi) is the set of all θ\theta-psh functions uu such that uu is more singular than ϕ\phi and Xθun=Xθϕn\int_{X}\theta_{u}^{n}=\int_{X}\theta_{\phi}^{n}.

Remark 3.8.

If u,v(X,θ,ϕ)u,v\in\mathcal{E}(X,\theta,\phi) then max(u,v)(X,θ,ϕ)\max(u,v)\in\mathcal{E}(X,\theta,\phi). Indeed, since both uu and vv are more singular than ϕ\phi, max(u,v)\max(u,v) is also more singular than ϕ\phi. By Theorem 3.3,

XθunXθmax(u,v)nXθϕn=Xθun,\int_{X}\theta_{u}^{n}\leq\int_{X}\theta_{\max(u,v)}^{n}\leq\int_{X}\theta_{\phi}^{n}=\int_{X}\theta_{u}^{n},

hence all the inequalities become equalities.

Definition 3.9.

A model potential is a θ\theta-psh function ϕ\phi such that Pθ[ϕ]=ϕP_{\theta}[\phi]=\phi and Xθϕn>0\int_{X}\theta_{\phi}^{n}>0.

Let us prove an important technical result from [DDL21a].

Theorem 3.10.

Assume that u,vPSH(X,θ)u,v\in{\rm PSH}(X,\theta), uvu\leq v, Xθun>0\int_{X}\theta_{u}^{n}>0 and b>1b>1 is such that

bnXθun>(bn1)Xθvn.b^{n}\int_{X}\theta_{u}^{n}>(b^{n}-1)\int_{X}\theta_{v}^{n}. (3.1)

Then Pθ(bu+(1b)v)PSH(X,θ)P_{\theta}(bu+(1-b)v)\in{\rm PSH}(X,\theta).

To be clear, Pθ(bu+(1b)v):=usc(sup{hPSH(X,θ) such that h+(b1)vbu})P_{\theta}(bu+(1-b)v):=\textup{usc}(\sup\{h\in\textup{PSH}(X,\theta)\textup{ such that }h+(b-1)v\leq bu\}), and the set of candidates for this supremum is typically empty. Observe that, if uu and vv have the same positive mass then (3.1) is trivially satisfied.

Proof.

Set ϕ:=Pθ[v]\phi:=P_{\theta}[v]. It suffices to prove that Pθ(bu(b1)ϕ)PSH(X,θ)P_{\theta}(bu-(b-1)\phi)\in{\rm PSH}(X,\theta) since vPθ[v]v\leq P_{\theta}[v], hence bu+(1b)vbu+(1b)Pθ[v]bu+(1-b)v\geq bu+(1-b)P_{\theta}[v]. By Remark 3.4 we have Xθvn=Xθϕn\int_{X}\theta_{v}^{n}=\int_{X}\theta_{\phi}^{n}.

For jj\in\mathbb{N} we set uj:=max(u,ϕj)u_{j}:=\max(u,\phi-j) and φj:=P(buj+(1b)ϕ)\varphi_{j}:=P(bu_{j}+(1-b)\phi). Observe that φj\varphi_{j} is a decreasing sequence of θ\theta-psh functions, having the same singularity type as ϕ\phi. The proof is finished if we can show that φ:=limjφj\varphi:=\lim_{j}\varphi_{j} is not identically -\infty. Assume by contradiction that supXφj\sup_{X}\varphi_{j}\to-\infty. Set ψj:=b1φj+(1b1)ϕ\psi_{j}:=b^{-1}\varphi_{j}+(1-b^{-1})\phi and Dj:={b1φj+(1b1)ϕ=uj}D_{j}:=\{b^{-1}\varphi_{j}+(1-b^{-1})\phi=u_{j}\}. Note that ψjuj\psi_{j}\leq u_{j} with equality on the contact set DjD_{j}. Lemma 2.9 thus ensures that

𝟏Djbnθφjn𝟏Djθψjn𝟏Djθujn.{\bf 1}_{D_{j}}b^{-n}\theta_{\varphi_{j}}^{n}\leq{\bf 1}_{D_{j}}\theta_{\psi_{j}}^{n}\leq{\bf 1}_{D_{j}}\theta_{u_{j}}^{n}. (3.2)

Also, observe that by Theorem 2.7, the Monge–Ampère measure θφjn\theta_{\varphi_{j}}^{n} is supported on {φj=buj+(1b)ϕ}=Dj\{\varphi_{j}=bu_{j}+(1-b)\phi\}=D_{j}.

Fix j>k>0j>k>0. We note that uj=uu_{j}=u on {u>ϕk}\{u>\phi-k\} and, since uju_{j} has the same singularity type as ϕ\phi, by Lemma 3.1 and the plurifine locality we have

{uϕk}θujn=Xθujn{u>ϕk}θujn=Xθϕn{u>ϕk}θun.\int_{\{u\leq\phi-k\}}\theta_{u_{j}}^{n}=\int_{X}\theta_{u_{j}}^{n}-\int_{\{u>\phi-k\}}\theta_{u_{j}}^{n}=\int_{X}\theta_{\phi}^{n}-\int_{\{u>\phi-k\}}\theta_{u}^{n}. (3.3)

Since {ujϕk}={uϕk}\{u_{j}\leq\phi-k\}=\{u\leq\phi-k\}, Theorem 2.7 and (3.2) we obtain

θφjn({φjϕbk})bn𝟏Djθujn({φjϕbk})bnθujn({buj+(1b)ϕϕbk})\displaystyle\theta_{\varphi_{j}}^{n}(\{\varphi_{j}\leq\phi-bk\})\leq b^{n}{\bf 1}_{D_{j}}\theta_{u_{j}}^{n}(\{\varphi_{j}\leq\phi-bk\})\leq b^{n}\theta_{u_{j}}^{n}(\{bu_{j}+(1-b)\phi\leq\phi-bk\})
=bnθujn({ujϕk})bnθujn({uϕk})bn(Xθϕn{u>ϕk}θun),\displaystyle=b^{n}\theta_{u_{j}}^{n}(\{u_{j}\leq\phi-k\})\leq b^{n}\theta_{u_{j}}^{n}(\{u\leq\phi-k\})\leq b^{n}\bigg{(}\int_{X}\theta_{\phi}^{n}-\int_{\{u>\phi-k\}}\theta_{u}^{n}\bigg{)}, (3.4)

where in the last inequality we have used (3.3). Since ϕ=Pθ[v]\phi=P_{\theta}[v], by Lemma 3.5 we have supX(φjϕ)=supXφj\sup_{X}(\varphi_{j}-\phi)=\sup_{X}\varphi_{j}\to-\infty. From this we see that {φjϕbk}=X\{\varphi_{j}\leq\phi-bk\}=X for jj0j\geq j_{0} large enough, kk being fixed. Thus,

Xθϕn=Xθφj0nbn(Xθϕn{u>ϕk}θun)\int_{X}\theta_{\phi}^{n}=\int_{X}\theta_{\varphi_{j_{0}}}^{n}\leq b^{n}\bigg{(}\int_{X}\theta_{\phi}^{n}-\int_{\{u>\phi-k\}}\theta_{u}^{n}\bigg{)}

where the first identity follows from the fact that φj0\varphi_{j_{0}} and ϕ\phi have the same singularity type. Letting j0j_{0}\to\infty in (3.2), and then kk\to\infty gives

Xθϕnbn(XθϕnXθun),\int_{X}\theta_{\phi}^{n}\leq b^{n}\left(\int_{X}\theta_{\phi}^{n}-\int_{X}\theta_{u}^{n}\right),

contradicting (3.1). Consequently, φj\varphi_{j} decreases to a function in PSH(X,θ){\rm PSH}(X,\theta), finishing the proof. ∎

In some instances the conclusion of the above result can be strengthened:

Lemma 3.11.

Assume u,vPSH(X,θ)u,v\in{\rm PSH}(X,\theta) and uvu\leq v. Assume also that, for all b>1b>1, Pθ(bu(b1)v)PSH(X,θ)P_{\theta}(bu-(b-1)v)\in{\rm PSH}(X,\theta). Then, for all b>1b>1,

X(θ+ddcPθ(bu(b1)v))n=Xθvn.\int_{X}(\theta+dd^{c}P_{\theta}(bu-(b-1)v))^{n}=\int_{X}\theta_{v}^{n}.
Proof.

Fix t>bt>b and observe that bu(b1)v=t1b(tu(t1)v)+(1t1b)vbu-(b-1)v=t^{-1}b(tu-(t-1)v)+(1-t^{-1}b)v, where t1b<1t^{-1}b<1. Hence

φ:=Pθ(bu(b1)v)t1bPθ(tu(t1)v)+(1t1b)v,\varphi:=P_{\theta}(bu-(b-1)v)\geq t^{-1}bP_{\theta}(tu-(t-1)v)+(1-t^{-1}b)v,

and by Lemma 2.12,

Pθ[φ]t1bPθ[Pθ(tu(t1)v)]+(1t1b)Pθ[v].P_{\theta}[\varphi]\geq t^{-1}bP_{\theta}[P_{\theta}(tu-(t-1)v)]+(1-t^{-1}b)P_{\theta}[v].

Set ψt:=Pθ[Pθ(tu(t1)v)]\psi_{t}:=P_{\theta}[P_{\theta}(tu-(t-1)v)]. Then supXψt=0\sup_{X}\psi_{t}=0 and ψtPSH(X,Aω)\psi_{t}\in{\rm PSH}(X,A\omega) for some fixed constant A>0A>0. It follows from [GZ05, Proposition 2.7] that the functions ψt\psi_{t} stay in a compact set of L1(X)L^{1}(X). Letting tt\to\infty we thus have

Pθ[φ]Pθ[v].P_{\theta}[\varphi]\geq P_{\theta}[v].

Since uvu\leq v, we also have φv\varphi\leq v. Combining this and the above inequality we see that Pθ[φ]=Pθ[v]P_{\theta}[\varphi]=P_{\theta}[v], and using Remark 3.4 we arrive at the result. ∎

Next we prove the domination principle of relative pluripotential theory, extending a result of Dinew from the Kähler case [BL12].

Theorem 3.12.

Assume ϕ\phi is a model potential and u,v(X,θ,ϕ)u,v\in\mathcal{E}(X,\theta,\phi). Then

θun({u<v})=0uv.\theta_{u}^{n}(\{u<v\})=0\Longrightarrow u\geq v.
Proof.

Since max(u,v)(X,θ,ϕ)\max(u,v)\in\mathcal{E}(X,\theta,\phi), we can assume that uv0u\leq v\leq 0. By Lemma 2.9 we have

θvn𝟏{u=v}θvn𝟏{u=v}θun=θun,\theta_{v}^{n}\geq{\bf 1}_{\{u=v\}}\theta_{v}^{n}\geq{\bf 1}_{\{u=v\}}\theta_{u}^{n}=\theta_{u}^{n},

where the last identity follows from the fact that θun(u<v)=0\theta_{u}^{n}(u<v)=0. Comparing the total mass we thus have θun=θvn\theta_{u}^{n}=\theta_{v}^{n} and the measure is concentrated on {u=v}\{u=v\}.

Fix b>1b>1 and set φb:=Pθ(bu(b1)v)\varphi_{b}:=P_{\theta}(bu-(b-1)v). Since the masses of uu and vv are equal the assumption in Theorem 3.10 is trivially satisfied, hence φbPSH(X,θ)\varphi_{b}\in{\rm PSH}(X,\theta). By Lemma 3.11 above, we have Xθφbn=Xθϕn>0\int_{X}\theta_{\varphi_{b}}^{n}=\int_{X}\theta_{\phi}^{n}>0.

Set ψb:=b1φb+(1b1)v\psi_{b}:=b^{-1}\varphi_{b}+(1-b^{-1})v and Db:={b1φb+(1b1)v=u}={φb=bu(b1)v}D_{b}:=\{b^{-1}\varphi_{b}+(1-b^{-1})v=u\}=\{\varphi_{b}=bu-(b-1)v\}. Note that ψbu\psi_{b}\leq u with equality on the contact set DbD_{b}. Lemma 2.9 and Theorem 2.7 thus ensure that

bnθφbnbnθφbn+𝟏Db(1b1)nθvn=𝟏Dbbnθφbn+𝟏Db(1b1)nθvn𝟏Dbθψbn𝟏Dbθun.b^{-n}\theta_{\varphi_{b}}^{n}\leq b^{-n}\theta_{\varphi_{b}}^{n}+{\bf 1}_{D_{b}}(1-b^{-1})^{n}\theta_{v}^{n}={\bf 1}_{D_{b}}b^{-n}\theta_{\varphi_{b}}^{n}+{\bf 1}_{D_{b}}(1-b^{-1})^{n}\theta_{v}^{n}\leq{\bf 1}_{D_{b}}\theta_{\psi_{b}}^{n}\leq{\bf 1}_{D_{b}}\theta_{u}^{n}.

We then infer

bnθφbnθun,b^{-n}\theta_{\varphi_{b}}^{n}\leq\theta_{u}^{n},

hence θφbn\theta_{\varphi_{b}}^{n} is concentrated on {φb=bu(b1)v=u}\{\varphi_{b}=bu-(b-1)v=u\}. Since φbu\varphi_{b}\leq u, it thus follows from Lemma 2.9 that

θφbn=𝟏{φb=u}θφbnθun.\theta_{\varphi_{b}}^{n}={\bf 1}_{\{\varphi_{b}=u\}}\theta_{\varphi_{b}}^{n}\leq\theta_{u}^{n}.

Comparing the total mass, we obtain θφbn=θun\theta_{\varphi_{b}}^{n}=\theta_{u}^{n}, hence

Xeφbθun=Xeφbθφbn=Xeuθun>0.\int_{X}e^{\varphi_{b}}\theta_{u}^{n}=\int_{X}e^{\varphi_{b}}\theta_{\varphi_{b}}^{n}=\int_{X}e^{u}\theta_{u}^{n}>0.

Note also that, since uvu\leq v, φb\varphi_{b} is decreasing in bb. The above thus implies that φ:=limbφb\varphi:=\lim_{b\to\infty}\varphi_{b} is not identically -\infty. Moreover, for any x{u<v}x\in\{u<v\}, we have φb(x)bu(x)(b1)v(x)v(x)+b(u(x)v(x))\varphi_{b}(x)\leq bu(x)-(b-1)v(x)\leq v(x)+b(u(x)-v(x)). Letting bb\to\infty, we see that φ(x)=\varphi(x)=-\infty. This implies that {u<v}\{u<v\} is pluripolar, hence empty. It then follows that uvu\geq v as desired. ∎

The following result states that the non-pluripolar Monge–Ampère measure determines the potential within a relative full mass class. This extends a result of Dinew from the Kähler case [Din09].

Theorem 3.13.

Assume ϕ\phi is a model potential and u,v(X,θ,ϕ)u,v\in\mathcal{E}(X,\theta,\phi). Then

θun=θvnuvis constant.\theta_{u}^{n}=\theta_{v}^{n}\Longrightarrow u-v\;\text{is constant}.
Proof.

We normalize u,vu,v by supXu=supXv=0\sup_{X}u=\sup_{X}v=0. The goal is then to prove that u=vu=v. Then max(u,v)(X,θ,ϕ)\max(u,v)\in\mathcal{E}(X,\theta,\phi) and by Lemma 2.9, θmax(u,v)nμ:=θun\theta_{\max(u,v)}^{n}\geq\mu:=\theta_{u}^{n}. Comparing the total mass we see that θmax(u,v)n=μ\theta_{\max(u,v)}^{n}=\mu. Thus, replacing vv with max(u,v)\max(u,v), we can assume that uvu\leq v.

Fix b>1b>1 and set φb:=Pθ(bu(b1)v)\varphi_{b}:=P_{\theta}(bu-(b-1)v). Since the masses of uu and vv are equal, Theorem 3.10 ensures that φbPSH(X,θ)\varphi_{b}\in{\rm PSH}(X,\theta). By Lemma 3.11, we have Xθφbn=Xθϕn>0\int_{X}\theta_{\varphi_{b}}^{n}=\int_{X}\theta_{\phi}^{n}>0. Since φbu\varphi_{b}\leq u, we infer that φb(X,θ,ϕ)\varphi_{b}\in\mathcal{E}(X,\theta,\phi).

Set ψb:=b1φb+(1b1)v\psi_{b}:=b^{-1}\varphi_{b}+(1-b^{-1})v and Db:={b1φb+(1b1)v=u}D_{b}:=\{b^{-1}\varphi_{b}+(1-b^{-1})v=u\}. Note that ψbu\psi_{b}\leq u with equality on the contact set DbD_{b}. Lemma 2.9 and Theorem 2.7 thus ensure that

𝟏Dbθψbn𝟏Dbθun=𝟏Dbμ,{\bf 1}_{D_{b}}\theta_{\psi_{b}}^{n}\leq{\bf 1}_{D_{b}}\theta_{u}^{n}={\bf 1}_{D_{b}}\mu,

hence

bnθφbn+𝟏Db(1b1)nθvn=𝟏Dbbnθφbn+𝟏Db(1b1)nθvn𝟏Dbθψbn𝟏Dbθun.b^{-n}\theta_{\varphi_{b}}^{n}+{\bf 1}_{D_{b}}(1-b^{-1})^{n}\theta_{v}^{n}={\bf 1}_{D_{b}}b^{-n}\theta_{\varphi_{b}}^{n}+{\bf 1}_{D_{b}}(1-b^{-1})^{n}\theta_{v}^{n}\leq{\bf 1}_{D_{b}}\theta_{\psi_{b}}^{n}\leq{\bf 1}_{D_{b}}\theta_{u}^{n}.

It thus follows that θφbn=fμ\theta_{\varphi_{b}}^{n}=f\mu for some fL1(μ)f\in L^{1}(\mu) whose support is contained in DbD_{b}. The mixed Monge–Ampère inequalities, [BEGZ10, Proposition 1.11] yield

θφbjθvnjfj/nμ,\theta_{\varphi_{b}}^{j}\wedge\theta_{v}^{n-j}\geq f^{j/n}\mu,

hence

θψbn\displaystyle\theta_{\psi_{b}}^{n} =j=0n(nj)(b1θφb)j((1b1)θv)nj\displaystyle=\sum_{j=0}^{n}\binom{n}{j}(b^{-1}\theta_{\varphi_{b}})^{j}\wedge\left((1-b^{-1})\theta_{v}\right)^{n-j}
j=0n(nj)bj(1b1)njfjnμ\displaystyle\geq\sum_{j=0}^{n}\binom{n}{j}b^{-j}(1-b^{-1})^{n-j}f^{\frac{j}{n}}\mu
=(b1f1/n+1b1)nμ.\displaystyle=(b^{-1}f^{1/n}+1-b^{-1})^{n}\mu. (3.5)

Since 𝟏Dbμ𝟏Dbθψbn{\bf 1}_{D_{b}}\mu\geq{\bf 1}_{D_{b}}\theta_{\psi_{b}}^{n}, it follows from the above that f1/n1f^{1/n}\leq 1 on DbD_{b}. Since ff is supported in DbD_{b} we get f1f\leq 1, and so θφbnμ\theta_{\varphi_{b}}^{n}\leq\mu. Comparing the total masses (noting that φb(X,θ,ϕ)\varphi_{b}\in\mathcal{E}(X,\theta,\phi)) gives f=1f=1. Hence θφbn=μ\theta_{\varphi_{b}}^{n}=\mu, and from (3.2) we also get θψbnμ\theta_{\psi_{b}}^{n}\geq\mu. Since ψbu\psi_{b}\leq u, we then infer, via Theorem 3.3, that θφbn=θψbn=μ\theta_{\varphi_{b}}^{n}=\theta_{\psi_{b}}^{n}=\mu is concentrated on {φb=bu(b1)v}\{\varphi_{b}=bu-(b-1)v\}. Now, {ψb<u}={φb<bu(b1)v}\{\psi_{b}<u\}=\{\varphi_{b}<bu-(b-1)v\}, hence

θψbn(ψb<u)=0.\theta_{\psi_{b}}^{n}(\psi_{b}<u)=0.

Invoking the domination principle (Theorem 3.12) we obtain ψb=u\psi_{b}=u, hence φb=bu(b1)v\varphi_{b}=bu-(b-1)v. Since uv0u\leq v\leq 0 and supXu=0\sup_{X}u=0, it follows that there exists xXx\in X with u(x)=v(x)=0u(x)=v(x)=0. We then infer supXφb=0\sup_{X}\varphi_{b}=0, hence the functions φb=v+b(uv)\varphi_{b}=v+b(u-v) decrease to some φPSH(X,θ)\varphi\in{\rm PSH}(X,\theta) which is -\infty in {u<v}\{u<v\}. It follows that {u<v}\{u<v\} is a pluripolar set, hence uvu\geq v almost everywhere. Since u,vu,v are quasi-psh functions we can conclude that uvu\geq v everywhere. ∎

Maximality of model potentials.

Based on our previous findings, one wonders if the following set of potentials has a maximal element:

Fϕ:={vPSH(X,θ):ϕv0andXθvn=Xθϕn}.F_{\phi}:=\bigg{\{}v\in{\rm PSH}(X,\theta)\;:\;\phi\leq v\leq 0\ \textrm{and}\ \int_{X}\theta_{v}^{n}=\int_{X}\theta_{\phi}^{n}\bigg{\}}.

In other words, does there exist a least singular potential that is less singular than ϕ\phi but has the same full mass as ϕ\phi. As shown in the following result, if Xθϕn>0\int_{X}\theta_{\phi}^{n}>0, this is indeed the case, moreover this maximal potential is equal to Pθ[ϕ]P_{\theta}[\phi].

Theorem 3.14.

Assume that ϕPSH(X,θ)\phi\in{\rm PSH}(X,\theta) and Xθϕn>0\int_{X}\theta_{\phi}^{n}>0. Then

Pθ[ϕ]=supvFϕv.P_{\theta}[\phi]=\sup_{v\in F_{\phi}}v.

In particular, Pθ[ϕ]=Pθ[Pθ[ϕ]]P_{\theta}[\phi]=P_{\theta}[P_{\theta}[\phi]].

As shown in Remark 3.4, Pθ[ϕ]FϕP_{\theta}[\phi]\in F_{\phi}, hence by Theorem 3.14, Pθ[ϕ]P_{\theta}[\phi] is the maximal element of FϕF_{\phi}.

Proof.

Let vFϕv\in F_{\phi}. By Theorem 3.6 we have

θPθ[ϕ]n({Pθ[ϕ]<v})\displaystyle\theta_{P_{\theta}[\phi]}^{n}(\{P_{\theta}[\phi]<v\}) 𝟏{Pθ[ϕ]=0}θn({Pθ[ϕ]<v})𝟏{Pθ[ϕ]=0}θn({Pθ[ϕ]<0})=0.\displaystyle\leq{\bf 1}_{\{P_{\theta}[\phi]=0\}}\theta^{n}(\{P_{\theta}[\phi]<v\})\leq{\bf 1}_{\{P_{\theta}[\phi]=0\}}\theta^{n}(\{P_{\theta}[\phi]<0\})=0.

As Xθϕn=Xθvn\int_{X}\theta_{\phi}^{n}=\int_{X}\theta_{v}^{n}, by Remark 3.4 we have

XθPθ[ϕ]n=Xθϕn=Xθvn=XθPθ[v]n>0.\int_{X}\theta^{n}_{P_{\theta}[\phi]}=\int_{X}\theta_{\phi}^{n}=\int_{X}\theta_{v}^{n}=\int_{X}\theta^{n}_{P_{\theta}[v]}>0.

Consequently, Pθ[ϕ],v(X,θ,Pθ[v])P_{\theta}[\phi],v\in\mathcal{E}(X,\theta,P_{\theta}[v]) and Theorem 3.12 now ensures that Pθ[ϕ]vP_{\theta}[\phi]\geq v, hence Pθ[ϕ]supvFϕvP_{\theta}[\phi]\geq\sup_{v\in F_{\phi}}v. As Pθ[ϕ]FϕP_{\theta}[\phi]\in F_{\phi}, it follows that Pθ[ϕ]=supvFϕvP_{\theta}[\phi]=\sup_{v\in F_{\phi}}v.

For the last statement notice that Pθ[ϕ]=supvFϕvsupvFPθ[ϕ]v=Pθ[Pθ[ϕ]]P_{\theta}[\phi]=\sup_{v\in F_{\phi}}v\geq\sup_{v\in F_{P_{\theta}[\phi]}}v=P_{\theta}[P_{\theta}[\phi]], since FϕFPθ[ϕ]F_{\phi}\supset F_{P_{\theta}[\phi]}. The reverse inequality is trivial. ∎

As a consequence of this last result, we obtain the following characterization of membership in (X,θ,ϕ)\mathcal{E}(X,\theta,\phi).

Theorem 3.15.

Suppose ϕPSH(X,θ)\phi\in\textup{PSH}(X,\theta) with Xθϕn>0\int_{X}\theta_{\phi}^{n}>0 and ϕ0\phi\leq 0. The following are equivalent:
(i) u(X,θ,ϕ)u\in\mathcal{E}(X,\theta,\phi).
(ii) ϕ\phi is less singular than uu, and Pθ[u](ϕ)=ϕP_{\theta}[u](\phi)=\phi.
(iii) ϕ\phi is less singular than uu, and Pθ[u]=Pθ[ϕ]P_{\theta}[u]=P_{\theta}[\phi].

As a consequence of the equivalence between (i) and (iii), we see that the potential Pθ[u]P_{\theta}[u] stays the same for all u(X,θ,ϕ)u\in\mathcal{E}(X,\theta,\phi), i.e., it is an invariant of this class. In particular, since (X,θ,ϕ)(X,θ,Pθ[ϕ])\mathcal{E}(X,\theta,\phi)\subset\mathcal{E}(X,\theta,P_{\theta}[\phi]), by the last statement of Theorem 3.14, it seems natural to only consider potentials ϕ\phi that are in the image of the operator ψPθ[ψ]\psi\to P_{\theta}[\psi], when studying classes of relative full mass (X,θ,ϕ)\mathcal{E}(X,\theta,\phi). What is more, in the next section it will be clear that considering such ϕ\phi is not just more natural, but also necessary when trying to solve complex Monge–Ampère equations with prescribed singularity.

Proof.

Assume that (i) holds. By Theorem 3.6 it follows that Pθ[u](ϕ)ϕP_{\theta}[u](\phi)\geq\phi a.e. with respect to θPθ[u](ϕ)n\theta^{n}_{P_{\theta}[u](\phi)}. Theorem 3.12 gives Pθ[u](ϕ)=ϕP_{\theta}[u](\phi)=\phi, hence (ii) holds.

Suppose (ii) holds. We can assume that uϕ0u\leq\phi\leq 0. Then Pθ[u]Pθ[u](ϕ)=ϕP_{\theta}[u]\geq P_{\theta}[u](\phi)=\phi. By the last statement of the previous theorem, this implies that

Pθ[u]=Pθ[Pθ[u]]Pθ[ϕ].P_{\theta}[u]=P_{\theta}[P_{\theta}[u]]\geq P_{\theta}[\phi].

As the reverse inequality is trivial, (iii) follows.

Lastly, assume that (iii) holds. By Remark 3.4 it follows that

Xθun=XθPθ[u]n=XθPθ[ϕ]n=Xθϕn,\int_{X}\theta_{u}^{n}=\int_{X}\theta_{P_{\theta}[u]}^{n}=\int_{X}\theta_{P_{\theta}[\phi]}^{n}=\int_{X}\theta_{\phi}^{n},

hence (i) holds. ∎

Corollary 3.16.

Suppose ϕPSH(X,θ)\phi\in\textup{PSH}(X,\theta) such that Xθϕn>0\int_{X}\theta_{\phi}^{n}>0. Then (X,θ,ϕ)\mathcal{E}(X,\theta,\phi) is convex. Moreover, given ψ1,,ψn(X,θ,ϕ)\psi_{1},\ldots,\psi_{n}\in\mathcal{E}(X,\theta,\phi) we have

Xθψ1s1θψnsn=Xθϕn,\int_{X}\theta_{\psi_{1}}^{s_{1}}\wedge\ldots\wedge\theta_{\psi_{n}}^{s_{n}}=\int_{X}\theta_{\phi}^{n}, (3.6)

where sj0s_{j}\geq 0 are integers such that j=1nsj=n\sum_{j=1}^{n}s_{j}=n.

Proof.

Let u,v(X,θ,ϕ)u,v\in\mathcal{E}(X,\theta,\phi) and fix t(0,1)t\in(0,1). It follows from Theorem 3.15 that Pθ[v](ϕ)=Pθ[u](ϕ)=ϕP_{\theta}[v](\phi)=P_{\theta}[u](\phi)=\phi. By Lemma 2.12,

Pθ[tv+(1t)u](ϕ)tPθ[v](ϕ)+(1t)Pθ[u](ϕ)=ϕ.P_{\theta}[tv+(1-t)u](\phi)\geq tP_{\theta}[v](\phi)+(1-t)P_{\theta}[u](\phi)=\phi.

As the reverse inequality is trivial, another application of Theorem 3.15 gives tv+(1t)u(X,θ,ϕ)tv+(1-t)u\in\mathcal{E}(X,\theta,\phi).

We prove the last statement. Since (X,θ,ϕ)\mathcal{E}(X,\theta,\phi) is convex, given ψ1,,ψn(X,θ,ϕ)\psi_{1},\ldots,\psi_{n}\in\mathcal{E}(X,\theta,\phi) we know that any convex combination ψ:=j=1nsjψj\psi:=\sum_{j=1}^{n}s_{j}\psi_{j} with 0sj10\leq s_{j}\leq 1 and jsj=n\sum_{j}s_{j}=n, belongs to (X,θ,ϕ)\mathcal{E}(X,\theta,\phi). Hence

X(jsjθψj)n=Xθψn=Xθϕn=X(jsjθϕ)n.\int_{X}\bigg{(}\sum_{j}s_{j}\theta_{\psi_{j}}\bigg{)}^{n}=\int_{X}\theta_{\psi}^{n}=\int_{X}\theta_{\phi}^{n}=\int_{X}\bigg{(}\sum_{j}s_{j}\theta_{\phi}\bigg{)}^{n}.

