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

A transcendental approach to non-Archimedean metrics of pseudoeffective classes

Tamás Darvas, Mingchen Xia and Kewei Zhang
Abstract

We introduce the concept of non-Archimedean metrics attached to a transcendental pseudoeffective cohomology class on a compact Kähler manifold. This is obtained via extending the Ross–Witt Nyström correspondence to the relative case, and we point out that our construction agrees with that of Boucksom–Jonsson when the class is induced by a pseudoeffective \mathbb{Q}-line bundle.

We introduce the notion of a flag configuration attached to a transcendental big class, recovering the notion of a test configuration in the ample case. We show that non-Archimedean finite energy metrics are approximable by flag configurations, and very general versions of the radial Ding energy are continuous, a novel result even in the ample case. As applications, we characterize the delta invariant as the Ding semistability threshold of flag configurations and filtrations, and prove a YTD type existence theorem for Kähler–Einstein metrics in terms of flag configurations.

1 Introduction and results

In recent years, especially since the appearance of [BBJ21], the non-Archimedean approach to K-stability has gained attention, as evidenced by recent works [BJ21, BE21, BHJ17, BJ18, DL20a, Li20, BJ22a], to name only a few papers in the fast expanding literature.

Studying K-stability for degenerate classes is a natural extension of the much studied ample case [CDS15, DS16, Tia15, TW20, CSW18, BBJ21, Zha21]. Recent works aim to understand existence of canonical metrics in this context as well. Degenerate Kähler–Einstein metrics have taken the forefront [LTW21, LTW21a, LXZ22, DZ22, DR22, Xu22], but one would think that the notion of constant scalar curvature Kähler metrics also extends to big classes, as minimzers of the K-energy functional. With the study still in its infancy [Zhe22, DL22], a better understanding of K-stability of big classes is needed before substantial progress can be made. Our work tries to fill some of this void.

We propose a transcendental approach to non-Archimedean metrics, that allows for their treatment even for a transcendental pseudoeffective class. Our analytic treatment has a number of advantages. The envelope conjecture of Boucksom–Jonsson naturally holds in our context, immediately recovering the main result of [BJ22], as a particular case. We show that very general versions of the radial Ding energy are continuous, and that finite energy non-Archimedean metrics can be approximated by algebraic objects called flag configurations (that reduce to test configurations in the ample case). The latter two properties allow to characterize the analytically defined delta invariant using algebraic data, and provide a YTD type existence theorem for big Kähler–Einstein (KE) metrics, without using the minimal model program, connecting with recent work of the first and third authors [DZ22].

Transcendental non-Archimedean metrics.

To state our main results, we fix some terminology. Let XX be a compact Kähler manifold and θ\theta a smooth closed real (1,1)(1,1)-form representing a pseudoeffective cohomology class {θ}\{\theta\}. We denote by PSH(X,θ)\textup{PSH}(X,\theta) the space of θ\theta-psh functions, and all complex Monge–Ampère measures in this work will be assumed to be pluripolar in the sense of [BEGZ10].

A relative test curve is a map τψτPSH(X,θ)\mathbb{R}\ni\tau\mapsto\psi_{\tau}\in\textup{PSH}(X,\theta) that is τ\tau-decreasing, τ\tau-concave, and τ\tau-usc. Moreover, ψτ\psi_{\tau}\equiv-\infty for all τ\tau big enough, and the limit ψ=limτψτPSH(X,θ)\psi_{-\infty}=\lim_{\tau\to-\infty}\psi_{\tau}\in\textup{PSH}(X,\theta) exists. Such an object will be denoted {ψτ}τ\{\psi_{\tau}\}_{\tau}, and we say that {ψτ}τ\{\psi_{\tau}\}_{\tau} is a test curve relative to ψ\psi_{-\infty}.

The relative test curve {ψτ}τ\{\psi_{\tau}\}_{\tau} is \mathcal{I}-maximal if ψτ=P[ψτ]\psi_{\tau}=P[\psi_{\tau}]_{\mathcal{I}} for all τ\tau\in\mathbb{R}, where P[]P[\cdot]_{\mathcal{I}} is the following envelope:

P[u]:=sup{w PSH(X,θ),w0,(tw)(tu),t0}.P[u]_{\mathcal{I}}:=\sup\{w\in\textup{ PSH}(X,\theta),w\leq 0,\mathcal{I}(tw)\subseteq\mathcal{I}(tu),t\geq 0\}.

Here (tu)\mathcal{I}(tu) is the multiplier ideal sheaf of uu, locally generated by holomorphic functions ff such that |f|2etu|f|^{2}e^{-tu} is integrable.

To exclude pathological behaviour, we need to avoid vanishing mass. For any big class {θ}\{\theta\}, by MTC(X,θ)\textup{MTC}_{\mathcal{I}}(X,\theta) we denote the set of maximal test curves that additionally satisfy Xθψn>0\int_{X}\theta_{\psi_{-\infty}}^{n}>0. As we will see in (28), given any Kähler metric ω\omega from the Kähler cone 𝒦\mathcal{K}, there is a natural map

MTC(X,θ)MTC(X,θ+ω)\textup{MTC}_{\mathcal{I}}(X,\theta)\to\textup{MTC}_{\mathcal{I}}(X,\theta+\omega)

This allows to define the space of non-Archimedean metrics associated with a pseudoeffective class {θ}\{\theta\} as the following projective limit in the category of sets:

PSHNA(θ):=limω𝒦MTC(X,θ+ω).\mathrm{PSH}^{\mathrm{NA}}(\theta):=\varprojlim_{\omega\in\mathcal{K}}\textup{MTC}_{\mathcal{I}}(X,\theta+\omega).

We refer to Definition 3.3 for the precise details. Given fC(X)f\in C^{\infty}(X), there is an obvious identification between PSHNA(θ)\mathrm{PSH}^{\mathrm{NA}}(\theta) and PSHNA(θ+ddcf)\mathrm{PSH}^{\mathrm{NA}}(\theta+\mathrm{dd}^{\mathrm{c}}f) allowing us to regard PSHNA(θ)\mathrm{PSH}^{\mathrm{NA}}(\theta) as an invariant of the cohomology class {θ}\{\theta\}.

When {θ}\{\theta\} is the first Chern class of a \mathbb{Q}-line bundle LL, Boucksom and Jonsson defined the space of non-Archimedean metrics PSHNA(Lan)\textup{PSH}^{\mathrm{NA}}(L^{\mathrm{an}}) in [BJ21] using algebraic tools (see Section 2.3). In our first result we show that our construction agrees with theirs in this important particular case.

Theorem 1.1.

(=Theorem 3.14) When {θ}\{\theta\} is the first Chern class of a pseudoeffective \mathbb{Q}-line bundle LL, there is a natural bijection between PSHNA(θ)\mathrm{PSH}^{\mathrm{NA}}(\theta) and PSH(Lan)\textup{PSH}(L^{\mathrm{an}}).

The same holds for a pseudoeffective (1,1)(1,1)-class lying in the Néron–Severi group with real coefficients, as we explain in Remark 3.15.

For the construction of the natural map between PSHNA(θ)\mathrm{PSH}^{\mathrm{NA}}(\theta) and PSH(Lan)\textup{PSH}(L^{\mathrm{an}}), we refer to (39). Recently Boucksom–Jonsson have showed that the so-called envelope conjecture holds in case {θ}\{\theta\} is the first Chern class of a \mathbb{Q}-line bundle LL [BJ22]. We will show that the same property holds in the transcendental case as well, thus recovering the Boucksom–Jonsson result as a particular case, via the above isomorphism theorem. Let us note here an equivalent formulation of the envelope conjecture, which is easy to state:

Theorem 1.2.

(=Theorem 3.4) Suppose that {θ}\{\theta\} is a pseudoeffective class. Any bounded from above increasing net of PSHNA(θ)\mathrm{PSH}^{\mathrm{NA}}(\theta) has a supremum inside PSHNA(θ)\mathrm{PSH}^{\mathrm{NA}}(\theta).

We refer to Theorem 3.4 for the precise statement, and see Conjecture 2.13 for the equivalent form of the envelope conjecture in non-Archimedean geometry. Moreover, in Corollary 3.16 we explain how the above result yields [BJ22, Theorem A].

For general transcendental classes {θ}\{\theta\} we cannot interpret PSHNA(θ)\mathrm{PSH}^{\mathrm{NA}}(\theta) as metrics on certain Berkovich spaces, but an analogous study is still possible, and we will treat this thoroughly in a separate paper.

Approximation of finite energy non-Archimedean metrics.

Let us assume for the rest of this introduction that {θ}\{\theta\} is a big class. A maximal test curve {ψτ}τ\{\psi_{\tau}\}_{\tau} is of finite energy (notation: {ψτ}τ1(X,θ)\{\psi_{\tau}\}_{\tau}\in\mathcal{R}^{1}(X,\theta)) if

τψ+(XθψτnXθVθn)dτ>,\int_{-\infty}^{\tau^{+}_{\psi}}\left(\int_{X}\theta_{\psi_{\tau}}^{n}-\int_{X}\theta_{V_{\theta}}^{n}\right)\,\mathrm{d}\tau>-\infty\,,

where VθPSH(X,θ)V_{\theta}\in\mathrm{PSH}(X,\theta) is the potential with minimal singularity. Such test curves are in a one-to-one correspondence with finite energy geodesic rays, per the Ross–Witt Nyström correspondence, elaborated in [DZ22, DL20] (see Theorem 2.6(iii) and (62)). As proved in [DDNL21, Theorem 2.14] there is a chordal metric d1cd_{1}^{c} on 1(X,θ)\mathcal{R}^{1}(X,\theta) making it a complete metric space (see Section 2.2).

We identify the space of finite energy non-Archimedean metrics 1,NA(X,θ)\mathcal{E}^{1,\mathrm{NA}}(X,\theta) with 1(X,θ)\mathcal{R}^{1}_{\mathcal{I}}(X,\theta), the space of \mathcal{I}-maximal finite energy test curves. The definition of 1,NA(X,θ)\mathcal{E}^{1,\mathrm{NA}}(X,\theta) agrees with that of Boucksom–Jonsson in the ample case, and one can approximate any finite energy non-Archimedean metric using test configurations, by [DX22, Theorem 1.1]. We show in this paper that the analogous result holds in the big case as well.

To state the result, let us define a flag configuration of a big cohomology class {θ}\{\theta\} to be a (partial) flag of coherent analytic ideal sheaves

𝔞0𝔞1𝔞N=(Vθ).\mathfrak{a}_{0}\subseteq\mathfrak{a}_{1}\subseteq\dots\subseteq\mathfrak{a}_{N}=\mathcal{I}(V_{\theta}).

By convention, 𝔞:=0\mathfrak{a}_{\ell}:=0 for <0\ell\in\mathbb{Z}_{<0} and 𝔞:=𝔞N\mathfrak{a}_{\ell}:=\mathfrak{a}_{N} if N\ell\in\mathbb{Z}_{\geq N}. A flag configuration will be conveniently denoted as an analytic coherent ideal sheaf on the product X×X\times\mathbb{C}:

𝔞:=𝔞0+𝔞1s++𝔞N1sN1+𝔞N(sN)𝒪(X×),\mathfrak{a}:=\mathfrak{a}_{0}+\mathfrak{a}_{1}s+\dots+\mathfrak{a}_{N-1}s^{N-1}+\mathfrak{a}_{N}(s^{N})\subseteq\mathcal{O}(X\times\mathbb{C}),

where ss denotes the coordinate on =Spec[s]\mathbb{C}=\operatorname{Spec}\mathbb{C}[s]. In this transcendental big setting, we avoid calling flag configurations actual test configurations, to avoid confusion with the concepts introduced in [DR22] for big line bundles, which are not yet proved to be equivalent with ours. In addition, in the Kähler case, there is notion of transcendental test configuration introduced in [DR17, Dyr16]. Further investigations are needed to find the correct analogue of this notion is in the big case, and prove possible equivalency with our flag configurations.

As we point out in (47), in case {θ}\{\theta\} is the first Chern class of a big line bundle LL, to a flag configuration one can associate a natural filtration 𝔞\mathcal{F}^{\mathfrak{a}} of the section ring R(X,L)R(X,L). Extending a construction of Phong–Sturm/Ross–Witt Nyström [PS07, RWN14], we show that even in the big case, one can associate a natural geodesic ray/maximal test curve to any filtration (see Definition 4.4).

Next we show that finite energy non-Archimedean metrics can be approximated by flag configurations or filtrations, extending [DX22, Theorem 1.1] to the big case, making contact with the ideas of [BBJ21, Section 5].

Theorem 1.3.

(=Theorem 4.14) When {θ}\{\theta\} is the first Chern class of a big line bundle LL, elements of 1(X,θ)=1,NA(X,Lan)\mathcal{R}^{1}_{\mathcal{I}}(X,\theta)=\mathcal{E}^{1,\mathrm{NA}}(X,L^{\mathrm{an}}) can be d1cd_{1}^{c}–approximated by Phong–Sturm rays of flag configurations.

As far as we are aware, approximations results of this nature have not been explored in the big case before. However in the Kähler case, perhaps the first result of this nature appeared in [Ber19, Corollary 1.3], where Berman devised an approximation scheme for geodesic rays coming from deformations to the normal cone. Another related result from the Kähler case is [DL20, Theorem 1.5]. With the usual difficulties associated with big classes in place, techniques from the Kähler case don’t seem to translate to our setting.

Continuity of the radial Ding functional.

Staying with an arbitrary big class {θ}\{\theta\}, we consider qpsh functions ψ,χ\psi,\chi on XX, with χ\chi having analytic singularity type. After adding some constants to either χ\chi and ψ\psi, one can attach to (χ,ψ)(\chi,\psi) a Radon probability measure, following [BBEGZ19]:

μ:=eχψωn.\mu:=e^{\chi-\psi}\omega^{n}. (1)

Next, following [Din88], one defines the λ\lambda-Ding functional for λ>0\lambda>0. This is 𝒟μλ:1(X,θ)\mathcal{D}_{\mu}^{\lambda}:\mathcal{E}^{1}(X,\theta)\to\mathbb{R}, the λ\lambda-Ding functional:

𝒟μλ(φ)=1λlogXeλφdμIθ(φ) for φ1(X,θ),\displaystyle\mathcal{D}_{\mu}^{\lambda}(\varphi)=-\frac{1}{\lambda}\log\int_{X}e^{-\lambda\varphi}\mathrm{d}\mu-I_{\theta}(\varphi)\text{ for }\varphi\in\mathcal{E}^{1}(X,\theta), (2)

where Iθ()I_{\theta}(\cdot) is the Monge-Ampère energy. For the above definition to make sense, we assume that cμ[Vθ]:=sup{γ0:XeγVθdμ<}>λc_{\mu}[V_{\theta}]:=\sup\{\gamma\geq 0:\int_{X}e^{-\gamma V_{\theta}}\,\mathrm{d}\mu<\infty\}>\lambda.

The Euler–Lagrange equation of the λ\lambda-Ding functional is the following twisted Monge–Ampère equation [DZ22]:

(θ+ddcu)n=eλu+χψωn.(\theta+\mathrm{dd}^{\mathrm{c}}u)^{n}=e^{-\lambda u+\chi-\psi}\omega^{n}. (3)

Solutions to this equation, represent potentials along the continuity method for a twisted KE metric, that solves the above equation in the particular case λ=1\lambda=1 [DZ22].

Despite lack of convexity of 𝒟μλ\mathcal{D}^{\lambda}_{\mu}, [DZ22, Theorem 1.4] gives a formula for the slope of the λ\lambda-Ding functional along subgeodesic rays. Let (0,)tut1(X,θ)(0,\infty)\ni t\mapsto u_{t}\in\mathcal{E}^{1}(X,\theta) be a sublinear subgeodesic ray [DZ22, §3]. Then

lim¯t𝒟μλ(ut)t=limtIθ(ut)t+sup{τ:cμ[u^τ]λ},\varliminf_{t\to\infty}\frac{\mathcal{D}_{\mu}^{\lambda}(u_{t})}{t}=-\lim_{t\to\infty}\frac{I_{\theta}(u_{t})}{t}+\sup\{\tau\ \in\mathbb{R}\ :\ c_{\mu}[\hat{u}_{\tau}]\geq\lambda\}, (4)

where cμ[u]:=sup{γ0:Xeγudμ<}c_{\mu}[u]:=\sup\{\gamma\geq 0:\int_{X}e^{-\gamma u}\,\mathrm{d}\mu<\infty\}. On the heels of the above result it is convenient to introduce

𝒟μλ{ut}:=lim¯t𝒟μλ(ut)t,\mathcal{D}_{\mu}^{\lambda}\{u_{t}\}:=\varliminf_{t\to\infty}\frac{\mathcal{D}_{\mu}^{\lambda}(u_{t})}{t},

and we will call this expression the radial λ\lambda-Ding functional of the subgeodesic {ut}t\{u_{t}\}_{t}. This should be viewed as the transcendental analogue of the βδ\beta_{\delta}-functional in [XZ20, §4].

In our next main result we show that the radial λ\lambda-Ding functional is d1cd_{1}^{c}-continuous.

Theorem 1.4.

(=Theorem 5.9) For any λ(0,cμ[Vθ])\lambda\in(0,c_{\mu}[V_{\theta}]), the functional 𝒟μλ:1(X,θ)\mathcal{D}_{\mu}^{\lambda}:\mathcal{R}^{1}(X,\theta)\to\mathbb{R} is continuous.

In case of numerical classes, there is a precise estimate due to Boucksom–Jonsson for the non-Archimedean Ding functional [BJ22a, Proposition 3.15], implying continuity. We wonder if this estimate can be extended to our radial transcendental setting.

Stability and KE metrics.

We discuss applications to delta invariants and KE metrics for a big class {θ}\{\theta\}. Based on the Blum–Jonsson interpretation of the Fujita–Odaka delta invariant [BJ20, BJ18, FO18], rooted in the non-Archimedean approach to K-stability (see [BBJ21, Definition 7.2]), we recall the definition of the twisted delta invariant of our data [DZ22]:

δμ({θ}):=infEAχ,ψ(E)Sθ(E).\delta_{\mu}(\{\theta\}):=\inf_{E}\frac{A_{\chi,\psi}(E)}{S_{\theta}(E)}. (5)

Here the infimum is taken over all prime divisors EE over XX, i.e., EE is a prime divisor on a modification π:YX\pi:Y\to X of XX (cf. [BBJ21, §B.5]). Also, Aχ,ψ(E):=AX(E)+ν(χ,E)ν(ψ,E)A_{\chi,\psi}(E):=A_{X}(E)+\nu(\chi,E)-\nu(\psi,E), AX(E)A_{X}(E) is the log discrepancy of EE and ν(ψ,E)\nu(\psi,E) denotes the Lelong number of πψ\pi^{*}\psi along EE (cf. [DZ22, (13)]). The expected Lelong number Sθ(E)S_{\theta}(E) of {θ}\{\theta\} along EE is defined by

Sθ(E):=1vol({θ})0τθ(E)vol({πθ}x{E})dx,S_{\theta}(E):=\frac{1}{\operatorname{vol}(\{\theta\})}\int_{0}^{\tau_{\theta}(E)}\operatorname{vol}(\{\pi^{*}\theta\}-x\{E\})\,\mathrm{d}x,

where τθ(E):=sup{τ:{πθ}τ{E} is big}\tau_{\theta}(E):=\sup\{\tau\in\mathbb{R}:\{\pi^{*}\theta\}-\tau\{E\}\textup{ is big}\} is the pseudoeffective threshold, and the volume function vol()\operatorname{vol}(\cdot) is understood in the sense of [BEGZ10].