We have an identity of two homogeneous polynomials of degree nn. All the coefficients of these polynomials have to be equal, giving (3.6). ∎

Lemma 3.17.

Assume ϕPSH(X,θ)\phi\in{\rm PSH}(X,\theta) is a model potential. If b>1b>1 and u,v(X,θ,ϕ)u,v\in\mathcal{E}(X,\theta,\phi) then Pθ(bu(b1)v)(X,θ,ϕ)P_{\theta}(bu-(b-1)v)\in\mathcal{E}(X,\theta,\phi).

Proof.

The proof is similar to that of [LN22, Corollary 3.20]. Since u,v(X,θ,ϕ)u,v\in\mathcal{E}(X,\theta,\phi), both uu and vv are more singular than ϕ\phi, we can assume that max(u,v)ϕ\max(u,v)\leq\phi. It follows from Theorem 3.10 (and the comment below it) that Pθ(bu(b1)ϕ)PSH(X,θ)P_{\theta}(bu-(b-1)\phi)\in{\rm PSH}(X,\theta) for all b>1b>1. Lemma 3.11 ensures that Pθ(bu(b1)ϕ)(X,θ,ϕ)P_{\theta}(bu-(b-1)\phi)\in\mathcal{E}(X,\theta,\phi) for all b>1b>1. Set ψ:=Pθ(bu(b1)v)\psi:=P_{\theta}(bu-(b-1)v) and φ:=Pθ(bu(b1)ϕ)\varphi:=P_{\theta}(bu-(b-1)\phi). Then ψφ\psi\geq\varphi and φ(X,θ,ϕ)\varphi\in\mathcal{E}(X,\theta,\phi). We also have

b1ψ+(1b1)vu,b^{-1}\psi+(1-b^{-1})v\leq u,

which, by Lemma 2.12, gives

b1Pθ[ψ]+(1b1)Pθ[v]Pθ[b1ψ+(1b1)v]Pθ[u]Pθ[ϕ]=ϕ.b^{-1}P_{\theta}[\psi]+(1-b^{-1})P_{\theta}[v]\leq P_{\theta}[b^{-1}\psi+(1-b^{-1})v]\leq P_{\theta}[u]\leq P_{\theta}[\phi]=\phi.

Since Pθ[v]=ϕP_{\theta}[v]=\phi, we can thus conclude that Pθ[ψ]ϕP_{\theta}[\psi]\leq\phi. On the other hand ϕ=Pθ[φ]Pθ[ψ]\phi=P_{\theta}[\varphi]\leq P_{\theta}[\psi]. Thus Pθ[ψ]=ϕP_{\theta}[\psi]=\phi and ψ(X,θ,ϕ)\psi\in\mathcal{E}(X,\theta,\phi). ∎

Lemma 3.18.

Assume that u,v,wPSH(X,θ)u,v,w\in\textup{PSH}(X,\theta) are such that Xθun+Xθvn>Xθwn\int_{X}\theta_{u}^{n}+\int_{X}\theta_{v}^{n}>\int_{X}\theta_{w}^{n} and max(u,v)w\max(u,v)\leq w. Then Pθ(u,v)PSH(X,θ)P_{\theta}(u,v)\in{\rm PSH}(X,\theta).

Proof.

We can assume without loss of generality that u,v,w0u,v,w\leq 0. Replacing ww with Pθ[εVθ+(1ε)w]P_{\theta}[\varepsilon V_{\theta}+(1-\varepsilon)w] for small enough ε>0\varepsilon>0 we can also assume that ww is a model potential.

For j0j\geq 0 we set uj:=max(u,wj),vj:=max(v,wj)u_{j}:=\max(u,w-j),v_{j}:=\max(v,w-j), hj:=P(uj,vj)h_{j}:=P(u_{j},v_{j}). Observe that uj,vj,hju_{j},v_{j},h_{j} have the same singularity type as ww. Fix s>0s>0 big enough, such that for all j>sj>s, we have

{u>ws}θujn+{v>ws}θvjn={u>ws}θun+{v>ws}θvn>Xθwn,\int_{\{u>w-s\}}\theta_{u_{j}}^{n}+\int_{\{v>w-s\}}\theta_{v_{j}}^{n}=\int_{\{u>w-s\}}\theta_{u}^{n}+\int_{\{v>w-s\}}\theta_{v}^{n}>\int_{X}\theta_{w}^{n},

where in the equality above we used Lemma 2.1. Observe that such ss exists thanks to the assumption.
It follows from Theorem 3.6 and the above estimate, that for j>sj>s,

{hjws}θhjn\displaystyle\int_{\{h_{j}\leq w-s\}}\theta_{h_{j}}^{n} {ujws}θujn+{vjws}θvjn\displaystyle\leq\int_{\{u_{j}\leq w-s\}}\theta_{u_{j}}^{n}+\int_{\{v_{j}\leq w-s\}}\theta_{v_{j}}^{n}
=2Xθwn{u>ws}θun{v>ws}θvn<Xθwn,\displaystyle=2\int_{X}\theta_{w}^{n}-\int_{\{u>w-s\}}\theta_{u}^{n}-\int_{\{v>w-s\}}\theta_{v}^{n}<\int_{X}\theta_{w}^{n},

where in the identity above we used the fact that {ujws}={uws}\{u_{j}\leq w-s\}=\{u\leq w-s\} and that uj,vju_{j},v_{j} and ww have the same singularity type. Since uj,vju_{j},v_{j} decrease to u,vu,v respectively, it follows that hjP(u,v)h_{j}\searrow P(u,v). We now rule out the possibility that P(u,v)P(u,v)\equiv-\infty. Indeed, suppose supXhj\sup_{X}h_{j} decreases to -\infty. From Lemma 3.5 we obtain that supXhj=supX(hjw)\sup_{X}h_{j}=\sup_{X}(h_{j}-w)\searrow-\infty. But then, for jj large enough the set {hjws}\{h_{j}\leq w-s\} coincides with XX, contradicting our last integral estimate, since each hjh_{j} has the same singularity type as ww. ∎

Corollary 3.19.

If ϕPSH(X,θ)\phi\in{\rm PSH}(X,\theta) be a model potential and u,v(X,θ,ϕ)u,v\in\mathcal{E}(X,\theta,\phi) then Pθ(u,v)(X,θ,ϕ)P_{\theta}(u,v)\in\mathcal{E}(X,\theta,\phi).

Proof.

Fixing b>1b>1, from Lemma 3.17 we infer that ub:=Pθ(bu(b1)ϕ)u_{b}:=P_{\theta}(bu-(b-1)\phi) and vb:=Pθ(bv(b1)ϕ)v_{b}:=P_{\theta}(bv-(b-1)\phi) belong to (X,θ,ϕ)\mathcal{E}(X,\theta,\phi). Lemma 3.18 then ensures that Pθ(ub,vb)PSH(X,θ)P_{\theta}(u_{b},v_{b})\in{\rm PSH}(X,\theta). We also have

Pθ(u,v)b1Pθ(ub,vb)+(1b1)ϕ.P_{\theta}(u,v)\geq b^{-1}P_{\theta}(u_{b},v_{b})+(1-b^{-1})\phi.

Comparing the total mass via Theorem 3.3 and letting bb\to\infty, we obtain the result. ∎

Plainly speaking, by the next lemma, the fixed point set of the map ψP[ψ]\psi\to P[\psi] is stable under the operation (ψ,ϕ)P(ψ,ϕ)(\psi,\phi)\to P(\psi,\phi).

Lemma 3.20.

Suppose u0,u1PSH(X,θ)u_{0},u_{1}\in\textup{PSH}(X,\theta) are such that P(u0,u1)PSH(X,θ)P(u_{0},u_{1})\in\textup{PSH}(X,\theta), and P[u0]=u0P[u_{0}]=u_{0} and P[u1]=u1P[u_{1}]=u_{1}. Then P[P(u0,u1)]=P(u0,u1).P[P(u_{0},u_{1})]=P(u_{0},u_{1}).

Proof.

As P(u0,u1)min(u0,u1)0P(u_{0},u_{1})\leq\min(u_{0},u_{1})\leq 0 and P[P(u0,u1)]min(P[u0],P[u1])P[P(u_{0},u_{1})]\leq\min(P[u_{0}],P[u_{1}]), it follows that

P(u0,u1)P[P(u0,u1)]P(P[u0],P[u1])=P(u0,u1).P(u_{0},u_{1})\leq P[P(u_{0},u_{1})]\leq P(P[u_{0}],P[u_{1}])=P(u_{0},u_{1}).

This shows that all the inequalities above are in fact equalities. ∎

Proposition 3.21.

Let ϕ,ψPSH(X,θ)\phi,\psi\in\textup{PSH}(X,\theta) be such that ϕ=P[ϕ]\phi=P[\phi], ψ=P[ψ]\psi=P[\psi], and P(ϕ,ψ)PSH(X,θ)P(\phi,\psi)\in{\rm PSH}(X,\theta). If u(X,θ,ϕ)u\in\mathcal{E}(X,\theta,\phi), v(X,θ,ψ)v\in\mathcal{E}(X,\theta,\psi) and XθP(ϕ,ψ)n>0\int_{X}\theta_{P(\phi,\psi)}^{n}>0 then

P(u,v)(X,θ,P(ϕ,ψ)).P(u,v)\in\mathcal{E}(X,\theta,P(\phi,\psi)).
Proof.

We can assume that uϕu\leq\phi and vψv\leq\psi.

Step 1. We first prove that P(u,ψ)(X,θ,P(ϕ,ψ))P(u,\psi)\in\mathcal{E}(X,\theta,P(\phi,\psi)). By assumption we have

Xθun+XθP(ϕ,ψ)n>Xθϕn,\int_{X}\theta_{u}^{n}+\int_{X}\theta_{P(\phi,\psi)}^{n}>\int_{X}\theta_{\phi}^{n},

and Lemma 3.18 gives P(u,ψ)=P(u,P(ϕ,ψ))PSH(X,θ)P(u,\psi)=P(u,P(\phi,\psi))\in{\rm PSH}(X,\theta). Also, it follows from Theorem 3.10 that ub:=Pθ(bu(b1)ϕ)PSH(X,θ)u_{b}:=P_{\theta}(bu-(b-1)\phi)\in{\rm PSH}(X,\theta) for all b>1b>1. By definition, for 1<b<t1<b<t we have

ϕubbt1ut+(1bt1)ϕ.\phi\geq u_{b}\geq bt^{-1}u_{t}+(1-bt^{-1})\phi.

Comparing the total mass via Theorem 3.3 and letting tt\to\infty we see that ub(X,θ,ϕ)u_{b}\in\mathcal{E}(X,\theta,\phi). As above, Lemma 3.18 ensures that P(ub,ψ)PSH(X,θ)P(u_{b},\psi)\in{\rm PSH}(X,\theta). On the other hand we also have

ϕub1ub+(1b1)ϕ,\phi\geq u\geq b^{-1}u_{b}+(1-b^{-1})\phi,

therefore P(ϕ,ψ)P(u,ψ)b1P(ub,ψ)+(1b1)P(ϕ,ψ)P(\phi,\psi)\geq P(u,\psi)\geq b^{-1}P(u_{b},\psi)+(1-b^{-1})P(\phi,\psi). Comparing the total mass via Theorem 3.3 and letting bb\to\infty we arrive at XθP(ϕ,ψ)nXθP(u,ψ)nXθP(ϕ,ψ)n\int_{X}\theta_{P(\phi,\psi)}^{n}\geq\int_{X}\theta_{P(u,\psi)}^{n}\geq\int_{X}\theta_{P(\phi,\psi)}^{n}, hence the conclusion.

Step 2. We prove that P(u,v)PSH(X,θ)P(u,v)\in{\rm PSH}(X,\theta). It follows from Theorem 3.3, the assumption v(X,θ,ψ)v\in\mathcal{E}(X,\theta,\psi), and the first step that

XθP(u,ψ)n+Xθvn=XθP(ϕ,ψ)n+Xθψn>Xθψn.\int_{X}\theta_{P(u,\psi)}^{n}+\int_{X}\theta_{v}^{n}=\int_{X}\theta_{P(\phi,\psi)}^{n}+\int_{X}\theta_{\psi}^{n}>\int_{X}\theta_{\psi}^{n}.

Since max(P(u,ψ),v)ψ\max(P(u,\psi),v)\leq\psi, Lemma 3.18 can be applied giving P(u,v)=P(P(u,ψ),v)PSH(X,θ)P(u,v)=P(P(u,\psi),v)\in{\rm PSH}(X,\theta).

Step 3. We conclude the proof. It follows from Theorem 3.10 that vb:=Pθ(bv(b1)ψ)PSH(X,θ)v_{b}:=P_{\theta}(bv-(b-1)\psi)\in{\rm PSH}(X,\theta) for all b>1b>1. For 1<b<t1<b<t we have

ψvbbt1vt+(1bt1)ψ.\psi\geq v_{b}\geq bt^{-1}v_{t}+(1-bt^{-1})\psi.

Comparing the total mass via Theorem 3.3 and letting tt\to\infty we see that vb(X,θ,ψ)v_{b}\in\mathcal{E}(X,\theta,\psi). By the second step we have that P(u,vb)PSH(X,θ)P(u,v_{b})\in{\rm PSH}(X,\theta). On the other hand we also have

ψvb1vb+(1b1)ψ,\psi\geq v\geq b^{-1}v_{b}+(1-b^{-1})\psi,

therefore P(u,ψ)P(u,v)b1P(u,vb)+(1b1)P(u,ψ)P(u,\psi)\geq P(u,v)\geq b^{-1}P(u,v_{b})+(1-b^{-1})P(u,\psi). Comparing the total mass via Theorem 3.3 and letting bb\to\infty we arrive at XθP(u,ψ)nXθP(u,v)nXθP(u,ψ)n\int_{X}\theta_{P(u,\psi)}^{n}\geq\int_{X}\theta_{P(u,v)}^{n}\geq\int_{X}\theta_{P(u,\psi)}^{n}. Combining this and the first step we arrive at the conclusion. ∎

Comparison principle.

We note the partial comparison principle for functions s of relative full mass, generalizing a result of Dinew from [Din09]:

Proposition 3.22.

Suppose ψkPSH(X,θk),k=1,,jn\psi_{k}\in\textup{PSH}(X,\theta^{k}),k=1,\ldots,j\leq n and ϕPSH(X,θ)\phi\in{\rm PSH}(X,\theta) is a model potential. If u,v(X,θ,ϕ)u,v\in\mathcal{E}(X,\theta,\phi) then

{u<v}θvnjθψ11θψjj{u<v}θunjθψ11θψjj.\int_{\{u<v\}}\theta_{v}^{n-j}\wedge\theta^{1}_{\psi_{1}}\wedge\ldots\wedge\theta^{j}_{\psi_{j}}\leq\int_{\{u<v\}}\theta_{u}^{n-j}\wedge\theta^{1}_{\psi_{1}}\wedge\ldots\wedge\theta^{j}_{\psi_{j}}.
Proof.

The proof follows the argument of [BEGZ10, Proposition 2.2] with a vital ingredient from Theorem 3.3.

Since max(u,v)(X,θ,ϕ)\max(u,v)\in\mathcal{E}(X,\theta,\phi), we have, by Theorem 3.15, Pθ[max(u,v)]=Pθ[u]=Pθ[v]=ϕP_{\theta}[\max(u,v)]=P_{\theta}[u]=P_{\theta}[v]=\phi. It thus follows from Theorem 3.3 and Remark 3.4 that

Xθϕnjθψ11θψjj\displaystyle\int_{X}\theta_{\phi}^{n-j}\wedge\theta^{1}_{\psi_{1}}\wedge\ldots\wedge\theta^{j}_{\psi_{j}} =\displaystyle= Xθvnjθψ11θψjj\displaystyle\int_{X}\theta_{v}^{n-j}\wedge\theta^{1}_{\psi_{1}}\wedge\ldots\wedge\theta^{j}_{\psi_{j}}
\displaystyle\leq Xθmax(u,v)njθψ11θψjj\displaystyle\int_{X}\theta_{\max(u,v)}^{n-j}\wedge\theta^{1}_{\psi_{1}}\wedge\ldots\wedge\theta^{j}_{\psi_{j}}
=\displaystyle= Xθϕnjθψ11θψjj.\displaystyle\int_{X}\theta_{\phi}^{n-j}\wedge\theta^{1}_{\psi_{1}}\wedge\ldots\wedge\theta^{j}_{\psi_{j}}.

Hence the inequality above is in fact equality. By locality of the non-pluripolar product we then can write:

Xθmax(u,v)njθψ11θψjj\displaystyle\int_{X}\theta_{\max(u,v)}^{n-j}\wedge\theta^{1}_{\psi_{1}}\wedge...\wedge\theta^{j}_{\psi_{j}} \displaystyle\geq {u>v}θunjθψ11θψjj+{v>u}θvnjθψ11θψjj\displaystyle\int_{\{u>v\}}\theta_{u}^{n-j}\wedge\theta^{1}_{\psi_{1}}\wedge...\wedge\theta^{j}_{\psi_{j}}+\int_{\{v>u\}}\theta_{v}^{n-j}\wedge\theta^{1}_{\psi_{1}}\wedge...\wedge\theta^{j}_{\psi_{j}}
=\displaystyle= Xθunjθψ11θψjj{uv}θunjθψ11θψjj\displaystyle\int_{X}\theta_{u}^{n-j}\wedge\theta^{1}_{\psi_{1}}\wedge...\wedge\theta^{j}_{\psi_{j}}-\int_{\{u\leq v\}}\theta_{u}^{n-j}\wedge\theta^{1}_{\psi_{1}}\wedge...\wedge\theta^{j}_{\psi_{j}}
+{v>u}θvnjθψ11θψjj\displaystyle+\int_{\{v>u\}}\theta_{v}^{n-j}\wedge\theta^{1}_{\psi_{1}}\wedge...\wedge\theta^{j}_{\psi_{j}}
=\displaystyle= Xθmax(u,v)njθψ11θψjj{uv}θunjθψ11θψjj\displaystyle\int_{X}\theta_{\max(u,v)}^{n-j}\wedge\theta^{1}_{\psi_{1}}\wedge...\wedge\theta^{j}_{\psi_{j}}-\int_{\{u\leq v\}}\theta_{u}^{n-j}\wedge\theta^{1}_{\psi_{1}}\wedge...\wedge\theta^{j}_{\psi_{j}}
+{v>u}θvnjθψ11θψjj.\displaystyle+\int_{\{v>u\}}\theta_{v}^{n-j}\wedge\theta^{1}_{\psi_{1}}\wedge...\wedge\theta^{j}_{\psi_{j}}.

We thus get

{u<v}θvnjθψ11θψjj{uv}θunjθψ11θψjj.\displaystyle\int_{\{u<v\}}\theta_{v}^{n-j}\wedge\theta^{1}_{\psi_{1}}\wedge\ldots\wedge\theta^{j}_{\psi_{j}}\leq\int_{\{u\leq v\}}\theta_{u}^{n-j}\wedge\theta^{1}_{\psi_{1}}\wedge\ldots\wedge\theta^{j}_{\psi_{j}}.

Replacing uu with u+εu+\varepsilon in the above inequality, and letting ε0\varepsilon\searrow 0, by the monotone convergence theorem we arrive at the result. ∎

The above result yields the following important consequence, generalizing [BEGZ10, Corollary 2.3].:

Corollary 3.23.

Suppose ϕPSH(X,θ)\phi\in\textup{PSH}(X,\theta) is a model potential and assume that u,v(X,θ,ϕ)u,v\in\mathcal{E}(X,\theta,\phi). Then

{u<v}θvn{u<v}θun and {uv}θvn{uv}θun.\int_{\{u<v\}}\theta_{v}^{n}\leq\int_{\{u<v\}}\theta_{u}^{n}\quad\textup{ and }\quad\int_{\{u\leq v\}}\theta_{v}^{n}\leq\int_{\{u\leq v\}}\theta_{u}^{n}.

The second inequality follows from the first inequality applied to uu and v+εv+\varepsilon and ε0\varepsilon\searrow 0.

Chapter 4 Generalized Monge–Ampère capacities and integration by parts

4.1 Generalized Monge–Ampère capacities

In this section we prove a comparison of Monge–Ampère capacities which will be used in the proof of the integration by parts formula in the next section. We first start with a version of the Chern-Levine-Nirenberg inequality.

Lemma 4.1.

Let u,v,ψPSH(X,ω)u,v,\psi\in{\rm PSH}(X,\omega). Assume that vuv+Bv\leq u\leq v+B for some positive constant BB. Then

XψωunXψωvnnBXωn.\int_{X}\psi\omega_{u}^{n}\geq\int_{X}\psi\omega_{v}^{n}-nB\int_{X}\omega^{n}.
Proof.

By the monotone convergence theorem we can assume that ψ\psi is bounded. By subtracting a constant we can assume that u0u\leq 0. We first prove the lemma under the assumption that u=vu=v on the open set

U:={min(u,v)=v<C},U:=\{\min(u,v)=v<-C\},

for some positive constant CC.
We approximate uu and vv by ut:=max(u,t)u^{t}:=\max(u,-t) and vt:=max(v,t)v^{t}:=\max(v,-t). For t>0t>0 we apply the integration by parts formula for bounded ω\omega-psh functions, which is a consequence of Stokes theorem, to get

Xψ(ωutnωvtn)=X(utvt)𝑑dcψSt,\int_{X}\psi(\omega_{u^{t}}^{n}-\omega_{v^{t}}^{n})=\int_{X}(u^{t}-v^{t})dd^{c}\psi\wedge S^{t},

where St:=k=0n1ωutkωvtn1kS^{t}:=\sum_{k=0}^{n-1}\omega_{u^{t}}^{k}\wedge\omega_{v^{t}}^{n-1-k}. Note that XωSt=nXωn\int_{X}\omega\wedge S^{t}=n\int_{X}\omega^{n} as there are nn terms in the sum and each of them is equal to Xωn\int_{X}\omega^{n} by Stokes’ theorem. Since utvtu^{t}\geq v^{t} we can continue the above estimate and obtain

Xψ(ωutnωvtn)=X(utvt)(ωψStωSt)BnXωn.\int_{X}\psi(\omega_{u^{t}}^{n}-\omega_{v^{t}}^{n})=\int_{X}(u^{t}-v^{t})(\omega_{\psi}\wedge S^{t}-\omega\wedge S^{t})\geq-Bn\int_{X}\omega^{n}.

For t>B+Ct>B+C we have that ut=vtu^{t}=v^{t} on the open set UU which contains {ut}={vt}\{u\leq-t\}=\{v\leq-t\}. It thus follows that 𝟏Uωutn=𝟏Uωvtn{\bf 1}_{U}\omega_{u^{t}}^{n}={\bf 1}_{U}\omega_{v^{t}}^{n}. Thus, for t>B+Ct>B+C we have

{v>t}ψ(ωunωvn)={v>t}ψ(ωutnωvtn)\displaystyle\int_{\{v>-t\}}\psi(\omega_{u}^{n}-\omega_{v}^{n})=\int_{\{v>-t\}}\psi(\omega_{u_{t}}^{n}-\omega_{v_{t}}^{n}) =\displaystyle= Xψ(ωutnωvtn)BnXωn.\displaystyle\int_{X}\psi(\omega_{u^{t}}^{n}-\omega_{v^{t}}^{n})\geq-Bn\int_{X}\omega^{n}.

Letting tt\to\infty we finish the first step.

We now treat the general case. By approximating ψ\psi from above by smooth ω\omega-psh functions, see [Dem94], [BK07], we can assume that ψ\psi is smooth (in fact, we only need the continuity of ψ\psi). We fix a(0,1)a\in(0,1) and set va:=avv_{a}:=av, ua:=max(u,va)u_{a}:=\max(u,v_{a}). Setting C:=a(1a)1BC:=a(1-a)^{-1}B we have that ua=vau_{a}=v_{a} on U={min(ua,va)=va<C}U=\{\min(u_{a},v_{a})=v_{a}<-C\} (see arguments in Step 1 of the proof of Lemma 3.1). We can thus apply the first step to get

XψωuanXψωvannBXωn.\int_{X}\psi\omega_{u_{a}}^{n}\geq\int_{X}\psi\omega_{v_{a}}^{n}-nB\int_{X}\omega^{n}.

Observe that vavv_{a}\searrow v and uauu_{a}\searrow u as a1a\nearrow 1. Also, by the multilinearity of non-pluripolar products, we have

lima1X(ω+ddcva)n=lima1X((1a)ω+aωv)n=X(ω+ddcv)n.\lim_{a\to 1^{-}}\int_{X}(\omega+dd^{c}v_{a})^{n}=\lim_{a\to 1^{-}}\int_{X}((1-a)\omega+a\omega_{v})^{n}=\int_{X}(\omega+dd^{c}v)^{n}.

Recalling that vu0v\leq u\leq 0, we have uua=max(u,av)max(u,au)=auu\leq u_{a}=\max(u,av)\leq\max(u,au)=au. By Theorem 3.3 we thus have

XωunXωuanX(ω+addcu)n.\int_{X}\omega_{u}^{n}\leq\int_{X}\omega_{u_{a}}^{n}\leq\int_{X}(\omega+add^{c}u)^{n}.

Using the multilinearity of non-pluripolar products, we then have

X(ω+ddcu)nlima1X(ω+ddcua)nlima1X(ω+addcu)n=X(ω+ddcu)n.\int_{X}(\omega+dd^{c}u)^{n}\leq\lim_{a\to 1^{-}}\int_{X}(\omega+dd^{c}u_{a})^{n}\leq\lim_{a\to 1^{-}}\int_{X}(\omega+add^{c}u)^{n}=\int_{X}(\omega+dd^{c}u)^{n}.

Hence the above inequalities are equalities. It then follows from Theorem 2.6 that the positive measures ωuan,ωvan\omega_{u_{a}}^{n},\omega_{v_{a}}^{n} converge respectively to ωun,ωvn\omega_{u}^{n},\omega_{v}^{n} in the weak sense of Radon measures as a1a\nearrow 1. Since ψ\psi is continuous on XX we thus obtain

XψωunXψωvnnBXωn.\int_{X}\psi\omega_{u}^{n}\geq\int_{X}\psi\omega_{v}^{n}-nB\int_{X}\omega^{n}.

We next use the Chern-Levine-Nirenberg inequality to compare Monge–Ampère capacities.

Definition 4.2.

Given ϕPSH(X,θ)\phi\in{\rm PSH}(X,\theta) and EXE\subset X a Borel subset we define

Capθ,ϕ(E):=sup{Eθun:uPSH(X,θ),ϕ1uϕ}.{\rm Cap}_{\theta,\phi}(E):=\sup\left\{\int_{E}\theta_{u}^{n}\;:\;u\in{\rm PSH}(X,\theta),\;\phi-1\leq u\leq\phi\right\}.

Note that in the Kähler case a related notion of capacity has been studied in [DL17, DL15]. In the case when ϕ=Vθ\phi=V_{\theta} we recover the Monge–Ampère capacity used in [BEGZ10, Section 4.1].

Lemma 4.3.

The relative Monge–Ampère capacity Capθ,ϕ{\rm Cap}_{\theta,\phi} is inner regular, i.e.

Capθ,ϕ(E)=sup{Capθ,ϕ(K):KE,Kis compact}.{\rm Cap}_{\theta,\phi}(E)=\sup\{{\rm Cap}_{\theta,\phi}(K)\;:\;K\subset E,\;K\;\textrm{is compact}\}.
Proof.

By definition Capθ,ϕ(E)Capθ,ϕ(K){\rm Cap}_{\theta,\phi}(E)\geq{\rm Cap}_{\theta,\phi}(K) for any compact set KEK\subset E. Fix ε>0\varepsilon>0. There exists uPSH(X,θ)u\in{\rm PSH}(X,\theta) such that ϕ1uϕ\phi-1\leq u\leq\phi and

EθunCapθ,ϕ(E)ε.\int_{E}\theta_{u}^{n}\geq{\rm Cap}_{\theta,\phi}(E)-\varepsilon.

Since θun\theta_{u}^{n} is an inner regular Borel measure it follows that there exists a compact set KEK\subset E such that KθunEθunεCapθ,ϕ(E)2ε\int_{K}\theta_{u}^{n}\geq\int_{E}\theta_{u}^{n}-\varepsilon\geq{\rm Cap}_{\theta,\phi}(E)-2\varepsilon. Hence Capθ,ϕ(K)Capθ,ϕ(E)2ε{\rm Cap}_{\theta,\phi}(K)\geq{\rm Cap}_{\theta,\phi}(E)-2\varepsilon. Taking the supremum over all the compact set KEK\subset E, we arrive at the conclusion. ∎

Proposition 4.4.

Assume ϕPSH(X,ω)\phi\in{\rm PSH}(X,\omega) is a model potential. There exists a constant C>0C>0 depending on X,ω,nX,\omega,n such that, for all Borel set EE, we have

Capω,ϕ(E)CCapω(E)1/n.{\rm Cap}_{\omega,\phi}(E)\leq C{\rm Cap}_{\omega}(E)^{1/n}.
Proof.

By inner regularity of the capacities, we can assume that E=KE=K is compact. We can also assume that KK is non-pluripolar, otherwise the inequality is trivial. Let VKV_{K}^{*} be the global extremal function which it is defined as

VK:=sup{uPSH(X,ω),u0onK}.V_{K}:=\sup\{u\in{\rm PSH}(X,\omega),\;\;u\leq 0\;{\rm on}\;K\}.

We recall that VK0V_{K}^{*}\geq 0 and that let MK:=supXVK<M_{K}:=\sup_{X}V_{K}^{*}<\infty if and only if KK is non-pluripolar. By [GZ05, Theorem 5.2], ωVKn\omega_{V_{K}^{*}}^{n} is concentrated on KK. If MK<1M_{K}<1 then

Xωn=XωVKn=KωVKn=KωVK1nCapω(K),\int_{X}\omega^{n}=\int_{X}\omega_{V_{K}^{*}}^{n}=\int_{K}\omega_{V_{K}^{*}}^{n}=\int_{K}\omega_{V_{K^{*}}-1}^{n}\leq{\rm Cap}_{\omega}(K),

while Capω,ϕ(K)Xωn{\rm Cap}_{\omega,\phi}(K)\leq\int_{X}\omega^{n}. Hence for C(Xωn)11/nC\geq(\int_{X}\omega^{n})^{1-1/n}, the desired inequality holds.

Assume now that MK1M_{K}\geq 1. Then v:=MK1VK1v:=M_{K}^{-1}V_{K}^{*}-1 is ω\omega-psh and takes values in [1,0][-1,0], hence

MKnXωn=MKnXωVKn=MKnKωVKnKωvnCapω(K).M_{K}^{-n}\int_{X}\omega^{n}=M_{K}^{-n}\int_{X}\omega_{V_{K}^{*}}^{n}=M_{K}^{-n}\int_{K}\omega_{V_{K}^{*}}^{n}\leq\int_{K}\omega_{v}^{n}\leq{\rm Cap}_{\omega}(K).

Set ψ:=VKMK\psi:=V_{K}^{*}-M_{K} and let uu be a ω\omega-psh function such that ϕ1uϕ\phi-1\leq u\leq\phi. Since ψ=MK\psi=-M_{K} on KK modulo a pluripolar set, it follows from Lemma 4.1 that

MKKωunX(ψ)ωunX(ψ)ωϕn+nXωnX(ψ)ωn+nXωnA,M_{K}\int_{K}\omega_{u}^{n}\leq\int_{X}(-\psi)\omega_{u}^{n}\leq\int_{X}(-\psi)\omega_{\phi}^{n}+n\int_{X}\omega^{n}\leq\int_{X}(-\psi)\omega^{n}+n\int_{X}\omega^{n}\leq A,

for a constant A>0A>0 depending on X,ω,nX,\omega,n. We have used above the fact that, ωϕn𝟏{ϕ=0}ωn\omega_{\phi}^{n}\leq{\bf 1}_{\{\phi=0\}}\omega^{n} since ϕ\phi is a model potential (see Theorem 3.6), hence X(ψ)ωϕnX(ψ)ωn\int_{X}(-\psi)\omega_{\phi}^{n}\leq\int_{X}(-\psi)\omega^{n}. Taking the supremum over all such uu yields

Capω,ϕ(K)AMK1CCapω(K)1/n,{\rm Cap}_{\omega,\phi}(K)\leq AM_{K}^{-1}\leq C{\rm Cap}_{\omega}(K)^{1/n},

where the last inequality follows from [GZ05, Theorem 7.1]. ∎

Lemma 4.5.

Fix φ,ψPSH(X,θ)\varphi,\psi\in{\rm PSH(X,\theta)} such that ψφ\psi\leq\varphi and Xθφn=Xθψn\int_{X}\theta_{\varphi}^{n}=\int_{X}\theta_{\psi}^{n}. Then there exists a continuous function g:[0,)[0,)g:[0,\infty)\rightarrow[0,\infty) with g(0)=0g(0)=0 such that, for all Borel sets EE,

Capθ,ψ(E)g(Capθ,φ(E)).{\rm Cap}_{\theta,\psi}(E)\leq g\left({\rm Cap}_{\theta,\varphi}(E)\right).

Our proof uses an idea from [GLZ19].

Proof.

We can assume that φ0\varphi\leq 0. We claim that if vPSH(X,θ)v\in{\rm PSH}(X,\theta) with φtvφ\varphi-t\leq v\leq\varphi (for t0t\geq 0) then for any Borel set EE we have

Eθvnmax(t,1)nCapθ,φ(E).\int_{E}\theta_{v}^{n}\leq\max(t,1)^{n}{\rm Cap}_{\theta,\varphi}(E).

If t[0,1]t\in[0,1] then vv is a candidate defining the capacity Capθ,φ{\rm Cap}_{\theta,\varphi}, hence the desired inequality holds. For t>1t>1, the function vt:=t1v+(1t1)φv_{t}:=t^{-1}v+(1-t^{-1})\varphi is θ\theta-psh and φ1vtφ\varphi-1\leq v_{t}\leq\varphi. Since non-pluripolar products are multilinear, we thus have

tnEθvnEθvtnCapθ,φ(E),t^{-n}\int_{E}\theta_{v}^{n}\leq\int_{E}\theta_{v_{t}}^{n}\leq{\rm Cap}_{\theta,\varphi}(E),

yielding the claim.

Let uu be a θ\theta-psh function such that ψ1uψ\psi-1\leq u\leq\psi. Fix t>1t>1 and set ut:=max(u,φ2t)u_{t}:=\max(u,\varphi-2t), Et:=E{u>φ2t}E_{t}:=E\cap\{u>\varphi-2t\}, Ft:=E{uφ2t}F_{t}:=E\cap\{u\leq\varphi-2t\}. Observe that φ2tutφ\varphi-2t\leq u_{t}\leq\varphi. By plurifine locality and the claim we have that

Etθun=Etθutn(2t)nCapθ,φ(Et)(2t)nCapθ,φ(E).\int_{E_{t}}\theta_{u}^{n}=\int_{E_{t}}\theta_{u_{t}}^{n}\leq(2t)^{n}{\rm Cap}_{\theta,\varphi}(E_{t})\leq(2t)^{n}{\rm Cap}_{\theta,\varphi}(E).

On the other hand, using the inclusions

Ft{ψ1u+φ2t}{ψ1φt}{ψ1t}F_{t}\subset\left\{\psi-1\leq\frac{u+\varphi}{2}-t\right\}\subset\{\psi-1\leq\varphi-t\}\subset\{\psi-1\leq-t\}

and the comparison principle, Corollary 3.23, we infer

Ftθun\displaystyle\int_{F_{t}}\theta_{u}^{n} {ψ1u+φ2t}θun2n{ψ1u+φ2t}θu+φ2n\displaystyle\leq\int_{\{\psi-1\leq\frac{u+\varphi}{2}-t\}}\theta_{u}^{n}\leq 2^{n}\int_{\{\psi-1\leq\frac{u+\varphi}{2}-t\}}\theta_{\frac{u+\varphi}{2}}^{n}
2n{ψφt+1}θψn2n{ψt+1}θψn.\displaystyle\leq 2^{n}\int_{\{\psi\leq\varphi-t+1\}}\theta_{\psi}^{n}\leq 2^{n}\int_{\{\psi\leq-t+1\}}\theta_{\psi}^{n}.

Taking the supremum over all candidates uu we obtain

Capθ,ψ(E)(2t)nCapθ,φ(E)+2n{ψt+1}θψn.{\rm Cap}_{\theta,\psi}(E)\leq(2t)^{n}{\rm Cap}_{\theta,\varphi}(E)+2^{n}\int_{\{\psi\leq-t+1\}}\theta_{\psi}^{n}.

Set t:=(Capθ,φ(E))1/2nt:=({\rm Cap}_{\theta,\varphi}(E))^{-1/2n}. If t>1t>1 we get Capθ,ψ(E)g(Capθ,φ(E)){\rm Cap}_{\theta,\psi}(E)\leq g\left({\rm Cap}_{\theta,\varphi}(E)\right), where gg is defined on [0,)[0,\infty) by

g(s):=(2n+Vol({θ}))s1/2+2n{ψs1/(2n)+1}θψn.g(s):=(2^{n}+{\rm Vol}(\{\theta\}))s^{1/2}+2^{n}\int_{\{\psi\leq-s^{-1/(2n)}+1\}}\theta_{\psi}^{n}.

Observe that g(0)=0g(0)=0 since θψn\theta_{\psi}^{n} does not charge the pluripolar set {ψ=}\{\psi=-\infty\}.

If t1t\leq 1 then s:=Capθ,φ(E)1s:={\rm Cap}_{\theta,\varphi}(E)\geq 1, and by the choice of gg above we have

Capθ,ψ(E)Vol({θ})g(Capθ,φ(E)),{\rm Cap}_{\theta,\psi}(E)\leq{\rm Vol}(\{\theta\})\leq g\left({\rm Cap}_{\theta,\varphi}(E)\right),

finishing the proof. ∎

Theorem 4.6.

Assume that ψPSH(X,θ)\psi\in{\rm PSH}(X,\theta). Then there exists a continuous function f:[0,)[0,)f:[0,\infty)\rightarrow[0,\infty) with f(0)=0f(0)=0 such that, for any Borel set EE,

Capθ,ψ(E)f(Capω(E)).{\rm Cap}_{\theta,\psi}(E)\leq f\left({\rm Cap}_{\omega}(E)\right).
Proof.

We can assume that θω\theta\leq\omega. Since PSH(X,θ)PSH(X,ω){\rm PSH}(X,\theta)\subseteq{\rm PSH}(X,\omega), for any Borel set EE we have Capθ,ψ(E)Capω,ψ(E){\rm Cap}_{\theta,\psi}(E)\leq{\rm Cap}_{\omega,\psi}(E). Also, by Proposition 4.4 and Lemma 4.5 we get

Capθ,ψ(E)Capω,ψ(E)g(Capω,Pω[ψ](E))g(CCapω(E)1/n).{\rm Cap}_{\theta,\psi}(E)\leq{\rm Cap}_{\omega,\psi}(E)\leq g\left({\rm Cap}_{\omega,P_{\omega}[\psi]}(E)\right)\leq g\left(C{\rm Cap}_{\omega}(E)^{1/n}\right).

4.2 Integration by parts

The integration by parts formula was recently obtained [Xia19] using Witt Nyström’s construction. We give an argument here following [Lu21] (see [Vu21] for a more general result). We first start with the following key lemma.

Lemma 4.7.

Let φ1,φ2,ψ1,ψ2PSH(X,θ)\varphi_{1},\varphi_{2},\psi_{1},\psi_{2}\in{\rm PSH}(X,\theta) be such that [φ1]=[φ2][\varphi_{1}]=[\varphi_{2}] and [ψ1]=[ψ2][\psi_{1}]=[\psi_{2}]. Then

X(φ1φ2)(θψ1nθψ2n)=X(ψ1ψ2)(S1S2),\int_{X}(\varphi_{1}-\varphi_{2})\left(\theta_{\psi_{1}}^{n}-\theta_{\psi_{2}}^{n}\right)=\int_{X}(\psi_{1}-\psi_{2})(S_{1}-S_{2}),

where Sj:=k=0n1θφjθψ1kθψ2nk1S_{j}:=\sum_{k=0}^{n-1}\theta_{\varphi_{j}}\wedge\theta_{\psi_{1}}^{k}\wedge\theta_{\psi_{2}}^{n-k-1}, j=1,2j=1,2.

Proof.

It follows from Proposition 3.2 that

X(θψ1nθψ2n)=X(S1S2)=0.\int_{X}(\theta_{\psi_{1}}^{n}-\theta_{\psi_{2}}^{n})=\int_{X}(S_{1}-S_{2})=0.

By adding a constant we can assume that φ1,φ2,ψ1,ψ2\varphi_{1},\varphi_{2},\psi_{1},\psi_{2} are negative and that φ1φ2\varphi_{1}\leq\varphi_{2} and ψ1ψ2\psi_{1}\leq\psi_{2}. Let B>0B>0 be a constant such that

φ2φ1+B,ψ2ψ1+B.\varphi_{2}\leq\varphi_{1}+B,\quad\psi_{2}\leq\psi_{1}+B.

Step 1. We assume θ\theta is Kähler and ψ1,ψ2,φ1,φ2\psi_{1},\psi_{2},\varphi_{1},\varphi_{2} are λθ\lambda\theta-psh for some λ(0,1)\lambda\in(0,1). Observe that the last condition ensures that their Monge–Ampère mass is strictly positive. Step 1.1. We also assume that there exists C>0C>0 such that ψ1=ψ2\psi_{1}=\psi_{2} on U:={min(ψ1,ψ2)<C}U:=\{\min(\psi_{1},\psi_{2})<-C\} and φ1=φ2\varphi_{1}=\varphi_{2} on V:={min(φ1,φ2)<C}V:=\{\min(\varphi_{1},\varphi_{2})<-C\}.

For a function uPSH(X,θ)u\in{\rm PSH}(X,\theta) we consider its canonical approximant ut:=max(u,t)u^{t}:=\max(u,-t), t>0t>0. It follows from Stokes’ theorem that

X(φ1tφ2t)(θψ1tnθψ2tn)=X(ψ1tψ2t)(S1tS2t),\int_{X}(\varphi_{1}^{t}-\varphi_{2}^{t})\left(\theta_{\psi_{1}^{t}}^{n}-\theta_{\psi_{2}^{t}}^{n}\right)=\int_{X}(\psi_{1}^{t}-\psi_{2}^{t})(S_{1}^{t}-S_{2}^{t}),

where Sjt:=k=0n1θφjtθψ1tkθψ2tnk1S_{j}^{t}:=\sum_{k=0}^{n-1}\theta_{\varphi_{j}^{t}}\wedge\theta_{\psi_{1}^{t}}^{k}\wedge\theta_{\psi_{2}^{t}}^{n-k-1}, j=1,2j=1,2. We now consider the limit as tt\to\infty in the above equality.

Fix t>Ct>C. By assumption we have φ1t=φ2t\varphi_{1}^{t}=\varphi_{2}^{t} on VV and {φ1t}={φ2t}V\{\varphi_{1}\leq-t\}=\{\varphi_{2}\leq-t\}\subset V. The same properties for ψ1,ψ2\psi_{1},\psi_{2} also hold: ψ1t=ψ2t\psi_{1}^{t}=\psi_{2}^{t} in the open set UU and {ψ1t}={ψ2t}U\{\psi_{1}\leq-t\}=\{\psi_{2}\leq-t\}\subset U. It thus follows that

𝟏Uθψ1tn=𝟏Uθψ2tnand𝟏VS1t=𝟏VS2t,{\bf 1}_{U}\theta_{\psi_{1}^{t}}^{n}={\bf 1}_{U}\theta_{\psi_{2}^{t}}^{n}\quad\text{and}\quad{\bf 1}_{V}S_{1}^{t}={\bf 1}_{V}S_{2}^{t},

hence multiplying with the characteristic functions 𝟏{ψ1t}{\bf 1}_{\{\psi_{1}\leq-t\}} and 𝟏{φ1t}{\bf 1}_{\{\varphi_{1}\leq-t\}} respectively gives

𝟏{ψ1t}θψ1tn=𝟏{ψ1t}θψ2tnand𝟏{φ1t}S1t=𝟏{φ1t}S2t.{\bf 1}_{\{\psi_{1}\leq-t\}}\theta_{\psi_{1}^{t}}^{n}={\bf 1}_{\{\psi_{1}\leq-t\}}\theta_{\psi_{2}^{t}}^{n}\quad\text{and}\quad{\bf 1}_{\{\varphi_{1}\leq-t\}}S_{1}^{t}={\bf 1}_{\{\varphi_{1}\leq-t\}}S_{2}^{t}.

By plurifine locality of the non-pluripolar product we thus have

X(φ1tφ2t)(θψ1tnθψ2tn)\displaystyle\int_{X}(\varphi_{1}^{t}-\varphi_{2}^{t})\left(\theta_{\psi_{1}^{t}}^{n}-\theta_{\psi_{2}^{t}}^{n}\right) ={ψ1>t}{φ1>t}(φ1tφ2t)(θψ1tnθψ2tn)\displaystyle=\int_{\{\psi_{1}>-t\}\cap\{\varphi_{1}>-t\}}(\varphi_{1}^{t}-\varphi_{2}^{t})(\theta_{\psi_{1}^{t}}^{n}-\theta_{\psi_{2}^{t}}^{n})
={ψ1>t}{φ1>t}(φ1φ2)(θψ1nθψ2n),\displaystyle=\int_{\{\psi_{1}>-t\}\cap\{\varphi_{1}>-t\}}(\varphi_{1}-\varphi_{2})(\theta_{\psi_{1}}^{n}-\theta_{\psi_{2}}^{n}),

and

X(ψ1tψ2t)(S1tS2t)={ψ1>t}{φ1>t}(ψ1ψ2)(S1S2).\int_{X}(\psi_{1}^{t}-\psi_{2}^{t})(S_{1}^{t}-S_{2}^{t})=\int_{\{\psi_{1}>-t\}\cap\{\varphi_{1}>-t\}}(\psi_{1}-\psi_{2})(S_{1}-S_{2}).

Since φ1φ2\varphi_{1}-\varphi_{2} and ψ1ψ2\psi_{1}-\psi_{2} are bounded, using the dominated convergence theorem we finish Step 1.1.

Step 1.2. We remove the assumption made in Step 1.1.
Recall that ψ1,ψ2,φ1,φ2\psi_{1},\psi_{2},\varphi_{1},\varphi_{2} are λθ\lambda\theta-psh for some λ(0,1)\lambda\in(0,1). For each ε(0,1λ)\varepsilon\in(0,1-\lambda) we define

ψ2,ε:=max(ψ1,(1+ε)ψ2);φ2,ε:=max(φ1,(1+ε)φ2).\psi_{2,\varepsilon}:=\max(\psi_{1},(1+\varepsilon)\psi_{2})\ ;\ \varphi_{2,\varepsilon}:=\max(\varphi_{1},(1+\varepsilon)\varphi_{2}).

Since θ\theta is assumed to be Kähler, and ε(0,1λ)\varepsilon\in(0,1-\lambda), the functions (1+ε)ψ2(1+\varepsilon)\psi_{2} and (1+ε)φ2(1+\varepsilon)\varphi_{2} are still θ\theta-psh. Also, ψ1ψ2,ε(1+ε)(ψ1+B)\psi_{1}\leq\psi_{2,\varepsilon}\leq(1+\varepsilon)(\psi_{1}+B) and φ1φ2,ε(1+ε)(φ1+B)\varphi_{1}\leq\varphi_{2,\varepsilon}\leq(1+\varepsilon)(\varphi_{1}+B). These are θ\theta-psh functions satisfying the assumptions in Step 1.1 with C=B+Bε1C=B+B\varepsilon^{-1}. Indeed, if φ1(x)<C\varphi_{1}(x)<-C then

(1+ε)φ2(x)=φ2(x)+εφ2(x)φ1(x)+B+ε(BC)=φ1(x).(1+\varepsilon)\varphi_{2}(x)=\varphi_{2}(x)+\varepsilon\varphi_{2}(x)\leq\varphi_{1}(x)+B+\varepsilon(B-C)=\varphi_{1}(x).

It then follows that φ2,ε=φ1\varphi_{2,\varepsilon}=\varphi_{1} on V={φ1=min(φ1,φ2,ε)<C}V=\{\varphi_{1}=\min(\varphi_{1},\varphi_{2,\varepsilon})<-C\}. Similarly we have that ψ2,ε=ψ1\psi_{2,\varepsilon}=\psi_{1} on U={ψ1=min(ψ1,ψ2,ε)<C}U=\{\psi_{1}=\min(\psi_{1},\psi_{2,\varepsilon})<-C\}, with the same CC.
We can thus apply Step 1.1 to ψ1\psi_{1}, ψ2,ε\psi_{2,\varepsilon}, φ1\varphi_{1},φ2,ε\varphi_{2,\varepsilon} to obtain

X(φ1φ2,ε)(θψ1nθψ2,εn)=X(ψ1ψ2,ε)(S1,εS2,ε),\int_{X}(\varphi_{1}-\varphi_{2,\varepsilon})\left(\theta_{\psi_{1}}^{n}-\theta_{\psi_{2,\varepsilon}}^{n}\right)=\int_{X}\left(\psi_{1}-\psi_{2,\varepsilon}\right)(S_{1,\varepsilon}-S_{2,\varepsilon}),

where S1,ε:=k=0n1θφ1θψ1kθψ2,εnk1S_{1,\varepsilon}:=\sum_{k=0}^{n-1}\theta_{\varphi_{1}}\wedge\theta_{\psi_{1}}^{k}\wedge\theta_{\psi_{2,\varepsilon}}^{n-k-1} and S2,ε:=k=0n1θφ2,εθψ1kθψ2,εnk1S_{2,\varepsilon}:=\sum_{k=0}^{n-1}\theta_{\varphi_{2,\varepsilon}}\wedge\theta_{\psi_{1}}^{k}\wedge\theta_{\psi_{2,\varepsilon}}^{n-k-1}.
By Theorem 4.6 there exists a continuous function f:[0,)[0,)f:[0,\infty)\rightarrow[0,\infty) with f(0)=0f(0)=0 such that for every Borel set EE,

Capθ,ψ(E)f(Capθ(E)),{\rm Cap}_{\theta,\psi}(E)\leq f({\rm Cap}_{\theta}(E)),

where

ψ:=φ1+φ2+ψ1+ψ25B.\psi:=\frac{\varphi_{1}+\varphi_{2}+\psi_{1}+\psi_{2}}{5}-B.

Note that ψ\psi is θ\theta-psh and Xθψn>0\int_{X}\theta_{\psi}^{n}>0. Indeed, recalling that in this step θ\theta is Kähler, we have

Xθψn=5nX(θ+θφ1+θφ2+θψ1+θψ2)n5nθn>0.\int_{X}\theta_{\psi}^{n}=5^{-n}\int_{X}(\theta+\theta_{\varphi_{1}}+\theta_{\varphi_{2}}+\theta_{\psi_{1}}+\theta_{\psi_{2}})^{n}\geq 5^{-n}\theta^{n}>0.

Since we have assumed that ψ2Bψ1ψ20\psi_{2}-B\leq\psi_{1}\leq\psi_{2}\leq 0 and φ2Bφ1φ20\varphi_{2}-B\leq\varphi_{1}\leq\varphi_{2}\leq 0, we get

φ2Bφ1φ2,ε=max(φ1,(1+ε)φ2)max(φ1,φ2)=φ2,\varphi_{2}-B\leq\varphi_{1}\leq\varphi_{2,\varepsilon}=\max(\varphi_{1},(1+\varepsilon)\varphi_{2})\leq\max(\varphi_{1},\varphi_{2})=\varphi_{2},

and

ψ2Bψ1ψ2,ε=max(ψ1,(1+εψ2))max(ψ1,ψ2)=ψ2.\psi_{2}-B\leq\psi_{1}\leq\psi_{2,\varepsilon}=\max(\psi_{1},(1+\varepsilon\psi_{2}))\leq\max(\psi_{1},\psi_{2})=\psi_{2}.

In particular ψ1ψ2ψ2,ε\psi_{1}\simeq\psi_{2}\simeq\psi_{2,\varepsilon} and φ1φ2φ2,ε\varphi_{1}\simeq\varphi_{2}\simeq\varphi_{2,\varepsilon}.
Set uε:=15(φ1+φ2,ε+ψ2,ε+ψ1)u_{\varepsilon}:=\frac{1}{5}(\varphi_{1}+\varphi_{2,\varepsilon}+\psi_{2,\varepsilon}+\psi_{1}). Using the above inequalities we get

ψ\displaystyle\psi φ1+ψ1+φ2+ψ252B5\displaystyle\leq\frac{\varphi_{1}+\psi_{1}+\varphi_{2}+\psi_{2}}{5}-\frac{2B}{5}
φ1+ψ1+φ2,ε+ψ2,ε5=uε\displaystyle\leq\frac{\varphi_{1}+\psi_{1}+\varphi_{2,\varepsilon}+\psi_{2,\varepsilon}}{5}=u_{\varepsilon}
φ1+ψ1+φ2+ψ25=ψ+B.\displaystyle\leq\frac{\varphi_{1}+\psi_{1}+\varphi_{2}+\psi_{2}}{5}=\psi+B.

Also observe that there exists a positive constant CC^{\prime} such that

Sj,εC(5θ+ddc(φ1+φ2,ε+ψ2,ε+ψ1))n=C5nθuεnS_{j,\varepsilon}\leq C^{\prime}(5\theta+dd^{c}(\varphi_{1}+\varphi_{2,\varepsilon}+\psi_{2,\varepsilon}+\psi_{1}))^{n}=C^{\prime}5^{n}\theta_{u_{\varepsilon}}^{n}

We then obtain that for any Borel set EE and any ε(0,1λ),j=1,2\varepsilon\in(0,1-\lambda),\;j=1,2,

ESj,εC5nBnCapθ,ψ(E)C′′f(Capθ(E)),\int_{E}S_{j,\varepsilon}\leq C^{\prime}5^{n}B^{n}{\rm Cap}_{\theta,\psi}(E)\leq C^{\prime\prime}f({\rm Cap}_{\theta}(E)),

where the first inequality follows from the arguments at the beginning of the proof of Lemma 4.5, and the second follows from Theorem 4.6.

Since ψ2,ε\psi_{2,\varepsilon} and φ2,ε\varphi_{2,\varepsilon} are increasing to ψ2\psi_{2} and φ2\varphi_{2} respectively, for each j{1,2}j\in\{1,2\} we also have that Sj,εSjS_{j,\varepsilon}\to S_{j} and θψ2,εnθψ2n\theta_{\psi_{2,\varepsilon}}^{n}\to\theta_{\psi_{2}}^{n} as ε0\varepsilon\to 0 in the weak sense of Radon measures (see Theorem 2.6 and Remark 3.4). By the above, these measures are uniformly dominated by Capθ{\rm Cap}_{\theta}.
Note also that φ1φ2,ε,φ1φ2,ψ2,εψ1,ψ2ψ1\varphi_{1}-\varphi_{2,\varepsilon},\varphi_{1}-\varphi_{2},\psi_{2,\varepsilon}-\psi_{1},\psi_{2}-\psi_{1} are uniformly bounded and quasi-continuous (because difference of quasi-psh functions). Moreover, ψ2,εψ1ψ2ψ1\psi_{2,\varepsilon}-\psi_{1}\to\psi_{2}-\psi_{1}, and φ1φ2,εφ1φ2\varphi_{1}-\varphi_{2,\varepsilon}\to\varphi_{1}-\varphi_{2} in capacity as ε0\varepsilon\to 0. It thus follows from Lemma 2.5 that

limε0X(φ1φ2,ε)(θψ1nθψ2,εn)=X(φ1φ2)(θψ1nθψ2n)\lim_{\varepsilon\to 0}\int_{X}(\varphi_{1}-\varphi_{2,\varepsilon})\left(\theta_{\psi_{1}}^{n}-\theta_{\psi_{2,\varepsilon}}^{n}\right)=\int_{X}(\varphi_{1}-\varphi_{2})\left(\theta_{\psi_{1}}^{n}-\theta_{\psi_{2}}^{n}\right)

and

limε0X(ψ1ψ2,ε)(S1,εS2,ε)=X(ψ1ψ2)(S1S2),\lim_{\varepsilon\to 0}\int_{X}\left(\psi_{1}-\psi_{2,\varepsilon}\right)(S_{1,\varepsilon}-S_{2,\varepsilon})=\int_{X}\left(\psi_{1}-\psi_{2}\right)(S_{1}-S_{2}),

finishing the proof of Step 1.2.

Step 2. We merely assume that {θ}\{\theta\} is big. We can assume that ωθω-\omega\leq\theta\leq\omega. For s>2s>2 we consider θs:=θ+sω\theta_{s}:=\theta+s\omega, which is Kähler, and we observe that φ1,φ2,ψ1,ψ2\varphi_{1},\varphi_{2},\psi_{1},\psi_{2} are λθs\lambda\theta_{s}-psh for some λ(0,1)\lambda\in(0,1). We can thus apply the first step to get

Xu((θs+ddcψ1)n(θs+ddcψ2)n)=XvTs,\int_{X}u\left((\theta_{s}+dd^{c}\psi_{1})^{n}-(\theta_{s}+dd^{c}\psi_{2})^{n}\right)=\int_{X}vT_{s},

where u=φ1φ2u=\varphi_{1}-\varphi_{2}, v=ψ1ψ2v=\psi_{1}-\psi_{2} and

Ts=\displaystyle T_{s}= k=0n1(θs+ddcφ1)(θs+ddcψ1)k(θs+ddcψ2)nk1\displaystyle\sum_{k=0}^{n-1}(\theta_{s}+dd^{c}\varphi_{1})\wedge(\theta_{s}+dd^{c}\psi_{1})^{k}\wedge(\theta_{s}+dd^{c}\psi_{2})^{n-k-1}
\displaystyle- k=0n1(θs+ddcφ2)(θs+ddcψ1)k(θs+ddcψ2)nk1.\displaystyle\sum_{k=0}^{n-1}(\theta_{s}+dd^{c}\varphi_{2})\wedge(\theta_{s}+dd^{c}\psi_{1})^{k}\wedge(\theta_{s}+dd^{c}\psi_{2})^{n-k-1}.

We thus obtain an equality between two polynomials in ss. Identifying the coefficients we arrive at the conclusion. ∎

Next we prove the integration by parts formula, extending the one in [BEGZ10] which only applies to the case of potentials with small unbounded locus.

Theorem 4.8.

Let u,vL(X)u,v\in L^{\infty}(X) be differences of quasi-psh functions, and ϕjPSH(X,θj)\phi_{j}\in\textup{PSH}(X,\theta^{j}), j{1,,n1}j\in\{1,\ldots,n-1\} with {θj}\{\theta^{j}\} big. Then

Xu𝑑dcvθϕ11θϕn1n1=Xv𝑑dcuθϕ11θϕn1n1.\int_{X}udd^{c}v\wedge\theta^{1}_{\phi_{1}}\wedge\ldots\wedge\theta^{n-1}_{\phi_{n-1}}=\int_{X}vdd^{c}u\wedge\theta^{1}_{\phi_{1}}\wedge\ldots\wedge\theta^{n-1}_{\phi_{n-1}}.
Proof.