In [DZ22, Theorem 1.5] it was shown that δμ\delta_{\mu} can be computed as the geodesic semistability threshold of the λ\lambda-Ding functionals. In our next main result we additionally argue that δμ\delta_{\mu} can be computed as the non-Archimedean Ding semistability threshold as well. For the definition of the non-Archimedean Ding functional 𝒟μλ,NA\mathcal{D}^{\lambda,\mathrm{NA}}_{\mu} we refer to (66).

Theorem 1.5.

(=Theorem 5.6, Corollary 5.8) We have the following identities:

δμ\displaystyle\delta_{\mu} =sup{λ>0|𝒟μλ{ut}0 for all sublinear subgeodesic ray ut1(X,θ)}\displaystyle=\sup\{\lambda>0\ |\ \mathcal{D}_{\mu}^{\lambda}\{u_{t}\}\geq 0\text{ for all sublinear subgeodesic ray }u_{t}\in\mathcal{E}^{1}(X,\theta)\} (6)
=sup{λ>0:𝒟μλ{ut}0,{ut}t1(X,θ)}\displaystyle=\sup\left\{\lambda>0:\mathcal{D}_{\mu}^{\lambda}\{u_{t}\}\geq 0,\ \{u_{t}\}_{t}\in\mathcal{R}^{1}(X,\theta)\right\}
=sup{λ>0:𝒟μλ{ut}0,{ut}t1(X,θ)}\displaystyle=\sup\left\{\lambda>0:\mathcal{D}_{\mu}^{\lambda}\{u_{t}\}\geq 0,\ \{u_{t}\}_{t}\in\mathcal{R}^{1}_{\mathcal{I}}(X,\theta)\right\}

When {θ}=c1(L)\{\theta\}=c_{1}(L) for some big line bundle LL on XX, we further have

δμ\displaystyle\delta_{\mu} =sup{λ>0:𝒟μλ{ut}0, rays {ut}t induced by filtrations}\displaystyle=\sup\left\{\lambda>0:\mathcal{D}_{\mu}^{\lambda}\{u_{t}\}\geq 0,\ \forall\textup{ rays }\{u_{t}\}_{t}\textup{ induced by filtrations}\right\}
=sup{λ>0:𝒟μλ{ut}0, rays {ut}t induced by flag configurations}\displaystyle=\sup\left\{\lambda>0:\mathcal{D}_{\mu}^{\lambda}\{u_{t}\}\geq 0,\ \forall\textup{ rays }\{u_{t}\}_{t}\textup{ induced by flag configurations}\right\}
=sup{λ>0:𝒟μλ,NA(u)0,u1,NA(X,θ)}.\displaystyle=\sup\left\{\lambda>0:\mathcal{D}_{\mu}^{\lambda,\mathrm{NA}}(u)\geq 0,\ u\in\mathcal{E}^{1,\mathrm{NA}}(X,\theta)\right\}.

Lastly, we turn to stability and assume that {θ}=c1(KX)\{\theta\}=c_{1}(-K_{X}). For simplicity we only treat the untwisted case, where ψ=0\psi=0 and χ:=fC(X)\chi:=f\in C^{\infty}(X) satisfying θ+ddcf=Ricω\theta+\mathrm{dd}^{\mathrm{c}}f=\operatorname{Ric}\omega.

Definition 1.6.

We say that (X,KX)(X,-K_{X}) is uniformly Ding stable with respect to flag configurations, if there exists ε>0\varepsilon>0 such that

𝒟μ1,NA(u)ε𝒥NA(u),\mathcal{D}_{\mu}^{1,\mathrm{NA}}(u)\geq\varepsilon\mathcal{J}^{\mathrm{NA}}(u), (7)

for any u={ut}t1,NA(X,θ)u=\{u_{t}\}_{t}\in\mathcal{E}^{1,\mathrm{NA}}(X,\theta) induced by flag configurations.

For the definition of the non-Archimedean JJ functional 𝒥μNA\mathcal{J}^{\mathrm{NA}}_{\mu} we refer to (44). We refer to Section 5.4 for more details, and specifically to (70) for an algebraic/valuative interpretation of uniform Ding stability.

Combining Theorems 1.3, 1.4 and 1.5 with the results of [DZ22], we prove a uniform YTD type existence theorem for KE metrics.

Theorem 1.7.

(=Theorem 5.9) Suppose KX-K_{X} is big. If (X,KX)(X,-K_{X}) is uniformly Ding stable with respect to flag configurations then there exists a KE metric. Specifically, there exists uPSH(X,θ)u\in\textup{PSH}(X,\theta) having minimal singularity type such that θu=Ricθu\theta_{u}=\operatorname{Ric}\theta_{u}, i.e.,

θun=efuωn.\theta_{u}^{n}=e^{f-u}\omega^{n}.

It follows from [Xu22, Remark 1.3] and the main result of [LTW21a] that finite generation of the anticanonical section ring and uniform Ding stability with respect to test configurations (as defined by Dervan–Reboulet [DR22] in the big case) also imply existence of KE metrics (for details see [DZ22, Section 6.2]). One advantage of our Theorem 1.7 is that we don’t impose any conditions on finite generation (or KX-K_{X} being klt) apriori. Similar to [DZ22], our approach is purely analytical, but on the heels of [Xu22] it is natural to ask if one could give another proof of Theorem 1.7 using techniques of the minimal model program.

Around the same time our paper appeared on the arXiv, the very intriguing work [Tru23] was published, proving a YTD type existence theorem for Kähler–Einstein metrics with prescribed singularity type. With a different motivation, Trusiani independently developed the theory of relative test curves, overlapping with our results in Section 2.2.

Acknowledgments.

We would like to thank Ruadhai Dervan, Mattias Jonsson, Yaxiong Liu and Pietro Piccione for their comments. The first named author was partially supported by an Alfred P. Sloan Fellowship and National Science Foundation grant DMS–1846942. The second named author is supported by Knut och Alice Wallenbergs Stiftelse grant KAW 2021.0231. The third named author is supported by NSFC grants 12101052, 12271040, and the Fundamental Research Funds 2021NTST10.

2 Preliminaries

2.1 Finite energy pluripotential theory

We give a very brief account of finite energy pluripotential theory in the big case. For a more complete treatment we refer to the papers [BEGZ10], [DDNL18], [DX21, §3], [DZ22, §3] and the textbook [GZ17].

Finite energy pluripotential theory.

Let (X,ω)(X,\omega) be a connected compact Kähler manifold of dimension nn and θ\theta a smooth closed real (1,1)(1,1)-form. A function u:X{}u:X\rightarrow\mathbb{R}\cup\{-\infty\} is called quasi-plurisubharmonic (qpsh) if locally u=ρ+φu=\rho+\varphi, where ρ\rho is smooth and φ\varphi is a plurisubharmonic (psh) function. We say that uu is θ\theta-plurisubharmonic (θ\theta-psh) if it is qpsh and θu:=θ+ddcu0\theta_{u}:=\theta+\mathrm{dd}^{\mathrm{c}}u\geq 0 as currents, with ddc=1¯/2π\mathrm{dd}^{\mathrm{c}}=\sqrt{-1}\partial\bar{\partial}/2\pi. We let PSH(X,θ)\mathrm{PSH}(X,\theta) denote the space of θ\theta-psh functions on XX.

The class {θ}\{\theta\} is big if there exists ψPSH(X,θ)\psi\in\textup{PSH}(X,\theta) satisfying θψεω\theta_{\psi}\geq\varepsilon\omega for some ε>0\varepsilon>0. The class {θ}\{\theta\} is pseudoeffective if {θ+εω}\{\theta+\varepsilon\omega\} is big for any ε>0\varepsilon>0. We assume that {θ}\{\theta\} is pseudoeffective in this section, unless specified otherwise.

Given u,vPSH(X,θ)u,v\in\textup{PSH}(X,\theta), uu is more singular than vv, (notation: uvu\preceq v) if there exists CC\in\mathbb{R} such that uv+Cu\leq v+C. The potential uu has the same singularity as vv (notation: uvu\simeq v) if uvu\preceq v and vuv\preceq u. The classes [u][u] of this latter equivalence relation are called singularity types. When {θ}\{\theta\} is merely big, all elements of PSH(X,θ)\textup{PSH}(X,\theta) are very singular, and we distinguish the potential with minimal singularity:

Vθ:=sup{uPSH(X,θ):u0}.V_{\theta}:=\sup\{u\in\textup{PSH}(X,\theta):u\leq 0\}. (8)

A function uPSH(X,θ)u\in\mathrm{PSH}(X,\theta) is said to have minimal singularity if it has the same singularity type as VθV_{\theta}, i.e., [u]=[Vθ][u]=[V_{\theta}].

We say that [u][u] is an analytic singularity type if it has a representative uPSH(X,θ)u\in\textup{PSH}(X,\theta) that locally can be written as u=clog(j|fj|2)+gu=c\log(\sum_{j}|f_{j}|^{2})+g, where c>0c>0, gg is a bounded function and the fjf_{j} are a finite set of holomorphic functions. By a Demailly’s approximation theorem there are plenty of θ\theta-psh functions with analytic singularity type if {θ}\{\theta\} is big [Dem12, Section 14].

In [BEGZ10] the authors introduce the non-pluripolar Borel measure θun\theta_{u}^{n} of an element uPSH(X,θ)u\in\textup{PSH}(X,\theta), satisfying XθunXθVθn=:vol({θ})\int_{X}\theta_{u}^{n}\leq\int_{X}\theta_{V_{\theta}}^{n}=:\operatorname{vol}(\{\theta\}). It was proved in [WN19, Theorem 1.2] that for any u,vPSH(X,θ)u,v\in\textup{PSH}(X,\theta) we have the following monotonicity result for the masses: if vuv\preceq u then XθvnXθun.\int_{X}\theta_{v}^{n}\leq\int_{X}\theta_{u}^{n}.

We say that uPSH(X,θ)u\in\textup{PSH}(X,\theta) is a full mass potential (u(X,θ)u\in\mathcal{E}(X,\theta)) if Xθun=XθVθn.\int_{X}\theta_{u}^{n}=\int_{X}\theta_{V_{\theta}}^{n}. Moreover, we say that u(X,θ)u\in\mathcal{E}(X,\theta) has finite energy (u1(X,θ)u\in\mathcal{E}^{1}(X,\theta)) if X|uVθ|θun<.\int_{X}|u-V_{\theta}|\theta_{u}^{n}<\infty.

The class 1(X,θ)\mathcal{E}^{1}(X,\theta) plays a central role in the variational theory of complex Monge–Ampère equations, as detailed in [BBEGZ19, BEGZ10] and later works. Here we only mention that the Monge–Ampère energy IθI_{\theta} naturally extends to this space with the usual formula:

Iθ(u)=1vol({θ})(n+1)j=0nX(uVθ)θujθVθnj,u1(X,θ).I_{\theta}(u)=\frac{1}{\operatorname{vol}(\{\theta\})(n+1)}\sum_{j=0}^{n}\int_{X}(u-V_{\theta})\,\theta_{u}^{j}\wedge\theta_{V_{\theta}}^{n-j},\ \ \ \ u\in\mathcal{E}^{1}(X,\theta). (9)

It is upper semi-continuous (usc) with respect to the L1L^{1} topology on PSH(X,θ)\textup{PSH}(X,\theta).

Given any f:X[,+]f:X\to[-\infty,+\infty] one can consider the envelope Pθ(f):=usc(sup{vPSH(X,θ),vf})P_{\theta}(f):=\textup{usc}(\sup\{v\in\textup{PSH}(X,\theta),\ v\leq f\}), where usc\mathrm{usc} denotes the upper-semicontinuous regularization. Then, for u,vPSH(X,θ)u,v\in\textup{PSH}(X,\theta) we can introduce the rooftop envelope Pθ(u,v):=Pθ(min(u,v))P_{\theta}(u,v):=P_{\theta}(\min(u,v)).

With the help of these envelopes one can define a complete metric on 1(X,θ)\mathcal{E}^{1}(X,\theta). Indeed, as pointed out in [DDNL18b, Theorem 2.10], for u,v1(X,θ)u,v\in\mathcal{E}^{1}(X,\theta) we have that P(u,v)1(X,θ)P(u,v)\in\mathcal{E}^{1}(X,\theta) and the following expression defines a complete metric on 1(X,θ)\mathcal{E}^{1}(X,\theta) [DDNL18, Theorem 1.1]:

d1(u,v)=Iθ(u)+Iθ(v)Iθ(P(u,v)).d_{1}(u,v)=I_{\theta}(u)+I_{\theta}(v)-I_{\theta}(P(u,v)).

In addition, d1d_{1}-convergence implies L1L^{1}-convergence of qpsh potentials [DDNL18, Theorem 3.11]. Moreover, any two points of u0,u11(X,θ)u_{0},u_{1}\in\mathcal{E}^{1}(X,\theta) can be connected by a special d1d_{1}-geodesic [0,1]tut1(X,θ)[0,1]\ni t\mapsto u_{t}\in\mathcal{E}^{1}(X,\theta), called the finite energy geodesic segment between u0,u1u_{0},u_{1}.

Envelopes of singularity type.

We now discuss envelopes attached to singularity types, going back to [RWN14] in the Kähler case: given u,vPSH(X,θ)u,v\in\mathrm{PSH}(X,\theta),

Pθ[u](v):=usc(limC+Pθ(u+C,v))=usc(sup{w PSH(X,θ):wv,[w][u]}).P_{\theta}[u](v):=\textup{usc}\Big{(}\lim_{C\to+\infty}P_{\theta}(u+C,v)\Big{)}=\textup{usc}\big{(}\sup\{w\in\textup{ PSH}(X,\theta):w\leq v,\ [w]\preceq[u]\}\big{)}.

It is easy to see that Pθ[u](v)P_{\theta}[u](v) depends only on the singularity type [u][u]. When v=Vθv=V_{\theta}, we simply write P[u]:=Pθ[u]:=Pθ[u](Vθ)P[u]:=P_{\theta}[u]:=P_{\theta}[u](V_{\theta}) and call this potential the envelope of the singularity type [u][u]. It follows from [DDNL18a, Theorem 3.8] that θP[u]n𝟙{P[u]=0}θn\theta_{P[u]}^{n}\leq\mathbbm{1}_{\{P[u]=0\}}\theta^{n}. Also, by [DDNL18a, Proposition 2.3 and Remark 2.5] we have that XθP[u]n=Xθun\int_{X}\theta_{P[u]}^{n}=\int_{X}\theta_{u}^{n}.

An algebraic counterpart of P[u](v)P[u](v) is the operator P[u](v)PSH(X,θ)P[u]_{\mathcal{I}}(v)\in\textup{PSH}(X,\theta) defined by

P[u](v):=sup{w PSH(X,θ):wv,(tw)(tu),t0}.P[u]_{\mathcal{I}}(v):=\sup\left\{w\in\textup{ PSH}(X,\theta):w\leq v,\mathcal{I}(tw)\subseteq\mathcal{I}(tu),t\geq 0\right\}.

Here (tu)\mathcal{I}(tu) is a multiplier ideal sheaf, locally generated by holomorphic functions ff such that |f|2etu|f|^{2}e^{-tu} is integrable. This envelope is implicit in [KS20], and was introduced and studied in detail in [DX22, §2.4], [Tru20], [DX21]. Again, P[u]:=Pθ[u](Vθ)P[u]_{\mathcal{I}}:=P_{\theta}[u]_{\mathcal{I}}(V_{\theta}), and it is not difficult to see that P[u]P[u]P[u]_{\mathcal{I}}\geq P[u] for any uPSH(X,θ)u\in\textup{PSH}(X,\theta).

We say uPSH(X,θ)u\in\textup{PSH}(X,\theta) is a model potential if Pθ[u]=uP_{\theta}[u]=u; it is an \mathcal{I}-model potential if Pθ[u]=uP_{\theta}[u]_{\mathcal{I}}=u. By [DX22, Proposition 2.18(i)], an \mathcal{I}-model potential is always a model potential.

We note the following technical result about model potentials that will be needed later.

Lemma 2.1.

Let φjPSH(X,θ)\varphi^{j}\in\mathrm{PSH}(X,\theta) be a decreasing sequence of model potentials. Let φjφPSH(X,θ)\varphi^{j}\searrow\varphi\in\textup{PSH}(X,\theta). If φ\varphi has positive mass then for any prime divisor EE over XX,

limjν(φj,E)=ν(φ,E).\lim_{j\to\infty}\nu(\varphi^{j},E)=\nu(\varphi,E).
Proof.

By [DDNL21, Proposition 4.6] we have that XθφjnXθφn.\int_{X}\theta_{\varphi^{j}}^{n}\searrow\int_{X}\theta_{\varphi}^{n}. Hence, by [DDNL21, Lemma 4.3], for any ε(0,1)\varepsilon\in(0,1) and jj big enough there exists ψjPSH(X,θ)\psi^{j}\in\mathrm{PSH}(X,\theta) such that (1ε)φj+εψjφ(1-\varepsilon)\varphi^{j}+\varepsilon\psi^{j}\leq\varphi. This implies that for jj big enough we have

(1ε)ν(φj,E)+εν(ψj,E)ν(φ,E)ν(φj,E).(1-\varepsilon)\nu(\varphi^{j},E)+\varepsilon\nu(\psi^{j},E)\geq\nu(\varphi,E)\geq\nu(\varphi^{j},E)\,.

However, for EE fixed, ν(χ,E)\nu(\chi,E) is uniformly bounded (see, e.g., [Bou02, Lemma 2.10]) for any χPSH(X,θ)\chi\in\mathrm{PSH}(X,\theta). So letting ε0\varepsilon\searrow 0 we conclude. ∎

Finally, we state an effective version of Guan–Zhou’s strong openness theorem that will be used multiple times in this work (see [GZ15], [BBJ21], c.f. [DZ22, Theorem 2.2]).

Theorem 2.2.

Suppose that there are qpsh functions on XX such that ujuu_{j}\nearrow u a.e. If f(u)f\in\mathcal{I}(u), then f(uj)f\in\mathcal{I}(u_{j}) for all jj big enough. More generally, if χ\chi is a qpsh function with analytic singularities satisfying eχuL1(X)e^{\chi-u}\in L^{1}(X), then eχujL1(X)e^{\chi-u_{j}}\in L^{1}(X) for all jj big enough.