We first assume that θ\theta is Kähler, u=φ1φ2u=\varphi_{1}-\varphi_{2} and v=ψ1ψ2v=\psi_{1}-\psi_{2} where ψ1,ψ2,φ1,φ2\psi_{1},\psi_{2},\varphi_{1},\varphi_{2} are θ\theta-psh. Fix ϕPSH(X,θ)\phi\in{\rm PSH}(X,\theta) and for each s[0,1]s\in[0,1], j=1,2j=1,2, we set ψj,s:=sψj+(1s)ϕ\psi_{j,s}:=s\psi_{j}+(1-s)\phi. Note that ψ1,sψ2,s\psi_{1,s}\simeq\psi_{2,s}. It follows from Lemma 4.7 that for any s[0,1]s\in[0,1],

Xu(θψ1,snθψ2,sn)=X(ψ1,sψ2,s)Ts=XsvTs,\int_{X}u\left(\theta_{\psi_{1,s}}^{n}-\theta_{\psi_{2,s}}^{n}\right)=\int_{X}(\psi_{1,s}-\psi_{2,s})T_{s}=\int_{X}svT_{s},

where

Ts:=k=0n1θφ1θψ1,skθψ2,snk1k=0n1θφ2θψ1,skθψ2,snk1.T_{s}:=\sum_{k=0}^{n-1}\theta_{\varphi_{1}}\wedge\theta_{\psi_{1,s}}^{k}\wedge\theta_{\psi_{2,s}}^{n-k-1}-\sum_{k=0}^{n-1}\theta_{\varphi_{2}}\wedge\theta_{\psi_{1,s}}^{k}\wedge\theta_{\psi_{2,s}}^{n-k-1}.

We thus have an identity between two polynomials in ss. We then compute the first derivative in s=0s=0 and we find that for j=1,2j=1,2

s(sθψj+(1s)θϕ)n|s=0=nθψjθϕn1nθϕn\frac{\partial}{\partial s}(s\theta_{\psi_{j}}+(1-s)\theta_{\phi})^{n}|_{s=0}=n\theta_{\psi_{j}}\wedge\theta_{\phi}^{n-1}-n\theta_{\phi}^{n}

and

s(sTs)|s=0=T0.\frac{\partial}{\partial s}(sT_{s})|_{s=0}=T_{0}.

Noticing that T0=nddcuθϕn1T_{0}=ndd^{c}u\wedge\theta_{\phi}^{n-1}, we obtain

Xu𝑑dcvθϕn1=Xv𝑑dcuθϕn1.\int_{X}udd^{c}v\wedge\theta_{\phi}^{n-1}=\int_{X}vdd^{c}u\wedge\theta_{\phi}^{n-1}.

For the general case, i.e. {θ}\{\theta\} is merely big, we can write u=φ1φ2u=\varphi_{1}-\varphi_{2} and v=ψ1ψ2v=\psi_{1}-\psi_{2}, where ψ1,ψ2,φ1,φ2\psi_{1},\psi_{2},\varphi_{1},\varphi_{2} are AωA\omega-psh, for some A>0A>0 large enough. We apply the first step with θ\theta replaced by θ+tω\theta+t\omega, for t>At>A to get

Xu𝑑dcv(tω+θϕ)n1=Xv𝑑dcu(tω+θϕ)n1.\int_{X}udd^{c}v\wedge(t\omega+\theta_{\phi})^{n-1}=\int_{X}vdd^{c}u\wedge(t\omega+\theta_{\phi})^{n-1}.

Identifying the coefficients of these two polynomials in tt we obtain

Xu𝑑dcvθϕn1=Xv𝑑dcuθϕn1.\int_{X}udd^{c}v\wedge\theta_{\phi}^{n-1}=\int_{X}vdd^{c}u\wedge\theta_{\phi}^{n-1}.

We now consider θ=s1θ1+.+sn1θn1\theta=s_{1}\theta_{1}+....+s_{n-1}\theta_{n-1}, ϕ:=s1ϕ1++sn1ϕn1\phi:=s_{1}\phi_{1}+...+s_{n-1}\phi_{n-1} with s1,,sn1[0,1]s_{1},...,s_{n-1}\in[0,1] and j=1n1sj=1\sum_{j=1}^{n-1}s_{j}=1. We obtain an identity between two polynomials in (s1,,sn1)(s_{1},...,s_{n-1}), and identifying the coefficients we arrive at the result. ∎

Chapter 5 Complex Monge–Ampère equations with prescribed singularity type

Let θ\theta be a smooth closed real (1,1)(1,1)-form on XX such that {θ}\{\theta\} is big and ϕPSH(X,θ)\phi\in\textup{PSH}(X,\theta). By PSH(X,θ,ϕ)\mathrm{PSH}(X,\theta,\phi) we denote the set of θ\theta-psh functions that are more singular than ϕ\phi. We say that vPSH(X,θ,ϕ)v\in\textup{PSH}(X,\theta,\phi) has relatively minimal singularity type if vv has the same singularity type as ϕ\phi. Clearly, (X,θ,ϕ)PSH(X,θ,ϕ)\mathcal{E}(X,\theta,\phi)\subset\textup{PSH}(X,\theta,\phi).

Let μ\mu be a non-pluripolar positive measure on XX such that μ(X)=Xθϕn>0\mu(X)=\int_{X}\theta_{\phi}^{n}>0. Our aim is to study existence and uniqueness of solutions to the following equation of complex Monge–Ampère type:

θun=μ,u(X,θ,ϕ).\theta_{u}^{n}=\mu,\ \ \ u\in\mathcal{E}(X,\theta,\phi). (5.1)

It is not hard to see that this equation does not have a solution for arbitrary ϕ\phi. Indeed, suppose for the moment that θ=ω\theta=\omega, and choose ϕ(X,ω):=(X,ω,0)\phi\in\mathcal{E}(X,\omega):=\mathcal{E}(X,\omega,0) unbounded. It is clear that (X,ω,ϕ)(X,ω,0)\mathcal{E}(X,\omega,\phi)\subsetneq\mathcal{E}(X,\omega,0). By [BEGZ10, Theorem A], the (trivial) equation ωun=ωn,u(X,ω,0)\omega_{u}^{n}=\omega^{n},\ u\in\mathcal{E}(X,\omega,0) is only solved by potentials uu that are constant over XX, hence we cannot have u(X,ω,ϕ)u\notin\mathcal{E}(X,\omega,\phi).

This simple example suggests that we need to be more selective in our choice of ϕ\phi, to make (5.1) well posed. As it turns out, the natural choice is to take ϕ\phi such that Pθ[ϕ]=ϕP_{\theta}[\phi]=\phi (see Theorem 5.22 for concrete evidence). Therefore, for the rest of this section we ask that ϕ\phi additionally satisfies:

ϕ=Pθ[ϕ].\phi=P_{\theta}[\phi]. (5.2)

We recall that such a potential ϕ\phi is model, and [ϕ][\phi] is a model type singularity. As Vθ=Pθ[Vθ]V_{\theta}=P_{\theta}[V_{\theta}], one can think of such ϕ\phi as generalizations of VθV_{\theta}, the potential with minimal singularity from [BEGZ10].

One wonders if maybe model type potentials (those that satisfy (5.2)) always have small unbounded locus. Sadly, this is not the case, as we now point out. Suppose θ\theta is a Kähler form, and {xj}jX\{x_{j}\}_{j}\subset X is a dense countable subset. Also let vjPSH(X,θ)v_{j}\in\textup{PSH}(X,\theta) such that vj<0v_{j}<0, X(vj)θn=1\int_{X}(-v_{j})\theta^{n}=1, and vjv_{j} has a positive Lelong number at xjx_{j}. Then u=j12jvjPSH(X,θ)u=\sum_{j}\frac{1}{2^{j}}v_{j}\in\textup{PSH}(X,\theta) has positive Lelong numbers at all xjx_{j}. As we have argued in Lemma 5.1 below, the Lelong numbers of Pθ[u]P_{\theta}[u] are the same as those of uu, hence the model type potential Pθ[u]P_{\theta}[u] cannot have small unbounded locus.

Lemma 5.1.

Suppose that uPSH(X,θ)u\in\textup{PSH}(X,\theta). Then (tu)=(tPθ[u])\mathcal{I}(tu)=\mathcal{I}(tP_{\theta}[u]) for all t>0t>0. In particular, all the Lelong numbers of uu and Pθ[u]P_{\theta}[u] are the same.

Recall that (v)\mathcal{I}(v) is the multiplier ideal sheaf of a quasi-psh function vv, whose germs are defined by

(v,x):={f𝒪x:U|f|2evωn<+forsomeopensubsetUX,xU}.\mathcal{I}(v,x):=\left\{f\in{\mathcal{O}}_{x}\ :\ \int_{U}|f|^{2}e^{-v}\omega^{n}<+\infty\quad{\rm for\ some\ open\ subset\ }U\subset X,x\in U\right\}.

We use the valuative criteria of integrability of Boucksom–Favre–Jonsson. For a more elementary argument see the proof of [DDL18b, Theorem 1.1].

Proof.

Since Pθ(u+C,Vθ)Pθ[u]P_{\theta}(u+C,V_{\theta})\nearrow P_{\theta}[u] a.e., as CC\nearrow\infty, from the Guan–Zhou openess theorem [GZ15] it follows that (tPθ(u+C,Vθ))=(tPθ[u])\mathcal{I}(tP_{\theta}(u+C,V_{\theta}))=\mathcal{I}(tP_{\theta}[u]) for C>0C>0 big enough. However [Pθ(u+C,Vθ)]=[u][P_{\theta}(u+C,V_{\theta})]=[u] for all C>0C>0, so we get that (tu)=(tPθ[u])\mathcal{I}(tu)=\mathcal{I}(tP_{\theta}[u]). The conclusion about Lelong numbers follows from the equivalence between (i) (applied with the proper modification π=id\pi=id) and (ii) in [BFJ08, Theorem A]. ∎

5.1 The relative finite energy class

To develop the variational approach to (5.1), we study the relative version of the Monge–Ampère energy, and its bounded locus 1(X,θ,ϕ)\mathcal{E}^{1}(X,\theta,\phi). For u(X,θ,ϕ)u\in\mathcal{E}(X,\theta,\phi) with relatively minimal singularity type, we define the Monge–Ampère energy of uu relative to ϕ\phi as

Iϕ(u):=1n+1k=0nX(uϕ)θukθϕnk.\mathrm{I}_{\phi}(u):=\frac{1}{n+1}\sum_{k=0}^{n}\int_{X}(u-\phi)\theta_{u}^{k}\wedge\theta_{\phi}^{n-k}.

Before we study this energy closely, we note the following inequality:

Lemma 5.2.

Assume u,v(X,θ,ϕ)u,v\in\mathcal{E}(X,\theta,\phi) have the same singularity type. Then

X(uv)θukθvnkX(uv)θuk1θvnk+1.\int_{X}(u-v)\theta_{u}^{k}\wedge\theta_{v}^{n-k}\leq\int_{X}(u-v)\theta_{u}^{k-1}\wedge\theta_{v}^{n-k+1}.
Proof.

Adding a constant to uu does not affect the inequality (by Theorem 3.3), so we can assume that vuv\leq u. By the partial comparison principle (Proposition 3.22), we have

X(uv)θukθvnk\displaystyle\int_{X}(u-v)\theta_{u}^{k}\wedge\theta_{v}^{n-k} =0θukθvnk(u>v+t)dt\displaystyle=\int_{0}^{\infty}\theta_{u}^{k}\wedge\theta_{v}^{n-k}(u>v+t)dt
0θuk1θvnk+1(u>v+t)dt\displaystyle\leq\int_{0}^{\infty}\theta_{u}^{k-1}\wedge\theta_{v}^{n-k+1}(u>v+t)dt
=X(uv)θuk1θvnk+1.\displaystyle=\int_{X}(u-v)\theta_{u}^{k-1}\wedge\theta_{v}^{n-k+1}.

In the next theorem we collect basic properties of the Monge–Ampère energy:

Theorem 5.3.

Suppose u,v(X,θ,ϕ)u,v\in\mathcal{E}(X,\theta,\phi) have relatively minimal singularity type. Then:
(i) Iϕ(u)Iϕ(v)=1n+1k=0nX(uv)θukθvnk.\mathrm{I}_{\phi}(u)-\mathrm{I}_{\phi}(v)=\frac{1}{n+1}\sum_{k=0}^{n}\int_{X}(u-v)\theta_{u}^{k}\wedge\theta_{v}^{n-k}. In particular Iϕ(u)Iϕ(v)\mathrm{I}_{\phi}(u)\leq\mathrm{I}_{\phi}(v) if uvu\leq v.
(ii) If uϕu\leq\phi then, X(uϕ)θunIϕ(u)1n+1X(uϕ)θun.\int_{X}(u-\phi)\theta_{u}^{n}\leq I_{\phi}(u)\leq\frac{1}{n+1}\int_{X}(u-\phi)\theta_{u}^{n}.
(iii) Iϕ\mathrm{I}_{\phi} is concave along affine curves. Also, the following estimates hold:

X(uv)θunIϕ(u)Iϕ(v)X(uv)θvn.\int_{X}(u-v)\theta_{u}^{n}\leq I_{\phi}(u)-I_{\phi}(v)\leq\int_{X}(u-v)\theta_{v}^{n}.
Proof.

Using Theorem 4.8, it is possible to repeat the arguments of [BEGZ10, Proposition 2.8], almost word for word. As a courtesy, we present a detailed proof.

We compute the derivative of f(t):=Iϕ(ut),t[0,1]f(t):=I_{\phi}(u_{t}),t\in[0,1], where ut:=tu+(1t)vu_{t}:=tu+(1-t)v. By the multi-linearity property of the non-pluripolar product we see that f(t)f(t) is a polynomial in tt. Using integration by parts (Theorem 4.8), one can check the following formula:

f(t)\displaystyle f^{\prime}(t) =\displaystyle= 1n+1(k=0nX(uv)θutkθϕnk+k=1nXk(utϕ)𝑑dc(uv)θutk1θϕnk)\displaystyle\frac{1}{n+1}\bigg{(}\sum_{k=0}^{n}\int_{X}(u-v)\theta_{u_{t}}^{k}\wedge\theta_{\phi}^{n-k}+\sum_{k=1}^{n}\int_{X}k(u_{t}-\phi)dd^{c}(u-v)\wedge\theta_{u_{t}}^{k-1}\wedge\theta_{\phi}^{n-k}\bigg{)}
=\displaystyle= 1n+1(k=0nX(uv)θutkθϕnk+k=1nXk(uv)(θutθϕ)θutk1θϕnk)\displaystyle\frac{1}{n+1}\bigg{(}\sum_{k=0}^{n}\int_{X}(u-v)\theta_{u_{t}}^{k}\wedge\theta_{\phi}^{n-k}+\sum_{k=1}^{n}\int_{X}k(u-v)(\theta_{u_{t}}-\theta_{\phi})\wedge\theta_{u_{t}}^{k-1}\wedge\theta_{\phi}^{n-k}\bigg{)}
=\displaystyle= X(uv)θutn.\displaystyle\int_{X}(u-v)\theta_{u_{t}}^{n}.

Computing one more derivative, we arrive at

f′′(t)=nX(uv)𝑑dc(uv)θutn1=nX(uv)(θuθv)θutn10,f^{\prime\prime}(t)=n\int_{X}(u-v)dd^{c}(u-v)\wedge\theta_{u_{t}}^{n-1}=n\int_{X}(u-v)(\theta_{u}-\theta_{v})\wedge\theta_{u_{t}}^{n-1}\leq 0,

where the inequality follows from Lemma 5.2.

Now, the function tf(t)t\mapsto f^{\prime}(t) is continuous on [0,1][0,1], thanks to convergence property of the Monge–Ampère operator (see Theorem 2.6). It thus follows that

Iϕ(u1)Iϕ(u0)=01f(t)𝑑t=01X(uv)θutn𝑑t.I_{\phi}(u_{1})-I_{\phi}(u_{0})=\int_{0}^{1}f^{\prime}(t)dt=\int_{0}^{1}\int_{X}(u-v)\theta_{u_{t}}^{n}dt.

Using the multi-linearity of the non-pluripolar product again, we get that

01X(uv)θutn𝑑t\displaystyle\int_{0}^{1}\int_{X}(u-v)\theta_{u_{t}}^{n}dt =\displaystyle= k=0n(01(nk)tk(1t)nk𝑑t)X(uv)θukθvnk\displaystyle\sum_{k=0}^{n}\left(\int_{0}^{1}\binom{n}{k}t^{k}(1-t)^{n-k}dt\right)\int_{X}(u-v)\theta_{u}^{k}\wedge\theta_{v}^{n-k}
=\displaystyle= 1n+1k=0nX(uv)θukθvnk.\displaystyle\frac{1}{n+1}\sum_{k=0}^{n}\int_{X}(u-v)\theta_{u}^{k}\wedge\theta_{v}^{n-k}.

This verifies (i), and another application of Lemma 5.2 finishes the proof of (iii).

For (ii) we observe that since uϕu-\phi\leq each term X(uϕ)θukθϕnk\int_{X}(u-\phi)\theta_{u}^{k}\wedge\theta_{\phi}^{n-k} is negative, hence Iϕ(u)X(uϕ)θunI_{\phi}(u)\leq\int_{X}(u-\phi)\theta_{u}^{n}. The left-hand side inequality of (ii) follows from Lemma 5.2. ∎

Lemma 5.4.

Suppose uj,u(X,θ,ϕ)u_{j},u\in\mathcal{E}(X,\theta,\phi) have relatively minimal singularity type such that ujuu_{j}\searrow u. Then Iϕ(uj)Iϕ(u){\rm I}_{\phi}(u_{j})\searrow\mathrm{I}_{\phi}(u).

Proof.

From Theorem 5.3 it follows that

|Iϕ(uj)Iϕ(u)|=Iϕ(uj)Iϕ(u)X(uju)θun.|{\rm I}_{\phi}(u_{j})-{\rm I}_{\phi}(u)|={\rm I}_{\phi}(u_{j})-{\rm I}_{\phi}(u)\leq\int_{X}(u_{j}-u)\theta_{u}^{n}.

An application of the dominated convergence theorem finishes the argument. ∎

We can now define the Monge–Ampère energy for arbitrary uPSH(X,θ,ϕ)u\in\mathrm{PSH}(X,\theta,\phi) using a familiar formula:

Iϕ(u):=inf{Iϕ(v)|v(X,θ,ϕ),vhas relatively minimal singularity type, and uv}.{\rm I}_{\phi}(u):=\inf\{{\rm I}_{\phi}(v)\ |\ v\in\mathcal{E}(X,\theta,\phi),\;v\ \textrm{has relatively minimal singularity type, and }u\leq v\}.
Lemma 5.5.

If uPSH(X,θ,ϕ)u\in{\rm PSH}(X,\theta,\phi) then Iϕ(u)=limtIϕ(max(u,ϕt)){\rm I}_{\phi}(u)=\lim_{t\to\infty}{\rm I}_{\phi}(\max(u,\phi-t)).

Proof.

It follows from the above definition that Iϕ(u)limtIϕ(max(u,ϕt)){\rm I}_{\phi}(u)\leq\lim_{t\to\infty}{\rm I}_{\phi}(\max(u,\phi-t)). Assume now that vPSH(X,θ,ϕ)v\in{\rm PSH}(X,\theta,\phi) is such that uvu\leq v, and vv has the same singularity type as ϕ\phi (i.e. vv is a candidate in the definition of Iϕ(u)I_{\phi}(u)). Then for tt large enough we have max(u,ϕt)v\max(u,\phi-t)\leq v, hence the other inequality follows from monotonicity of Iϕ{\rm I}_{\phi}. ∎

Let 1(X,θ,ϕ)\mathcal{E}^{1}(X,\theta,\phi) be the set of all uPSH(X,θ,ϕ)u\in{\rm PSH}(X,\theta,\phi) such that Iϕ(u){\rm I}_{\phi}(u) is finite. As a result of Lemma 5.5 and Theorem 5.3(i)(i) we observe that Iϕ{\rm I}_{\phi} is non-decreasing in PSH(X,θ,ϕ){\rm PSH}(X,\theta,\phi). Consequently, 1(X,θ,ϕ)\mathcal{E}^{1}(X,\theta,\phi) is stable under the max operation, moreover we have the following familiar characterization of 1(X,θ,ϕ)\mathcal{E}^{1}(X,\theta,\phi):

Lemma 5.6.

Let uPSH(X,θ,ϕ)u\in\mathrm{PSH}(X,\theta,\phi). Then u1(X,θ,ϕ)u\in\mathcal{E}^{1}(X,\theta,\phi) if and only if u(X,θ,ϕ)u\in\mathcal{E}(X,\theta,\phi) and X(uϕ)θun>\int_{X}(u-\phi)\theta_{u}^{n}>-\infty.

Proof.

Let u1(X,θ,ϕ)u\in\mathcal{E}^{1}(X,\theta,\phi). We can assume that uϕu\leq\phi. For each C>0C>0 we set uC:=max(u,ϕC)u^{C}:=\max(u,\phi-C).

If Iϕ(u)>I_{\phi}(u)>-\infty (say Iϕ(u)>AI_{\phi}(u)>-A), then by the monotonicity property we have Iϕ(uC)Iϕ(u)I_{\phi}(u^{C})\geq I_{\phi}(u). Since uCϕu^{C}\leq\phi, an application of Theorem 5.3(ii)(ii) gives that X(uCϕ)θuCnA,C\int_{X}(u^{C}-\phi)\theta_{u^{C}}^{n}\geq-A,\ \forall C, for some A>0A>0. From this we obtain that

{uϕC}θuCn{uϕC}ϕuCθuCnXϕuCCθuCnAC0,\int_{\{u\leq\phi-C\}}\theta_{u^{C}}^{n}\leq\int_{\{u\leq\phi-C\}}\frac{\phi-u}{C}\,\theta_{u^{C}}^{n}\leq\int_{X}\frac{\phi-u^{C}}{C}\,\theta_{u^{C}}^{n}\leq\frac{A}{C}\to 0,

as CC\to\infty. By plurifine locality and the above, we have

{u>ϕC}θun={u>ϕC}θuCn=XθuCn{uϕC}θuCn=Xθϕn{uϕC}θuCnXθϕn\int_{\{u>\phi-C\}}\theta_{u}^{n}=\int_{\{u>\phi-C\}}\theta_{u^{C}}^{n}=\int_{X}\theta_{u^{C}}^{n}-\int_{\{u\leq\phi-C\}}\theta_{u^{C}}^{n}=\int_{X}\theta_{\phi}^{n}-\int_{\{u\leq\phi-C\}}\theta_{u^{C}}^{n}\to\int_{X}\theta_{\phi}^{n}

as CC\to\infty, showing that u(X,θ,ϕ)u\in\mathcal{E}(X,\theta,\phi). Moreover,

X(uCϕ)θuCn{u>ϕC}(uϕ)θun.\int_{X}(u^{C}-\phi)\theta_{u^{C}}^{n}\leq\int_{\{u>\phi-C\}}(u-\phi)\theta_{u}^{n}.

Letting CC\to\infty we see that X(uϕ)θun>A\int_{X}(u-\phi)\theta_{u}^{n}>-A.

To prove the converse statement, assume that u(X,θ,ϕ)u\in\mathcal{E}(X,\theta,\phi) and X(uϕ)θun>\int_{X}(u-\phi)\theta_{u}^{n}>-\infty. For each C>0C>0 the measures θun\theta_{u}^{n} and θuCn\theta_{u^{C}}^{n} have the same total mass and they coincide on {u>ϕC}\{u>\phi-C\}. It follows that {uϕC}θuCn={uϕC}θun\int_{\{u\leq\phi-C\}}\theta_{u^{C}}^{n}=\int_{\{u\leq\phi-C\}}\theta_{u}^{n}, and from this we deduce that

X(uCϕ)θuCn\displaystyle\int_{X}(u^{C}-\phi)\theta_{u^{C}}^{n} =\displaystyle= {uϕC}Cθun+{u>ϕC}(uϕ)θunX(uϕ)θun>A.\displaystyle-\int_{\{u\leq\phi-C\}}C\theta_{u}^{n}+\int_{\{u>\phi-C\}}(u-\phi)\theta_{u}^{n}\geq\int_{X}(u-\phi)\theta_{u}^{n}>-A.

It thus follows from Theorem 5.3(ii)(ii) that Iϕ(uC)I_{\phi}(u^{C}) is uniformly bounded. Finally, it follows from Lemma 5.5 that Iϕ(uC)Iϕ(u)I_{\phi}(u^{C})\searrow I_{\phi}(u) as CC\to\infty, finishing the proof. ∎

We finish this section with a series of standard results listing various properties of the class 1(X,θ,ϕ)\mathcal{E}^{1}(X,\theta,\phi).

Lemma 5.7.

Assume that uj,u1(X,θ,ϕ)u_{j},u\in\mathcal{E}^{1}(X,\theta,\phi) such that ujuu_{j}\searrow u. Then Iϕ(uj){\rm I}_{\phi}(u_{j}) decreases to Iϕ(u){\rm I}_{\phi}(u).

Proof.

Without loss of generality we can assume that ujϕu_{j}\leq\phi for all jj. For each C>0C>0 we set ujC:=max(uj,ϕC)u_{j}^{C}:=\max(u_{j},\phi-C) and uC:=max(u,ϕC)u^{C}:=\max(u,\phi-C). Note that ujC,uCu_{j}^{C},u^{C} have the same singularity type as ϕ\phi. Then Lemma 5.4 insures that limjIϕ(ujC)=Iϕ(uC)\lim_{j}I_{\phi}(u_{j}^{C})=I_{\phi}(u^{C}). Monotonicity of IϕI_{\phi} gives now that Iϕ(u)limjIϕ(uj)limjIϕ(ujC)=Iϕ(uC)I_{\phi}(u)\leq\lim_{j}I_{\phi}(u_{j})\leq\lim_{j}I_{\phi}(u_{j}^{C})=I_{\phi}(u^{C}). Letting CC\to\infty, the result follows from Lemma 5.5. ∎

Lemma 5.8.

Assume that {uj}j1(X,θ,ϕ)\{u_{j}\}_{j}\subset\mathcal{E}^{1}(X,\theta,\phi) is decreasing, such that Iϕ(uj){\rm I}_{\phi}(u_{j}) is uniformly bounded. Then the limit u:=limjuju:=\lim_{j}u_{j} belongs to 1(X,θ,ϕ)\mathcal{E}^{1}(X,\theta,\phi) and Iϕ(uj){\rm I}_{\phi}(u_{j}) decreases to Iϕ(u){\rm I}_{\phi}(u).

Proof.

We can assume that ujϕu_{j}\leq\phi for all jj. As all the terms in the definition of Iϕ(uj)I_{\phi}(u_{j}) are negative, we notice that C(n+1)Iϕ(uj)X(ujϕ)θϕn0-C\leq(n+1){\rm I}_{\phi}(u_{j})\leq\int_{X}(u_{j}-\phi)\theta_{\phi}^{n}\leq 0 for some C>0C>0.

If uu\equiv-\infty, then supXuj=supX(ujϕ)\sup_{X}u_{j}=\sup_{X}(u_{j}-\phi)\to-\infty (Lemma 3.5). This implies that X(ujϕ)θϕnsupX(ujϕ)Xθϕn\int_{X}(u_{j}-\phi)\theta_{\phi}^{n}\leq\sup_{X}(u_{j}-\phi)\int_{X}\theta_{\phi}^{n}\to-\infty, a contradiction. Hence uPSH(X,θ)u\in\textup{PSH}(X,\theta).

By continuity along decreasing sequences (Lemma 5.7) we have

limjIϕ(max(uj,ϕC))=Iϕ(max(u,ϕC)).\lim_{j\rightarrow\infty}{\rm I}_{\phi}(\max(u_{j},\phi-C))={\rm I}_{\phi}(\max(u,\phi-C)).

It follows that Iϕ(max(u,ϕC)){\rm I}_{\phi}(\max(u,\phi-C)) is uniformly bounded. Lemma 5.5 then insures that Iϕ(u){\rm I}_{\phi}(u) is finite, i.e., u1(X,θ,ϕ)u\in\mathcal{E}^{1}(X,\theta,\phi). ∎

Corollary 5.9.

Iϕ{\rm I}_{\phi} is concave along affine curves in PSH(X,θ,ϕ){\rm PSH}(X,\theta,\phi). In particular, the set 1(X,θ,ϕ)\mathcal{E}^{1}(X,\theta,\phi) is convex.

Proof.

Let u,vPSH(X,θ,ϕ)u,v\in{\rm PSH}(X,\theta,\phi) and ut:=tu+(1t)v,t(0,1)u_{t}:=tu+(1-t)v,t\in(0,1). If one of u,vu,v is not in 1(X,θ,ϕ)\mathcal{E}^{1}(X,\theta,\phi) then the conclusion is obvious. So, we can assume that both uu and vv belong to 1(X,θ,ϕ)\mathcal{E}^{1}(X,\theta,\phi). For each C>0C>0 we set utC:=tmax(u,ϕC)+(1t)max(v,ϕC)u_{t}^{C}:=t\max(u,\phi-C)+(1-t)\max(v,\phi-C). By Theorem 5.3(iii)(iii), tIϕ(utC)t\to{\rm I}_{\phi}(u_{t}^{C}) is concave. Since utCu_{t}^{C} decreases to utu_{t} as CC\to\infty, Lemma 5.8 gives the conclusion. ∎

Lemma 5.10.

Suppose u,v1(X,θ,ϕ)u,v\in\mathcal{E}^{1}(X,\theta,\phi) have the same singularity type. Then

X(uv)θunIϕ(u)Iϕ(v)X(uv)θvn.\int_{X}(u-v)\theta_{u}^{n}\leq{\rm I}_{\phi}(u)-{\rm I}_{\phi}(v)\leq\int_{X}(u-v)\theta_{v}^{n}.

With a bit of extra work, one can also get rid of the assumption that uu and vv have the same singularity type [DDL18, Proposition 2.5].

Proof.

First, note that these estimates hold for uC:=max(u,ϕC),vC:=max(v,ϕC)u^{C}:=\max(u,\phi-C),v^{C}:=\max(v,\phi-C), by Theorem 5.3(iii)(iii). It is easy to see that uCvCu^{C}-v^{C} is uniformly bounded and converges in capacity to uvu-v. Putting these last two facts together, Theorem 2.6 gives that

|X(uCvC)θvCnX(uv)θvn|0,|X(uCvC)θuCnX(uv)θun|0.\displaystyle\bigg{|}\int_{X}(u^{C}-v^{C})\theta_{v^{C}}^{n}-\int_{X}(u-v)\theta_{v}^{n}\bigg{|}\to 0,\ \ \ \bigg{|}\int_{X}(u^{C}-v^{C})\theta_{u^{C}}^{n}-\int_{X}(u-v)\theta_{u}^{n}\bigg{|}\to 0.

The result follows from Lemma 5.5. ∎

5.2 Monge–Ampère equations with prescribed singularity type

Throughout this section ϕ\phi is a θ\theta-psh function satisfying ϕ=Pθ[ϕ]\phi=P_{\theta}[\phi], and Xθϕ>0\int_{X}\theta_{\phi}>0. We additionally assume the normalization

Xθϕn=1.\int_{X}\theta_{\phi}^{n}=1.

This can always be achieved by rescaling our big class {θ}\{\theta\}. In this section we consider the following complex Monge–Ampère equation in our prescribed setting:

θun=eλuμ,u(X,θ,ϕ).\theta_{u}^{n}=e^{\lambda u}\mu,\ u\in\mathcal{E}(X,\theta,\phi). (5.3)

where λ0\lambda\geq 0, μ\mu is a positive non-pluripolar probability measure on XX.

When λ>0\lambda>0, we adapt the variational method in [BBGZ13] to our needs. When λ=0\lambda=0, it is also possible to use the variational method to solve the equation, but it requires a detailed study of the relative Monge–Ampère capacities, as carried out in [DDL18a]. In this survey we provide a simpler approach by using the solutions for the case λ>0\lambda>0 and letting λ0+\lambda\to 0^{+}.

Proposition 5.11.

Iϕ:1(X,θ,ϕ){\rm I}_{\phi}:\mathcal{E}^{1}(X,\theta,\phi)\to\mathbb{R} is upper semicontinuous with respect to the weak L1L^{1} topology of potentials.

Proof.

Assume that {uj}j\{u_{j}\}_{j} is a sequence in 1(X,θ,ϕ)\mathcal{E}^{1}(X,\theta,\phi) L1L^{1}-converging to u1(X,θ,ϕ)u\in\mathcal{E}^{1}(X,\theta,\phi). We can assume that uj0u_{j}\leq 0 for all jj. For each k,k,\ell\in\mathbb{N} we set vk,:=max(uk,,uk+)v_{k,\ell}:=\max(u_{k},...,u_{k+\ell}). As 1(X,θ,ϕ)\mathcal{E}^{1}(X,\theta,\phi) is stable under the max operation, we have that vk,1(X,θ,ϕ)v_{k,\ell}\in\mathcal{E}^{1}(X,\theta,\phi).

Moreover vk,φk:=usc(supjkuj)v_{k,\ell}\nearrow\varphi_{k}:=\textup{usc}\left(\sup_{j\geq k}u_{j}\right), hence by the monotonicity property we get Iϕ(φk)Iϕ(vk,)Iϕ(uk)>{\rm I}_{\phi}(\varphi_{k})\geq{\rm I}_{\phi}(v_{k,\ell})\geq{\rm I}_{\phi}(u_{k})>-\infty. As a result, φk1(X,θ,ϕ)\varphi_{k}\in\mathcal{E}^{1}(X,\theta,\phi). By Hartogs’ lemma φku\varphi_{k}\searrow u as kk\to\infty. By Lemma 5.7 it follows that Iϕ(φk){\rm I}_{\phi}(\varphi_{k}) decreases to Iϕ(u){\rm I}_{\phi}(u). Thus, using the monotonicity of Iϕ{\rm I}_{\phi} we get Iϕ(u)=limkIϕ(φk)lim supkIϕ(uk),{\rm I}_{\phi}(u)=\lim_{k\rightarrow\infty}{\rm I}_{\phi}(\varphi_{k})\geq\limsup_{k\rightarrow\infty}{\rm I}_{\phi}(u_{k}), finishing the proof. ∎

Next we describe the first order variation of IϕI_{\phi}, shadowing a result from [BB10]:

Proposition 5.12.

Let u1(X,θ,ϕ)u\in\mathcal{E}^{1}(X,\theta,\phi) and χ\chi be a continuous function on XX. For each tt\in\mathbb{R} set ut:=Pθ(u+tχ)u_{t}:=P_{\theta}(u+t\chi). Then ut1(X,θ,ϕ)u_{t}\in\mathcal{E}^{1}(X,\theta,\phi), tIϕ(ut)t\mapsto{\rm I}_{\phi}(u_{t}) is differentiable, and its derivative is given by

ddtIϕ(ut)=Xχθutn,t.\frac{d}{dt}{\rm I}_{\phi}(u_{t})=\int_{X}\chi\theta_{u_{t}}^{n},\ t\in\mathbb{R}.
Proof.

For t0t\geq 0 the potential u+tinfXχu+t\inf_{X}\chi is a candidate in each envelope, hence u+tinfXχutu+t\inf_{X}\chi\leq u_{t}. Monotonicity of IϕI_{\phi} now implies that ut1(X,θ,ϕ)u_{t}\in\mathcal{E}^{1}(X,\theta,\phi). A similar argument implies that ut1(X,θ,ϕ)u_{t}\in\mathcal{E}^{1}(X,\theta,\phi) for t0t\leq 0.

Let tt\in\mathbb{R} and s>0s>0. As the singularity type of each utu_{t} is the same of that of uu, we can apply Lemma 5.10 and conclude:

X(ut+sut)θut+snIϕ(ut+s)Iϕ(ut)X(ut+sut)θutn.\int_{X}(u_{t+s}-u_{t})\theta_{u_{t+s}}^{n}\leq{\rm I}_{\phi}(u_{t+s})-{\rm I}_{\phi}(u_{t})\leq\int_{X}(u_{t+s}-u_{t})\theta_{u_{t}}^{n}.

It follows from Theorem 2.7 that θutn\theta_{u_{t}}^{n} is supported on {ut=u+tχ}\{u_{t}=u+t\chi\}. We thus have

X(ut+sut)θutn=X(ut+sutχ)θutnXsχθutn,\int_{X}(u_{t+s}-u_{t})\theta_{u_{t}}^{n}=\int_{X}(u_{t+s}-u-t\chi)\theta_{u_{t}}^{n}\leq\int_{X}s\chi\theta_{u_{t}}^{n},

since ut+su+(t+s)χu_{t+s}\leq u+(t+s)\chi. Similarly we have

X(ut+sut)θut+sn=X(u+(t+s)χut)θut+snXsχθut+sn.\int_{X}(u_{t+s}-u_{t})\theta_{u_{t+s}}^{n}=\int_{X}(u+(t+s)\chi-u_{t})\theta_{u_{t+s}}^{n}\geq\int_{X}s\chi\theta_{u_{t+s}}^{n}.

Since ut+su_{t+s} converges uniformly to utu_{t} as s0s\to 0, by Theorem 2.6 it follows that θut+sn\theta_{u_{t+s}}^{n} converges weakly to θutn\theta_{u_{t}}^{n}. As χ\chi is continuous, dividing by s>0s>0 and letting s0+s\to 0^{+} we see that the right derivative of Iϕ(ut){\rm I}_{\phi}(u_{t}) at tt is Xχθutn\int_{X}\chi\theta_{u_{t}}^{n}. The same argument applies for the left derivative. ∎

The case λ>0\lambda>0.

It suffices to treat the case λ=1\lambda=1 as the other cases can be done similarly.

We introduce the following functional on 1(X,θ,ϕ)\mathcal{E}^{1}(X,\theta,\phi):

F(u):=Fμ(u):=Iϕ(u)Lμ(u),u1(X,θ,ϕ),F(u):=F_{\mu}(u):={\rm I}_{\phi}(u)-L_{\mu}(u),\ u\in\mathcal{E}^{1}(X,\theta,\phi),

where Lμ(u):=Xeu𝑑μL_{\mu}(u):=\int_{X}e^{u}d\mu.

Theorem 5.13.

Assume that u1(X,θ,ϕ)u\in\mathcal{E}^{1}(X,\theta,\phi) maximizes FF on 1(X,θ,ϕ)\mathcal{E}^{1}(X,\theta,\phi). Then uu solves the equation (5.3).

Proof.

Let χ\chi be an arbitrary continuous function on XX and set ut:=Pθ(u+tχ)u_{t}:=P_{\theta}(u+t\chi). It follows from Proposition 5.12 that ut1(X,θ,ϕ)u_{t}\in\mathcal{E}^{1}(X,\theta,\phi) for all tt\in\mathbb{R}, that the function

g(t):=Iϕ(ut)Lμ(u+tχ)g(t):={\rm I}_{\phi}(u_{t})-L_{\mu}(u+t\chi)

is differentiable on \mathbb{R}, and its derivative is given by

g(t)=XχθutnXχe(u+tχ)𝑑μ.g^{\prime}(t)=\int_{X}\chi\theta_{u_{t}}^{n}-\int_{X}\chi e^{(u+t\chi)}d\mu.

Moreover, as utu+tχu_{t}\leq u+t\chi, we have

g(t)Iϕ(ut)Lμ(ut)=Fμ(ut)sup1(X,θ,ϕ)Fμ=F(u)=g(0).g(t)\leq{\rm I}_{\phi}(u_{t})-L_{\mu}(u_{t})=F_{\mu}(u_{t})\leq\sup_{\mathcal{E}^{1}(X,\theta,\phi)}F_{\mu}=F(u)=g(0).

This means that gg attains a maximum at 0, hence g(0)=0g^{\prime}(0)=0. Since χC0(X)\chi\in C^{0}(X) is arbitrary it follows that θun=eλuμ\theta_{u}^{n}=e^{\lambda u}\mu. ∎

Having computed the first order variation of the Monge–Ampère energy, we establish the following existence and uniqueness result.

Theorem 5.14.

Assume that μ\mu is a positive non-pluripolar measure on XX. Then there exists a unique u1(X,θ,ϕ)u\in\mathcal{E}^{1}(X,\theta,\phi) such that

θun=euμ.\theta_{u}^{n}=e^{u}\mu. (5.4)
Proof.

We use the variational method. Let {uj}j\{u_{j}\}_{j} be a sequence in 1(X,θ,ϕ)\mathcal{E}^{1}(X,\theta,\phi) such that limjF(uj)=sup1(X,θ,ϕ)F>\lim_{j}F(u_{j})=\sup_{\mathcal{E}^{1}(X,\theta,\phi)}F>-\infty. We claim that supXuj\sup_{X}u_{j} is uniformly bounded from above. Indeed, assume that it were not the case. Then by relabeling the sequence we can assume that supXuj\sup_{X}u_{j} increase to \infty. By the compactness property [GZ05, Proposition 2.7] it follows that the sequence ψj:=ujsupXuj\psi_{j}:=u_{j}-\sup_{X}u_{j} converges in L1(X,ωn)L^{1}(X,\omega^{n}) to some ψPSH(X,θ)\psi\in{\rm PSH}(X,\theta) such that supXψ=0\sup_{X}\psi=0. In particular, Xeψ𝑑μ>0\int_{X}e^{\psi}d\mu>0. It thus follows that

Xeuj𝑑μ=esupXujXeψj𝑑μcesupXuj\int_{X}e^{u_{j}}d\mu=e^{\sup_{X}u_{j}}\int_{X}e^{\psi_{j}}d\mu\geq c\ e^{\sup_{X}u_{j}} (5.5)

for some c>0c>0. Note also that ψjϕ\psi_{j}\leq\phi since ψj(X,θ,ϕ)\psi_{j}\in\mathcal{E}(X,\theta,\phi) and ψj0\psi_{j}\leq 0 and ϕ\phi is the maximal function with these properties (see Theorem 3.14). It then follows that

Iϕ(uj)=Iϕ(ψj)+supXujsupXuj.{\rm I}_{\phi}(u_{j})={\rm I}_{\phi}(\psi_{j})+\sup_{X}u_{j}\leq\sup_{X}u_{j}. (5.6)

From (5.5) and (5.6) we arrive at

limjF(uj)limj(supXujcesupXuj)=,\lim_{j\to\infty}F(u_{j})\leq\lim_{j\to\infty}\left(\sup_{X}u_{j}-ce^{\sup_{X}u_{j}}\right)=-\infty,

which is a contradiction. Thus supXuj\sup_{X}u_{j} is bounded from above as claimed. Since F(uj)Iϕ(uj)supXujF(u_{j})\leq{\rm I}_{\phi}(u_{j})\leq\sup_{X}u_{j} it follows that Iϕ(uj){\rm I}_{\phi}(u_{j}) and hence supXuj\sup_{X}u_{j} is also bounded from below. It follows again from [GZ05, Proposition 2.7] that a subsequence of uju_{j} (still denoted by uju_{j}) converges in L1(X,ωn)L^{1}(X,\omega^{n}) to some uPSH(X,θ)u\in{\rm PSH}(X,\theta). Since Iϕ{\rm I}_{\phi} is upper semicontinuous we have

lim supjIϕ(uj)Iϕ(u),\limsup_{j\to\infty}{\rm I}_{\phi}(u_{j})\leq{\rm I}_{\phi}(u),

hence u1(X,θ,ϕ)u\in\mathcal{E}^{1}(X,\theta,\phi). Moreover, by continuity of φXeφ𝑑μ\varphi\mapsto\int_{X}e^{\varphi}d\mu we get that F(u)sup1(X,θ,ϕ)FF(u)\geq\sup_{\mathcal{E}^{1}(X,\theta,\phi)}F. Hence uu maximizes FF on 1(X,θ,ϕ)\mathcal{E}^{1}(X,\theta,\phi). Now Theorem 5.13 shows that uu solves the desired complex Monge–Ampère equation. The next lemma address the uniqueness question. ∎

Lemma 5.15.

Assume that u(X,θ,ϕ)u\in\mathcal{E}(X,\theta,\phi) is a solution of (5.4) and v(X,θ,ϕ)v\in\mathcal{E}(X,\theta,\phi) is a subsolution, i.e., θvnevμ.\theta_{v}^{n}\geq e^{v}\mu. Then uvu\geq v on XX.

Proof.

By the comparison principle for the class (X,θ,ϕ)\mathcal{E}(X,\theta,\phi) (Corollary 3.23) we have

{u<v}θvn{u<v}θun.\int_{\{u<v\}}\theta_{v}^{n}\leq\int_{\{u<v\}}\theta_{u}^{n}.

As uu is a solution and vv is a subsolution to (5.4) we then have

{u<v}ev𝑑μ{u<v}θvn{u<v}θun={u<v}eu𝑑μ{u<v}ev𝑑μ.\int_{\{u<v\}}e^{v}d\mu\leq\int_{\{u<v\}}\theta_{v}^{n}\leq\int_{\{u<v\}}\theta_{u}^{n}=\int_{\{u<v\}}e^{u}d\mu\leq\int_{\{u<v\}}e^{v}d\mu.

It follows that all inequalities above are equalities, hence μ({u<v})=0\mu(\{u<v\})=0. Since μ=euθun\mu=e^{u}\theta_{u}^{n}, it follows that θun({u<v})=0\theta_{u}^{n}(\{u<v\})=0. By the domination principle (Theorem 3.12) we get that uvu\geq v everywhere on XX. ∎

The case λ=0\lambda=0.

It immediately follows from Lemma 2.9 that subsolutions are preserved under taking maximums. In addition to this, the L1L^{1}-limit of subsolutions is also a subsolution:

Lemma 5.16.

Let (uj)(u_{j}) be a sequence of θ\theta-psh functions such that θujnfjμ\theta_{u_{j}}^{n}\geq f_{j}\mu, where fjL1(X,μ)f_{j}\in L^{1}(X,\mu) and μ\mu is a positive non-pluripolar Borel measure on XX. Assume that fjf_{j} converge in L1(X,μ)L^{1}(X,\mu) to fL1(X,μ)f\in L^{1}(X,\mu), and uju_{j} converge in L1(X,ωn)L^{1}(X,\omega^{n}) to uPSH(X,θ)u\in{\rm PSH}(X,\theta). Then θunfμ\theta_{u}^{n}\geq f\mu.

Proof.

By extracting a subsequence if necessary, we can assume that fjf_{j} converge μ\mu-a.e. to ff. For each kk we set vk:=usc(supjkuj)v_{k}:=\textup{usc}(\sup_{j\geq k}u_{j}). Then vkv_{k} decreases pointwise to uu and Lemma 2.9 together with Theorem 2.6 gives

θvkn(infjkfj)μ.\theta_{v_{k}}^{n}\geq\left(\inf_{j\geq k}f_{j}\right)\mu.

Indeed max(uk,,uk+l)vk\max(u_{k},\ldots,u_{k+l})\nearrow v_{k} a.e., as ll\to\infty, making Theorem 2.6 applicable due to Remark 3.4.

To explain our notation below, for t>0t>0 and a function gg we set gt:=max(g,Vθt)g^{t}:=\max(g,V_{\theta}-t).

Note that {u>Vθt}{vk>Vθt}\{u>V_{\theta}-t\}\subset\{v_{k}>V_{\theta}-t\}. Multiplying both sides of the above estimate with 𝟏{u>Vθt}{\bf 1}_{\{u>V_{\theta}-t\}}, t>0t>0, and using the locality of the complex Monge–Ampère operator with respect to the plurifine topology we arrive at

θvktn𝟏{u>Vθt}(infjkfj)μ.\theta_{v_{k}^{t}}^{n}\geq{\bf 1}_{\{u>V_{\theta}-t\}}\left(\inf_{j\geq k}f_{j}\right)\mu.

Note that for t>0t>0 fixed, vktv_{k}^{t} decreases to utu^{t}, all having minimal singularity type (in particular they all have full mass); also the sequence (infjkfj)\left(\inf_{j\geq k}f_{j}\right) is increasing to ff. Letting kk\to\infty and using Theorem 2.6 we obtain

θutn𝟏{u>Vθt}fμ,t>0.\theta_{u^{t}}^{n}\geq{\bf 1}_{\{u>V_{\theta}-t\}}f\mu,\ t>0.

Again, multiplying both sides with 𝟏{u>Vθt}{\bf 1}_{\{u>V_{\theta}-t\}}, t>0t>0, and using the locality of the complex Monge–Ampère operator with respect to the plurifine topology we arrive at

𝟏{u>Vθt}θun𝟏{u>Vθt}fμ.{\bf 1}_{\{u>V_{\theta}-t\}}\theta_{u}^{n}\geq{\bf 1}_{\{u>V_{\theta}-t\}}f\mu.

Finally, letting tt\to\infty we obtain the result. ∎

Theorem 5.17.

Assume that μ\mu is a non-pluripolar positive measure such that μ(X)=1\mu(X)=1. Then there exists a unique u(X,θ,ϕ)u\in\mathcal{E}(X,\theta,\phi) such that θun=μ\theta_{u}^{n}=\mu and supXu=0\sup_{X}u=0.

Proof.

The uniqueness follows from Theorem 3.13. For each j>0j>0, using Theorem 5.14 we solve

θvjn=ej1vjμ,vj(X,θ,ϕ).\theta_{v_{j}}^{n}=e^{j^{-1}v_{j}}\mu,\quad v_{j}\in\mathcal{E}(X,\theta,\phi).

We set uj:=vjsupXvju_{j}:=v_{j}-\sup_{X}v_{j}, so that supXuj=0\sup_{X}u_{j}=0. Up to extracting a subsequence, we can assume that ujuPSH(X,θ)u_{j}\to u\in{\rm PSH}(X,\theta) in L1L^{1} and almost everywhere. Since ujϕu_{j}\leq\phi, we have uϕu\leq\phi. Observe also that j1ujj^{-1}u_{j} converge in capacity to 0. Indeed, for any fixed ε>0\varepsilon>0 we have

Capω(j1uj<ε)=Capω(uj<εj)0{\rm Cap}_{\omega}(j^{-1}u_{j}<-\varepsilon)={\rm Cap}_{\omega}(u_{j}<-\varepsilon j)\to 0

as follows from [GZ05, Proposition 3.6].

Hence the functions ej1uje^{j^{-1}u_{j}} converge in capacity to 11. It thus follows from Theorem 2.6 that

limjXej1uj𝑑μ=μ(X)=1.\lim_{j\to\infty}\int_{X}e^{j^{-1}u_{j}}d\mu=\mu(X)=1.

Since θujn=ej1supXvjej1ujμ\theta_{u_{j}}^{n}=e^{j^{-1}\sup_{X}v_{j}}e^{j^{-1}u_{j}}\mu, it follows from the above that j1supXvj0j^{-1}\sup_{X}v_{j}\to 0 as jj\to\infty, hence ej1vje^{j^{-1}v_{j}} converges to 11 in L1(μ)L^{1}(\mu) and almost everywhere. It thus follows from Lemma 5.16 that θunμ\theta_{u}^{n}\geq\mu. Since uϕu\leq\phi, Theorem 3.3 ensures that XθunXθϕn=1=μ(X)\int_{X}\theta_{u}^{n}\leq\int_{X}\theta_{\phi}^{n}=1=\mu(X). Hence θun=μ\theta_{u}^{n}=\mu as desired. ∎

Log concavity of non-pluripolar masses.

We give a proof for the following result from [DDL21], initially conjectured in [BEGZ10, Conjecture 1.23]:

Theorem 5.18.

Let T1,,TnT_{1},...,T_{n} be positive (1,1)(1,1)-currents on a compact Kähler manifold XX. Then

XT1Tn(XT1n)1n(XTnn)1n.\int_{X}\langle T_{1}\wedge...\wedge T_{n}\rangle\geq\bigg{(}\int_{X}\langle T_{1}^{n}\rangle\bigg{)}^{\frac{1}{n}}...\bigg{(}\int_{X}\langle T_{n}^{n}\rangle\bigg{)}^{\frac{1}{n}}.
Proof.

We can assume that the classes of TjT_{j} are big and their total masses are non-zero. Otherwise the right-hand side of the inequality to be proved is zero. Consider smooth closed real (1,1)(1,1)-forms θj\theta^{j}, and ujPSH(X,θj)u_{j}\in{\rm PSH}(X,\theta^{j}) such that Tj=θujjT_{j}=\theta^{j}_{u_{j}}.

We can assume that Xωn=1\int_{X}\omega^{n}=1. For each j=1,,nj=1,...,n, Theorem 5.17 ensures that there exists a normalizing constant cj>0c_{j}>0 and φj(X,θj,Pθj[uj])\varphi_{j}\in\mathcal{E}(X,\theta^{j},P_{\theta^{j}}[u_{j}]) such that (θφjj)n=cjωn\big{(}\theta^{j}_{\varphi_{j}}\big{)}^{n}=c_{j}\omega^{n}.

Thus we have

cj=X(θφjj)n=X(θPθj[uj]j)n=X(θujj)n=XTjn.c_{j}=\int_{X}\big{(}\theta^{j}_{\varphi_{j}}\big{)}^{n}=\int_{X}\big{(}\theta^{j}_{P_{\theta^{j}}[u_{j}]}\big{)}^{n}=\int_{X}\big{(}\theta^{j}_{u_{j}}\big{)}^{n}=\int_{X}\langle T_{j}^{n}\rangle.

Remark 3.4 and Theorem 3.15 then gives

Xθφ11θφnn=XθPθ1[u1]1θPθn[un]n=Xθu11θunn=XT1Tn.\int_{X}\theta^{1}_{\varphi_{1}}\wedge...\wedge\theta^{n}_{\varphi_{n}}=\int_{X}\theta^{1}_{P_{\theta^{1}}[u_{1}]}\wedge...\wedge\theta^{n}_{P_{\theta^{n}}[u_{n}]}=\int_{X}\theta^{1}_{u_{1}}\wedge...\wedge\theta^{n}_{u_{n}}=\int_{X}\langle T_{1}\wedge\ldots\wedge T_{n}\rangle.

An application of the mixed Monge–Ampère inequalities ([BEGZ10, Proposition 1.11]) gives that θφ11θφnnc11/ncn1/nωn\theta^{1}_{\varphi_{1}}\wedge\ldots\wedge\theta^{n}_{\varphi_{n}}\geq c_{1}^{1/n}\ldots c_{n}^{1/n}\omega^{n}. The result follows from integrating this estimate. ∎

5.3 Relative boundedness of solutions

Recall that we are working with ϕPSH(X,θ)\phi\in{\rm PSH}(X,\theta) such that Pθ[ϕ]=ϕP_{\theta}[\phi]=\phi, and Xθϕn>0\int_{X}\theta_{\phi}^{n}>0. Let fLp(ωn)f\in L^{p}(\omega^{n}) with f0f\geq 0. In the previous section we have shown that the equation

θun=fωn,u(X,θ,ϕ)\theta_{u}^{n}=f\omega^{n},\ \ u\in\mathcal{E}(X,\theta,\phi)

has a unique solution. In this section we will show that this solution has the same singularity type as ϕ\phi. This generalizes [BEGZ10, Theorem B], that treats the particular case of solutions with minimal singularity type in a big class. Analogous results will be obtained for equations of the type (5.4) as well.

Our arguments follow the one in [GL21] which relies on quasi-psh envelopes. The original argument in [DDL18a, DDL21] was given using a Kolodziej type estimate, inspired from [Koł98].

In our study we make use of the following lemma multiple times:

Lemma 5.19.

Let v(X,θ,ϕ)v\in\mathcal{E}(X,\theta,\phi), vϕv\leq\phi, and γ:+{}+{}\gamma:\mathbb{R}^{+}\cup\{\infty\}\mapsto\mathbb{R}^{+}\cup\{\infty\} denote a concave continuous increasing function with γ1\gamma^{\prime}\leq 1. Then χ:=γ(ϕv)+ϕPSH(X,θ,ϕ)\chi:=-\gamma(\phi-v)+\phi\in\textup{PSH}(X,\theta,\phi) and χ\chi satisfies

θχn(γ(ϕv))nθvn.\theta_{\chi}^{n}\geq(\gamma^{\prime}(\phi-v))^{n}\theta_{v}^{n}. (5.7)

To be precise, in the above statement the function χ\chi is equal to -\infty on the pluripolar set {ϕ=}\{\phi=-\infty\} and χ=γ(ϕv)+ϕ\chi=-\gamma(\phi-v)+\phi otherwise.

Proof.

Notice that there exists γk:+{}+{}\gamma_{k}:\mathbb{R}^{+}\cup\{\infty\}\mapsto\mathbb{R}^{+}\cup\{\infty\} smooth on [0,)[0,\infty), such that γkγ\gamma_{k}\nearrow\gamma pointwise and γk1\gamma^{\prime}_{k}\leq 1. As a result, it will be enough to prove the theorem for γ\gamma smooth.

First we want to show that χ=γ(ϕv)+ϕPSH(X,θ,ϕ)\chi=-\gamma(\phi-v)+\phi\in\textup{PSH}(X,\theta,\phi). Observe that χϕ\chi\leq\phi since γ0-\gamma\leq 0. In order to prove that χ\chi is θ\theta-psh, we will show that it is the decreasing limit of θ\theta-psh functions.

Since the problem is local, we can work in a local chart VXV\subset X where we write θ=ddcg\theta=dd^{c}g in VV. After slightly shrinking VV, the mollifications vj:=ρ1/j(v+g)gv_{j}:=\rho_{1/j}*(v+g)-g and ϕj:=ρ1/j(ϕ+g)g\phi_{j}:=\rho_{1/j}*(\phi+g)-g are smooth approximants of vv and ϕ\phi on VV. Moreover, they are θ\theta-psh on VV, vjvv_{j}\searrow v, ϕjϕ\phi_{j}\searrow\phi, and vjϕjv_{j}\leq\phi_{j}. Let χj:=γ(ϕjvj)+ϕj\chi_{j}:=-\gamma(\phi_{j}-v_{j})+\phi_{j}. Then for any jj, χj\chi_{j} is a smooth θ\theta-psh function in VV since

θ+ddcχj\displaystyle\theta+dd^{c}\chi_{j} =\displaystyle= γ(ϕjvj)θvj+(1γ(ϕjvj))θϕjγ′′(ϕjvj)d(ϕjvj)dc(ϕjvj)\displaystyle\gamma^{\prime}(\phi_{j}-v_{j})\theta_{v_{j}}+(1-\gamma^{\prime}(\phi_{j}-v_{j}))\theta_{\phi_{j}}-\gamma^{\prime\prime}(\phi_{j}-v_{j})d(\phi_{j}-v_{j})\wedge d^{c}(\phi_{j}-v_{j}) (5.8)
\displaystyle\geq γ(ϕjvj)θvj,\displaystyle\gamma^{\prime}(\phi_{j}-v_{j})\theta_{v_{j}},

where we used that γ\gamma is concave and that γ1\gamma^{\prime}\leq 1.

We claim that χj\chi_{j} is decreasing in jj. Indeed, for ss\in\mathbb{R} fixed, the function tγ(ts)+tt\mapsto-\gamma(t-s)+t is increasing in [s,+)[s,+\infty) because 1γ(ts)01-\gamma^{\prime}(t-s)\geq 0. Using also that γ\gamma is increasing we find that for jkj\leq k we have

χj=γ(ϕjvj)+ϕjγ(ϕjvk)+ϕjγ(ϕkvk)+ϕk=χk.\chi_{j}=-\gamma(\phi_{j}-v_{j})+\phi_{j}\geq-\gamma(\phi_{j}-v_{k})+\phi_{j}\geq-\gamma(\phi_{k}-v_{k})+\phi_{k}=\chi_{k}.

It follows that χj\chi_{j} decreases to some θ\theta-psh function in VV which has to be χ\chi. Indeed, the pointwise convergence χjχ\chi_{j}\to\chi is seen to hold on the set {ϕ>}\{\phi>-\infty\}. Since we have χjϕj\chi_{j}\leq\phi_{j}\to-\infty on {ϕ>}\{\phi>-\infty\}, pointwise convergence holds everywhere on XX. It follows that χ\chi is θ\theta-psh on XX.