2.2 The theory of relative test curves

Let XX be a connected compact Kähler manifold of dimension nn. Let θ\theta be a closed smooth real (1,1)(1,1)-form on XX representing a big cohomology class {θ}\{\theta\}. Set V=XθVθnV=\int_{X}\theta_{V_{\theta}}^{n}.

Test curves.

A map τψτPSH(X,θ)\mathbb{R}\ni\tau\mapsto\psi_{\tau}\in\mathrm{PSH}(X,\theta) is a relative test curve, denoted {ψτ}τ\{\psi_{\tau}\}_{\tau}, if τψτ(x)\tau\mapsto\psi_{\tau}(x) is concave, decreasing and usc for any xXx\in X. Moreover, ψτ\psi_{\tau}\equiv-\infty for all τ\tau big enough and the weak L1L^{1} limit ψ:=limτψτPSH(X,θ)\psi_{-\infty}:=\lim_{\tau\to-\infty}\psi_{\tau}\in\textup{PSH}(X,\theta) exists. We say that {ψτ}τ\{\psi_{\tau}\}_{\tau} is a relative test curve relative to ψ\psi_{-\infty}. If ψ=Vθ\psi_{-\infty}=V_{\theta} we simply call {ψτ}τ\{\psi_{\tau}\}_{\tau} a test curve, and this particular case was studied in detail in [DDNL18, DZ22], having its origins in [RWN14], in the ample case.

The above definition allows to introduce the following constant:

τψ+:=inf{τ:ψτ}.\tau_{\psi}^{+}:=\inf\{\tau\in\mathbb{R}:\psi_{\tau}\equiv-\infty\}\,. (10)

Using the convention P[]=P[-\infty]=-\infty, a relative test curve {ψτ}τ\{\psi_{\tau}\}_{\tau} can have the following properties:
(i) {ψτ}τ\{\psi_{\tau}\}_{\tau} is maximal if P[ψτ]=ψτP[\psi_{\tau}]=\psi_{\tau} for any τ\tau\in\mathbb{R}.
(ii) A relative test curve {ψτ}τ\{\psi_{\tau}\}_{\tau} is a finite energy test curve if

Iθ{ψτ}:=τψ++1Vτψ+(XθψτnXθVθn)dτ>.I_{\theta}\{\psi_{\tau}\}:=\tau^{+}_{\psi}+\frac{1}{V}\int_{-\infty}^{\tau^{+}_{\psi}}\left(\int_{X}\theta_{\psi_{\tau}}^{n}-\int_{X}\theta_{V_{\theta}}^{n}\right)\,\mathrm{d}\tau>-\infty\,. (11)

(iii) We say that {ψτ}τ\{\psi_{\tau}\}_{\tau} is bounded if ψτ=ψ\psi_{\tau}=\psi_{-\infty} for all τ\tau small enough.
(iv) {ψτ}τ\{\psi_{\tau}\}_{\tau} is \mathcal{I}-maximal if P[ψτ]=ψτP[\psi_{\tau}]_{\mathcal{I}}=\psi_{\tau} for any τ\tau\in\mathbb{R}. Since ψτP[ψτ]P[ψτ]\psi_{\tau}\leq P[\psi_{\tau}]\leq P[\psi_{\tau}]_{\mathcal{I}}, we notice that {ψτ}τ\{\psi_{\tau}\}_{\tau} is maximal if it is \mathcal{I}-maximal.

Subgeodesic rays.

A psh subgeodesic ray, denoted {ut}t\{u_{t}\}_{t}, is an assignment (0,)tutPSH(X,θ)(0,\infty)\ni t\mapsto u_{t}\in\mathrm{PSH}(X,\theta) such that

Φ(x,ξ):=ulog|ξ|2(x)\Phi(x,\xi):=u_{-\log|\xi|^{2}}(x) (12)

is p1θp_{1}^{*}\theta-plurisubharmonic on X×ΔX\times\Delta^{*}, where Δ\Delta\subset\mathbb{C} is the unit disk and Δ=Δ{0}\Delta^{*}=\Delta\setminus\{0\}, and p1:X×ΔXp_{1}:X\times\Delta\rightarrow X is the projection.

A sublinear subgeodesic ray is a psh subgeodesic ray {ut}t\{u_{t}\}_{t} such that the weak L1L^{1} limit u0:=limt0+utPSH(X,θ)u_{0}:=\lim_{t\to 0+}u_{t}\in\textup{PSH}(X,\theta) exists and for some CC\in\mathbb{R} we have

utu0+Ct.u_{t}\leq u_{0}+Ct.

In this case we say that {ut}t\{u_{t}\}_{t} is a sublinear subgeodesic emanating from u0u_{0}.

A psh geodesic segment in PSH(X,θ)\mathrm{PSH}(X,\theta) from φ\varphi to ψ\psi is a map [a,b]tutPSH(X,θ)[a,b]\ni t\mapsto u_{t}\in\mathrm{PSH}(X,\theta) such that Φ(x,z):=ulog|z|2(x)\Phi(x,z):=u_{-\log|z|^{2}}(x) is πθ\pi^{*}\theta-psh on X×{z:exp(b)<|z|2<exp(a)}X\times\{z\in\mathbb{C}:\exp(-b)<|z|^{2}<\exp(-a)\}. Moreover, for any S1S^{1}-invariant p1θp_{1}^{*}\theta-psh function Ψ\Psi on X×{z:exp(b)<|z|2<exp(a)}X\times\{z\in\mathbb{C}:\exp(-b)<|z|^{2}<\exp(-a)\} such that

lim¯tbΨ(,eb)ψ,lim¯ta+Ψ(,ea)φ\varlimsup_{t\to b-}\Psi(\cdot,\mathrm{e}^{-b})\leq\psi,\quad\varlimsup_{t\to a+}\Psi(\cdot,\mathrm{e}^{-a})\leq\varphi (13)

almost everywhere, we have ΨΦ\Psi\leq\Phi. Furthermore, Φ\Phi assumes the correct boundary values:

limtbΦ(,eb)=ψ,limta+Φ(,ea)=φ.\lim_{t\to b-}\Phi(\cdot,\mathrm{e}^{-b})=\psi,\quad\lim_{t\to a+}\Phi(\cdot,\mathrm{e}^{-a})=\varphi.

A psh geodesic ray is a sublinear subgeodesic ray tutt\mapsto u_{t} such that [a,b]tutPSH(X,θ)[a,b]\ni t\mapsto u_{t}\in\textup{PSH}(X,\theta) is a psh geodesic segment for all a,b(0,)a,b\in(0,\infty).

The geometry of finite energy rays

A finite energy geodesic ray [0,)tut1(X,θ)[0,\infty)\ni t\mapsto u_{t}\in\mathcal{E}^{1}(X,\theta) is simply a psh geodesic ray with u0=Vθu_{0}=V_{\theta}. The space of such finite energy rays, denoted 1(X,θ)\mathcal{R}^{1}(X,\theta), was studied in [DDNL21] (see [DL20] for the Kähler case).

Using linearity/convexity one can define the chordal metric between two rays:

d1c({ut}t,{vt}t)=limtd1(ut,vt)t,{ut}t,{vt}t1(X,θ).d_{1}^{c}(\{u_{t}\}_{t},\{v_{t}\}_{t})=\lim_{t\to\infty}\frac{d_{1}(u_{t},v_{t})}{t},\quad\{u_{t}\}_{t},\{v_{t}\}_{t}\in\mathcal{R}^{1}(X,\theta).

It was shown in [DDNL21, Theorem 2.14] that (1(X,θ),d1c)(\mathcal{R}^{1}(X,\theta),d_{1}^{c}) is a complete metric space. Given any ray {ut}t1(X,θ)\{u_{t}\}_{t}\in\mathcal{R}^{1}(X,\theta), we define its radial energy as

Iθ{ut}:=limtIθ(ut)t.I_{\theta}\{u_{t}\}:=\lim_{t\to\infty}\frac{I_{\theta}(u_{t})}{t}. (14)

The limit exists as Iθ(ut)I_{\theta}(u_{t}) is linear in tt.

As in the case of the metric space structure of (1(X,θ),d1)(\mathcal{E}^{1}(X,\theta),d_{1}), for any rays {ut}t,{vt}t1(X,θ)\{u_{t}\}_{t},\{v_{t}\}_{t}\in\mathcal{R}^{1}(X,\theta) one can define {utvt}t1(X,θ)\{u_{t}\land v_{t}\}_{t}\in\mathcal{R}^{1}(X,\theta) (resp. {utvt}t1(X,θ)\{u_{t}\lor v_{t}\}_{t}\in\mathcal{R}^{1}(X,\theta)), the biggest (resp. smallest) ray satisfying utvtmin(ut,vt)u_{t}\wedge v_{t}\leq\min(u_{t},v_{t}) (resp. utvtmax(ut,vt)u_{t}\wedge v_{t}\geq\max(u_{t},v_{t})) for all t0t\geq 0.

These two special rays satisfy the Pythagorean formula and Pythagorean inequality respectively ([Xia21, Theorem 1.3], [DDNL21, Proposition 2.15]), for C=C(n)C=C(n):

d1c({ut}t,{vt}t)=\displaystyle d_{1}^{c}(\{u_{t}\}_{t},\{v_{t}\}_{t})= d1c({ut}t,{utvt}t)+d1c({utvt}t,{vt}t),\displaystyle d_{1}^{c}(\{u_{t}\}_{t},\{u_{t}\wedge v_{t}\}_{t})+d_{1}^{c}(\{u_{t}\wedge v_{t}\}_{t},\{v_{t}\}_{t}), (15)
Cd1c({ut}t,{vt}t)\displaystyle Cd_{1}^{c}(\{u_{t}\}_{t},\{v_{t}\}_{t})\geq d1c({ut}t,{utvt}t)+d1c({utvt}t,{vt}t).\displaystyle d_{1}^{c}(\{u_{t}\}_{t},\{u_{t}\vee v_{t}\}_{t})+d_{1}^{c}(\{u_{t}\vee v_{t}\}_{t},\{v_{t}\}_{t}).

We note the following result, that is the radial analogue of [BDL17, Proposition 2.6]:

Proposition 2.3.

Let {utk}t,{ut}1(X,θ)\{u^{k}_{t}\}_{t},\{u_{t}\}\in\mathcal{R}^{1}(X,\theta) such that d1c({utk}t,{ut}t)0d_{1}^{c}(\{u^{k}_{t}\}_{t},\{u_{t}\}_{t})\to 0. Then there exists a subsequence, again denoted {utk}t\{u^{k}_{t}\}_{t} and {vtk}t,{wtk}t1(X,θ)\{v^{k}_{t}\}_{t},\{w^{k}_{t}\}_{t}\in\mathcal{R}^{1}(X,\theta) such that:
(i) wtkutkvtkw^{k}_{t}\leq u^{k}_{t}\leq v^{k}_{t} for any fixed t0t\geq 0,
(ii) wtkutw^{k}_{t}\nearrow u_{t} and vtkutv^{k}_{t}\searrow u_{t} for any fixed t0t\geq 0,
(iii) d1c({wtk}t,{ut}t)0d_{1}^{c}(\{w^{k}_{t}\}_{t},\{u_{t}\}_{t})\to 0 and d1c({vtk}t,{ut}t)0d_{1}^{c}(\{v^{k}_{t}\}_{t},\{u_{t}\}_{t})\to 0.

Proof.

The proof follows the argument of [BDL17, Proposition 2.6] in case of potentials. In fact, the sequence {wtk}\{w^{k}_{t}\} is constructed using the exact same ideas, with (15) used in the appropriate places.

The sequence of {vtk}\{v^{k}_{t}\} is constructed following the argument of [DDNL21, Proposition 4.2] word-for-word, using the radial Pythagorean inequality (15) in the appropriate places. ∎

The Ross–Witt Nyström correspondence.

We adopt the following convention: relative test curves will always be parametrized by τ\tau, whereas rays will be parametrized by tt. Hence, {ψt}t\{\psi_{t}\}_{t} will always refer to some kind of ray, whereas {ϕτ}τ\{\phi_{\tau}\}_{\tau} will refer to some type of test curve. As we will now point out, rays and test curves are dual to each other, so one should think of the parameters tt and τ\tau to be dual to each other as well.

To any sublinear subgeodesic ray {ϕt}t\{\phi_{t}\}_{t} (relative test curve {ψτ}τ\{\psi_{\tau}\}_{\tau}) we can associate its (inverse) partial Legendre transform at xXx\in X as

ϕ^τ(x):=inft>0(ϕt(x)tτ),\displaystyle\hat{\phi}_{\tau}(x):=\inf_{t>0}(\phi_{t}(x)-t\tau)\,, τ,\displaystyle\quad\tau\in\mathbb{R}\,, (16)
ψˇt(x):=supτ(ψτ(x)+tτ),\displaystyle\check{\psi}_{t}(x):=\sup_{\tau\in\mathbb{R}}(\psi_{\tau}(x)+t\tau)\,, t>0.\displaystyle\quad t>0\,.

We say that a ray {ut}t1(X,θ)\{u_{t}\}_{t}\in\mathcal{R}^{1}(X,\theta) is \mathcal{I}-maximal if {u^τ}τ\{\hat{u}_{\tau}\}_{\tau} is \mathcal{I}-maximal. The set of \mathcal{I}-maximal finite energy rays is denoted by 1(X,θ)\mathcal{R}^{1}_{\mathcal{I}}(X,\theta). As we will see later, there is a bijection between 1(X,θ)\mathcal{R}^{1}_{\mathcal{I}}(X,\theta) and the set of non-Archimedean finite energy potentials.

Lemma 2.4.

Let {ψτ}τ\{\psi_{\tau}\}_{\tau} be a test curve relative to ϕ\phi. Then supτ(ψτ(x)+tτ)\sup_{\tau}(\psi_{\tau}(x)+t\tau) is usc with respect to (t,x)(0,)×X(t,x)\in(0,\infty)\times X. In particular, ψˇtPSH(X,θ)\check{\psi}_{t}\in\mathrm{PSH}(X,\theta) for all t>0t>0, and {ψˇt}t\{\check{\psi}_{t}\}_{t} is a sublinear subgeodesic ray, emanating from ϕ\phi.

Proof.

The argument is exactly the same as in the particular case when ϕ=Vθ\phi=V_{\theta}, proved in [DZ22, Proposition 3.6]. ∎

We note the following simple result, that will have important consequences in the sequel.

Lemma 2.5.

Let {ψτ}τ\{\psi_{\tau}\}_{\tau} be a test curve relative to ϕ\phi, with Xθϕn>0\int_{X}\theta_{\phi}^{n}>0. Then P[ψˇt]=P[ϕ]P[\check{\psi}_{t}]=P[\phi] for all t>0t>0.

Proof.

We may assume that ϕ0\phi\leq 0, and we observe that ψτ+tτψˇtϕ+tτψ+\psi_{\tau}+t\tau\leq\check{\psi}_{t}\leq\phi+t\tau^{+}_{\psi} for all τ<τψ+\tau<\tau^{+}_{\psi} and t>0t>0. Taking envelopes with respect to singularity type, we find

P[ψτ]P[ψˇt]P[ϕ].P[\psi_{\tau}]\leq P[\check{\psi}_{t}]\leq P[\phi]. (17)

But for almost all xXx\in X, ϕ(x)=limτψτ(x)limτP[ψτ].\phi(x)=\lim_{\tau\to-\infty}\psi_{\tau}(x)\leq\lim_{\tau\to-\infty}P[\psi_{\tau}]. It follows that for almost all xXx\in X, ϕ(x)P[ψˇt](x).\phi(x)\leq P[\check{\psi}_{t}](x). As both sides are θ\theta-psh, it follows that the inequality holds everywhere. As both sides have positive mass, we get that P[ϕ]P[ψˇt]P[\phi]\leq P[\check{\psi}_{t}] due to [DDNL18a, Theorem 3.12]. Together with (17), we conclude. ∎

Lastly we state the versions of the Ross–Witt Nyström correspondence between test curves and rays that will be needed in this work:

Theorem 2.6.

Let ϕPSH(X,θ)\phi\in\textup{PSH}(X,\theta) with Xθϕn>0\int_{X}\theta_{\phi}^{n}>0. The following hold:
(i) The map {ψτ}{ψˇt}t\{\psi_{\tau}\}\mapsto\{\check{\psi}_{t}\}_{t} gives a bijection from the set of test curves relative to ϕ\phi to the set of sublinear subgeodesic rays in PSH(X,θ)\mathrm{PSH}(X,\theta) emanating from ϕ\phi, with inverse {ut}t{u^τ}τ\{u_{t}\}_{t}\mapsto\{\hat{u}_{\tau}\}_{\tau}.
(ii) The map of (i) gives a bijection between the set of maximal test curves relative to ϕ\phi, and the set of psh geodesic rays in PSH(X,θ)\mathrm{PSH}(X,\theta) emanating from ϕ\phi.
(iii) The map of (i) gives a bijection between the set of finite energy maximal test curves, and the set of finite energy psh geodesic rays 1(X,θ)\mathcal{R}^{1}(X,\theta). In addition, under this correspondence, the radial IθI_{\theta} functional (14) is equal to the energy of the test curve (11).

Proof.

Using Proposition 2.4, the proof of the first and second part are carried out in exactly the same manner as the particular case ϕ=Vθ\phi=V_{\theta} [DZ22, Theorem 3.7(i)(ii)]. The last part of the theorem is simply [DZ22, Theorem 3.9]. ∎

Test curves relative to potentials ϕ\phi with Xθϕn=0\int_{X}\theta_{\phi}^{n}=0 exhibit pathological behaviour (c.f. [DDNL21, Section 4]). To exclude these we introduce:

MTC(X,θ)={{ϕτ}τPSH(X,θ) relative test with Xϕn>0},\textup{MTC}(X,\theta)=\left\{\{\phi_{\tau}\}_{\tau}\subset\textup{PSH}(X,\theta)\textup{ relative test with }\int_{X}\phi_{-\infty}^{n}>0\right\}, (18)

with “MTC” indicating maximal test curve. Similarly, the test curves in MTC(X,θ)\textup{MTC}(X,\theta) that are \mathcal{I}-maximal will be denoted by MTC(X,θ)\textup{MTC}_{\mathcal{I}}(X,\theta). For now, we record the following simple observation, that already uses the positive mass assumption crucially.

Lemma 2.7.

Suppose that {ψτ}τMTC(X,θ)\{\psi_{\tau}\}_{\tau}\in\textup{MTC}(X,\theta). Then ψ=P[ψ]\psi_{-\infty}=P[\psi_{-\infty}].

Proof.

We have that ψτψ\psi_{\tau}\nearrow\psi_{-\infty} a.e., as τ\tau\to-\infty and ψτ=P[ψτ]\psi_{\tau}=P[\psi_{\tau}]. By [DDNL21, Corollary 4.7] we conclude that ψ=P[ψ]\psi_{-\infty}=P[\psi_{-\infty}]. ∎

The next two results show \mathcal{I}-maximal test curves are preserved under monotone limits.