Next we now show that

θ+ddcχγ(ϕv)θv.\displaystyle\theta+dd^{c}\chi\geq\gamma^{\prime}(\phi-v)\theta_{v}. (5.9)

For t>0t>0, let ϕt:=max(ϕ,Vθt)\phi^{t}:=\max(\phi,V_{\theta}-t), vt:=max(v,Vθt)v^{t}:=\max(v,V_{\theta}-t) and χt:=γ(ϕtvt)+ϕt\chi^{t}:=-\gamma({\phi^{t}-v_{t}})+\phi^{t}. Note that all of these potentials are locally bounded on Amp({θ})\textup{Amp}(\{\theta\}). As before, after slightly shrinking VV, the mollifications vjtv^{t}_{j} and ϕjt\phi^{t}_{j} are smooth approximants of vtv^{t} and ϕt\phi^{t} on VV. The same computations give that χjt:=γ(ϕjtvjt)+ϕjt\chi^{t}_{j}:=-\gamma({\phi^{t}_{j}-v^{t}_{j}})+\phi^{t}_{j} satisfies (5.8). By Proposition 2.2, letting jj\rightarrow\infty we obtain

θ+ddcχtγ(ϕtvt)(θ+ddcvt)inV.\displaystyle\theta+dd^{c}\chi^{t}\geq\gamma^{\prime}(\phi^{t}-v^{t})(\theta+dd^{c}v^{t})\quad{\rm in}\;V.

Let Ut={v>Vθt}={ϕ>Vθt}{v>Vθt}U_{t}=\{v>V_{\theta}-t\}=\{\phi>V_{\theta}-t\}\cap\{v>V_{\theta}-t\}. Using plurifine locality, we get that

𝟏Ut(θ+ddcχ)=𝟏Ut(θ+ddcχt)𝟏Utγ(ϕv)(θ+ddcv)inV.\displaystyle{\bf 1}_{U_{t}}(\theta+dd^{c}\chi)={\bf 1}_{U_{t}}(\theta+dd^{c}\chi_{t})\geq{\bf 1}_{U_{t}}\gamma^{\prime}(\phi-v)(\theta+dd^{c}v)\quad{\rm in}\;V.

Letting tt\to\infty, we arrive at (5.9).

Unfortunately (5.9) does not directly imply (5.7) since γ(ϕv)θv\gamma^{\prime}(\phi-v)\theta_{v} is not a closed form. However the above approximation process allows to conclude on the chart VV considered above.
Since χjt\chi^{t}_{j} and vjtv^{t}_{j} are smooth, from (5.9) we get that

(θ+ddcχjt)n(γ(ϕjtvjt))n(θ+ddcvjt)ninV.\displaystyle(\theta+dd^{c}\chi^{t}_{j})^{n}\geq(\gamma^{\prime}(\phi^{t}_{j}-v^{t}_{j}))^{n}(\theta+dd^{c}v^{t}_{j})^{n}\quad{\rm in}\;V.

We now conclude similarly. By Proposition 2.2, letting jj\rightarrow\infty we obtain

(θ+ddcχt)n(γ(ϕtvt))n(θ+ddcvt)ninV.\displaystyle(\theta+dd^{c}\chi^{t})^{n}\geq(\gamma^{\prime}(\phi^{t}-v^{t}))^{n}(\theta+dd^{c}v^{t})^{n}\quad{\rm in}\;V.

Let Ut={ϕ>Vθt}{v>Vθt}U_{t}=\{\phi>V_{\theta}-t\}\cap\{v>V_{\theta}-t\}. Using plurifine locality, we get that

𝟏Ut(θ+ddcχ)n𝟏Ut(γ(ϕv))n(θ+ddcv)ninV.\displaystyle{\bf 1}_{U_{t}}(\theta+dd^{c}\chi)^{n}\geq{\bf 1}_{U_{t}}(\gamma^{\prime}(\phi-v))^{n}(\theta+dd^{c}v)^{n}\quad{\rm in}\;V.

Letting tt\to\infty, we arrive at the conclusion. ∎

Theorem 5.20.

Let u(X,θ,ϕ)u\in\mathcal{E}(X,\theta,\phi) with supXu=0\sup_{X}u=0. If θun=fωn\theta_{u}^{n}=f\omega^{n} for some fLp(ωn),p>1,f\in L^{p}(\omega^{n}),\ p>1, then uu has the same singularity type as ϕ\phi. More precisely:

ϕC(fLp,p,ω,θ,Xθϕn)uϕ.\phi-C\Big{(}\|f\|_{L^{p}},p,\omega,\theta,\int_{X}\theta_{\phi}^{n}\Big{)}\leq u\leq\phi.
Proof.

To simplify the notation, we set μ=fωn\mu=f\omega^{n}.

A priori estimate. We assume at the moment that uϕu-\phi is bounded, hence

Tmax:=sup{t>0:μ(u<ϕt)>0}<.T_{\max}:=\sup\{t>0\;:\;\mu(u<\phi-t)>0\}<\infty.

Our goal is to establish a uniform bound on TmaxT_{\max}. Since fLp(ωn)f\in L^{p}(\omega^{n}) and PSH(X,ω)Lq(ωn){\rm PSH}(X,\omega)\subset L^{q}(\omega^{n}) for any q>0q>0, by Hölder inequality, PSH(X,ω)Lr(μ){\rm PSH}(X,\omega)\subset L^{r}(\mu), for any r>0r>0 and the quantity

Ar(μ):=sup{X(h)r𝑑μ:hPSH(X,ω),supXh=0}A_{r}(\mu):=\sup\left\{\int_{X}(-h)^{r}d\mu\;:\;h\in{\rm PSH}(X,\omega),\;\sup_{X}h=0\right\}

is finite and it depends on an upper bound for fp\|f\|_{p}.

By definition, uϕTmaxu\geq\phi-T_{\max} almost everywhere with respect to μ=θun\mu=\theta_{u}^{n}, hence everywhere by the domination principle (Theorem 3.12), providing the desired a priori bound. In particular |uϕ|Tmax|u-\phi|\leq T_{\max}.

We let χ:++\chi:\mathbb{R}^{+}\rightarrow\mathbb{R}^{+} denote a convex increasing function such that χ(0)=0\chi(0)=0 and χ(0)=1\chi^{\prime}(0)=1 (hence χ(t)1\chi^{\prime}(t)\geq 1). Let γ:++\gamma:\mathbb{R}^{+}\rightarrow\mathbb{R}^{+} denote the inverse function of χ\chi, which is concave and increasing. We set ψ=χ(ϕu)+ϕ\psi=-\chi(\phi-u)+\phi, v=Pθ(ψ)v=P_{\theta}(\psi), and observe that

φ:=γ(ϕv)+ϕγ(ϕψ)+ϕ=γ(χ(ϕu))+ϕu\varphi:=-\gamma(\phi-v)+\phi\leq-\gamma(\phi-\psi)+\phi=-\gamma(\chi(\phi-u))+\phi\leq u

with equality on the contact set {v=ψ}\{v=\psi\}. Since uu has the same singularity type as ϕ\phi, so does vv. In particular, by Proposition 3.2

Xθvn=Xθϕn=Xθun.\int_{X}\theta_{v}^{n}=\int_{X}\theta_{\phi}^{n}=\int_{X}\theta_{u}^{n}.

Observe that γ1\gamma^{\prime}\leq 1 since χ(γ(t))γ(t)=1\chi^{\prime}(\gamma(t))\gamma^{\prime}(t)=1 and that ϕu=γ(ϕv)\phi-u=\gamma(\phi-v) on {v=ψ}\{v=\psi\}. Lemmas 5.19 and 2.9 thus give

𝟏{v=ψ}(χ(ϕu))nθvn=𝟏{v=ψ}(γ(ϕv))nθvn𝟏{v=ψ}θφn𝟏{v=ψ}θunθun,{\bf 1}_{\{v=\psi\}}(\chi^{\prime}(\phi-u))^{-n}\theta_{v}^{n}={\bf 1}_{\{v=\psi\}}(\gamma^{\prime}(\phi-v))^{n}\theta_{v}^{n}\leq{\bf 1}_{\{v=\psi\}}\theta_{\varphi}^{n}\leq{\bf 1}_{\{v=\psi\}}\theta_{u}^{n}\leq\theta_{u}^{n},

hence 𝟏{v=ψ}θvn(χ(ϕu))nμ{\bf 1}_{\{v=\psi\}}\theta_{v}^{n}\leq(\chi^{\prime}(\phi-u))^{n}\mu. By Theorem 2.7, we can infer θvn(χ(ϕu))nμ\theta_{v}^{n}\leq(\chi^{\prime}(\phi-u))^{n}\mu.

Step 1: Controlling the energy of vv. The convexity of χ\chi and the normalization χ(0)=0\chi(0)=0 yields χ(t)tχ(t)\chi(t)\leq t\chi^{\prime}(t). Since v=χ(ϕu)+ϕv=-\chi(\phi-u)+\phi on {v=ψ}\{v=\psi\}, the above inequality, the convexity of χ\chi and Hölder inequality (with p=n+2p=n+2 and q=(n+2)/(n+1)q=(n+2)/(n+1)) yields

X(ϕv)θvn\displaystyle\int_{X}(\phi-v)\theta_{v}^{n} =Xχ(ϕu)θvnXχ(ϕu)(χ(ϕu))n𝑑μ\displaystyle=\int_{X}\chi(\phi-u)\theta_{v}^{n}\leq\int_{X}\chi(\phi-u)(\chi^{\prime}(\phi-u))^{n}d\mu
X(ϕu)(χ(ϕu))n+1𝑑μ\displaystyle\leq\int_{X}(\phi-u)(\chi^{\prime}(\phi-u))^{n+1}d\mu
(X(ϕu)n+2𝑑μ)1n+2(X(χ(ϕu))n+2𝑑μ)n+1n+2\displaystyle\leq\left(\int_{X}(\phi-u)^{n+2}d\mu\right)^{\frac{1}{n+2}}\left(\int_{X}(\chi^{\prime}(\phi-u))^{n+2}d\mu\right)^{\frac{n+1}{n+2}}
An+2(μ)1n+2.(X(χ(ϕu))n+2𝑑μ)n+1n+2.\displaystyle\leq A_{n+2}(\mu)^{\frac{1}{n+2}}.\left(\int_{X}(\chi^{\prime}(\phi-u))^{n+2}d\mu\right)^{\frac{n+1}{n+2}}.

Step 2: Controlling the norms uLm||u||_{L^{m}}. To simplify the notation we set m=n+3m=n+3. We are going to choose below the weight χ\chi in such a way that X(χ(ϕu))n+2𝑑μ2\int_{X}(\chi^{\prime}(\phi-u))^{n+2}d\mu\leq 2. This provides a uniform lower bound on supXv\sup_{X}v as we now explain. Indeed,

0(supXv)μ(X)=supX(vϕ)XθvnX(ϕv)θvn2An+2(μ)1n+2=C1,0\leq(-\sup_{X}v)\mu(X)=-\sup_{X}(v-\phi)\int_{X}\theta_{v}^{n}\leq\int_{X}(\phi-v)\theta_{v}^{n}\leq 2A_{n+2}(\mu)^{\frac{1}{n+2}}=C_{1},

where the first identity follows from Lemma 3.5. This yields supXvC1μ(X)1\sup_{X}v\geq-C_{1}\mu(X)^{-1}. We infer that vv belongs to a compact set of θ\theta-psh functions, hence its norm vLm(μ)||v||_{L^{m}(\mu)} is under control. Since vψv\leq\psi it follows that ϕvχ(ϕu)0\phi-v\geq\chi(\phi-u)\geq 0 and hence

X(χ(ϕu))m𝑑μ\displaystyle\int_{X}(\chi(\phi-u))^{m}d\mu X(ϕv)m𝑑μ\displaystyle\leq\int_{X}(\phi-v)^{m}d\mu
X|vsupXv+supXv|m𝑑μ\displaystyle\leq\int_{X}|v-\sup_{X}v+\sup_{X}v|^{m}d\mu
2m1X(|vsupXv|m+|supXv|m)𝑑μ\displaystyle\leq 2^{m-1}\int_{X}(|v-\sup_{X}v|^{m}+|\sup_{X}v|^{m})d\mu
2m1Am(μ)+2m1|supXv|mμ(X)C2.\displaystyle\leq 2^{m-1}A_{m}(\mu)+2^{m-1}|\sup_{X}v|^{m}\mu(X)\leq C_{2}.

Chebyshev’s inequality thus yields

μ(u<ϕt)1(χ(t))mX(χ(ϕu))m𝑑μC2(χ(t))m.{\mu}(u<\phi-t)\leq\frac{1}{(\chi(t))^{m}}\int_{X}(\chi(\phi-u))^{m}d\mu\leq\frac{C_{2}}{(\chi(t))^{m}}. (5.10)

Step 3: Choice of χ\chi. If g:++g:\mathbb{R}^{+}\rightarrow\mathbb{R}^{+} is strictly increasing with g(0)=1g(0)=1, Lebesgue’s formula gives

Xg(ϕu)𝑑μ=μ(X)+0Tmaxg(t)μ(u<ϕt)𝑑t.\int_{X}g(\phi-u)d\mu=\mu(X)+\int_{0}^{T_{\max}}g^{\prime}(t)\mu(u<\phi-t)dt. (5.11)

Fix 0<T0<Tmax0<T_{0}<T_{\max}. Setting g(t)=(χ(t))n+2g(t)=(\chi^{\prime}(t))^{n+2} we define χ\chi by imposing χ(0)=0\chi(0)=0, χ(0)=1\chi^{\prime}(0)=1, and

g(t)={1(1+t)2μ(u<ϕt),iftT01(1+t)2 ift>T0.g^{\prime}(t)=\begin{cases}\dfrac{1}{(1+t)^{2}{\mu}(u<\phi-t)},\;\text{if}\;t\leq T_{0}\\ \;\\ \frac{1}{(1+t)^{2}}\;\;\;\;\text{ if}\;t>T_{0}\end{cases}.

This choice guarantees that χ:++\chi:\mathbb{R}^{+}\rightarrow\mathbb{R}^{+} is convex increasing with χ1\chi^{\prime}\geq 1, and by (5.11) that

X(χ(ϕu))n+2𝑑μμ(X)+0dt(1+t)2=2.\int_{X}(\chi^{\prime}(\phi-u))^{n+2}d\mu\leq\mu(X)+\int_{0}^{\infty}\frac{dt}{(1+t)^{2}}=2.

Step 4: Conclusion. Observe that g(t)g(0)=1g(t)\geq g(0)=1, hence χ(t)=(g(t))1n+21\chi^{\prime}(t)=(g(t))^{\frac{1}{n+2}}\geq 1. This yields

χ(t)=0tχ(s)𝑑st.\chi(t)=\int_{0}^{t}\chi^{\prime}(s)ds\geq t. (5.12)

In particular χ(t)t\chi(t)\geq t. Together with (5.10), our choice of χ\chi yields, for all t[0,T0]t\in[0,T_{0}],

1(1+t)2g(t)=μ(u<ϕt)C2(χ(t))m.\frac{1}{(1+t)^{2}g^{\prime}(t)}={\mu}(u<\phi-t)\leq\frac{C_{2}}{(\chi(t))^{m}}.

This reads

(χ(t))mC2(1+t)2g(t)=(n+2)C2(1+t)2χ′′(t)(χ(t))n+1,t[0,T0].(\chi(t))^{m}\leq C_{2}(1+t)^{2}g^{\prime}(t)=(n+2)C_{2}(1+t)^{2}\chi^{\prime\prime}(t)(\chi^{\prime}(t))^{n+1},\quad\forall t\in[0,T_{0}].

Multiplying by χ1\chi^{\prime}\geq 1, integrating between 0 and tt, we get that for all t[0,T0]t\in[0,T_{0}]

(χ(t))m+1m+1\displaystyle\frac{(\chi(t))^{m+1}}{m+1} \displaystyle\leq (n+2)C20t(1+s)2χ′′(s)(χ(s))n+2𝑑s\displaystyle(n+2)C_{2}\int_{0}^{t}(1+s)^{2}\chi^{\prime\prime}(s)(\chi^{\prime}(s))^{n+2}\ ds
\displaystyle\leq (n+2)C20t[(1+s)2χ′′(s)(χ(s))n+2+2(1+s)((χ(s))n+31)]𝑑s\displaystyle(n+2)C_{2}\int_{0}^{t}[(1+s)^{2}\chi^{\prime\prime}(s)(\chi^{\prime}(s))^{n+2}+2(1+s)((\chi^{\prime}(s))^{n+3}-1)]\ ds
=\displaystyle= (n+2)C2(1+t)2n+3(1+s)2((χ(s))n+31)|s=0s=t\displaystyle\frac{(n+2)C_{2}(1+t)^{2}}{n+3}(1+s)^{2}\left((\chi^{\prime}(s))^{n+3}-1\right)\big{|}_{s=0}^{s=t}
\displaystyle\leq (n+2)C2(1+t)2n+3((χ(t))n+31)\displaystyle\frac{(n+2)C_{2}(1+t)^{2}}{n+3}\left((\chi^{\prime}(t))^{n+3}-1\right)
\displaystyle\leq C3(1+t)2(χ(t))n+3.\displaystyle C_{3}(1+t)^{2}(\chi^{\prime}(t))^{n+3}.

Recall that we choose m=n+3m=n+3 so that α:=m+1>β:=n+3>2.\alpha:=m+1>\beta:=n+3>2. The previous inequality then reads

(1+t)2βC4χ(t)χ(t)αβ.(1+t)^{-\frac{2}{\beta}}\leq C_{4}{\chi^{\prime}(t)}{\chi(t)^{-\frac{\alpha}{\beta}}}.

Since α>β>2\alpha>\beta>2 and χ(1)1\chi(1)\geq 1 (by (5.12)), integrating the above inequality between 11 and T0T_{0} we obtain

β2β(1+t)β2β|1T0C4βαβχ(t)α+ββ|1T0C4αββχ(1)α+ββC5.\frac{\beta-2}{\beta}(1+t)^{\frac{\beta-2}{\beta}}\big{|}_{1}^{T_{0}}\leq C_{4}\frac{\beta-\alpha}{\beta}\chi(t)^{\frac{-\alpha+\beta}{\beta}}\big{|}_{1}^{T_{0}}\leq C_{4}\frac{\alpha-\beta}{\beta}\chi(1)^{\frac{-\alpha+\beta}{\beta}}\leq C_{5}.

It then follows from the above that T0C6,T_{0}\leq C_{6}, for some uniform constant C6>0C_{6}>0. Since T0T_{0} was chosen arbitrarily in (0,Tmax)(0,T_{\max}) the uniform estimate for TmaxT_{\max} follows.

Relative boundedness of uu.

To finish the proof we finally prove that uϕu-\phi is bounded. We fix 1<q<p1<q<p and a constant ε>0\varepsilon>0 so small that eεhfLq(X)e^{-\varepsilon h}f\in L^{q}(X) for all hPSH(X,ω)h\in{\rm PSH}(X,\omega). For each jj we solve

uj(X,θ,ϕ),θujn=𝟏{u>ϕj}eε(ujmax(u,ϕj)μ.u_{j}\in\mathcal{E}(X,\theta,\phi),\;\theta_{u_{j}}^{n}={\bf 1}_{\{u>\phi-j\}}e^{\varepsilon(u_{j}-\max(u,\phi-j)}\mu.

Observe that for any fixed jj, max(u,ϕj)\max(u,\phi-j) is a subsolution of the above equation because

θmax(u,ϕj)n𝟏{u>ϕj}θun=𝟏{u>ϕj}eε(max(u,ϕj)max(u,ϕj))μ.\theta_{\max(u,\phi-j)}^{n}\geq{\bf 1}_{\{u>\phi-j\}}\theta_{u}^{n}={\bf 1}_{\{u>\phi-j\}}e^{\varepsilon(\max(u,\phi-j)-\max(u,\phi-j))}\mu.

Lemma 5.15 then gives ujmax(u,ϕj)u_{j}\geq\max(u,\phi-j), hence supXuj0\sup_{X}u_{j}\geq 0 and ujϕu_{j}-\phi is bounded. If j<kj<k, then

θukn=𝟏{u>ϕk}𝟏{u>ϕk}eε(ukmax(u,ϕk)μ𝟏{u>ϕj}eε(ukmax(u,ϕj)μ.\theta_{u_{k}}^{n}={\bf 1}_{\{u>\phi-k\}}{\bf 1}_{\{u>\phi-k\}}e^{\varepsilon(u_{k}-\max(u,\phi-k)}\mu\geq{\bf 1}_{\{u>\phi-j\}}e^{\varepsilon(u_{k}-\max(u,\phi-j)}\mu.

Invoking again Lemma 5.15 we obtain ujuku_{j}\geq u_{k}. The measures

μj:=𝟏{u>ϕj}eε(ujmax(u,ϕj)μ=fjωn\mu_{j}:={\bf 1}_{\{u>\phi-j\}}e^{\varepsilon(u_{j}-\max(u,\phi-j)}\mu=f_{j}\omega^{n}

have densities fjeεu1eεmax(u,ϕj)feεsupXu1eεmax(u,ϕj)ff_{j}\leq e^{\varepsilon u_{1}}e^{-\varepsilon\max(u,\phi-j)}f\leq e^{\varepsilon\sup_{X}u_{1}}e^{-\varepsilon\max(u,\phi-j)}f which are uniformly bounded in Lq(X)L^{q}(X). By the above a priori estimate, we have a uniform bound ujϕCu_{j}\geq\phi-C, hence v:=limjujv:=\lim_{j\to\infty}u_{j} satisfies vϕCv\geq\phi-C and θvn=eε(vu)μ\theta_{v}^{n}=e^{\varepsilon(v-u)}\mu. Since θvn=eε(uu)μ\theta_{v}^{n}=e^{\varepsilon(u-u)}\mu, Lemma 5.15 ensures u=vu=v. Thus uϕu-\phi is bounded as desired. This finishes the proof. ∎

Corollary 5.21.

If λ>0\lambda>0 and u(X,θ,ϕ)u\in\mathcal{E}(X,\theta,\phi), θun=eλufωn\theta_{u}^{n}=e^{\lambda u}f\omega^{n} for some fLp(ωn),p>1,f\in L^{p}(\omega^{n}),\ p>1, then uu has the same singularity type as ϕ\phi.

Proof.

Since uu is bounded from above on XX and λ>0\lambda>0 it follows that eλufLp(X,ωn)e^{\lambda u}f\in L^{p}(X,\omega^{n}), p>1p>1. The result follows from Theorem 5.20. ∎

5.4 Naturality of model type singularities and examples

One may still wonder if our choice of model potentials is a natural one in our pursuit of complex Monge–Ampère equations with prescribed singularity. We address these doubts in the next well posedness result.

Theorem 5.22.

Suppose that ψPSH(X,θ)\psi\in\textup{PSH}(X,\theta) and the equation

θun=fωn\theta_{u}^{n}=f\omega^{n}

has a solution uPSH(X,θ)u\in\textup{PSH}(X,\theta) with the same singularity type as ψ\psi, for all fL(X),f0f\in L^{\infty}(X),\ f\geq 0 satisfying Xθψn=Xfωn>0\int_{X}\theta_{\psi}^{n}=\int_{X}f\omega^{n}>0. Then ψ\psi has model type singularity.

Proof.

Suppose that [ψ][\psi] is not of model type. Then Pθ[ψ]P_{\theta}[\psi] is strictly less singular than ψ\psi. On the other hand (X,θ,ψ)(X,θ,Pθ[ψ])\mathcal{E}(X,\theta,\psi)\subset\mathcal{E}(X,\theta,P_{\theta}[\psi]) as Xθψn=XθPθ[ψ]n\int_{X}\theta_{\psi}^{n}=\int_{X}\theta_{P_{\theta}[\psi]}^{n}.

By Theorem 3.6, there exists gLg\in L^{\infty} such that θPθ[ψ]n=gωn.\theta^{n}_{P_{\theta}[\psi]}=g\omega^{n}. By uniqueness (Theorem 3.13), Pθ[ψ]P_{\theta}[\psi] is the only solution of this last equation inside (X,θ,Pθ[ψ])\mathcal{E}(X,\theta,P_{\theta}[\psi]).

Since (X,θ,ψ)(X,θ,Pθ[ψ])\mathcal{E}(X,\theta,\psi)\subset\mathcal{E}(X,\theta,P_{\theta}[\psi]), but Pθ[ψ](X,θ,ψ)P_{\theta}[\psi]\notin\mathcal{E}(X,\theta,\psi), we get that θun=gωn\theta_{u}^{n}=g\omega^{n} cannot have any solution that has the same singularity type as ψ\psi. ∎

Next we point out a simple way to construct potential with model singularity types:

Proposition 5.23.

Suppose that ψPSH(X,θ)\psi\in\textup{PSH}(X,\theta) and θψn=fωn\theta_{\psi}^{n}=f\omega^{n} for some fLp(ωn),p>1f\in L^{p}(\omega^{n}),\ p>1 with Xfωn>0\int_{X}f\omega^{n}>0. Then ψ\psi has model type singularity.

Proof.

We first observe that ψ(X,θ,Pθ[ψ])\psi\in\mathcal{E}(X,\theta,P_{\theta}[\psi]). Since θψn\theta_{\psi}^{n} has LpL^{p} density with p>1p>1, it thus follows from Theorem 5.20 that ψPθ[ψ]\psi-P_{\theta}[\psi] is bounded on XX, hence [ψ]=[Pθ[ψ]][\psi]=[P_{\theta}[\psi]], implying that ψ\psi has model singularity type. ∎

As noticed in [RW14] (see [RS05] for the local case), all analytic singularity types are of model type.

Proposition 5.24.

Suppose ψPSH(X,θ)\psi\in\textup{PSH}(X,\theta) has analytic singularity type, i.e. ψ\psi can be locally written as c2log(j|fj|2)+g\frac{c}{2}\log\big{(}\sum_{j}|f_{j}|^{2}\big{)}+g, where fjf_{j} are holomorphic, c>0c>0 and gg is bounded. Then [ψ][\psi] is of model type.

We give an argument that is different from the elementary one given in [DDL18a, Section 4.5]. Following [RW14], we use resolution of singularities, and get a slightly more general result (with gg being bounded instead of smooth).

Proof.

Let \mathcal{I} be the coherent sheaf of holomorphic functions ff satisfying |f|Ae2ψc|f|\leq Ae^{\frac{2\psi}{c}} for some A>0A>0 and π:YX\pi:Y\to X be a resolution of singularities of \mathcal{I}, as in [Dem12, Remark 5.9]. We obtain that the singularity type of ψπ\psi\circ\pi is modelled by an snc divisor D=jαjDjYD=\sum_{j}\alpha_{j}D_{j}\subset Y, where αj>0\alpha_{j}>0. That is to say, the Lelong numbers of ψπ\psi\circ\pi along each DjD_{j} is αj\alpha_{j}, i.e., ψπ\psi\circ\pi has αjlog|z|\alpha_{j}\log|z| type poles along each DjD_{j}.

By Lemma 5.1 we know that the Lelong numbers of Pπθ[ψπ]P_{\pi^{*}\theta}[\psi\circ\pi] and ψπ\psi\circ\pi are the same. In particular, we obtain that [Pπθ[ψπ]]=[ψπ][P_{\pi^{*}\theta}[\psi\circ\pi]]=[\psi\circ\pi]. Since any πθ\pi^{*}\theta-psh function can be written as uπu\circ\pi, for some θ\theta-psh function uu, it follows that for any C>0C>0, Pπθ(ψπ+C,0)=Pθ(ψ+C,0)πP_{\pi^{*}\theta}(\psi\circ\pi+C,0)=P_{\theta}(\psi+C,0)\circ\pi. This means that Pπθ[ψπ]=Pθ[ψ]πP_{\pi^{*}\theta}[\psi\circ\pi]=P_{\theta}[\psi]\circ\pi. Combining the above and pushing forward to XX we obtain that [Pθ[ψ]]=[ψ][P_{\theta}[\psi]]=[\psi], as desired. ∎

Chapter 6 The finite energy range of the complex Monge–Ampère operator

In this chapter we give a characterization of the Borel measures μ\mu that are equal to the complex Monge–Ampere measure of some uχ(X,θ,ϕ)u\in\mathcal{E}_{\chi}(X,\theta,\phi), with χ\chi having polynomial growth, and ϕ\phi a model potential (ϕ=Pθ[ϕ]\phi=P_{\theta}[\phi]) with Xθϕn>0\int_{X}\theta_{\phi}^{n}>0.

Before we can achieve this we need to develop more potential theory. A weight is a continuous increasing function χ:[0,)[0,)\chi:[0,\infty)\rightarrow[0,\infty) such that χ(0)=0\chi(0)=0 and χ()=\chi(\infty)=\infty. Denote by χ1\chi^{-1} its inverse function, i.e. such that χ(χ1(t))=t\chi(\chi^{-1}(t))=t for all t0t\geq 0.

Throughout this section we assume that the weight χ\chi satisfies the following condition

t0,λ1,χ(λt)λMχ(t),\forall t\geq 0,\;\forall\lambda\geq 1,\;\chi(\lambda t)\leq\lambda^{M}\chi(t), (6.1)

where M1M\geq 1 is a fixed constant. Observe that from (6.1) it follows that

t0,γ<1,χ(γt)γMχ(t).\forall t\geq 0,\;\forall\gamma<1,\;\chi(\gamma t)\geq\gamma^{M}\chi(t). (6.2)

We fix ϕ\phi a model potential and we let χ(X,θ,ϕ)\mathcal{E}_{\chi}(X,\theta,\phi) denote the set of all u(X,θ,ϕ)u\in\mathcal{E}(X,\theta,\phi) such that

Eχ(u,ϕ):=Xχ(|uϕ|)θun<.E_{\chi}(u,\phi):=\int_{X}\chi(|u-\phi|)\theta_{u}^{n}<\infty.

When ϕ=Vθ\phi=V_{\theta}, we denote (X,θ)=(X,θ,Vθ)\mathcal{E}(X,\theta)=\mathcal{E}(X,\theta,V_{\theta}), χ(X,θ)=χ(X,θ,Vθ\mathcal{E}_{\chi}(X,\theta)=\mathcal{E}_{\chi}(X,\theta,V_{\theta}) and Eχ(u)=Eχ(u,Vθ)E_{\chi}(u)=E_{\chi}(u,V_{\theta}). Compared to [GZ07], we have changed the sign of the weight, but the weighted classes are the same. The following simple observation shows that the class (X,θ,ϕ)\mathcal{E}(X,\theta,\phi) is stable under adding a constant.

Lemma 6.1.

If uu belongs to χ(X,θ,ϕ)\mathcal{E}_{\chi}(X,\theta,\phi) then so does u+Cu+C for any constant CC.

Proof.

Since χ\chi satisfies (6.1), for t>0t>0 and s>0s>0 we have

χ(t+s)=χ(2(t+s)/2)2Mχ((t+s)/2)2Mχ(max(t,s))2Mmax(χ(t),χ(s)),\chi(t+s)=\chi(2(t+s)/2)\leq 2^{M}\chi((t+s)/2)\leq 2^{M}\chi(\max(t,s))\leq 2^{M}\max(\chi(t),\chi(s)),

where the last two inequalities follow from the fact that χ\chi is increasing. Then

Xχ(|u+Cϕ|)θun\displaystyle\int_{X}\chi(|u+C-\phi|)\theta_{u}^{n} Xχ(|uϕ|+|C|)θun\displaystyle\leq\int_{X}\chi(|u-\phi|+|C|)\theta_{u}^{n}
2MXmax(χ(|uϕ|),χ(|C|))θun\displaystyle\leq 2^{M}\int_{X}\max(\chi(|u-\phi|),\chi(|C|))\theta_{u}^{n}
2Mmax(Xχ(|uϕ|)θun,χ(|C|)Xθun)<.\displaystyle\leq 2^{M}\max\left(\int_{X}\chi(|u-\phi|)\theta_{u}^{n},\ \chi(|C|)\int_{X}\theta_{u}^{n}\right)<\infty.

Next we prove a few of technical results that will be often used.

Lemma 6.2.

There exists a uniform constant C>0C>0 such that for all uPSH(X,θ,ϕ)u\in{\rm PSH}(X,\theta,\phi) normalized with supXu=0\sup_{X}u=0 we have

Xχ(ϕu)θϕnC.\int_{X}\chi(\phi-u)\theta_{\phi}^{n}\leq C.
Proof.

We recall that by Theorem 3.6, θϕn𝟏{ϕ=0}θnAωn\theta_{\phi}^{n}\leq{\bf 1}_{\{\phi=0\}}\theta^{n}\leq A\omega^{n}, for some A>0A>0. Also, thanks to (6.1), if ϕu1\phi-u\geq 1 we have χ(ϕu)(ϕu)Mχ(1)\chi(\phi-u)\leq(\phi-u)^{M}\chi(1); otherwise (since χ\chi is increasing) we have χ(ϕu)χ(1)\chi(\phi-u)\leq\chi(1). In both cases we can infer that

Xχ(ϕu)θϕnCX(|ϕ|M+|u|M+1)ωn.\int_{X}\chi(\phi-u)\theta_{\phi}^{n}\leq C^{\prime}\int_{X}(|\phi|^{M}+|u|^{M}+1)\omega^{n}.

The conclusion then follows from the fact that X|h|Mωn\int_{X}|h|^{M}\omega^{n} is uniformly bounded for hPSH(X,θ)h\in{\rm PSH}(X,\theta) with supXh=0\sup_{X}h=0. ∎

Lemma 6.3.

Let u(X,θ,ϕ)u\in\mathcal{E}(X,\theta,\phi) with supXu=0\sup_{X}u=0. Then, for any j{1,,n}j\in\{1,...,n\}

Xχ(ϕu)θujθϕnjXχ(ϕu)θun.\int_{X}\chi(\phi-u)\theta_{u}^{j}\wedge\theta_{\phi}^{n-j}\leq\int_{X}\chi(\phi-u)\theta_{u}^{n}.
Proof.

By Lemma 3.5 we have that uϕ0u\leq\phi\leq 0. By the partial comparison principle (Proposition 3.22) we have that for any j{1,,n}j\in\{1,...,n\}

Xχ(ϕu)θujθϕnj\displaystyle\int_{X}\chi(\phi-u)\theta_{u}^{j}\wedge\theta_{\phi}^{n-j} =0θujθϕnj(u<ϕχ1(t))dt\displaystyle=\int_{0}^{\infty}\theta_{u}^{j}\wedge\theta_{\phi}^{n-j}(u<\phi-\chi^{-1}(t))dt
0θun(u<ϕχ1(t))𝑑t=Xχ(ϕu)θun.\displaystyle\leq\int_{0}^{\infty}\theta_{u}^{n}(u<\phi-\chi^{-1}(t))dt=\int_{X}\chi(\phi-u)\theta_{u}^{n}.

Lemmas 6.4, 6.5 below are essentially known by [GZ07], but we repeat the simple proof for the reader’s convenience. Similar simplifications can also be found in [Gup22].

Lemma 6.4.

If u,v(X,θ,ϕ)u,v\in\mathcal{E}(X,\theta,\phi) and u,v0u,v\leq 0, then

Xχ(ϕu)θvn2n+MEχ(u,ϕ)+Eχ(v,ϕ).\int_{X}\chi(\phi-u)\theta_{v}^{n}\leq 2^{n+M}E_{\chi}(u,\phi)+E_{\chi}(v,\phi).
Proof.

By Lemma 3.5 we have that uϕ0u\leq\phi\leq 0. By the comparison principle, Corollary 3.23, we have

Xχ(ϕu)θvn\displaystyle\int_{X}\chi(\phi-u)\theta_{v}^{n} =0θvn(u<ϕχ1(t))𝑑t\displaystyle=\int_{0}^{\infty}\theta_{v}^{n}(u<\phi-\chi^{-1}(t))dt
0θvn(2u<v+ϕχ1(t))𝑑t+0θvn(v<ϕχ1(t))𝑑t\displaystyle\leq\int_{0}^{\infty}\theta_{v}^{n}(2u<v+\phi-\chi^{-1}(t))dt+\int_{0}^{\infty}\theta_{v}^{n}(v<\phi-\chi^{-1}(t))dt
2n0(θ+ddcv+ϕ2)n(u<v+ϕχ1(t)2)𝑑t+Eχ(v,ϕ)\displaystyle\leq 2^{n}\int_{0}^{\infty}\left(\theta+dd^{c}{\frac{v+\phi}{2}}\right)^{n}\left(u<\frac{v+\phi-\chi^{-1}(t)}{2}\right)dt+E_{\chi}(v,\phi)
2n0θun(2u<2ϕχ1(t))𝑑t+Eχ(v,ϕ)\displaystyle\leq 2^{n}\int_{0}^{\infty}\theta_{u}^{n}(2u<2\phi-\chi^{-1}(t))dt+E_{\chi}(v,\phi)
=2nXχ(2ϕ2u)θun+Eχ(v,ϕ)\displaystyle=2^{n}\int_{X}\chi(2\phi-2u)\theta_{u}^{n}+E_{\chi}(v,\phi)
2n+MEχ(u,ϕ)+Eχ(v,ϕ),\displaystyle\leq 2^{n+M}E_{\chi}(u,\phi)+E_{\chi}(v,\phi),

where in the second line we have used the trivial inclusion

{2uv+ϕχ1(t)}{vϕχ1(t)}{uϕχ1(t)},\{2u\geq v+\phi-\chi^{-1}(t)\}\cap\{v\geq\phi-\chi^{-1}(t)\}\subseteq\{u\geq\phi-\chi^{-1}(t)\},

in the third one we have used

θvn2n(θ+ddcv+ϕ2)n,\theta_{v}^{n}\leq 2^{n}\left(\theta+dd^{c}\frac{v+\phi}{2}\right)^{n},

and in the last line we have used (6.1) with λ=2\lambda=2. ∎

Lemma 6.5.

For all u,v(X,θ,ϕ)u,v\in\mathcal{E}(X,\theta,\phi) with uv0u\leq v\leq 0 we have

Xχ(ϕv)θvnXχ(ϕu)θvn2n+MEχ(u,ϕ).\int_{X}\chi(\phi-v)\,\theta_{v}^{n}\leq\int_{X}\chi(\phi-u)\,\theta_{v}^{n}\leq 2^{n+M}E_{\chi}(u,\phi).
Proof.

By the comparison principle, Corollary 3.23, and the fact that uvϕu\leq v\leq\phi, we have

Xχ(ϕu)θvn\displaystyle\int_{X}\chi(\phi-u)\,\theta_{v}^{n} =0θvn(u<ϕχ1(t))𝑑t\displaystyle=\int_{0}^{\infty}\theta_{v}^{n}(u<\phi-\chi^{-1}(t))dt
0θvn(2u<v+ϕχ1(t))𝑑t\displaystyle\leq\int_{0}^{\infty}\theta_{v}^{n}(2u<v+\phi-\chi^{-1}(t))dt
2n0θv+ϕ2n(u<v+ϕχ1(t)2)𝑑t\displaystyle\leq 2^{n}\int_{0}^{\infty}\theta_{\frac{v+\phi}{2}}^{n}\left(u<\frac{v+\phi-\chi^{-1}(t)}{2}\right)dt
2n0θun(2u<2ϕχ1(t))𝑑t\displaystyle\leq 2^{n}\int_{0}^{\infty}\theta_{u}^{n}(2u<2\phi-\chi^{-1}(t))dt
=2nXχ(2ϕ2u)θun2n+MEχ(u,ϕ).\displaystyle=2^{n}\int_{X}\chi(2\phi-2u)\theta_{u}^{n}\leq 2^{n+M}E_{\chi}(u,\phi).

Proposition 6.6.

If uχ(X,θ,ϕ)u\in\mathcal{E}_{\chi}(X,\theta,\phi) and uvu\leq v then vχ(X,θ,ϕ)v\in\mathcal{E}_{\chi}(X,\theta,\phi). Moreover, the class χ(X,θ,ϕ)\mathcal{E}_{\chi}(X,\theta,\phi) is convex.

Proof.

In view of Lemma 6.1 we can assume that v0v\leq 0. Lemma 6.5 then yields

Eχ(v,ϕ)=Xχ(ϕv)θvn2n+MEχ(u,ϕ).E_{\chi}(v,\phi)=\int_{X}\chi(\phi-v)\,\theta_{v}^{n}\leq 2^{n+M}E_{\chi}(u,\phi).

We next prove that the class χ(X,θ,ϕ)\mathcal{E}_{\chi}(X,\theta,\phi) is convex. Assume uu and vv are in χ(X,θ,ϕ)\mathcal{E}_{\chi}(X,\theta,\phi). It follows from Corollary 3.19 that w:=Pθ(u,v)w:=P_{\theta}(u,v) belongs to (X,θ,ϕ)\mathcal{E}(X,\theta,\phi). From Theorem 3.6 we also have

Xχ(ϕw)θwnXχ(ϕu)θun+Xχ(ϕv)θvn<,\int_{X}\chi(\phi-w)\theta_{w}^{n}\leq\int_{X}\chi(\phi-u)\theta_{u}^{n}+\int_{X}\chi(\phi-v)\theta_{v}^{n}<\infty,

hence wχ(X,θ,ϕ)w\in\mathcal{E}_{\chi}(X,\theta,\phi). For t[0,1]t\in[0,1], since wtu+(1t)vw\leq tu+(1-t)v, it follows from Lemma 6.5 that tu+(1t)vχ(X,θ,ϕ)tu+(1-t)v\in\mathcal{E}_{\chi}(X,\theta,\phi). ∎

Lemma 6.7.

Assume (uj)(u_{j}) is a sequence in χ(X,θ,ϕ)\mathcal{E}_{\chi}(X,\theta,\phi) converging in L1L^{1} to uPSH(X,θ,ϕ)u\in{\rm PSH}(X,\theta,\phi). If supjEχ(uj,ϕ)<\sup_{j}E_{\chi}(u_{j},\phi)<\infty, then uχ(X,θ,ϕ)u\in\mathcal{E}_{\chi}(X,\theta,\phi).

Proof.

We can assume uj0u_{j}\leq 0 for all jj. Define vk:=(supjkuj)PSH(X,θ,ϕ)v_{k}:=(\sup_{j\geq k}u_{j})^{*}\in{\rm PSH}(X,\theta,\phi). Then vkuv_{k}\searrow u and uu is more singular than ϕ\phi. By Theorem 3.3 we then have XθunXθϕn\int_{X}\theta_{u}^{n}\leq\int_{X}\theta_{\phi}^{n}.
Since 0vkuk0\geq v_{k}\geq u_{k}, we have χ(ϕvk)χ(ϕuk)\chi(\phi-v_{k})\leq\chi(\phi-u_{k}). Lemma 6.5 then ensures that Eχ(vk,ϕ)2n+MEχ(uk,ϕ)E_{\chi}(v_{k},\phi)\leq 2^{n+M}E_{\chi}(u_{k},\phi) and the latter quantity is uniformly bounded by assumption.

Fix t>0t>0 and consider vk,t:=max(vk,ϕt)v_{k,t}:=\max(v_{k},\phi-t). Note that vk,tmax(u,ϕt)v_{k,t}\searrow\max(u,\phi-t) as kk\to\infty and that by Lemma 6.5 we have Eχ(vk,t,ϕ)2n+MEχ(vk,ϕ)E_{\chi}(v_{k,t},\phi)\leq 2^{n+M}E_{\chi}(v_{k},\phi). This means that Eχ(vk,t,ϕ)E_{\chi}(v_{k,t},\phi) has a uniform upper bound. Moreover, the functions χ(ϕvk,t)\chi(\phi-v_{k,t}) are quasi-continuous and uniformly bounded on XX. It thus follows from Theorem 2.6 that

lim infkXχ(ϕvk,t)(θ+ddcvk,t)nXχ(ϕmax(u,ϕt))(θ+ddcmax(u,ϕt))n.\liminf_{k\to\infty}\int_{X}\chi(\phi-v_{k,t})(\theta+dd^{c}v_{k,t})^{n}\geq\int_{X}\chi(\phi-\max(u,\phi-t))(\theta+dd^{c}\max(u,\phi-t))^{n}.

The above estimate then gives a uniform upper bound for Eχ(max(u,ϕt),ϕ)E_{\chi}(\max(u,\phi-t),\phi). We can thus infer that there exists a constant C>0C>0 such that

CXχ(ϕmax(u,ϕt))θmax(u,ϕt)nχ(t){uϕt}θmax(u,ϕt)n.C\geq\int_{X}\chi(\phi-\max(u,\phi-t))\theta_{\max(u,\phi-t)}^{n}\geq\chi(t)\int_{\{u\leq\phi-t\}}\theta_{\max(u,\phi-t)}^{n}.

Letting tt\to\infty yields that {uϕt}θmax(u,ϕt)n0\int_{\{u\leq\phi-t\}}\theta_{\max(u,\phi-t)}^{n}\rightarrow 0 (since χ(t)\chi(t)\rightarrow\infty). Hence, using the plurifine property of the Monge–Ampère operator and the fact that uu is more singular than ϕ\phi, we get

Xθϕn=limtXθmax(u,ϕt)n=limt({uϕt}θmax(u,ϕt)n+{u>ϕt}θun)XθunXθϕn.\int_{X}\theta_{\phi}^{n}=\lim_{t\to\infty}\int_{X}\theta_{\max(u,\phi-t)}^{n}=\lim_{t\to\infty}\bigg{(}\int_{\{u\leq\phi-t\}}\theta_{\max(u,\phi-t)}^{n}+\int_{\{u>\phi-t\}}\theta_{u}^{n}\bigg{)}\leq\int_{X}\theta_{u}^{n}\leq\int_{X}\theta_{\phi}^{n}.

This means that u(X,θ,ϕ)u\in\mathcal{E}(X,\theta,\phi). Using plurifine locality, we also obtain

{u>ϕt}χ(ϕu)θunC,\int_{\{u>\phi-t\}}\chi(\phi-u)\theta_{u}^{n}\leq C,

and letting tt\to\infty, we see that Xχ(ϕu)θunC\int_{X}\chi(\phi-u)\theta_{u}^{n}\leq C. Hence uχ(X,θ,ϕ)u\in\mathcal{E}_{\chi}(X,\theta,\phi). ∎

Lemma 6.8.

Let μ\mu be a positive Borel measure on XX. Assume that μ({ϕ=})=0\mu(\{\phi=-\infty\})=0 and χ(ϕu)L1(μ)\chi(\phi-u)\in L^{1}(\mu) for all uχ(X,θ,ϕ)u\in\mathcal{E}_{\chi}(X,\theta,\phi) and fix a constant A>0A>0. Then there exists a constant C=C(A)>0C=C(A)>0 such that, for all uχ(X,θ,ϕ)u\in\mathcal{E}_{\chi}(X,\theta,\phi) with supXu=0\sup_{X}u=0 and Eχ(u,ϕ)AE_{\chi}(u,\phi)\leq A we have

Xχ(ϕu)𝑑μC.\int_{X}\chi(\phi-u)d\mu\leq C.

Observe that the condition μ({ϕ=})=0\mu(\{\phi=-\infty\})=0 is necessary. Without it, μ\mu-integrability of χ(ϕu)\chi(\phi-u) can not be discussed, as the values of this function are not defined on the set {ϕ=}\{\phi=-\infty\}.

Proof.

Assume by contradiction that there exists a sequence (uj)χ(X,θ,ϕ)(u_{j})\subset\mathcal{E}_{\chi}(X,\theta,\phi) satisfying supXuj=0\sup_{X}u_{j}=0 and Eχ(uj,ϕ)AE_{\chi}(u_{j},\phi)\leq A such that

Xχ(ϕuj)𝑑μ4jM,\int_{X}\chi(\phi-u_{j})\,d\mu\geq 4^{jM},

where MM is the exponent appearing in (6.1).
Define vk:=Pθ(min1jk(2juj+(12j)ϕ))ϕv_{k}:=P_{\theta}(\min_{1\leq j\leq k}(2^{-j}u_{j}+(1-2^{-j})\phi))\leq\phi. Observe that vkv_{k} is decreasing and setting

v:=limkvk,ψ:=j12juj,v:=\lim_{k}v_{k},\quad\psi:=\sum_{j\geq 1}2^{-j}u_{j},

then, since ujϕu_{j}\leq\phi, we can check that ψ2juj+(12j)ϕϕ\psi\leq 2^{-j}u_{j}+(1-2^{-j})\phi\leq\phi. It then follows that vψPSH(X,θ,ϕ)v\geq\psi\in{\rm PSH}(X,\theta,\phi). We fix k2k\geq 2 and for 1lk1\leq l\leq k we set

D1:={vk=21u1+21ϕ}D_{1}:=\left\{v_{k}=2^{-1}u_{1}+2^{-1}\phi\right\}

and

Dl:={vk=2lul+(12l)ϕ<min1pl1(2pup+(12p)ϕ)}for2D_{l}:=\left\{v_{k}=2^{-l}u_{l}+(1-2^{-l})\phi<\min_{1\leq p\leq l-1}(2^{-p}u_{p}+(1-2^{-p})\phi)\right\}\quad{\rm for}\;\ell\geq 2

Note that the sets DlD_{l} are pairwise disjoint and

kl1Dl={vk=min1jk(2juj+(12j)ϕ)}.\bigcup_{k\geq l\geq 1}D_{l}=\{v_{k}=\min_{1\leq j\leq k}(2^{-j}u_{j}+(1-2^{-j})\phi)\}.

It follows from Theorem 2.7 and Lemma 2.9 that

Xχ(ϕvk)θvknl=1kDlχ(2l(ϕul))(2lθul+(12l)θϕ)n.\int_{X}\chi(\phi-v_{k})\theta_{v_{k}}^{n}\leq\sum_{l=1}^{k}\int_{D_{l}}\chi(2^{-l}(\phi-u_{l}))(2^{-l}\theta_{u_{l}}+(1-2^{-l})\theta_{\phi})^{n}.

We note that

(2lθul+(12l)θϕ)nθϕn+2lp=1n(np)θulpθϕnp.(2^{-l}\theta_{u_{l}}+(1-2^{-l})\theta_{\phi})^{n}\leq\theta_{\phi}^{n}+2^{-l}\sum_{p=1}^{n}\binom{n}{p}\theta_{u_{l}}^{p}\wedge\theta_{\phi}^{n-p}. (6.3)

Since the sets DlD_{l} are pairwise disjoint and χ\chi is increasing, it then follows that

Xχ(ϕvk)θvkn\displaystyle\int_{X}\chi(\phi-v_{k})\theta_{v_{k}}^{n} \displaystyle\leq l=1kDlχ(2l(ϕul))(2lθul+(12l)θϕ)n\displaystyle\sum_{l=1}^{k}\int_{D_{l}}\chi(2^{-l}(\phi-u_{l}))(2^{-l}\theta_{u_{l}}+(1-2^{-l})\theta_{\phi})^{n}
\displaystyle\leq l=1kDlχ(2l(ϕul))θϕn+2nl=1k2lXχ(ϕul)θuln\displaystyle\sum_{l=1}^{k}\int_{D_{l}}\chi(2^{-l}(\phi-u_{l}))\theta_{\phi}^{n}+2^{n}\sum_{l=1}^{k}2^{-l}\int_{X}\chi(\phi-u_{l})\theta_{u_{l}}^{n}
\displaystyle\leq Xχ(ϕψ)θϕn+2nl=1k2lXχ(ϕul)θuln\displaystyle\int_{X}\chi(\phi-\psi)\theta_{\phi}^{n}+2^{n}\sum_{l=1}^{k}2^{-l}\int_{X}\chi(\phi-u_{l})\theta_{u_{l}}^{n}
\displaystyle\leq C2+2nA.\displaystyle C_{2}+2^{n}A.

The second inequality follows from (6.3) together with Lemma 6.3. The third inequality follows from the fact 02l(ϕul)l12l(ϕul)=ϕψ0\leq 2^{-l}(\phi-u_{l})\leq\sum_{l\geq 1}2^{-l}(\phi-u_{l})=\phi-\psi. The last line follows from Lemma 6.2.
It then follows from Lemma 6.7 that vv, which is the decreasing limit of vkv_{k}, belongs to χ(X,θ,ϕ)\mathcal{E}_{\chi}(X,\theta,\phi). On the other hand, since v2kuk+(12k)ϕv\leq 2^{-k}u_{k}+(1-2^{-k})\phi for any k1k\geq 1, and thanks to (6.2) we have

Xχ(ϕv)𝑑μXχ(2k(ϕuk))𝑑μ2kMXχ(ϕuk)𝑑μ2kM,\int_{X}\chi(\phi-v)d\mu\geq\int_{X}\chi(2^{-k}(\phi-u_{k}))d\mu\geq 2^{-kM}\int_{X}\chi(\phi-u_{k})d\mu\geq 2^{kM}\to\infty,

as kk\to\infty, contradicting the assumption. ∎

Lemma 6.9.

Assume μ\mu is a positive Borel measure satisfying μ({ϕ=})=0\mu(\{\phi=-\infty\})=0 and χ(ϕu)L1(μ)\chi(\phi-u)\in L^{1}(\mu) for all uχ(X,θ,ϕ)u\in\mathcal{E}_{\chi}(X,\theta,\phi). Then there exists a constant C>0C>0 such that, for all uχ(X,θ,ϕ)u\in\mathcal{E}_{\chi}(X,\theta,\phi) with supXu=0\sup_{X}u=0, we have

Xχ(ϕu)𝑑μC(Eχ(u,ϕ)+1).\int_{X}\chi(\phi-u)d\mu\leq C(E_{\chi}(u,\phi)+1).
Proof.

We prove the lemma by contradiction, assuming there exists a sequence (uj)χ(X,θ,ϕ)(u_{j})\subset\mathcal{E}_{\chi}(X,\theta,\phi) with supXuj=0\sup_{X}u_{j}=0, but

Xχ(ϕuj)𝑑μj(Eχ(uj,ϕ)+1).\int_{X}\chi(\phi-u_{j})d\mu\geq j(E_{\chi}(u_{j},\phi)+1).

Set εj:=(Eχ(uj,ϕ)+1)1<1\varepsilon_{j}:=(E_{\chi}(u_{j},\phi)+1)^{-1}<1 and

ψj:=Pθ(χ1(εjχ(ϕuj))+ϕ).\psi_{j}:=P_{\theta}(-\chi^{-1}(\varepsilon_{j}\chi(\phi-u_{j}))+\phi).

Since θψjn\theta_{\psi_{j}}^{n} is supported on {ψj=χ1(εjχ(ϕuj))+ϕ}\{\psi_{j}=-\chi^{-1}(\varepsilon_{j}\chi(\phi-u_{j}))+\phi\} (Theorem 2.7), we have

Xχ(ϕψj)θψjnεjXχ(ϕuj)θψjnC1.\int_{X}\chi(\phi-\psi_{j})\theta_{\psi_{j}}^{n}\leq\varepsilon_{j}\int_{X}\chi(\phi-u_{j})\theta_{\psi_{j}}^{n}\leq C_{1}.

In the last inequality we have used the fact that ujψju_{j}\leq\psi_{j} and that χ\chi is increasing and Lemma 6.5. By Lemma 6.8 we thus have

Xχ(ϕψj)𝑑μC2.\int_{X}\chi(\phi-\psi_{j})d\mu\leq C_{2}.

But since ϕψjχ1(εjχ(ϕuj))\phi-\psi_{j}\geq\chi^{-1}(\varepsilon_{j}\chi(\phi-u_{j})), we have

Xχ(ϕψj)𝑑μXεjχ(ϕuj)𝑑μj,\int_{X}\chi(\phi-\psi_{j})d\mu\geq\int_{X}\varepsilon_{j}\chi(\phi-u_{j})d\mu\geq j,

which is a contradiction. ∎

Lemma 6.10.

There exists a constant C>0C>0 such that, for all u,vχ(X,θ,ϕ)u,v\in\mathcal{E}_{\chi}(X,\theta,\phi) with supXv=0\sup_{X}v=0 and u0u\leq 0, we have

Xχ(ϕv)θunC(1+Eχ(u,ϕ))Eχ(v,ϕ)M/(M+1)+C.\int_{X}\chi(\phi-v)\theta_{u}^{n}\leq C(1+E_{\chi}(u,\phi))E_{\chi}(v,\phi)^{M/(M+1)}+C. (6.4)

The above result is due to Duc-Thai Do and Duc-Viet Vu [DV21]. Their proof uses generalized non-pluripolar products. We give below a less technical argument that relies only on the comparison principle.

Proof.

We observe that for all t>1t>1, v+(t1)ϕ,tuχ(X,tθ,tϕ)v+(t-1)\phi,tu\in\mathcal{E}_{\chi}(X,t\theta,t\phi) and tϕ(v+(t1)ϕ)=ϕvt\phi-(v+(t-1)\phi)=\phi-v (where equality holds outside a puripolar set). Fixing t>1t>1, by Lemma 6.4 (regarding v+(t1)ϕv+(t-1)\phi and tutu as elements of (X,tθ,tϕ)\mathcal{E}(X,t\theta,t\phi)) we have

Xχ(ϕv)(tθ+ddctu)n\displaystyle\int_{X}\chi(\phi-v)(t\theta+dd^{c}tu)^{n} \displaystyle\leq 2n+MXχ(ϕv)(tθ+ddc(v+(t1)ϕ))n\displaystyle 2^{n+M}\int_{X}\chi(\phi-v)(t\theta+dd^{c}(v+(t-1)\phi))^{n}
+Xχ(tϕtu)(tθ+ddctu)n\displaystyle+\int_{X}\chi(t\phi-tu)(t\theta+dd^{c}tu)^{n}
\displaystyle\leq 2n+MtnXχ(ϕv)θϕn+tn122n+MEχ(v,ϕ)+tn+MEχ(u,ϕ),\displaystyle 2^{n+M}t^{n}\int_{X}\chi(\phi-v)\theta_{\phi}^{n}+t^{n-1}2^{2n+M}E_{\chi}(v,\phi)+t^{n+M}E_{\chi}(u,\phi),

where the last inequality follows from Lemma 6.3 after observing that

(tθ+ddc(v+(t1)ϕ))n=((t1)θϕ+θv)ntnθϕn+tn12nj=1nθvjθϕnj.(t\theta+dd^{c}(v+(t-1)\phi))^{n}=((t-1)\theta_{\phi}+\theta_{v})^{n}\leq t^{n}\theta_{\phi}^{n}+t^{n-1}2^{n}\sum_{j=1}^{n}\theta_{v}^{j}\wedge\theta_{\phi}^{n-j}.

Dividing by tnt^{n} and using Lemma 6.2 we thus get

Xχ(ϕv)θunC1+22n+Mt1Eχ(v,ϕ)+tMEχ(u,ϕ).\int_{X}\chi(\phi-v)\theta_{u}^{n}\leq C_{1}+2^{2n+M}t^{-1}E_{\chi}(v,\phi)+t^{M}E_{\chi}(u,\phi).

Choosing t=(1+Eχ(v,ϕ))1/(M+1)t=(1+E_{\chi}(v,\phi))^{{1}/{(M+1)}}, we finish the proof. ∎

Proposition 6.11.

Assume φ(X,θ,ϕ)\varphi\in\mathcal{E}(X,\theta,\phi) with supXφ=0\sup_{X}\varphi=0 satisfies θφnAθun\theta_{\varphi}^{n}\leq A\theta_{u}^{n} and [u]=[ϕ][u]=[\phi] for some A>0A>0 and uPSH(X,θ)u\in{\rm PSH}(X,\theta). Then there exists α>0\alpha>0 such that

Xeα(ϕφ)θφn<.\int_{X}e^{\alpha(\phi-\varphi)}\theta_{\varphi}^{n}<\infty.
Proof.