Lemma 2.8.

Let {ψτi}τMTC(θ)\{\psi^{i}_{\tau}\}_{\tau}\in\textup{MTC}_{\mathcal{I}}(\theta) be a decreasing net (in the sense that ψτi\psi^{i}_{\tau} is decreasing for each τ\tau). Let ψτ:=infiψτi\psi_{\tau}:=\inf_{i}\psi^{i}_{\tau}. Assume that for some τ\tau\in\mathbb{R}, ψτ\psi_{\tau} is not identically -\infty, and it has positive mass. Then {ψτ}τMTC(θ)\{\psi_{\tau}\}_{\tau}\in\textup{MTC}_{\mathcal{I}}(\theta).

Proof.

Since ψψτ\psi_{-\infty}\geq\psi_{\tau}, we get that ψ\psi_{-\infty} has positive mass by [WN19, Theorem 1.2]. It suffices to observe that ψτ\psi_{\tau} is \mathcal{I}-model whenever it is not -\infty. Indeed, ψτPθ[ψτ]infiPθ[ψτi]=infiψτi=ψτ.\psi_{\tau}\leq P_{\theta}[\psi_{\tau}]_{\mathcal{I}}\leq\inf_{i}P_{\theta}[\psi^{i}_{\tau}]_{\mathcal{I}}=\inf_{i}\psi^{i}_{\tau}=\psi_{\tau}.

Lemma 2.9.

Let {ψτi}τMTC(θ)\{\psi^{i}_{\tau}\}_{\tau}\in\textup{MTC}_{\mathcal{I}}(\theta) be an increasing net in ii. Assume that τψ+:=supiτψi,+<\tau^{+}_{\psi}:=\sup_{i}\tau^{i,+}_{\psi}<\infty. Let ψτ:=usc(supiψτi)\psi_{\tau}:=\textup{usc}\big{(}\sup_{i}\psi^{i}_{\tau}\big{)} for ττψ+\tau\neq\tau^{+}_{\psi} and ψτψ+:=limττψ+ψτ\psi_{\tau^{+}_{\psi}}:=\lim_{\tau\nearrow\tau^{+}_{\psi}}\psi_{\tau}. Then {ψτ}τMTC(θ)\{\psi_{\tau}\}_{\tau}\in\textup{MTC}_{\mathcal{I}}(\theta).

Proof.

Showing that ψτ\psi_{\tau} is \mathcal{I}-model and has positive mass for τ<τψ+\tau<\tau^{+}_{\psi} is a consequence of [WN19, Theorem 1.2] and [DX22, Lemma 2.21(iii)]. That ψτψ+\psi_{\tau^{+}_{\psi}} is \mathcal{I}-model follows from [DX22, Lemma 2.21(i)]. ∎

Maximization.

We adapt the maximization process of test curves from the works [RWN14, DDNL18, DZ22] to our relative setting. Let {ψτ}τ\{\psi_{\tau}\}_{\tau} be a relative test curve with Xθψn>0\int_{X}\theta_{\psi_{-\infty}}^{n}>0. The maximization of {ψτ}τ\{\psi_{\tau}\}_{\tau} is simply the relative test curve {ϕτ}τ\{\phi_{\tau}\}_{\tau} such that

ϕτ:={P[ψτ],if τ<τψ+;limττψ+ϕτ,if τ=τψ+;,if τ>τψ+.\phi_{\tau}:=\left\{\begin{aligned} P[\psi_{\tau}],&\quad\text{if }\tau<\tau^{+}_{\psi};\\ \lim_{\tau\nearrow\tau^{+}_{\psi}}\phi_{\tau},&\quad\text{if }\tau=\tau^{+}_{\psi};\\ -\infty,&\quad\text{if }\tau>\tau^{+}_{\psi}.\end{aligned}\right. (19)

The condition Xθψn>0\int_{X}\theta_{\psi_{-\infty}}^{n}>0, [WN19, Theorem 1.2] and τ\tau-concavity of τψτ\tau\mapsto\psi_{\tau} implies that Xθψτn>0\int_{X}\theta_{\psi_{\tau}}^{n}>0 for all τ<τψ+\tau<\tau^{+}_{\psi}. By [DDNL18a, Theorem 3.12] we obtain that {ϕτ}τ\{\phi_{\tau}\}_{\tau} is a maximal test curve, relative to ϕ=P[ψ]\phi_{-\infty}=P[\psi_{-\infty}] (by [DDNL21, Corollary 4.7]).

Lastly, along the same lines as above, we introduce the \mathcal{I}-maximization of a relative test curve {ψτ}τ\{\psi_{\tau}\}_{\tau} with Xθψn>0\int_{X}\theta_{\psi_{-\infty}}^{n}>0. The \mathcal{I}-maximization of {ψτ}τ\{\psi_{\tau}\}_{\tau} is the relative test curve {χτ}τ\{\chi_{\tau}\}_{\tau} defined by:

χτ:={P[ψτ],if τ<τψ+;limττψ+χτ,if τ=τψ+;,if τ>τψ+.\chi_{\tau}:=\left\{\begin{aligned} P[\psi_{\tau}]_{\mathcal{I}},&\quad\text{if }\tau<\tau^{+}_{\psi};\\ \lim_{\tau\nearrow\tau^{+}_{\psi}}\chi_{\tau},&\quad\text{if }\tau=\tau^{+}_{\psi};\\ -\infty,&\quad\text{if }\tau>\tau^{+}_{\psi}.\end{aligned}\right. (20)

By definition, {χτ}τ\{\chi_{\tau}\}_{\tau} is an \mathcal{I}-maximal test curve relative to χ=P[ψ]\chi_{-\infty}=P[\psi_{-\infty}]_{\mathcal{I}} by [DX22, Lemma 2.21(iii)].

The maximization/\mathcal{I}-maximization of a sublinear (sub)geodesic {ut}t\{u_{t}\}_{t} will simply be the inverse Legendre transform of the maximization/\mathcal{I}-maximization of {u^τ}τ\{\hat{u}_{\tau}\}_{\tau}.

2.3 The non-Archimedean formalism of Boucksom–Jonsson

In this section, we recall the basics of the space of non-Archimedean psh metrics, as defined by Boucksom–Jonsson [BJ21].

The space of valuations.

Let XX be an irreducible reduced variety over \mathbb{C} of dimension nn. We recall the notion of Berkovich analytification XanX^{\mathrm{an}} of XX with respect to the trivial valuation on \mathbb{C}. Recall that a (real-valued) valuation on XX (or a valuation of (X)\mathbb{C}(X)) is a map v:(X)(,]v:\mathbb{C}(X)\rightarrow(-\infty,\infty] satisfying
(i) For f(X)f\in\mathbb{C}(X), v(f)=v(f)=\infty if and only if f=0f=0.
(ii) For f,g(X)f,g\in\mathbb{C}(X), v(fg)=v(f)+v(g)v(fg)=v(f)+v(g).
(iii) For f,g(X)f,g\in\mathbb{C}(X), v(f+g)min{v(f),v(g)}v(f+g)\geq\min\{v(f),v(g)\}.
The set of valuations on XX is denoted by XvalX^{\mathrm{val}}. The center of a valuation vv is the scheme-theoretic point c=c(v)c=c(v) of XX such that v0v\geq 0 on 𝒪X,c\mathcal{O}_{X,c} and v>0v>0 on the maximal ideal 𝔪X,c\mathfrak{m}_{X,c} of 𝒪X,c\mathcal{O}_{X,c}. The center is unique if exists. It exists if XX is proper.

In the remaining of this section, we assume that XX is projective.

As a set, XanX^{\mathrm{an}} is the set of semi-valuations on XX, in other words, real-valued valuations vv on irreducible reduced subvarieties YY in XX that is trivial on \mathbb{C}. We call YY the support of the semi-valuation vv. In other words,

Xan=YYval.X^{\mathrm{an}}=\coprod_{Y}Y^{\mathrm{val}}.

We will write vtrivXanv_{\mathrm{triv}}\in X^{\mathrm{an}} for the trivial valuation on XX: vtriv(f)=0v_{\mathrm{triv}}(f)=0 for any f(X)×f\in\mathbb{C}(X)^{\times}. We endow XanX^{\mathrm{an}} with the coarsest topology such that :
(i) For any Zariski open subset UXU\subseteq X, the subset UanU^{\mathrm{an}} of XanX^{\mathrm{an}} consisting of semi-valuations whose supports meet UU is open.
(ii) For each Zariski open subset UXU\subseteq X and each fH0(U,𝒪X)f\in H^{0}(U,\mathcal{O}_{X}) (here 𝒪X\mathcal{O}_{X} is the sheaf of regular functions), the map |f|:Uan|f|:U^{\mathrm{an}}\rightarrow\mathbb{R} sending vv to exp(v(f))\exp(-v(f)) is continuous.
See [Ber93] for more details.

We will be most interested in divisorial valuations. Recall that a divisorial valuation on XX is a valuation of the form cordEc\operatorname{ord}_{E}, where c>0c\in\mathbb{Q}_{>0} and EE is a prime divisor over XX. The set of divisorial valuations on XX is denoted by XdivX^{\mathrm{div}}. When >0\mathbb{Q}_{>0} is replaced by >0\mathbb{R}_{>0}, we can similarly define a space XdivX^{\mathrm{div}}_{\mathbb{R}}.

Given any coherent ideal 𝔞\mathfrak{a} on XX and any vXanv\in X^{\mathrm{an}}, we define

v(𝔞):=min{v(f):f𝔞c(v)}[0,],v(\mathfrak{a}):=\min\{v(f):f\in\mathfrak{a}_{c(v)}\}\in[0,\infty], (21)

where c(v)c(v) is the center of the valuation vv on XX.

Given any valuation vv on XX, the Gauss extension of vv is a valuation σ(v)\sigma(v) on X×𝔸1X\times\mathbb{A}^{1}:

σ(v)(ifiti):=mini(v(fi)+i).\sigma(v)\left(\sum_{i}f_{i}t^{i}\right):=\min_{i}(v(f_{i})+i).

Here tt is the standard coordinate on 𝔸1=Spec[t]\mathbb{A}^{1}=\operatorname{Spec}\mathbb{C}[t]. The key property is that when vv is a divisorial valuation, then so it σ(v)\sigma(v). See [BHJ17, Lemma 4.2].

Non-Archimedean plurisubharmonic functions.

Let XX be an irreducible complex projective variety of dimension nn and LL be a holomorphic pseudoeffective \mathbb{Q}-line bundle on XX. Through the GAGA morphism XanXX^{\mathrm{an}}\rightarrow X of ringed spaces, LL can be pulled-back to an analytic line bundle LanL^{\mathrm{an}} on XX. The purpose of this section is to study the psh metrics on LanL^{\mathrm{an}}. We will follow the approach of [BJ21], which avoids the direct treatment of LanL^{\mathrm{an}} itself.

Following [BJ21, Definition 2.18], we define gf(L)\mathcal{H}^{\mathrm{gf}}_{\mathbb{Q}}(L), the set of (rational) generically finite Fubini–Study functions ϕ:Xan[,)\phi:X^{\mathrm{an}}\rightarrow[-\infty,\infty), that are of the following form:

ϕ=1mmaxj{log|sj|+λj}.\phi=\frac{1}{m}\max_{j}\{\log|s_{j}|+\lambda_{j}\}. (22)

Here m>0m\in\mathbb{Z}_{>0} is an integer such that LmL^{m} is a line bundle, the sjs_{j}’s are a finite collection of non-vanishing sections in H0(X,Lm)H^{0}(X,L^{m}), and λj\lambda_{j}\in\mathbb{Q}. We followed the convention of Boucksom–Jonsson by writing log|sj|(v)=v(sj)\log|s_{j}|(v)=-v(s_{j}).

Now we come to the main definition of this paragraph:

Definition 2.10 ([BJ21, Definition 4.1]).

A psh metric on LanL^{\mathrm{an}} is a function ϕ:Xan[,)\phi:X^{\mathrm{an}}\rightarrow[-\infty,\infty) that is not identically -\infty, and is the pointwise limit of a decreasing net (ϕi)iI(\phi_{i})_{i\in I}, where ϕigf(Lian)\phi_{i}\in\mathcal{H}^{\mathrm{gf}}_{\mathbb{Q}}(L_{i}^{\mathrm{an}}) for some \mathbb{Q}-line bundles LiL_{i} on XX satisfying c1(Li)c1(L)c_{1}(L_{i})\to c_{1}(L) in NS1(X)\mathrm{NS}^{1}(X)_{\mathbb{R}}.

The set of psh metrics on LanL^{\mathrm{an}} is denoted by PSH(Lan)\mathrm{PSH}(L^{\mathrm{an}}). We endow PSH(Lan)\mathrm{PSH}(L^{\mathrm{an}}) with the topology of pointwise convergence on XdivX^{\mathrm{div}}. This topology is Hausdorff as functions in PSH(Lan)\mathrm{PSH}(L^{\mathrm{an}}) are completely determined by their restriction on XdivX^{\mathrm{div}}:

Theorem 2.11 ([BJ21, Theorem 4.22]).

Let ϕPSH(Lan)\phi\in\mathrm{PSH}(L^{\mathrm{an}}) and ψ:Xan[,)\psi:X^{\mathrm{an}}\rightarrow[-\infty,\infty) be an usc function. Assume that ϕψ\phi\leq\psi on XdivX^{\mathrm{div}}, then the same holds on XanX^{\mathrm{an}}.

Next we note that we may use sequences instead of nets in the definition of PSH(Lan)\mathrm{PSH}(L^{\mathrm{an}}):

Theorem 2.12 ([BJ21, Corollary 12.18]).

Let SS be an ample line bundle on XX. Let ϕPSH(Lan)\phi\in\mathrm{PSH}(L^{\mathrm{an}}). Then there is a sequence of rational numbers εi0\varepsilon_{i}\searrow 0 and a decreasing sequence ϕigf((L+εiS)an)\phi_{i}\in\mathcal{H}^{\mathrm{gf}}_{\mathbb{Q}}((L+\varepsilon_{i}S)^{\mathrm{an}}) such that ϕ\phi is the pointwise limit of ϕi\phi_{i}, as ii\to\infty.

The space PSH(Lan)\mathrm{PSH}(L^{\mathrm{an}}) inherits most of the expected properties of (Archimedean) psh functions ([BJ21, Theorem 4.7]). However, the following compactness result is not known:

Conjecture 2.13 ([BJ21, §5]).

Assume that XX is unibranch, then every bounded from above increasing net of elements in PSH(Lan)\mathrm{PSH}(L^{\mathrm{an}}) converges in PSH(Lan)\mathrm{PSH}(L^{\mathrm{an}}).

This prediction is equivalent to so-called envelope conjecture [BJ21, Conjecture 5.14]: the regularized supremum of a bounded from above family of functions in PSH(Lan)\mathrm{PSH}(L^{\mathrm{an}}) lies in PSH(Lan)\mathrm{PSH}(L^{\mathrm{an}}). See [BJ21, Theorem 5.11] for the proof of the equivalence. This conjecture is proved when XX is smooth and LL is nef in [BJ21]. More recently, in [BJ22], Boucksom–Jonsson further established the case when XX is smooth and LL is pseudoeffective.

3 Non-Archimedean psh functions and the envelope conjecture

In this section, we study the space of non-Archimedean psh functions. The main technical difficulty is to find the correct definition of these objects for a big/pseudoeffective cohomology class. As our class {θ}\{\theta\} may be transcendental, we will give a completely analytic definition. However, we will point out that our choices coincide with the analogous algebraic notions of [BJ21], whenever {θ}\{\theta\} is the first Chern class of a \mathbb{Q}-line bundle.

3.1 Relative test curves and non-Archimedean metrics

Let XX be a connected projective manifold of dimension nn. Let θ\theta be a closed smooth real (1,1)(1,1)-form on XX representing a big cohomology class {θ}\{\theta\}.

For φPSH(X,θ)\varphi\in\mathrm{PSH}(X,\theta), we define the analytification φan:Xan[,0]\varphi^{\mathrm{an}}:X^{\mathrm{an}}\rightarrow[-\infty,0] as follows:

φan(v):=v(φ)=limk1kv((kφ)).\varphi^{\mathrm{an}}(v):=-v(\varphi)=-\lim_{k\to\infty}\frac{1}{k}v\left(\mathcal{I}(k\varphi)\right)\,. (23)

The quantity inside the limit is defined in (21). In addition, by the subadditivity of multiplier ideals, we have that ((k+k)φ)(kφ)(kφ).\mathcal{I}((k+k^{\prime})\varphi)\subseteq\mathcal{I}(k\varphi)\mathcal{I}(k^{\prime}\varphi). It follows that

v(((k+k)φ))v((kφ))+v((kφ)).v(\mathcal{I}((k+k^{\prime})\varphi))\geq v(\mathcal{I}(k\varphi))+v(\mathcal{I}(k^{\prime}\varphi)).

In particular, thanks to Fekete’s lemma, the limit in (23) exists.

When v=cordEv=c\operatorname{ord}_{E} for some prime divisor EE over XX, φan(v)=cν(φ,E)\varphi^{\mathrm{an}}(v)=-c\nu(\varphi,E), by [BBJ21, Lemma B.4]. Here ν(φ,E)\nu(\varphi,E) is the Lelong number of φ\varphi along EE (cf. [DZ22, (13)]).

The analytification of a relative test curve {ψτ}τ\{\psi_{\tau}\}_{\tau} is ψan:Xdiv[,0]\psi^{\mathrm{an}}:X^{\mathrm{div}}\to[-\infty,0], defined as

ψan(v):=supττψ+(ψτan(v)+τ).\psi^{\mathrm{an}}(v):=\sup_{\tau\leq\tau^{+}_{\psi}}(\psi_{\tau}^{\mathrm{an}}(v)+\tau). (24)

With the convention ()an(v)=(-\infty)^{\mathrm{an}}(v)=-\infty, we can even allow τ\tau\in\mathbb{R} in the supremum of (24). Since τψτ\tau\mapsto\psi_{\tau} is τ\tau-decreasing, we observe that it suffices to take supremum over τ<τψ+\tau<\tau_{\psi}^{+} in (24).

We point out that ψan\psi^{\mathrm{an}} can be computed using the subgeodesic corresponding to the test curve:

Proposition 3.1.

Let {ψτ}τ\{\psi_{\tau}\}_{\tau} be relative a test curve with τψ+0\tau^{+}_{\psi}\leq 0. Let Ψ\Psi be the potential on X×ΔX\times\Delta^{*} corresponding to {ψˇt}t\{\check{\psi}_{t}\}_{t} given by Ψ(x,ξ):=ψˇlog|ξ|2(x)\Psi(x,\xi):={\check{\psi}}_{-\log|\xi|^{2}}(x) (recall (12)). Since τψ+0\tau^{+}_{\psi}\leq 0, Ψ\Psi extends to a qpsh potential on X×ΔX\times\Delta. Moreover,

ψan(v)=σ(v)(Ψ)for vXdiv.\psi^{\mathrm{an}}(v)=-\sigma(v)(\Psi)\quad\text{for }v\in X^{\mathrm{div}}. (25)

We will often refer to Ψ\Psi as the potential of the test curve {ψτ}τ\{\psi_{\tau}\}_{\tau}. The right-hand side of (25) agrees with ψˇan(v)\check{\psi}^{\mathrm{an}}(v) as defined in [BBJ21], when θ=c1(L,h)\theta=c_{1}(L,h) for some Hermitian ample line bundle (L,h)(L,h) (see [BBJ21, Section 4.3]).

Proof.

The proof is the same as [DX22, Proposition 3.13], that deals with the ample case. We briefly recall the argument. We start with observing that

Ψ(x,δ)=supττψ+(ψτ(x)log|δ|2τ)for xX,δΔ.\Psi(x,\delta)=\sup_{\tau\leq\tau^{+}_{\psi}}(\psi_{\tau}(x)-\log|\delta|^{2}\tau)\quad\text{for }x\in X,\delta\in\Delta^{*}.

By definition of σ(v)(X×Δ)div\sigma(v)\in(X\times\Delta)^{\mathrm{div}} we have that σ(v)(ψτ(x)log|δ|2τ)=v(ψτ)τ.\sigma(v)(\psi_{\tau}(x)-\log|\delta|^{2}\tau)=v(\psi_{\tau})-\tau. Lastly, since σ(v)\sigma(v) is a divisorial valuation on X×ΔX\times\Delta, by [DX22, Lemma 3.14], we conclude that σ(v)(Ψ)=infττψ+(v(ψτ)τ),\sigma(v)(\Psi)=\inf_{\tau\leq\tau^{+}_{\psi}}(v(\psi_{\tau})-\tau)\,, finishing the proof. ∎

Piecewise linear curves.

Next we introduce the notion of a piecewise linear curve {ψτ}τ\{\psi_{\tau}\}_{\tau} in PSH(X,θ)\mathrm{PSH}(X,\theta) associated with ψτjPSH(X,θ)\psi_{\tau_{j}}\in\textup{PSH}(X,\theta), for a finite number of parameters τ0>τ1>>τN\tau_{0}>\tau_{1}>\dots>\tau_{N}. The piecewise linear curve is the affine interpolation of this data:
(i)(i) ψτ=ψτN\psi_{\tau}=\psi_{\tau_{N}} for ττN\tau\leq\tau_{N}.
(ii)(ii) For t(0,1)t\in(0,1), we have ψ(1t)τi+tτi+1=(1t)ψτi+tψτi+1.\psi_{(1-t)\tau_{i}+t\tau_{i+1}}=(1-t)\psi_{\tau_{i}}+t\psi_{\tau_{i+1}}\,.
(iii)(iii) ψτ=\psi_{\tau}=-\infty for τ>τ0\tau>\tau_{0}.

Observe that {ψτ}τ\{\psi_{\tau}\}_{\tau} is τ\tau-usc but may not be τ\tau-concave. Despite this we introduce the analytification ψan\psi^{\mathrm{an}} of {ψτ}τ\{\psi_{\tau}\}_{\tau} as follows:

ψan(v):=supττ0(ψτan(v)+τ)=maxτi(ψτian(v)+τi)for vXan.\psi^{\mathrm{an}}(v):=\sup_{\tau\leq\tau_{0}}(\psi_{\tau}^{\mathrm{an}}(v)+\tau)=\max_{\tau_{i}}(\psi_{\tau_{i}}^{\mathrm{an}}(v)+\tau_{i})\quad\text{for }v\in X^{\mathrm{an}}. (26)

Given a bounded from above usc function ff defined on an interval of \mathbb{R}, the concave envelope f~\tilde{f} of ff is the minimal concave function lying above ff. Recall that ff and f~\tilde{f} have the same inverse Legendre transform.

Lemma 3.2.

Let {ψτ}τ\{\psi_{\tau}\}_{\tau} be a piecewise linear curve in PSH(X,θ)\mathrm{PSH}(X,\theta), then the τ\tau-concave envelope {ψ~τ}τ\{\tilde{\psi}_{\tau}\}_{\tau} of {ψτ}τ\{\psi_{\tau}\}_{\tau} is a relative test curve. Moreover,

ψan=ψ~anon Xdiv.\psi^{\mathrm{an}}=\tilde{\psi}^{\mathrm{an}}\quad\text{on }X^{\mathrm{div}}. (27)
Proof.

For the first part, recall that {ψ~τ}τ\{\tilde{\psi}_{\tau}\}_{\tau} is the Legendre transform of the inverse Legendre transform

ψˇt:=supτ(ψτ+tτ)for t>0\check{\psi}_{t}:=\sup_{\tau\in\mathbb{R}}(\psi_{\tau}+t\tau)\quad\text{for }t>0

of {ψτ}τ\{\psi_{\tau}\}_{\tau}. So by Theorem 2.6, it suffices to show that {ψˇt}t\{\check{\psi}_{t}\}_{t} is a subgeodesic. This is clear because

ψˇt=maxj=0,,N(ψτj+tτj).\check{\psi}_{t}=\max_{j=0,\dots,N}(\psi_{\tau_{j}}+t\tau_{j}).

Each term in the maximum is clearly a subgeodesic, hence so is {ψˇt}t\{\check{\psi}_{t}\}_{t}.

In order to prove (27), we may assume that ψτ=\psi_{\tau}=-\infty when τ>0\tau>0. Note that by the same arguments as Proposition 3.1, ψan(v)=σ(v)(Ψ)\psi^{\mathrm{an}}(v)=-\sigma(v)(\Psi), where Ψ\Psi is the potential on X×ΔX\times\Delta corresponding to the subgeodesic {ψˇt}t\{\check{\psi}_{t}\}_{t}. So (27) follows from Proposition 3.1 applied to {ψ~τ}τ\{\tilde{\psi}_{\tau}\}_{\tau}, and τ\tau-Legendre duality. ∎

3.2 Transcendental non-Archimedean metrics

Let XX be a connected compact Kähler manifold of dimension nn. Let θ\theta be a closed smooth real (1,1)(1,1)-form on XX representing a big cohomology class {θ}\{\theta\}.

We are ready to introduce the set of non-Archimedean psh metrics for a general pseudoeffective transcendental class {θ}\{\theta\}. Let 𝒦\mathcal{K} denote the set of Kähler metrics on XX endowed with the partial order: ωω\omega\succeq\omega^{\prime} if ωω\omega\leq\omega^{\prime} as forms. Clearly 𝒦\mathcal{K} is a directed set.

For ωω\omega\leq\omega^{\prime} there is a natural transition map from MTC(θ+ω)MTC(θ+ω)\textup{MTC}_{\mathcal{I}}(\theta+\omega)\mapsto\textup{MTC}_{\mathcal{I}}(\theta+\omega^{\prime}) described as follows (recall (18)). To {ψτ}τMTC(θ+ω)\{\psi_{\tau}\}_{\tau}\in\textup{MTC}_{\mathcal{I}}(\theta+\omega) we associate {ψτ}τMTC(θ+ω)\{\psi^{\prime}_{\tau}\}_{\tau}\in\textup{MTC}_{\mathcal{I}}(\theta+\omega^{\prime}), where

ψτ:={Pθ+ω[ψτ],if τ<τψ+;limττψ+ψτ,if τ=τψ+;,if τ>τψ+.\psi^{\prime}_{\tau}:=\left\{\begin{aligned} P_{\theta+\omega^{\prime}}[\psi_{\tau}]_{\mathcal{I}},&\quad\text{if }\tau<\tau^{+}_{\psi};\\ \lim_{\tau\nearrow\tau^{+}_{\psi}}\psi^{\prime}_{\tau},&\quad\text{if }\tau=\tau^{+}_{\psi};\\ -\infty,&\quad\text{if }\tau>\tau^{+}_{\psi}.\end{aligned}\right. (28)

With this in hand, we can define the space of non-Archimedean functions:

Definition 3.3.

The space of non-Archimedean functions of a pseudoeffective class {θ}\{\theta\} is the following projective limit in the category of sets:

PSHNA(θ):=limω𝒦MTC(θ+ω).\mathrm{PSH}^{\mathrm{NA}}(\theta):=\varprojlim_{\omega\in\mathcal{K}}\textup{MTC}_{\mathcal{I}}(\theta+\omega).

Notice that the above limit is well-defined since Pθ+ω′′[Pθ+ω[ψτ]]=Pθ+ω′′[ψτ]P_{\theta+\omega^{\prime\prime}}[P_{\theta+\omega^{\prime}}[\psi_{\tau}]_{\mathcal{I}}]_{\mathcal{I}}=P_{\theta+\omega^{\prime\prime}}[\psi_{\tau}]_{\mathcal{I}} if ωωω′′\omega\leq\omega^{\prime}\leq\omega^{\prime\prime} and {ψτ}τMTC(θ+ω)\{\psi_{\tau}\}_{\tau}\in\textup{MTC}_{\mathcal{I}}(\theta+\omega).

We introduce a partial order on PSHNA(θ)\mathrm{PSH}^{\mathrm{NA}}(\theta): for ϕ={{ϕτω}τ}ω𝒦,ϕ={{ϕτω}τ}ω𝒦PSHNA(θ)\phi=\{\{\phi^{\omega}_{\tau}\}_{\tau}\}_{\omega\in\mathcal{K}},\ \phi^{\prime}=\{\{\phi^{\prime\omega}_{\tau}\}_{\tau}\}_{\omega\in\mathcal{K}}\in\mathrm{PSH}^{\mathrm{NA}}(\theta), we say ϕϕ\phi\leq\phi^{\prime} if ϕτωϕτω\phi^{\omega}_{\tau}\leq\phi^{\prime\omega}_{\tau} for all ω𝒦\omega\in\mathcal{K} and τ\tau\in\mathbb{R}.

Observe that for ϕ={{ϕτω}τ}ω𝒦PSHNA(θ)\phi=\{\{\phi^{\omega}_{\tau}\}_{\tau}\}_{\omega\in\mathcal{K}}\in\mathrm{PSH}^{\mathrm{NA}}(\theta), τϕω+\tau^{+}_{\phi^{\omega}} does not depend on the choice of ω\omega. We denote the common value by τϕ+\tau^{+}_{\phi}.

When {θ}\{\theta\} is big, note that there exists a natural inclusion MTC(θ)PSHNA(θ)\textup{MTC}_{\mathcal{I}}(\theta)\hookrightarrow\mathrm{PSH}^{\mathrm{NA}}(\theta). Indeed, with {ϕτ}τMTC(θ)\{\phi_{\tau}\}_{\tau}\in\textup{MTC}_{\mathcal{I}}(\theta) one simply associates {{ϕτω}τ}ωPSHNA(θ)\{\{\phi^{\omega}_{\tau}\}_{\tau}\}_{\omega}\in\mathrm{PSH}^{\mathrm{NA}}(\theta), where

ϕτω:={Pθ+ω[ϕτ],if τ<τϕ+;limττϕ+ϕτω,if τ=τϕ+;,if τ>τϕ+.\phi^{\omega}_{\tau}:=\left\{\begin{aligned} P_{\theta+\omega}[\phi_{\tau}]_{\mathcal{I}},&\quad\text{if }\tau<\tau^{+}_{\phi};\\ \lim_{\tau\nearrow\tau_{\phi}^{+}}\phi^{\omega}_{\tau},&\quad\text{if }\tau=\tau^{+}_{\phi};\\ -\infty,&\quad\text{if }\tau>\tau^{+}_{\phi}.\end{aligned}\right. (29)

Lastly, let us note that the counterpart of Conjecture 2.13 naturally holds in the setting of PSHNA(θ)\textup{PSH}^{\mathrm{NA}}(\theta).

Theorem 3.4.

Suppose that {ϕi}iIPSHNA(θ)\{\phi^{i}\}_{i\in I}\subset\textup{PSH}^{\mathrm{NA}}(\theta) is an increasing net with

supiIτϕi+<.\sup_{i\in I}\tau^{+}_{\phi^{i}}<\infty.

Then there exists ϕPSHNA(θ)\phi\in\textup{PSH}^{\mathrm{NA}}(\theta) such that for any ω𝒦\omega\in\mathcal{K}, ϕτi,ωϕτω\phi^{i,\omega}_{\tau}\nearrow\phi^{\omega}_{\tau} almost everywhere for all τ<τϕ+\tau<\tau^{+}_{\phi} and

τϕ+=supiIτϕi+.\tau^{+}_{\phi}=\sup_{i\in I}\tau^{+}_{\phi^{i}}.
Proof.

Let ω𝒦\omega\in\mathcal{K}. By Lemma 2.9 there exists {ϕτω}τMCT(θ)\{\phi^{\omega}_{\tau}\}_{\tau}\in\textup{MCT}_{\mathcal{I}}(\theta) such that ϕτi,ωϕτω\phi^{i,\omega}_{\tau}\nearrow\phi^{\omega}_{\tau} almost everywhere for all τ<limiτϕi+=τϕ+\tau<\lim_{i}\tau^{+}_{\phi^{i}}=\tau^{+}_{\phi} and ϕτϕ+ω=limττϕ+ϕτω\phi^{\omega}_{\tau^{+}_{\phi}}=\lim_{\tau\nearrow\tau^{+}_{\phi}}\phi^{\omega}_{\tau} almost everywhere.

Let ωω\omega\leq\omega^{\prime}. Since ϕτωϕτω\phi^{\omega}_{\tau}\simeq_{\mathcal{I}}\phi^{\omega^{\prime}}_{\tau} for all τ<τϕ+\tau<\tau^{+}_{\phi} [DX22, Proposition 2.18(ii)], we obtain that ϕτω=Pθ+ω[ϕτω]\phi^{\omega^{\prime}}_{\tau}=P_{\theta+\omega^{\prime}}[\phi^{\omega}_{\tau}]_{\mathcal{I}}, hence ϕ={{ϕτω}τ}ωPSHNA(θ)\phi=\{\{\phi^{\omega}_{\tau}\}_{\tau}\}_{\omega}\in\textup{PSH}^{\mathrm{NA}}(\theta). ∎

3.3 Comparison with Boucksom–Jonsson’s NA metrics

Let XX be a connected projective manifold of dimension nn. Assume furthermore that {θ}=c1(L)\{\theta\}=c_{1}(L) for some big \mathbb{Q}-line bundle LL on XX. We will compare the sets PSHNA(θ)\textup{PSH}^{\mathrm{NA}}(\theta) introduced in the previous section and the space PSH(Lan)\mathrm{PSH}(L^{\mathrm{an}}) recalled in Section 2.3.

Lemma 3.5.

For any φPSH(X,θ)\varphi\in\mathrm{PSH}(X,\theta) we have that φanPSH(Lan)\varphi^{\mathrm{an}}\in\mathrm{PSH}(L^{\mathrm{an}}).

Proof.

After replacing LL with a sufficiently high power, we may assume that LL is a line bundle. Take a very ample line bundle HH on XX. By Siu’s uniform global generation theorem [Siu98], [Dem12, Theorem 6.27] there exists b>0b>0 large enough so that HbLk(kφ)H^{b}\otimes L^{k}\otimes\mathcal{I}(k\varphi) is globally generated for all k>0k>0. Let {si}\{s_{i}\} be a finite set of global sections that generate the sheaf HbLk(kφ)H^{b}\otimes L^{k}\otimes\mathcal{I}(k\varphi). Then v((kφ))=mini(v(si))v(\mathcal{I}(k\varphi))=\min_{i}(v(s_{i})). It follows that vk1v((kφ))v\mapsto-k^{-1}v\left(\mathcal{I}(k\varphi)\right) lies in gf((L+bkH)an)\mathcal{H}^{\mathrm{gf}}_{\mathbb{Q}}((L+\frac{b}{k}H)^{\mathrm{an}}). Picking km:=2mk_{m}:=2^{m} and letting mm\to\infty we conclude that φanPSH(Lan)\varphi^{\mathrm{an}}\in\mathrm{PSH}(L^{\mathrm{an}}). ∎

Lemma 3.6.

Let {ψτ}τ\{\psi_{\tau}\}_{\tau} be a piecewise linear curve in PSH(X,θ)\mathrm{PSH}(X,\theta). Then ψan\psi^{\mathrm{an}}, as defined in (26), extends to an element ψanPSH(Lan)\psi^{\mathrm{an}}\in\mathrm{PSH}(L^{\mathrm{an}}).

Proof.

The result follows from definition (26), [BJ21, Theorem 4.7(ii)] and Lemma 3.5. ∎

Lemma 3.7.

Let RR be a commutative \mathbb{C}-algebra of finite type and II be an ideal of R[t]R[t]. If for any aS1a\in S^{1}, aIIa^{*}I\subseteq I, then II is stable under the \mathbb{C}^{*}-action. Moreover, there are ideals I0I1ImI_{0}\subseteq I_{1}\subseteq\dots\subseteq I_{m} in RR so that

I=I0+I1t++Im(tm),I=I_{0}+I_{1}t+\dots+I_{m}(t^{m}), (30)
Proof.

It suffices to argue that II can be expanded as in (30). To see this, assume that aIa\in I. We can write a=a0+a1t++amtma=a_{0}+a_{1}t+\dots+a_{m}t^{m} with aiRa_{i}\in R. Then our assumption implies that iaiρitiI\sum_{i}a_{i}\rho^{i}t^{i}\in I as well for all ρS1\rho\in S^{1}. So by the Lagrange interpolation formula, aitiIa_{i}t^{i}\in I for all ii. Therefore, we can write II as I0+I1t+I2t2+I_{0}+I_{1}t+I_{2}t^{2}+\dots for some ideals I0I1I_{0}\subseteq I_{1}\subseteq\dots in RR. But as RR is noetherian, there is m0m\geq 0 so that Im=ImI_{m^{\prime}}=I_{m} for m>mm^{\prime}>m. (30) follows. ∎

Lemma 3.8.

Let XX be a complex projective variety and p:X×Xp:X\times\mathbb{C}\rightarrow X be the natural projection. Assume that \mathcal{I} is an analytic coherent ideal sheaf on X×X\times\mathbb{C}. Assume that |X×=p𝒥\mathcal{I}|_{X\times\mathbb{C}^{*}}=p^{*}\mathcal{J} for some coherent ideal sheaf 𝒥\mathcal{J} on XX. Then \mathcal{I} is the analytification of an algebraic coherent ideal sheaf.

Proof.

Let q:X×(1{0})Xq:X\times(\mathbb{P}^{1}\setminus\{0\})\rightarrow X be the natural projection. As 1{0}\mathbb{C}^{*}\subset\mathbb{P}^{1}\setminus\{0\} we can glue q𝒥q^{*}\mathcal{J} with \mathcal{I} to get an analytic coherent ideal sheaf on X×1X\times\mathbb{P}^{1}. By the GAGA principle, this ideal sheaf is necessarily algebraic, hence so is its restriction to X×X\times\mathbb{C}. ∎

Next we point out a version of Siu’s uniform global generatedness lemma [Siu88] that we will need in the proof of our next theorem:

Lemma 3.9.