We can assume that supXφ=0\sup_{X}\varphi=0 and ϕ1uϕ\phi-1\leq u\leq\phi. We argue as in [DL17, Lemma 4.3]. Fix t>0t>0 and s>1s>1. Observe that away from the pluripolar set {ϕ=}\{\phi=-\infty\} we have that {φ<ϕts}{φ<ϕt+s(uϕ)}{φ<ϕt}\{\varphi<\phi-t-s\}\subseteq\{\varphi<\phi-t+s(u-\phi)\}\subseteq\{\varphi<\phi-t\}. By the assumption and the partial comparison principle (Proposition 3.22) we have

{φ<ϕts}θφn\displaystyle\int_{\{\varphi<\phi-t-s\}}\theta_{\varphi}^{n} A{s1φ+(1s1)ϕ<us1t}θun\displaystyle\leq A\int_{\{s^{-1}\varphi+(1-s^{-1})\phi<u-s^{-1}t\}}\theta_{u}^{n}
A{s1φ+(1s1)ϕ<us1t}(s1θφ+(1s1)θϕ)n\displaystyle\leq A\int_{\{s^{-1}\varphi+(1-s^{-1})\phi<u-s^{-1}t\}}(s^{-1}\theta_{\varphi}+(1-s^{-1})\theta_{\phi})^{n}
A{φ<ϕt}(s1θφ+(1s1)θϕ)n\displaystyle\leq A\int_{\{\varphi<\phi-t\}}(s^{-1}\theta_{\varphi}+(1-s^{-1})\theta_{\phi})^{n}
A{φ<ϕt}θϕn+Ak=1nsk(nk){φ<ϕt}θφkθϕnk\displaystyle\leq A\int_{\{\varphi<\phi-t\}}\theta_{\phi}^{n}+A\sum_{k=1}^{n}s^{-k}\binom{n}{k}\int_{\{\varphi<\phi-t\}}\theta_{\varphi}^{k}\wedge\theta_{\phi}^{n-k}
AC0{φ<ϕt}ωn+2nAs1{φ<ϕt}θφn\displaystyle\leq AC_{0}\int_{\{\varphi<\phi-t\}}\omega^{n}+2^{n}As^{-1}\int_{\{\varphi<\phi-t\}}\theta_{\varphi}^{n}
C2eat{φ<t}ea|φ|ωn+2nAs1{φ<ϕt}θφn\displaystyle\leq C_{2}e^{-at}\int_{\{\varphi<-t\}}e^{a|\varphi|}\omega^{n}+2^{n}As^{-1}\int_{\{\varphi<\phi-t\}}\theta_{\varphi}^{n}
C3eat+2nAs1{φ<ϕt}θφn.\displaystyle\leq C_{3}e^{-at}+2^{n}As^{-1}\int_{\{\varphi<\phi-t\}}\theta_{\varphi}^{n}.

In the fifth inequality we used Theorem 3.6. The last inequality holds for a choice of a>0a>0 very small, so that the uniform Skoda integrability theorem (see [Sko72], [GZ17]) ensures

XeaφωnC1\int_{X}e^{-a\varphi}\omega^{n}\leq C_{1}

is uniformly bounded.

We fix ss so large that 2nAs1<(2e)12^{n}As^{-1}<(2e)^{-1}. Then, for a=1/sa=1/s we have 2nAs1eas<1/22^{n}As^{-1}e^{as}<1/2. Setting

F(t):=eat{φ<ϕt}θφn,t>0,F(t):=e^{at}\int_{\{\varphi<\phi-t\}}\theta_{\varphi}^{n},\;t>0,

we then have

F(t+s)C4+F(t)2,F(t+s)\leq C_{4}+\frac{F(t)}{2},

from which we obtain, by induction on kk\in\mathbb{N},

F(t+ks)C4j=1k21j+2kF(t).F(t+ks)\leq C_{4}\sum_{j=1}^{k}2^{1-j}+2^{-k}F(t).

Any t0t\geq 0 can be written as t=t0+kst=t_{0}+ks for some t0[0,s]t_{0}\in[0,s] and kk\in\mathbb{N}. The above estimate thus gives a uniform bound F(t)2C4+sup[0,s]FC5F(t)\leq 2C_{4}+\sup_{[0,s]}F\leq C_{5}. Now, for α=a/2\alpha=a/2, we get

Xeα(ϕφ)θφn\displaystyle\int_{X}e^{\alpha(\phi-\varphi)}\theta_{\varphi}^{n} =0αeαtθφn({φ<ϕt})𝑑t=0a2ea2teatθφn({φ<ϕt})𝑑t,\displaystyle=\int_{0}^{\infty}\alpha e^{\alpha t}\theta_{\varphi}^{n}(\{\varphi<\phi-t\})dt=\int_{0}^{\infty}\frac{a}{2}e^{-{\frac{a}{2}}t}e^{at}\theta_{\varphi}^{n}(\{\varphi<\phi-t\})dt,
=0a2ea2tF(t)𝑑taC520ea2t𝑑t<.\displaystyle=\int_{0}^{\infty}\frac{a}{2}e^{-\frac{a}{2}t}F(t)dt\leq\frac{aC_{5}}{2}\int_{0}^{\infty}e^{-\frac{a}{2}t}dt<\infty.

Lemma 6.12.

If EXE\subset X is a pluripolar set then there exists uχ(X,θ,ϕ)u\in\mathcal{E}_{\chi}(X,\theta,\phi) such that E{u=}E\subset\{u=-\infty\}.

Proof.

If η(t)=tM,t0\eta(t)=t^{M},t\geq 0 then the inclusion η(X,θ,ϕ)χ(X,θ,ϕ)\mathcal{E}_{\eta}(X,\theta,\phi)\subset\mathcal{E}_{\chi}(X,\theta,\phi) holds because χ(t)χ(1)tM\chi(t)\leq\chi(1)t^{M} for all t1t\geq 1. We can thus assume that χ=η\chi=\eta is convex. We claim that there exists h1(X,θ,ϕ)h\in\mathcal{E}^{1}(X,\theta,\phi) such that E{h=}E\subset\{h=-\infty\}. Indeed, it follows from [BBGZ13, Corollary 2.11] that E{ψ=}E\subset\{\psi=-\infty\} for some ψ1(X,θ)\psi\in\mathcal{E}^{1}(X,\theta). We can assume supXψ=1\sup_{X}\psi=-1, which implies ψVθ\psi\leq V_{\theta}. Since Pθ(Vθ,ϕ)=ϕP_{\theta}(V_{\theta},\phi)=\phi, Proposition 3.21 ensures that the function h:=Pθ(ψ,ϕ)h:=P_{\theta}(\psi,\phi) belongs to (X,θ,ϕ)\mathcal{E}(X,\theta,\phi). By Theorem 3.6 we have

X(ϕh)θhn{h=ψ}(ϕψ)θψnX(Vθψ)θψn<.\int_{X}(\phi-h)\theta_{h}^{n}\leq\int_{\{h=\psi\}}(\phi-\psi)\theta_{\psi}^{n}\leq\int_{X}(V_{\theta}-\psi)\theta_{\psi}^{n}<\infty.

We thus have that h1(X,θ,ϕ)h\in\mathcal{E}^{1}(X,\theta,\phi) and E{ψ=}{h=}E\subset\{\psi=-\infty\}\subset\{h=-\infty\}, proving the claim.

Now, let γ:=χ1\gamma:=\chi^{-1} be the inverse of χ\chi (so γ(t)=t1/M\gamma(t)=t^{1/M} and γ(t)1\gamma^{\prime}(t)\leq 1 if t1t\geq 1), which is a concave weight. By adding a constant, we can assume that h1h\leq-1. Consider u:=γ(ϕh)+ϕPSH(X,θ,ϕ)u:=-\gamma(\phi-h)+\phi\in{\rm PSH}(X,\theta,\phi), as in Lemma 5.19 (see also the discussion following the statement). We observe also that E{u=}E\subset\{u=-\infty\} and uhu\geq h, hence u(X,θ,ϕ)u\in\mathcal{E}(X,\theta,\phi). By Lemma 6.5,

Xχ(ϕu)θun=X(ϕh)θun2nE1(h,ϕ)<.\int_{X}\chi(\phi-u)\theta_{u}^{n}=\int_{X}(\phi-h)\theta_{u}^{n}\leq 2^{n}E_{1}(h,\phi)<\infty.

We thus have that uχ(X,θ,ϕ)u\in\mathcal{E}_{\chi}(X,\theta,\phi), finishing the proof. ∎

Proposition 6.13.

Assume μ\mu is a positive Radon measure satisfying Xθϕn=μ(X)>0\int_{X}\theta_{\phi}^{n}=\mu(X)>0. Assume also that μ({ϕ=})=0\mu(\{\phi=-\infty\})=0 and

Xχ(ϕφ)𝑑μaEχ(φ,ϕ)+C,φχ(X,θ,ϕ),supXφ=0,\int_{X}\chi(\phi-\varphi)d\mu\leq aE_{\chi}(\varphi,\phi)+C,\quad\varphi\in\mathcal{E}_{\chi}(X,\theta,\phi),\;\sup_{X}\varphi=0, (6.5)

for some constants a(0,1)a\in(0,1), C>0C>0. Then μ=(θ+ddcu)n\mu=(\theta+dd^{c}u)^{n} for some uχ(X,θ,ϕ)u\in\mathcal{E}_{\chi}(X,\theta,\phi).

Proof.

We prove the proposition by an argument going back to Cegrell [Ceg98]. We first claim that μ\mu vanishes on pluripolar sets. Indeed, fix such a Borel set EE. It follows from Lemma 6.12 that E{h=}E\subset\{h=-\infty\} for some hχ(X,θ,ϕ)h\in\mathcal{E}_{\chi}(X,\theta,\phi). We can assume that hϕh\leq\phi.

Since Eχ(ϕh)𝑑μXχ(ϕh)𝑑μ<\int_{E}\chi(\phi-h)d\mu\leq\int_{X}\chi(\phi-h)d\mu<\infty, we get that μ(E{ϕ})=0\mu(E\cap\{\phi\neq-\infty\})=0. Since μ({ϕ=})=0\mu(\{\phi=-\infty\})=0, it follows that μ(E)=0\mu(E)=0.

We next claim that we can write

μ=fν,ν=(θ+ddcϕ0)n,\mu=f\nu,\;\nu=(\theta+dd^{c}\phi_{0})^{n},

for some ϕ0PSH(X,θ,ϕ)\phi_{0}\in{\rm PSH}(X,\theta,\phi) such that ϕ1ϕ0ϕ\phi-1\leq\phi_{0}\leq\phi, and some 0fL1(X,ν)0\leq f\in L^{1}(X,\nu). To see this, we first find ψ(X,θ,ϕ)\psi\in\mathcal{E}(X,\theta,\phi) such that (θ+ddcψ)n=μ(\theta+dd^{c}\psi)^{n}=\mu and supXψ=0\sup_{X}\psi=0 (in particular ψϕ\psi\leq\phi). The existence of ψ\psi follows from Theorem 5.17.

We set ϕ0:=eψϕ+ϕ\phi_{0}:=e^{\psi-\phi}+\phi as in Lemma 5.19 (see also the discussion following the statement). Then ϕ0PSH(X,θ,ϕ)\phi_{0}\in{\rm PSH}(X,\theta,\phi), [ϕ0]=[ϕ][\phi_{0}]=[\phi], and

(θ+ddcϕ0)nen(ψϕ)(θ+ddcψ)n=en(ψϕ)μ.\displaystyle(\theta+dd^{c}\phi_{0})^{n}\geq e^{n(\psi-\phi)}(\theta+dd^{c}\psi)^{n}=e^{n(\psi-\phi)}\mu. (6.6)

From (6.6) it follows that μ\mu is absolutely continuous with respect to (θ+ddcϕ0)n(\theta+dd^{c}\phi_{0})^{n}, proving the second claim.

Now, for each j>1j>1, let φj(X,θ,ϕ)\varphi_{j}\in\mathcal{E}(X,\theta,\phi) be the unique solution of

(θ+ddcφj)n=cjmin(f,j)(θ+ddcϕ0)n,supXφj=0,(\theta+dd^{c}\varphi_{j})^{n}=c_{j}\min(f,j)(\theta+dd^{c}\phi_{0})^{n},\;\sup_{X}\varphi_{j}=0,

where cjc_{j} is a normalization constant to have equality between the total masses of the left and right-hand side. For jj large enough we have cja<1c_{j}a<1, since cj1c_{j}\to 1 and by assumption a<1a<1. We can thus assume that cja<λ<1c_{j}a<\lambda<1. It follows from Proposition 6.11 that

Xeαj(ϕφj)(θ+ddcφj)n<,\int_{X}e^{\alpha_{j}(\phi-\varphi_{j})}(\theta+dd^{c}\varphi_{j})^{n}<\infty,

for some αj>0\alpha_{j}>0. In particular, since for any t>0t>0 χ(t)Ceαjt\chi(t)\leq Ce^{\alpha_{j}t} for some C>0C>0, we infer that Eχ(φj,ϕ)E_{\chi}(\varphi_{j},\phi) is finite. We claim that this bound is uniform in jj. Indeed, since θφjncjfθϕ0n\theta_{\varphi_{j}}^{n}\leq c_{j}f\theta_{\phi_{0}}^{n}, we have

Eχ(φj,ϕ)Xχ(ϕφj)cj𝑑μacjEχ(φj,ϕ)+CE_{\chi}(\varphi_{j},\phi)\leq\int_{X}\chi(\phi-\varphi_{j})c_{j}d\mu\leq ac_{j}E_{\chi}(\varphi_{j},\phi)+C

gives Eχ(φj,ϕ)C(1λ)1E_{\chi}(\varphi_{j},\phi)\leq C(1-\lambda)^{-1}. Extracting a subsequence we can assume that φjφ\varphi_{j}\to\varphi in L1L^{1}. It follows from Lemma 6.7 that φχ(X,θ,ϕ)\varphi\in\mathcal{E}_{\chi}(X,\theta,\phi). It then follows from Lemma 5.16 that (θ+ddcφ)nμ(\theta+dd^{c}\varphi)^{n}\geq\mu. Comparing the total mass, we obtain the equality, finishing the proof. ∎

We are now ready to prove our main result of this section.

Theorem 6.14.

Assume μ({ϕ=})=0\mu(\{\phi=-\infty\})=0 and χ(|ϕu|)L1(μ)\chi(|\phi-u|)\in L^{1}(\mu), for all uχ(X,θ,ϕ)u\in\mathcal{E}_{\chi}(X,\theta,\phi). Then μ=(θ+ddcφ)n\mu=(\theta+dd^{c}\varphi)^{n} for some φχ(X,θ,ϕ)\varphi\in\mathcal{E}_{\chi}(X,\theta,\phi).

Proof.

Let CC be the constant in Lemma 6.9. Let v(X,θ,ϕ)v\in\mathcal{E}(X,\theta,\phi) be the unique solution to

(θ+ddcv)n=(4C)1μ+bωn,supXv=1.(\theta+dd^{c}v)^{n}=(4C)^{-1}\mu+b\omega^{n},\;\sup_{X}v=-1.

Here b>0b>0 is a constant so that X(4C)1μ+bωn=Xθϕn\int_{X}(4C)^{-1}\mu+b\omega^{n}=\int_{X}\theta_{\phi}^{n}. The existence of vv follows from Theorem 5.17. By Lemma 6.9 there exists a constant C1>0C_{1}>0 such that, for all φχ(X,θ,ϕ)\varphi\in\mathcal{E}_{\chi}(X,\theta,\phi) with supXφ=0\sup_{X}\varphi=0, we have

Xχ(ϕφ)θvn14Eχ(φ,ϕ)+C1.\int_{X}\chi(\phi-\varphi)\theta_{v}^{n}\leq\frac{1}{4}E_{\chi}(\varphi,\phi)+C_{1}.

By Proposition 6.13, vχ(X,θ,ϕ)v\in\mathcal{E}_{\chi}(X,\theta,\phi). Since μ4C(θ+ddcv)n\mu\leq 4C(\theta+dd^{c}v)^{n}, it follows from Lemma 6.10 that μ\mu satisfies

Xχ(ϕu)μ4CXχ(ϕu)θvnC(1+Eχ(v,ϕ))Eχ(u,ϕ)M/(M+1)+C.\int_{X}\chi(\phi-u)\mu\leq 4C\int_{X}\chi(\phi-u)\theta_{v}^{n}\leq C^{\prime}(1+E_{\chi}(v,\phi))E_{\chi}(u,\phi)^{M/(M+1)}+C^{\prime}.

As a result, μ\mu also satisfies (6.5). We can thus use Lemma 6.13 to complete the proof. ∎

We summarize the findings of this chapter in the theorem below.

Theorem 6.15.

Fix a Radon measure μ\mu with μ({ϕ=})=0\mu(\{\phi=-\infty\})=0 and Xθϕn=μ(X)>0\int_{X}\theta_{\phi}^{n}=\mu(X)>0. Then the following are equivalent.

  1. (i)

    There exists a constant C>0C>0 such that, for all uχ(X,θ,ϕ)u\in\mathcal{E}_{\chi}(X,\theta,\phi) with supXu=0\sup_{X}u=0, we have

    Xχ(ϕu)𝑑μCEχ(u,ϕ)M/(M+1)+C.\int_{X}\chi(\phi-u)d\mu\leq CE_{\chi}(u,\phi)^{M/(M+1)}+C.
  2. (ii)

    χ(|ϕu|)L1(μ)\chi(|\phi-u|)\in L^{1}(\mu), for all uχ(X,θ,ϕ)u\in\mathcal{E}_{\chi}(X,\theta,\phi).

  3. (iii)

    μ=(θ+ddcφ)n\mu=(\theta+dd^{c}\varphi)^{n} for some φχ(X,θ,ϕ)\varphi\in\mathcal{E}_{\chi}(X,\theta,\phi), with supXφ=0\sup_{X}\varphi=0.

Proof.

(i) \Longrightarrow (ii) is immediate. The implication (ii) \Longrightarrow (iii) is Theorem 6.14. The implication (iii) \Longrightarrow (i) is Lemma 6.10. ∎

Bibliography

  • [BB10] Robert Berman and Sébastien Boucksom “Growth of balls of holomorphic sections and energy at equilibrium” In Invent. Math. 181.2, 2010, pp. 337–394 DOI: 10.1007/s00222-010-0248-9
  • [BBGZ13] R.. Berman, S. Boucksom, V. Guedj and A. Zeriahi “A variational approach to complex Monge-Ampère equations” In Publ. Math. Inst. Hautes Études Sci. 117, 2013, pp. 179–245 DOI: 10.1007/s10240-012-0046-6
  • [BEGZ10] S. Boucksom, P. Eyssidieux, V. Guedj and A. Zeriahi “Monge-Ampère equations in big cohomology classes” In Acta Math. 205.2, 2010, pp. 199–262 DOI: 10.1007/s11511-010-0054-7
  • [Ber19] R.. Berman “From Monge-Ampère equations to envelopes and geodesic rays in the zero temperature limit” In Math. Z. 291.1-2, 2019, pp. 365–394 DOI: 10.1007/s00209-018-2087-0
  • [BFJ08] S. Boucksom, C. Favre and M. Jonsson “Valuations and plurisubharmonic singularities” In Publications of the Research Institute for Mathematical Sciences 44.2, 2008, pp. 449–494
  • [BK07] Z. Błocki and S. Kołodziej “On regularization of plurisubharmonic functions on manifolds” In Proc. Amer. Math. Soc. 135.7, 2007, pp. 2089–2093 DOI: 10.1090/S0002-9939-07-08858-2
  • [BL12] T. Bloom and N. Levenberg “Pluripotential energy” In Potential Anal. 36.1, 2012, pp. 155–176 DOI: 10.1007/s11118-011-9224-2
  • [Bou04] S. Boucksom “Divisorial Zariski decompositions on compact complex manifolds” In Ann. Sci. École Norm. Sup. (4) 37.1, 2004, pp. 45–76 DOI: 10.1016/j.ansens.2003.04.002
  • [BT76] E. Bedford and B.. Taylor “The Dirichlet problem for a complex Monge-Ampère equation” In Invent. Math. 37.1, 1976, pp. 1–44 DOI: 10.1007/BF01418826
  • [BT82] E. Bedford and B.. Taylor “A new capacity for plurisubharmonic functions” In Acta Math. 149.1-2, 1982, pp. 1–40 DOI: 10.1007/BF02392348
  • [BT87] E. Bedford and B.. Taylor “Fine topology, Šilov boundary, and (ddc)n(dd^{c})^{n} In J. Funct. Anal. 72.2, 1987, pp. 225–251 DOI: 10.1016/0022-1236(87)90087-5
  • [Ceg98] U. Cegrell “Pluricomplex energy” In Acta Math. 180.2, 1998, pp. 187–217 DOI: 10.1007/BF02392899
  • [CG09] Dan Coman and Vincent Guedj “Quasiplurisubharmonic Green functions” In J. Math. Pures Appl. (9) 92.5, 2009, pp. 456–475
  • [Dar17] T. Darvas “The Mabuchi completion of the space of Kähler potentials” In Amer. J. Math. 139.5, 2017, pp. 1275–1313 DOI: 10.1353/ajm.2017.0032
  • [Dar19] T. Darvas “Geometric pluripotential theory on Kähler manifolds” In Advances in complex geometry 735, Contemp. Math. Amer. Math. Soc., Providence, RI, 2019, pp. 1–104 DOI: 10.1090/conm/735/14822
  • [DDL18] T. Darvas, E. Di Nezza and C.. Lu L1L^{1} metric geometry of big cohomology classes” In Ann. Inst. Fourier (Grenoble) 68.7, 2018, pp. 3053–3086 URL: http://aif.cedram.org.revues.math.u-psud.fr:2048/item?id=AIF_2018__68_7_3053_0
  • [DDL18a] T. Darvas, E. Di Nezza and C.. Lu “Monotonicity of nonpluripolar products and complex Monge-Ampère equations with prescribed singularity” In Anal. PDE 11.8, 2018, pp. 2049–2087 DOI: 10.2140/apde.2018.11.2049
  • [DDL18b] T. Darvas, E. Di Nezza and C.. Lu “On the singularity type of full mass currents in big cohomology classes” In Compos. Math. 154.2, 2018, pp. 380–409 DOI: 10.1112/S0010437X1700759X
  • [DDL21] T. Darvas, E. Di Nezza and C.. Lu “Log-concavity of volume and complex Monge-Ampère equations with prescribed singularity” In Math. Ann. 379.1-2, 2021, pp. 95–132 DOI: 10.1007/s00208-019-01936-y
  • [DDL21a] T. Darvas, E. Di Nezza and C.H. Lu “The metric geometry of singularity types” In J. Reine Angew. Math. 771, 2021, pp. 137–170
  • [Dem12] J.-P. Demailly “Analytic methods in algebraic geometry” 1, Surveys of Modern Mathematics International Press, Somerville, MA; Higher Education Press, Beijing, 2012, pp. viii+231
  • [Dem94] J.-P. Demailly “Regularization of closed positive currents of type (1,1)(1,1) by the flow of a Chern connection” In Contributions to complex analysis and analytic geometry, Aspects Math., E26 Friedr. Vieweg, Braunschweig, 1994, pp. 105–126
  • [Din09] S. Dinew “Uniqueness in (X,ω)\mathcal{E}(X,\omega) In J. Funct. Anal. 256.7, 2009, pp. 2113–2122 DOI: 10.1016/j.jfa.2009.01.019
  • [DL15] Eleonora Di Nezza and Chinh H. Lu “Generalized Monge-Ampère capacities” In Int. Math. Res. Not. IMRN, 2015, pp. 7287–7322 DOI: 10.1093/imrn/rnu166
  • [DL17] E. Di Nezza and C.. Lu “Complex Monge-Ampère equations on quasi-projective varieties” In J. Reine Angew. Math. 727, 2017, pp. 145–167
  • [Don11] S. Donaldson “Kähler metrics with cone singularities along a divisor” In arXiv:1102.1196, 2011
  • [DR16] T. Darvas and Y.A. Rubinstein “Kiselman’s principle, the Dirichlet problem for the Monge-Ampère equation, and rooftop obstacle problems” In J. Math. Soc. Japan 68.2, 2016, pp. 773–796 DOI: 10.2969/jmsj/06820773
  • [DT21] E. Di Nezza and S. Trapani “Monge-Ampère measures on contact sets” In Math. Res. Lett. 28.5, 2021, pp. 1337–1352
  • [DT23] E. Di Nezza and S. Trapani “The regularity of envelopes” In Annales scientifiques de l’ENS., 2023
  • [DV21] D.-T. Do and D.-V. Vu “Complex Monge-Ampère equations with solutions in finite energy classes” In arXiv:2010.08619, to appear in Math. Research Letters, 2021
  • [DV22] H.-S. Do and D.-V. Vu “Quantitative stability for the complex Monge-Ampere equations” In arXiv:2209.00248, 2022
  • [DX21] T. Darvas and M. Xia “The volume of pseudoeffective line bundles and partial equilibrium”, 2021 arXiv:2112.03827 [math.DG]
  • [DXZ23] T. Darvas, M. Xia and K. Zhang “A transcendental approach to non-Archimedean metrics of pseudoeffective classes”, 2023 arXiv:2302.02541 [math.AG]
  • [DZ22] T. Darvas and K. Zhang “Twisted Kähler–Einstein metrics in big classes”, 2022 arXiv:2208.08324 [math.DG]
  • [EGZ09] P. Eyssidieux, V. Guedj and A. Zeriahi “Singular Kähler-Einstein metrics” In J. Amer. Math. Soc. 22.3, 2009, pp. 607–639 DOI: 10.1090/S0894-0347-09-00629-8
  • [GL21] V. Guedj and C.. Lu “Quasi-plurisubharmonic envelopes 1: Uniform estimates on Kähler manifolds” In arxiv:2106.04273, 2021
  • [GLZ19] V. Guedj, C.. Lu and A. Zeriahi “Plurisubharmonic envelopes and supersolutions” In J. Differential Geom. 113.2, 2019, pp. 273–313 DOI: 10.4310/jdg/1571882428
  • [GP16] Henri Guenancia and Mihai Păun “Conic singularities metrics with prescribed Ricci curvature: general cone angles along normal crossing divisors” In J. Differential Geom. 103.1, 2016, pp. 15–57 URL: http://projecteuclid.org/euclid.jdg/1460463562
  • [Gup22] P. Gupta “A complete metric topology on relative low energy spaces” In preprint, 2022
  • [GZ05] V. Guedj and A. Zeriahi “Intrinsic capacities on compact Kähler manifolds” In J. Geom. Anal. 15.4, 2005, pp. 607–639 DOI: 10.1007/BF02922247
  • [GZ07] V. Guedj and A. Zeriahi “The weighted Monge-Ampère energy of quasiplurisubharmonic functions” In J. Funct. Anal. 250.2, 2007, pp. 442–482 DOI: 10.1016/j.jfa.2007.04.018
  • [GZ15] Q. Guan and X. Zhou “A proof of Demailly’s strong openness conjecture” In Ann. of Math. (2) 182.2, 2015, pp. 605–616
  • [GZ17] V. Guedj and A. Zeriahi “Degenerate complex Monge-Ampère equations” 26, EMS Tracts in Mathematics European Mathematical Society (EMS), Zürich, 2017, pp. xxiv+472 DOI: 10.4171/167
  • [JMR16] Thalia Jeffres, Rafe Mazzeo and Yanir A. Rubinstein “Kähler-Einstein metrics with edge singularities” In Ann. of Math. (2) 183.1, 2016, pp. 95–176 DOI: 10.4007/annals.2016.183.1.3
  • [Koł03] S. Kołodziej “The complex Monge-Ampère equation on compact Kähler manifolds” In Indiana Univ. Math. J. 52.3, 2003, pp. 667–686
  • [Koł98] S. Kołodziej “The complex Monge-Ampère equation” In Acta Math. 180.1, 1998, pp. 69–117 DOI: 10.1007/BF02392879
  • [LN22] C.. Lu and V.-D. Nguyên “Complex Hessian equations with prescribed singularity on compact Kähler manifolds” In Ann. Sc. Norm. Super. Pisa Cl. Sci. (5) 23.1, 2022, pp. 425–462
  • [Lu21] Chinh H. Lu “Comparison of Monge-Ampère capacities” In Ann. Polon. Math. 126.1, 2021
  • [PS14] D.. Phong and Jacob Sturm “On the singularities of the pluricomplex Green’s function” In Advances in analysis: the legacy of Elias M. Stein 50, Princeton Math. Ser. Princeton Univ. Press, Princeton, NJ, 2014, pp. 419–435
  • [RS05] Alexander Rashkovskii and Ragnar Sigurdsson “Green functions with singularities along complex spaces” In Internat. J. Math. 16.4, 2005
  • [RW14] J. Ross and D. Witt Nyström “Analytic test configurations and geodesic rays” In J. Symplectic Geom. 12.1, 2014, pp. 125–169 DOI: 10.4310/JSG.2014.v12.n1.a5
  • [Sko72] H. Skoda “Sous-ensembles analytiques d’ordre fini ou infini dans 𝐂n{\bf C}^{n} In Bull. Soc. Math. France 100, 1972, pp. 353–408 URL: http://www.numdam.org.revues.math.u-psud.fr:2048/item?id=BSMF_1972__100__353_0
  • [Tos18] V. Tosatti “Regularity of envelopes in Kähler classes” In Math. Res. Lett. 25.1, 2018, pp. 281–289
  • [Tru19] A. Trusiani L1L^{1} metric geometry of potentials with prescribed singularities on compact Kähler manifolds” In arXiv:1909.03897, 2019
  • [Tru20] A. Trusiani “Kähler-Einstein metrics with prescribed singularities on Fano manifolds”, 2020 arXiv:2006.09130
  • [Tru23] A. Trusiani “A relative Yau-Tian-Donaldson conjecture and stability thresholds”, 2023 arXiv:arXiv:2302.07213
  • [Vu20] D.-V. Vu “Lelong numbers of currents of full mass intersection” In arxiv:2008.09219, 2020
  • [Vu21] D.-V. Vu “Relative non-pluripolar product of currents” In Ann. Global Anal. Geom. 60.2, 2021, pp. 269–311
  • [Wit19] D. Witt Nyström “Monotonicity of non-pluripolar Monge-Ampère masses” In Indiana Univ. Math. J. 68.2, 2019, pp. 579–591 DOI: 10.1512/iumj.2019.68.7630
  • [Xia19] M. Xia “Integration by parts formula for non-pluripolar product” In arXiv:1907.06359, 2019
  • [Xia19a] M. Xia “Mabuchi geometry of big cohomology classes with prescribed singularities” In
    arXiv:1907.07234
    , 2019
  • [Xin96] Y. Xing “Continuity of the complex Monge-Ampère operator” In Proc. Amer. Math. Soc. 124.2, 1996
  • [Yau78] S.-T. Yau “On the Ricci curvature of a compact Kähler manifold and the complex Monge-Ampère equation. I” In Comm. Pure Appl. Math. 31.3, 1978, pp. 339–411 DOI: 10.1002/cpa.3160310304