Let LL be a big line bundle on XX such that c1(L)={θ}c_{1}(L)=\{\theta\} and ΦPSH(X×Δ,p1θ\Phi\in\textup{PSH}(X\times\Delta,p_{1}^{*}\theta), where Δ\Delta is the unit disk. Let GG be an ample line bundle on XX. Then there exists k>0k>0, only dependent on XX and GG such that p1(GkLm)(mΦ)p_{1}^{*}(G^{k}\otimes L^{m})\otimes\mathcal{I}(m\Phi) is globally generated for all mm\in\mathbb{N}.

Proof.

The argument for this is exactly the same as the one in [BBJ21, Lemma 5.6] with Nadal’s vanishing replaced by the family version proved by Matsumura in [Mat16, Theorem 1.7]. Alternatively, one can adapt the proof of [Dem12, Theorem 6.27] to our setting, with only one small change: instead of applying Nadel’s theorem directly, one needs to use [Dem12, Corollary 5.3] when solving the ¯\bar{\partial}-problem in the argument. ∎

Proposition 3.10.

Let ϕPSH(X,θ)\phi\in\mathrm{PSH}(X,\theta) be a model potential with positive mass. Let Φ\Phi be the p1θp_{1}^{*}\theta-psh function on X×ΔX\times\Delta corresponding to a psh geodesic ray {ϕt}t\{\phi_{t}\}_{t} in PSH(X,θ)\mathrm{PSH}(X,\theta) emanating from ϕ\phi, with supϕ10\sup\phi_{1}\leq 0. Then the function

vσ(v)(Φ)for vXdivv\mapsto-\sigma(v)(\Phi)\quad\text{for }v\in X^{\mathrm{div}}

admits a unique extension to an element in PSH(Lan)\mathrm{PSH}(L^{\mathrm{an}}).

Proof.

We may assume that LL is a line bundle. Observe that the extension is unique if it exists by Theorem 2.11.

Claim. For each m>0m\in\mathbb{Z}_{>0}, (mΦ)|X×Δ=p1(mϕ)|X×Δ\mathcal{I}(m\Phi)|_{X\times\Delta^{*}}=p_{1}^{*}\mathcal{I}(m\phi)|_{X\times\Delta^{*}}. In particular, (mΦ)\mathcal{I}(m\Phi) admits an extension to a coherent ideal sheaf on X×X\times\mathbb{C}.

To prove the claim it is enough to show that

(mΦ)|X×Δ=((mϕp1)|X×Δ)=p1(mϕ)|X×Δ.\mathcal{I}(m\Phi)|_{X\times\Delta^{*}}=\mathcal{I}((m\phi\circ p_{1})|_{X\times\Delta^{*}})=p_{1}^{*}\mathcal{I}(m\phi)|_{X\times\Delta^{*}}. (31)

We observe that ((mϕp1)|X×Δ)p1(mϕ)|X×Δ\mathcal{I}((m\phi\circ p_{1})|_{X\times\Delta^{*}})\subseteq p_{1}^{*}\mathcal{I}(m\phi)|_{X\times\Delta^{*}} by [Dem12, Proposition 14.3] and the reverse inclusion is obvious. So it suffices to establish the first equality in (31).

If we could prove this, then the \mathbb{C}^{*}-invariant extension of (mΦ)\mathcal{I}(m\Phi) to X×X\times\mathbb{C}^{*} would be simply ((mϕp1)|X×)\mathcal{I}((m\phi\circ p_{1})|_{X\times\mathbb{C}^{*}}).

Consider the annulus Δa,b={z:exp(b)<|z|2<exp(a)}Δ\Delta_{a,b}=\{z\in\mathbb{C}:\exp(-b)<|z|^{2}<\exp(-a)\}\subset\Delta^{*} for 0<a<b0<a<b. To argue (31), it suffices to show that (mΦ)|X×Δa,b=((mϕp1)|X×Δa,b)\mathcal{I}(m\Phi)|_{X\times\Delta_{a,b}}=\mathcal{I}((m\phi\circ p_{1})|_{X\times\Delta_{a,b}}) for arbitrary a,ba,b.

First we notice that due to convexity of (sub)geodesics in the time variable we have Φ(x,z)|X×Δa,bϕ(x)+C\Phi(x,z)|_{X\times\Delta_{a,b}}\leq\phi(x)+C for some C>0C>0.

Second, since ϕ\phi has positive mass, due to Lemma 2.5 and [DDNL18a, Theorem 1.3] we have that Φ(,exp(b)),Φ(,exp(b))(X,θ,ϕ)\Phi(\cdot,\exp(-b)),\Phi(\cdot,\exp(-b))\in\mathcal{E}(X,\theta,\phi). By [Gup, Theorem 2.9], there exists ψ:=Pθ(Φ(,exp(a)),Φ(,exp(b)))(X,θ,ϕ)\psi:=P_{\theta}(\Phi(\cdot,\exp(-a)),\Phi(\cdot,\exp(-b)))\in\mathcal{E}(X,\theta,\phi) such that ψ(z)Φ(x,z)|X×Δa,b\psi(z)\leq\Phi(x,z)|_{X\times\Delta_{a,b}}. This last inequality follows from the comparison principle built into the definition of psh geodesic segments, recalled in (13). We conclude that

((mψp1)|X×Δa,b)(mΦ)|X×Δa,b((mϕp1)|X×Δa,b).\mathcal{I}((m\psi\circ p_{1})|_{X\times\Delta_{a,b}})\subseteq\mathcal{I}(m\Phi)|_{X\times\Delta_{a,b}}\subseteq\mathcal{I}((m\phi\circ p_{1})|_{X\times\Delta_{a,b}}). (32)

By [DDNL18a, Theorem 1.3] we have that P(ψ+C,ϕ)P[ϕ]=ϕP(\psi+C,\phi)\nearrow P[\phi]=\phi a.e. on XX as CC\nearrow\infty. Since [P(ψ+C,ϕ)]=[ψ][P(\psi+C,\phi)]=[\psi], by Theorem 2.2 we conclude that ((mψp1)|X×Δa,b)=((mϕp1)|X×Δa,b)\mathcal{I}((m\psi\circ p_{1})|_{X\times\Delta_{a,b}})=\mathcal{I}((m\phi\circ p_{1})|_{X\times\Delta_{a,b}}). Together with (32) this finishes the proof of (31).

From the claim and Lemma 3.7 and Lemma 3.8 we get that

(mΦ)=𝔞0+𝔞1t++𝔞N1tN1+𝔞N(tN),\mathcal{I}(m\Phi)=\mathfrak{a}_{0}+\mathfrak{a}_{1}t+\dots+\mathfrak{a}_{N-1}t^{N-1}+\mathfrak{a}_{N}(t^{N})\,, (33)

where the 𝔞i\mathfrak{a}_{i}’s are coherent ideal sheaves on XX.

Using Lemma 3.9, there exists TXT\to X ample such that p1TLm(mΦ)p_{1}^{*}T\otimes L^{m}\otimes\mathcal{I}(m\Phi) is globally generated, which is equivalent to TLm𝔞iT\otimes L^{m}\otimes\mathfrak{a}_{i} being globally generated for all ii (in contrast with the case where ϕ\phi is bounded, explored in [BBJ21], 𝔞N𝒪X\mathfrak{a}_{N}\neq\mathcal{O}_{X} in general).

We define

φm(v):=1mσ(v)((mΦ))=1mmini(v(𝔞i)+i),vXdiv.\varphi_{m}(v):=-\frac{1}{m}\sigma(v)(\mathcal{I}(m\Phi))=-\frac{1}{m}\min_{i}(v(\mathfrak{a}_{i})+i),\quad v\in X^{\mathrm{div}}\,.

From the right-hand side of the formula, φm\varphi_{m} can be extended to an element in gf((L+m1T)an)\mathcal{H}^{\mathrm{gf}}_{\mathbb{Q}}((L+m^{-1}T)^{\mathrm{an}}), which we denote by the same symbol.

For vXdivv\in X^{\mathrm{div}}, it follows from the well-known argument using the Ohsawa–Takegoshi extension theorem ([BBJ21, Lemma B.4]) that

σ(v)(Φ)=limm12mσ(v)((2mΦ))=limmφ2m(v)-\sigma(v)(\Phi)=\lim_{m\to\infty}-\frac{1}{2^{m}}\sigma(v)(\mathcal{I}(2^{m}\Phi))=\lim_{m\to\infty}\varphi_{2^{m}}(v)

and the right-hand side defines an element in PSH(Lan)\mathrm{PSH}(L^{\mathrm{an}}) by definition, since {φ2m}m\{\varphi_{2^{m}}\}_{m} is decreasing. ∎

Corollary 3.11.

Let {ψτ}τ\{\psi_{\tau}\}_{\tau} be a test curve relative to ϕPSH(X,θ)\phi\in\textup{PSH}(X,\theta). Then ψan:Xdiv\psi^{\mathrm{an}}:X^{\mathrm{div}}\to\mathbb{R} admits a unique extension to ψanPSH(Lan)\psi^{\mathrm{an}}\in\mathrm{PSH}(L^{\mathrm{an}}).

Proof.

Observe that the extension is unique if it exists by Theorem 2.11. We may assume that τψ+=0\tau^{+}_{\psi}=0, without loss of generality.

Let us first assume that ϕ\phi has positive mass. If {χτ}τ\{\chi_{\tau}\}_{\tau} is the maximization of {ψτ}τ\{\psi_{\tau}\}_{\tau} then χan=ψan\chi^{\mathrm{an}}=\psi^{\mathrm{an}} by definition. Hence, we can assume that {ψτ}τ\{\psi_{\tau}\}_{\tau} is maximal, i.e., {ψˇt}t\{\check{\psi}_{t}\}_{t} is a psh geodesic emanating from P[ϕ]P[\phi]. The result now follows from Proposition 3.10 and Proposition 3.1.

In general, take an ample line bundle SS on XX. Then the previous case shows that ψanPSH((L+ϵS)an)\psi^{\mathrm{an}}\in\mathrm{PSH}((L+\epsilon S)^{\mathrm{an}}) for any rational ϵ>0\epsilon>0. It follows that ψanPSH(Lan)\psi^{\mathrm{an}}\in\mathrm{PSH}(L^{\mathrm{an}}) by [BJ21, Theorem 4.5]. ∎

Before we can prove Theorem 1.1, we deal with an intermediate case:

Theorem 3.12.

Assume that {θ}=c1(L)\{\theta\}=c_{1}(L) for some big line bundle LL on XX. The following hold:
(i) Let {ψτ}τ,{χτ}τMTC(θ)\{\psi_{\tau}\}_{\tau},\{\chi_{\tau}\}_{\tau}\in\textup{MTC}_{\mathcal{I}}(\theta). If ψanχan\psi^{\mathrm{an}}\geq\chi^{\mathrm{an}} then ψτχτ\psi_{\tau}\geq\chi_{\tau} for any τ\tau\in\mathbb{R}. In particular, the map {ψτ}τψan\{\psi_{\tau}\}_{\tau}\mapsto\psi^{\mathrm{an}} is an injection MTC(θ)PSH(Lan)\textup{MTC}_{\mathcal{I}}(\theta)\hookrightarrow\mathrm{PSH}(L^{\mathrm{an}}).
(ii) The image of the map MTC(θ){ψτ}τψanPSH(Lan)\textup{MTC}_{\mathcal{I}}(\theta)\ni\{\psi_{\tau}\}_{\tau}\mapsto\psi^{\mathrm{an}}\in\mathrm{PSH}(L^{\mathrm{an}}) contains PSH((LT)an)\mathrm{PSH}((L-T)^{\mathrm{an}}) for any ample \mathbb{Q}-line bundle TT on XX such that LTL-T is big.

Proof.

First observe that the map MTC(θ)PSH(Lan)\textup{MTC}_{\mathcal{I}}(\theta)\rightarrow\mathrm{PSH}(L^{\mathrm{an}}) is well-defined by Corollary 3.11.

We argue (i). Let vXdivv\in X^{\mathrm{div}} and t+t\in\mathbb{Q}_{+}. Then, using (24) we notice that

tψan(1tv)=supττψ+(ψτan(v)+tτ).t\psi^{\mathrm{an}}\Big{(}\frac{1}{t}v\Big{)}=\sup_{\tau\leq\tau^{+}_{\psi}}(\psi_{\tau}^{\mathrm{an}}(v)+t\tau). (34)

Using the condition ψanχan\psi^{\mathrm{an}}\geq\chi^{\mathrm{an}} we obtain that

supτ(ψτan(v)+tτ)supτ(χτan(v)+tτ).\sup_{\tau\in\mathbb{R}}(\psi_{\tau}^{\mathrm{an}}(v)+t\tau)\geq\sup_{\tau\in\mathbb{R}}(\chi_{\tau}^{\mathrm{an}}(v)+t\tau).

As a result, since +\mathbb{Q}_{+} is dense in +\mathbb{R}_{+}, the above inequality holds for all t0t\geq 0. Since τψτan(v),ψτan(v)\tau\mapsto\psi_{\tau}^{\mathrm{an}}(v),\psi_{\tau}^{\mathrm{an}}(v) are both concave, taking the tt-Legendre transform of both sides we conclude that ψτan(v)χτan(v)\psi^{\mathrm{an}}_{\tau}(v)\geq\chi^{\mathrm{an}}_{\tau}(v) for all τ\tau\in\mathbb{R}. Since the potentials ψτ,χτ\psi_{\tau},\chi_{\tau} are \mathcal{I}-model, from [DX22, Corollary 2.16] we obtain that ψτχτ\psi_{\tau}\geq\chi_{\tau} for all τ\tau\in\mathbb{R}.

To argue (ii), we take ϕPSH((LT)an)PSH(Lan)\phi\in\mathrm{PSH}((L-T)^{\mathrm{an}})\subset\mathrm{PSH}(L^{\mathrm{an}}), and we want to write it as ψan\psi^{\mathrm{an}} for some ψMTC(θ)\psi\in\textup{MTC}_{\mathcal{I}}(\theta). Before we deal with this, let us only consider ϕgf(L)\phi\in\mathcal{H}^{\mathrm{gf}}_{\mathbb{Q}}(L), say

ϕ=m1maxi(log|si|+λi),\phi=m^{-1}\max_{i}(\log|s_{i}|+\lambda_{i}), (35)

where sis_{i} are a finite number of sections of LmL^{m} and λi\lambda_{i}\in\mathbb{Q}.

We may assume that λ1λ2λN\lambda_{1}\leq\lambda_{2}\leq\dots\leq\lambda_{N}. Write IλI_{\lambda} for the set of ii such that λi=λ\lambda_{i}=\lambda. We denote the finitely many λ\lambda so that IλI_{\lambda} is non-empty as τ0>>τN\tau_{0}>\dots>\tau_{N}. Define a curve ψτ\psi_{\tau} as follows:

ψτi=1mmaxiIτi(log|si|hm2+τi).\psi_{\tau_{i}}=\frac{1}{m}\max_{i\in I_{\tau_{i}}}(\log|s_{i}|^{2}_{h^{m}}+\tau_{i})\,.

We define {ψτ}τ\{\psi_{\tau}\}_{\tau} to be the piecewise linear curve associated with the ψτi\psi_{\tau_{i}} (recall the definition preceding (26)). Let {ψτ}τ\{\psi^{\prime}_{\tau}\}_{\tau} be the τ\tau-concave envelope of {ψτ}τ\{\psi_{\tau}\}_{\tau}.

By Lemma 3.2, ψan=ψan=ϕ\psi^{\prime\mathrm{an}}=\psi^{\mathrm{an}}=\phi on XdivX^{\mathrm{div}}. By Lemma 3.6 and Theorem 2.11, the same holds on XanX^{\mathrm{an}}. We can replace ψτ\psi^{\prime}_{\tau} with P[ψτ]P[\psi^{\prime}_{\tau}]_{\mathcal{I}} when τ<τψ+\tau<\tau^{+}_{\psi} and ψτ+\psi^{\prime}_{\tau^{+}} with the limit value of P[ψτ]P[\psi^{\prime}_{\tau}]_{\mathcal{I}} as τ\tau increases to τψ+\tau^{+}_{\psi}. Defined this way, {ψτ}τ\{\psi^{\prime}_{\tau}\}_{\tau} is an \mathcal{I}-maximal test curve, and we still have ψan=ϕ\psi^{\prime\mathrm{an}}=\phi on XdivX^{\mathrm{div}}. However, we may not have that {ψτ}τMTC(θ)\{\psi^{\prime}_{\tau}\}_{\tau}\in\textup{MTC}_{\mathcal{I}}(\theta) as ψ\psi^{\prime}_{-\infty} may not have positive mass.

Now we consider ϕgf(LT)gf(L)\phi\in\mathcal{H}_{\mathbb{Q}}^{\mathrm{gf}}(L-T)\subset\mathcal{H}_{\mathbb{Q}}^{\mathrm{gf}}(L) for some ample \mathbb{Q}-line bundle TT such that LTL-T is still big. We may assume that TT is a line bundle. Fix a smooth strictly psh Hermitian metric hTh_{T} on TT with Chern form ω\omega.

Let {ψτ}τPSH(X,θω)\{\psi^{\prime}_{\tau}\}_{\tau}\subset\textup{PSH}(X,\theta-\omega) be the \mathcal{I}-maximal test curve with respect to θω\theta-\omega, such that ψan=ϕ\psi^{\prime\mathrm{an}}=\phi, constructed above. We define χτPSH(X,θ)\chi_{\tau}\in\textup{PSH}(X,\theta) in the following manner:

χτ:=Pθ[ψτ],τ<τψ+,χτψ+:=limττψ+χτ.\chi_{\tau}:=P_{\theta}[\psi^{\prime}_{\tau}]_{\mathcal{I}},\ \ \tau<\tau^{+}_{\psi^{\prime}},\ \ \ \ \chi_{\tau^{+}_{\psi^{\prime}}}:=\lim_{\tau\nearrow\tau^{+}_{\psi^{\prime}}}\chi_{\tau}.

We get that {χτ}τPSH(X,θ)\{\chi_{\tau}\}_{\tau}\subset\textup{PSH}(X,\theta) is \mathcal{I}-maximal and χan=ϕ\chi^{\mathrm{an}}=\phi.

Finally, we only need to argue that {χτ}τMTC(θ)\{\chi_{\tau}\}_{\tau}\in\textup{MTC}_{\mathcal{I}}(\theta). This follows from the fact that Pθ[ψτ]Pθ[ψτ]P_{\theta}[\psi^{\prime}_{\tau}]_{\mathcal{I}}\geq P_{\theta}[\psi^{\prime}_{\tau}] and [WN19, Theorem 1.2]:

Xθχτn=XθPθ[ψτ]nXθPθ[ψτ]n=X((θω)ψτ+ω)nXωn.\int_{X}\theta_{\chi_{\tau}}^{n}=\int_{X}\theta_{P_{\theta}[\psi^{\prime}_{\tau}]_{\mathcal{I}}}^{n}\geq\int_{X}\theta_{P_{\theta}[\psi^{\prime}_{\tau}]}^{n}=\int_{X}((\theta-\omega)_{\psi^{\prime}_{\tau}}+\omega)^{n}\geq\int_{X}\omega^{n}. (36)

Finally, we deal with the case when ϕPSH((LS)an)\phi\in\mathrm{PSH}((L-S)^{\mathrm{an}}) for some ample \mathbb{Q}-line bundle SS on XX. By Theorem 2.12, we can take an ample line bundle SS on XX, a decreasing sequence of rational numbers ci0c_{i}\to 0 and a decreasing sequence ϕigf((LS+ciS)an)gf(Lan)\phi_{i}\in\mathcal{H}^{\mathrm{gf}}_{\mathbb{Q}}((L-S+c_{i}S)^{\mathrm{an}})\subset\mathcal{H}^{\mathrm{gf}}_{\mathbb{Q}}(L^{\mathrm{an}}) converging to ϕ\phi. We will assume that ci1/2c_{i}\leq 1/2 for all ii.

Fix a smooth psh metric hSh_{S} on SS with ω=c1(S,hS)>0\omega=c_{1}(S,h_{S})>0. By the previous step, we can find {ψτi}τMTC(θ)\{\psi^{i}_{\tau}\}_{\tau}\in\textup{MTC}_{\mathcal{I}}(\theta) such that ψi,an=ϕi\psi^{i,\mathrm{an}}=\phi_{i}. Moreover, due to (36) we have that

Xθψτin(1ci)nXωn12nXωn,τ.\int_{X}\theta_{\psi^{i}_{\tau}}^{n}\geq(1-c_{i})^{n}\int_{X}\omega^{n}\geq\frac{1}{2^{n}}\int_{X}\omega^{n},\ \ \tau\in\mathbb{R}. (37)

Since (ϕi)an(ϕi+1)an(\phi^{i})^{\mathrm{an}}\geq(\phi^{i+1})^{\mathrm{an}}, due to Theorem 3.12(i), we obtain that {ψτi}τ\{\psi^{i}_{\tau}\}_{\tau} is ii-decreasing. Let ψτ=infiψτi\psi_{\tau}=\inf_{i}\psi^{i}_{\tau}. By [DDNL21, Proposition 4.6] and Lemma 2.8, {ψτ}τMTC(θ)\{\psi_{\tau}\}_{\tau}\in\textup{MTC}_{\mathcal{I}}(\theta) with Xθψτn12nXωn\int_{X}\theta^{n}_{\psi_{\tau}}\geq\frac{1}{2^{n}}\int_{X}\omega^{n}.

We need to show that ψan=ϕ\psi^{\mathrm{an}}=\phi. Using (37), again, Lemma 2.1 implies that ψτan=infiψτi,an\psi_{\tau}^{\mathrm{an}}=\inf_{i}\psi_{\tau}^{i,\mathrm{an}} when τ<τψ+\tau<\tau^{+}_{\psi}. To finish, we need to show that for any vXdivv\in X^{\mathrm{div}},

supτinfi(ψτi,an(v)+τ)=infisupτ(ψτi,an(v)+τ).\adjustlimits{\sup}_{\tau\in\mathbb{R}}{\inf}_{i}(\psi_{\tau}^{i,\mathrm{an}}(v)+\tau)=\adjustlimits{\inf}_{i}{\sup}_{\tau\in\mathbb{R}}(\psi_{\tau}^{i,\mathrm{an}}(v)+\tau). (38)

Due to Lemma 2.1, we have that τψτi,an(v)\tau\mapsto\psi_{\tau}^{i,\mathrm{an}}(v) is concave and usc on \mathbb{R}. So (38) follows from Lemma 3.13, proved below. ∎

We state the following result from convex analysis, a special case of [PZ04, Theorem 2]:

Lemma 3.13.

Let fi:[,)f_{i}:\mathbb{R}\rightarrow[-\infty,\infty) (iIi\in I) be a monotone net of proper usc concave functions and f:[,)f:\mathbb{R}\rightarrow[-\infty,\infty) another proper usc concave function. Assume that fi(τ)f(τ)f_{i}(\tau)\to f(\tau) for all τ\tau\in\mathbb{R}. Then

fˇ=limifˇi.\check{f}=\lim_{i}\check{f}_{i}.

Recall the definition of PSHNA(θ)\mathrm{PSH}^{\mathrm{NA}}(\theta) from Definition 3.3. When {θ}=c1(L)\{\theta\}=c_{1}(L) for some pseudoeffective \mathbb{Q}-line bundle LL we now define the map:

PSHNA(θ)ϕϕanPSH(Lan).\textup{PSH}^{\mathrm{NA}}(\theta)\ni\phi\mapsto\phi^{\mathrm{an}}\in\textup{PSH}(L^{\mathrm{an}}). (39)

Let ϕ={{ϕτω}τ}ω𝒦PSHNA(θ)\phi=\{\{\phi^{\omega}_{\tau}\}_{\tau}\}_{\omega\in\mathcal{K}}\in\textup{PSH}^{\mathrm{NA}}(\theta). Let ω𝒦\omega\in\mathcal{K} be such that {ω}=c1(T)\{\omega\}=c_{1}(T), for a \mathbb{Q}-ample line bundle TT. We get that {ϕτω}τMTC(θ+ω)\{\phi^{\omega}_{\tau}\}_{\tau}\in\textup{MTC}_{\mathcal{I}}(\theta+\omega), hence (ϕω)anPSH((L+T)an)(\phi^{\omega})^{\mathrm{an}}\in\textup{PSH}((L+T)^{\mathrm{an}}) by Theorem 3.12(i). We make the following preliminary definition:

ϕan:=(ϕω)an.\phi^{\mathrm{an}}:=(\phi^{\omega})^{\mathrm{an}}. (40)

Among other things, we need to show that this definition is independent of the choice of ω.\omega. Let ω𝒦\omega^{\prime}\in\mathcal{K}^{\prime} such that {ω}=c1(T)\{\omega^{\prime}\}=c_{1}(T^{\prime}), for some \mathbb{Q}-ample line bundle TT^{\prime} and ωω\omega\leq\omega^{\prime}. Then we have that

Pθ+ω[ϕτω]=ϕτω,τ<τϕ+.P_{\theta+\omega^{\prime}}[\phi^{\omega}_{\tau}]_{\mathcal{I}}=\phi^{\omega^{\prime}}_{\tau},\ \ \ \ \tau<\tau^{+}_{\phi}.

As a result, using [DX22, Proposition 2.18(ii)], we get that (ϕτω)an=(ϕτω)an,τ<τϕ+(\phi^{\omega}_{\tau})^{\mathrm{an}}=(\phi^{\omega^{\prime}}_{\tau})^{\mathrm{an}},\ \tau<\tau^{+}_{\phi}. Comparing with (24), we arrive at (ϕω)an=(ϕω)an(\phi^{\omega})^{\mathrm{an}}=(\phi^{\omega^{\prime}})^{\mathrm{an}}, as desired.

Moreover, our above analysis also shows that ϕanLPSH((L+L)an)\phi^{\mathrm{an}}\in\bigcap_{L^{\prime}}\mathrm{PSH}((L+L^{\prime})^{\mathrm{an}}), where LL^{\prime} runs over all ample \mathbb{Q}-line bundles on XX. The latter space is equal to PSH(Lan)\mathrm{PSH}(L^{\mathrm{an}}) by [BJ21, Theorem 4.5], so our map ϕϕan\phi\mapsto\phi^{\mathrm{an}} from (40) is indeed well-defined.

Observe that

ϕan(vtriv)=τϕ+,\phi^{\mathrm{an}}(v_{\mathrm{triv}})=\tau^{+}_{\phi}, (41)

where vtrivv_{\mathrm{triv}} denotes the trivial valuation.

We show that the map of (39) is actually a bijection, giving a transcendental interpretation of the non-Archimedean metrics of Boucksom–Jonsson:

Theorem 3.14.

Assume that {θ}=c1(L)\{\theta\}=c_{1}(L) for some pseudoeffective \mathbb{Q}-line bundle LL. The map of (39) is a bijection. In addition, let ϕ,χPSHNA(θ)\phi,\chi\in\mathrm{PSH}^{\mathrm{NA}}(\theta). If ϕanχan\phi^{\mathrm{an}}\leq\chi^{\mathrm{an}} then ϕχ\phi\leq\chi.

Proof.

By scaling, we may assume that LL is a line bundle. We first address the last statement. Let ϕ={{ϕτω}τ}ω\phi=\{\{\phi^{\omega}_{\tau}\}_{\tau}\}_{\omega} and χ={{χτω}τ}ω\chi=\{\{\chi^{\omega}_{\tau}\}_{\tau}\}_{\omega}. Given that ϕanχan\phi^{\mathrm{an}}\leq\chi^{\mathrm{an}}, Theorem 3.12(i) gives that ϕτωχτω,τ,ω𝒦\phi^{\omega}_{\tau}\leq\chi^{\omega}_{\tau},\ \tau\in\mathbb{R},\omega\in\mathcal{K}, implying that ϕχ\phi\leq\chi. This immediately gives that ϕϕan\phi\mapsto\phi^{\mathrm{an}} is injective.

To address surjectivity, let χPSH(Lan)\chi\in\textup{PSH}(L^{\mathrm{an}}). Let ω𝒦\omega\in\mathcal{K} such that {ω}=c1(T)\{\omega\}=c_{1}(T) for a \mathbb{Q}-ample line bundle TT. Then χPSH((L+T)an)\chi\in\textup{PSH}((L+T)^{\mathrm{an}}), and by Theorem 3.12(ii) there exists {ϕτω}τMTC(θ+ω)\{\phi^{\omega}_{\tau}\}\tau\in\textup{MTC}_{\mathcal{I}}(\theta+\omega) such that (ϕω)an=χ(\phi^{\omega})^{\mathrm{an}}=\chi.

Let ω𝒦\omega^{\prime}\in\mathcal{K} such that {ω}=c1(T)\{\omega^{\prime}\}=c_{1}(T^{\prime}), for some \mathbb{Q}-ample line bundle TT^{\prime} and ωω\omega\leq\omega^{\prime}. Due to injectivity of MTC(θ+ω)PSH((L+T)an)\textup{MTC}_{\mathcal{I}}(\theta+\omega^{\prime})\mapsto\mathrm{PSH}((L+T^{\prime})^{\mathrm{an}}) we get that Pθ+ω[ϕτω]=ϕτω,τ<τϕ+P_{\theta+\omega^{\prime}}[\phi^{\omega}_{\tau}]_{\mathcal{I}}=\phi^{\omega^{\prime}}_{\tau},\ \tau<\tau^{+}_{\phi}.

For a non-rational form ω′′𝒦\omega^{\prime\prime}\in\mathcal{K}, let ω𝒦\omega\in\mathcal{K} such that {ω}=c1(T)\{\omega\}=c_{1}(T) for a \mathbb{Q}-ample line bundle TT and ωω′′\omega\leq\omega^{\prime\prime}. Then we define

ϕτω′′=Pθ+ω′′[ϕτω],τ<τϕ+.\phi^{\omega^{\prime\prime}}_{\tau}=P_{\theta+\omega^{\prime\prime}}[\phi^{\omega}_{\tau}]_{\mathcal{I}},\ \ \ \tau<\tau^{+}_{\phi}.

It is immediate to see that the above definition is independent of the choice of ω\omega. We see that {{ϕτω}τ}ωPSHNA(θ)\{\{\phi^{\omega}_{\tau}\}_{\tau}\}_{\omega}\in\textup{PSH}^{\mathrm{NA}}(\theta), and ϕan=χ\phi^{\mathrm{an}}=\chi, proving surjectivity. ∎

Remark 3.15.

Boucksom–Jonsson’s theory extends to any pseudoeffective (1,1)(1,1)-class {θ}\{\theta\} in the Néron–Severi group of XX with real coefficients, giving a space of non-Archimedean metrics PSH({θ}NA)\mathrm{PSH}(\{\theta\}^{\mathrm{NA}}). In this case we still have a canonical identification

PSHNA(θ)PSH({θ}NA).\mathrm{PSH}^{\mathrm{NA}}(\theta)\xrightarrow{\sim}\mathrm{PSH}(\{\theta\}^{\mathrm{NA}}). (42)

To see this, let 𝒦\mathcal{K}^{\prime} be the directed set of (1,1)(1,1)-classes {θ}\{\theta^{\prime}\} in the Néron–Severi group of XX with rational coefficients such that {θ}{θ}\{\theta^{\prime}\}-\{\theta\} is a Kähler class. Then we have natural identifications

PSHNA(θ)limθ𝒦PSHNA(θ),PSH({θ}NA)limθ𝒦PSH({θ}NA).\mathrm{PSH}^{\mathrm{NA}}(\theta)\xrightarrow{\sim}\varprojlim_{\theta^{\prime}\in\mathcal{K}^{\prime}}\mathrm{PSH}^{\mathrm{NA}}(\theta^{\prime}),\quad\mathrm{PSH}(\{\theta\}^{\mathrm{NA}})\xrightarrow{\sim}\varprojlim_{\theta^{\prime}\in\mathcal{K}^{\prime}}\mathrm{PSH}(\{\theta^{\prime}\}^{\mathrm{NA}}).

The former follows immediately from our definition, the latter follows from [BJ21, Section 4.1, (PSH2)]. In particular, (42) follows from Theorem 3.14.

Using our analysis we conclude that the envelope conjecture holds in our setting, as confirmed by Boucksom–Jonsson using non-Archimedean methods [BJ22]:

Corollary 3.16.

Conjecture 2.13 holds for any pseudo-effective \mathbb{Q}-line bundle LL on XX.

Proof.

Take a smooth closed real (1,1)(1,1)-form θ\theta in c1(L)c_{1}(L). Let {φi}iI\{\varphi^{i}\}_{i\in I} be a bounded from increasing net in PSHNA(L)\mathrm{PSH}^{\mathrm{NA}}(L). By Theorem 3.14, φi=ϕi,an\varphi^{i}=\phi^{i,\mathrm{an}} with ϕiPSHNA(θ)\phi^{i}\in\mathrm{PSH}^{\mathrm{NA}}(\theta). By the same theorem, {φi}iI\{\varphi^{i}\}_{i\in I} is also an increasing net. Moreover, by (41),

supiIτφi+<.\sup_{i\in I}\tau^{+}_{\varphi^{i}}<\infty.

By Theorem 3.4, we can find ϕPSHNA(θ)\phi\in\mathrm{PSH}^{\mathrm{NA}}(\theta) such that for any ω𝒦\omega\in\mathcal{K}, ϕτi,ωϕτω\phi^{i,\omega}_{\tau}\nearrow\phi^{\omega}_{\tau} almost everywhere for any τ<τϕ+\tau<\tau^{+}_{\phi}.

We claim that ϕi,an\phi^{i,\mathrm{an}} converges to ϕan\phi^{\mathrm{an}}.

For any fixed ω𝒦\omega\in\mathcal{K}, the usc property of the Lelong number [Bou17, Lemma 2.6] gives

(ϕτω)an(v)=limi(ϕτi,ω)an(v)for vXdiv,τ<τψ+.(\phi^{\omega}_{\tau})^{\mathrm{an}}(v)=\lim_{i}(\phi^{i,\omega}_{\tau})^{\mathrm{an}}(v)\quad\text{for }v\in X^{\mathrm{div}},\tau<\tau_{\psi}^{+}.

It follows that

ϕan(v)=(ϕω)an(v)=supτ((ϕτω)an(v)+τ)=supτsupiI((ϕτi,ω)an(v)+τ)=supiIϕi,an(v)\phi^{\mathrm{an}}(v)=(\phi^{\omega})^{\mathrm{an}}(v)=\sup_{\tau\in\mathbb{R}}((\phi^{\omega}_{\tau})^{\mathrm{an}}(v)+\tau)=\sup_{\tau\in\mathbb{R}}\sup_{i\in I}((\phi^{i,\omega}_{\tau})^{\mathrm{an}}(v)+\tau)=\sup_{i\in I}\phi^{i,\mathrm{an}}(v)

for all vXdivv\in X^{\mathrm{div}}, finishing the proof. ∎

3.4 The non-Archimedean finite energy space

Let XX be a compact Kähler manifold of dimension nn and {θ}\{\theta\} be a big cohomology class on XX. Let V=XθVθnV=\int_{X}\theta_{V_{\theta}}^{n} denote the volume of {θ}\{\theta\}.

Recall that MTC(θ)\textup{MTC}_{\mathcal{I}}(\theta) naturally embeds into PSHNA(θ)\mathrm{PSH}^{\mathrm{NA}}(\theta) (29). For {ητ}τMTC(θ)\{\eta_{\tau}\}_{\tau}\in\textup{MTC}_{\mathcal{I}}(\theta) we define the Monge–Ampère energy IθNA:MTC(θ)[,)I^{\mathrm{NA}}_{\theta}:\textup{MTC}_{\mathcal{I}}(\theta)\rightarrow[-\infty,\infty) as

IθNA(η):=Iθ{ητ}=τη++1Vτη+(XθητnV)dτ.I_{\theta}^{\mathrm{NA}}(\eta):=I_{\theta}\{\eta_{\tau}\}=\tau^{+}_{\eta}+\frac{1}{V}\int_{-\infty}^{\tau^{+}_{\eta}}\left(\int_{X}\theta_{\eta_{\tau}}^{n}-V\right)\,\mathrm{d}\tau. (43)

We extend IθNAI^{\mathrm{NA}}_{\theta} to a function on PSHNA(θ)\mathrm{PSH}^{\mathrm{NA}}(\theta) by setting it to be -\infty on PSHNA(θ)MTC(θ)\mathrm{PSH}^{\mathrm{NA}}(\theta)\setminus\textup{MTC}_{\mathcal{I}}(\theta).

We then define the non-Archimedean finite energy space as

1,NA(X,θ):={ϕPSHNA(θ):IθNA(ϕ)>}.\mathcal{E}^{1,\mathrm{NA}}(X,\theta):=\left\{\phi\in\mathrm{PSH}^{\mathrm{NA}}(\theta):I_{\theta}^{\mathrm{NA}}(\phi)>-\infty\right\}\,.

Since VXθητnV\geq\int_{X}\theta_{\eta_{\tau}}^{n}, for (43) to be finite Xθητn\int_{X}\theta_{\eta_{\tau}}^{n} must converge to VV as τ\tau\searrow-\infty. By [DDNL18a, Theorem 2.3] this implies that Xθηn=V\int_{X}\theta_{\eta_{-\infty}}^{n}=V. Since η\eta_{-\infty} is maximal, we obtain that η=Vθ\eta_{-\infty}=V_{\theta}, i.e., {ητ}τ\{\eta_{\tau}\}_{\tau} is a test curve relative to VθV_{\theta}, whenever {ητ}τ1,NA(X,θ)\{\eta_{\tau}\}_{\tau}\in\mathcal{E}^{1,\mathrm{NA}}(X,\theta).

Similarly, the non-Archimedean 𝒥\mathcal{J} functional is introduced as

𝒥θNA(η)=τη+IθNA(η)=1Vτη+(VXθητn)dτ.\mathcal{J}_{\theta}^{\mathrm{NA}}(\eta)=\tau^{+}_{\eta}-I_{\theta}^{\mathrm{NA}}(\eta)=\frac{1}{V}\int_{-\infty}^{\tau^{+}_{\eta}}\left(V-\int_{X}\theta_{\eta_{\tau}}^{n}\right)\,\mathrm{d}\tau. (44)

When {θ}=c1(L)\{\theta\}=c_{1}(L) for some ample \mathbb{Q}-line bundle LL, it has been pointed out in [DX22, Theorem 1.1, Theorem 1.2] that our definitions of 1,NA(X,θ)\mathcal{E}^{1,\mathrm{NA}}(X,\theta) and 𝒥θNA,IθNA\mathcal{J}_{\theta}^{\mathrm{NA}},I_{\theta}^{\mathrm{NA}} coincide with the ones given in [BJ21].

In the literature, the space 1(Lan)\mathcal{E}^{1}(L^{\mathrm{an}}) has not yet been defined for a general big \mathbb{Q}-line bundle LL. Above we gave an analytic definition, but it is desirable to have a purely non-Archimedean/algebraic definition as well.

After defining finite energy non-Archimedean spaces, we can finally relate the \mathcal{I}-maximal finite energy rays to non-Archimedean potentials, as a direct consequence of Theorem 2.6:

Theorem 3.17.

There is a bijective function {ψτ}τ{ψˇt}t\{\psi_{\tau}\}_{\tau}\mapsto\{\check{\psi}_{t}\}_{t}, between 1(X,θ)\mathcal{R}^{1}_{\mathcal{I}}(X,\theta) and 1,NA(X,θ)\mathcal{E}^{1,\mathrm{NA}}(X,\theta). Moreover, under this correspondence, the radial Monge–Ampère energy corresponds to the non-Archimedean Monge–Ampère energy defined in (43).

4 Approximations of \mathcal{I}-maximal rays by filtrations and flag configurations

In this section XX is a connected projective manifold of dimension nn. The purpose of this section is to show that an \mathcal{I}-maximal finite energy ray can be approximated by simpler rays induced by flag configurations or filtrations defined below.

4.1 Filtrations of big line bundles

For this subsection let LL be a big line bundle on XX and hh be a smooth Hermitian metric. Let θ:=ddclogh\theta:=-\mathrm{dd}^{\mathrm{c}}\log h. We set vol(L):=XθVθn\operatorname{vol}(L):=\int_{X}\theta_{V_{\theta}}^{n}. Denote the section ring of LL by R(X,L)R(X,L):

R(X,L):=mRm,Rm:=H0(X,Lm).R(X,L):=\bigoplus_{m\in\mathbb{N}}R_{m},\quad R_{m}:=H^{0}(X,L^{m}).
Definition 4.1.

By a (bounded left-continuous multiplicative decreasing) filtration \mathcal{F} of the section ring R(X,L)R(X,L), we mean a family of linear subspaces {λRm}λ\{\mathcal{F}^{\lambda}R_{m}\}_{\lambda\in\mathbb{R}} of RmR_{m} for each mm\in\mathbb{N} such that
(i)λRmλRm\mathcal{F}^{\lambda}R_{m}\subseteq\mathcal{F}^{\lambda^{\prime}}R_{m} whenever λλ\lambda\geq\lambda^{\prime};
(ii) λRm=λ<λλRm\mathcal{F}^{\lambda}R_{m}=\bigcap_{\lambda^{\prime}<\lambda}\mathcal{F}^{\lambda^{\prime}}R_{m};
(iii)λ1Rm1λ2Rm2λ1+λ2Rm1+m2\mathcal{F}^{\lambda_{1}}R_{m_{1}}\cdot\mathcal{F}^{\lambda_{2}}R_{m_{2}}\subseteq\mathcal{F}^{\lambda_{1}+\lambda_{2}}R_{m_{1}+m_{2}} for λ1,λ2\lambda_{1},\lambda_{2}\in\mathbb{R} and m1,m2m_{1},m_{2}\in\mathbb{N};
(iv) there exists C>0C>0 such that CmRm=Rm\mathcal{F}^{-Cm}R_{m}=R_{m} and CmRm={0}\mathcal{F}^{Cm}R_{m}=\{0\} for all mm\in\mathbb{N}.

Given a filtration \mathcal{F} of R(X,L)R(X,L) and mm\in\mathbb{N}, set

τm():=max{λ:λRm{0}}.\tau_{m}(\mathcal{F}):=\max\{\lambda\in\mathbb{R}:\mathcal{F}^{\lambda}R_{m}\neq\{0\}\}.

Then clearly τm()+τk()τm+k()\tau_{m}(\mathcal{F})+\tau_{k}(\mathcal{F})\leq\tau_{m+k}(\mathcal{F}) for any m,km,k\in\mathbb{N}, so by Fekete’s lemma one can put

τL():=limmτm()m=supmτm()m.\tau_{L}(\mathcal{F}):=\lim_{m\rightarrow\infty}\frac{\tau_{m}(\mathcal{F})}{m}=\sup_{m\in\mathbb{N}}\frac{\tau_{m}(\mathcal{F})}{m}.

Moreover, for any τ<τ()\tau<\tau(\mathcal{F}), set

vol(R(τ)):=lim¯mdimτmRmmn/n!\operatorname{vol}(\mathcal{F}R^{(\tau)}):=\varlimsup_{m\rightarrow\infty}\frac{\dim\mathcal{F}^{\tau m}R_{m}}{m^{n}/n!}

and

SL():=τL()+1vol(L)τL()(vol(R(τ))vol(L))dτ.S_{L}(\mathcal{F}):=\tau_{L}(\mathcal{F})+\frac{1}{\operatorname{vol}(L)}\int_{-\infty}^{\tau_{L}(\mathcal{F})}\left(\operatorname{vol}(\mathcal{F}R^{(\tau)})-\operatorname{vol}(L)\right)\,\mathrm{d}\tau.

It is trivial to see that SL()τL()S_{L}(\mathcal{F})\leq\tau_{L}(\mathcal{F}) in general.

Let EE be any prime divisor over XX. It induces a natural filtration E\mathcal{F}_{E} on the section ring R(X,L)R(X,L). More precisely, for λ\lambda\in\mathbb{R} and m1m\geq 1 one puts

EλRm:={sRm=H0(X,Lm):ordE(s)λ}.\mathcal{F}_{E}^{\lambda}R_{m}:=\{s\in R_{m}=H^{0}(X,L^{m}):\operatorname{ord}_{E}(s)\geq\lambda\}.

In this case, we set for simplicity

τL(E):=τL(E) and SL(E):=SL(E).\tau_{L}(E):=\tau_{L}(\mathcal{F}_{E})\text{ and }S_{L}(E):=S_{L}(\mathcal{F}_{E}).
Lemma 4.2.

For any prime divisor EE over XX, one has SL(E)<τL(E)S_{L}(E)<\tau_{L}(E).

Proof.

Consider f(τ):=vol(ER(τ))=vol(LτE)f(\tau):=\operatorname{vol}(\mathcal{F}_{E}R^{(\tau)})=\operatorname{vol}(L-\tau E) for τ[0,τL(E)]\tau\in[0,\tau_{L}(E)]. This is a continuous and decreasing function with f(0)=vol(L)>0f(0)=\operatorname{vol}(L)>0 and f(τL(E))=0f(\tau_{L}(E))=0. So one must have

SL(E)=1vol(L)0τL(E)f(τ)dτ<1vol(L)0τL(E)f(0)dτ=τL(E).S_{L}(E)=\frac{1}{\operatorname{vol}(L)}\int_{0}^{\tau_{L}(E)}f(\tau)\,\mathrm{d}\tau<\frac{1}{\operatorname{vol}(L)}\int_{0}^{\tau_{L}(E)}f(0)\,\mathrm{d}\tau=\tau_{L}(E).

Definition 4.3.

Given any filtration \mathcal{F} of R(X,L)R(X,L), for τ<τL()\tau<\tau_{L}(\mathcal{F}) and m>0m\in\mathbb{Z}_{>0}, let

ϕτ,m:=uscsup{1mlog|s|hm2:sτmH0(X,Lm),|s|hm21} and ϕτ:=uscsupmϕτ,m.\phi^{\mathcal{F}}_{\tau,m}:=\operatorname{usc}\sup\left\{\frac{1}{m}\log|s|^{2}_{h^{m}}:s\in\mathcal{F}^{\tau m}H^{0}(X,L^{m}),\ \ |s|^{2}_{h^{m}}\leq 1\right\}\textup{ and }\phi^{\mathcal{F}}_{\tau}:=\operatorname{usc}\sup_{m}\phi^{\mathcal{F}}_{\tau,m}.

Observe that mϕτ,mm\phi^{\mathcal{F}}_{\tau,m} is super-additive, so by Fekete’s lemma,

ϕτ=usclimmϕτ,m.\phi^{\mathcal{F}}_{\tau}=\operatorname{usc}\lim_{m\to\infty}\phi^{\mathcal{F}}_{\tau,m}.

Setting ϕτ():=limττ()ϕτ\phi^{\mathcal{F}}_{\tau(\mathcal{F})}:=\lim_{{\tau\nearrow\tau(\mathcal{F})}}\phi_{\tau} and ϕτ=\phi^{\mathcal{F}}_{\tau}=-\infty if τ>τ()\tau>\tau(\mathcal{F}), we get that ϕ={ϕτ}τ\phi^{\mathcal{F}}=\{\phi^{\mathcal{F}}_{\tau}\}_{\tau} is a bounded test curve, called the test curve associated with \mathcal{F}.

When working with a divisor EE over XX we will use the notation

ϕτ,mE:=ϕτ,mE,ϕτE:=ϕτE.\phi^{E}_{\tau,m}:=\phi^{\mathcal{F}_{E}}_{\tau,m},\quad\phi^{E}_{\tau}:=\phi^{\mathcal{F}_{E}}_{\tau}.

But we need to point out that, as opposed to the ample setting studied in [RWN14, DX22], it is not clear if the test curves thus constructed are maximal in the big case. We conjecture this to be the case, but as we shall see, this uncertainty will not cause any issue for the discussions below.

Definition 4.4 ([PS07, RWN14]).

We define the Phong–Sturm ray {rt}t\{r^{\mathcal{F}}_{t}\}_{t} associated with a filtration \mathcal{F} of R(X,L)R(X,L) to be the inverse Legendre transform of the maximization of {ϕτ}τ\{\phi^{\mathcal{F}}_{\tau}\}_{\tau}, defined in (19).

Due to boundedness of \mathcal{F}, we immediately notice that r0=Vθr^{\mathcal{F}}_{0}=V_{\theta} and r^τ=Vθ\hat{r}_{\tau}^{\mathcal{F}}=V_{\theta} for τC\tau\leq-C. This implies that the potentials r^τ,ϕτ,τ<τ()\hat{r}^{\mathcal{F}}_{\tau},\phi^{\mathcal{F}}_{\tau},\ \tau<\tau(\mathcal{F}) have non-zero mass.

According to the next lemma, the Phong-Sturm ray is \mathcal{I}-maximal:

Lemma 4.5.

We have that r^τ=P[ϕτ]=P[ϕτ]\hat{r}^{\mathcal{F}}_{\tau}=P[\phi^{\mathcal{F}}_{\tau}]=P[\phi^{\mathcal{F}}_{\tau}]_{\mathcal{I}} for any τ<τL()\tau<\tau_{L}(\mathcal{F}).

Proof.

By Fekete’s lemma, ψm:=ϕτ,2mϕτP[ϕτ].\psi_{m}:=\phi^{\mathcal{F}}_{\tau,2^{m}}\nearrow\phi^{\mathcal{F}}_{\tau}\leq P[\phi^{\mathcal{F}}_{\tau}]. This implies that P[ψm]=P[ψm]P[ϕτ]P[\psi_{m}]_{\mathcal{I}}=P[\psi_{m}]\leq P[\phi^{\mathcal{F}}_{\tau}], by [DX22, Proposition 2.20].

Let χ:=limmP[ψm]\chi:=\lim_{m}P[\psi_{m}]_{\mathcal{I}}. Since τ<τL()\tau<\tau_{L}(\mathcal{F}) we have that χ=P[χ]\chi=P[\chi]_{\mathcal{I}}, by [DX22, Lemma 2.21(iii)]. Since ϕτχP[ϕτ]\phi^{\mathcal{F}}_{\tau}\leq\chi\leq P[\phi^{\mathcal{F}}_{\tau}], and \mathcal{I}-model potentials are model [DX22, Proposition 2.18] we obtain that χ=P[ϕτ]\chi=P[\phi^{\mathcal{F}}_{\tau}], hence P[ϕτ]=P[ϕτ]P[\phi^{\mathcal{F}}_{\tau}]=P[\phi^{\mathcal{F}}_{\tau}]_{\mathcal{I}}. ∎

Lastly, we note the following formula for the radial Monge–Ampère energy of the Phong–Sturm ray:

Proposition 4.6.

Let \mathcal{F} be a filtration of R(X,L)R(X,L). Then Iθ{rt}=SL()I_{\theta}\{r^{\mathcal{F}}_{t}\}=S_{L}(\mathcal{F}).

Proof.

First, by [DZ22, (25)],

Iθ{rt}=Iθ{ϕˇt}=τL()+1vol(L)τL()(Xθϕτnvol(L))dτ.I_{\theta}\{r^{\mathcal{F}}_{t}\}=I_{\theta}\{\check{\phi}^{\mathcal{F}}_{t}\}=\tau_{L}(\mathcal{F})+\frac{1}{\operatorname{vol}(L)}\int_{-\infty}^{\tau_{L}(\mathcal{F})}\left(\int_{X}\theta^{n}_{\phi^{\mathcal{F}}_{\tau}}-\operatorname{vol}(L)\right)\mathrm{d}\tau.

So it is enough to argue that Xθϕτn=vol(R(τ))\int_{X}\theta^{n}_{\phi^{\mathcal{F}}_{\tau}}=\operatorname{vol}(\mathcal{F}R^{(\tau)}) for any τ<τL()\tau<\tau_{L}(\mathcal{F}). This is a consequence of [His13, Theorem 1.3]. Indeed, consider the graded linear series R(τ):={τmRm}m\mathcal{F}R^{(\tau)}:=\{\mathcal{F}^{\tau m}R_{m}\}_{m\in\mathbb{N}}. It contains an ample linear series in the sense of [BC11, Lemma 1.6]. Then for all mm sufficiently divisible the natural map X(τmRm)X\dashrightarrow\mathbb{P}(\mathcal{F}^{\tau m}R_{m})^{*} is birational to its image. So [His13, Theorem 1.3] implies that Xθϕτn=vol(R(τ))\int_{X}\theta^{n}_{\phi^{\mathcal{F}}_{\tau}}=\operatorname{vol}(\mathcal{F}R^{(\tau)}), as wished. ∎

4.2 Flag configurations

Definition 4.7.

A flag configuration of a big cohomology class θ\theta is a (partial) flag of coherent analytic ideal sheaves

𝔞0𝔞1𝔞N=(Vθ).\mathfrak{a}_{0}\subseteq\mathfrak{a}_{1}\subseteq\dots\subseteq\mathfrak{a}_{N}=\mathcal{I}(V_{\theta}).

By convention, 𝔞:=0\mathfrak{a}_{\ell}:=0 for <0\ell\in\mathbb{Z}_{<0} and 𝔞:=𝔞N\mathfrak{a}_{\ell}:=\mathfrak{a}_{N} if N\ell\in\mathbb{Z}_{\geq N}.

A flag configuration will be sometimes conveniently denoted as an analytic coherent ideal sheaf on the product X×X\times\mathbb{C}:

𝔞:=𝔞0+𝔞1s+𝔞N1sN1+𝔞N(sN)𝒪(X×).\mathfrak{a}:=\mathfrak{a}_{0}+\mathfrak{a}_{1}s+\dots\mathfrak{a}_{N-1}s^{N-1}+\mathfrak{a}_{N}(s^{N})\subseteq\mathcal{O}(X\times\mathbb{C}).

For rr\in\mathbb{N} and λ\lambda\in\mathbb{N} we introduce the following coherent sheaves associated with a flag configuration 𝔞\mathfrak{a}:

𝔞r,λ:=λ1++λr=λλi𝔞λ1𝔞λ2𝔞λr=jβj=rjjβj=λβj for jj=0𝔞jβj.\mathfrak{a}_{r,\lambda}:=\sum_{\begin{subarray}{c}\lambda_{1}+\dots+\lambda_{r}=\lambda\\ \lambda_{i}\in\mathbb{N}\end{subarray}}\mathfrak{a}_{\lambda_{1}}\mathfrak{a}_{\lambda_{2}}\dots\mathfrak{a}_{\lambda_{r}}=\sum_{\begin{subarray}{c}\sum_{j}\beta_{j}=r\\ \sum_{j}j\beta_{j}=\lambda\\ \beta_{j}\in\mathbb{N}\text{ for }j\in\mathbb{N}\end{subarray}}\prod_{j=0}^{\infty}\mathfrak{a}_{j}^{\beta_{j}}. (45)

We notice that λ𝔞λr\lambda\mapsto\mathfrak{a}^{r}_{\lambda} is increasing and

𝔞r,λ𝔞r,λ𝔞r+r,λ+λ.\mathfrak{a}_{r,\lambda}\cdot\mathfrak{a}_{r^{\prime},\lambda^{\prime}}\subseteq\mathfrak{a}_{r+r^{\prime},\lambda+\lambda^{\prime}}. (46)

For λ>0\lambda\in\mathbb{R}_{>0}\setminus\mathbb{Z}, we formally set 𝔞r,λ=𝔞r,λ\mathfrak{a}_{r,\lambda}=\mathfrak{a}_{r,\lfloor\lambda\rfloor}. For λ<0\lambda<0, we set 𝔞r,λ=0\mathfrak{a}_{r,\lambda}=0.

Returning to the case when {θ}=c1(L)\{\theta\}=c_{1}(L), to a flag configuration 𝔞\mathfrak{a}, one can associate a filtration of R(X,L)=rH0(X,Lr)R(X,L)=\bigoplus_{r\in\mathbb{N}}H^{0}(X,L^{r}), following ideas from [Oda13]:

𝔞λH0(X,Lr):=H0(X,Lr𝔞r,λ)H0(X,Lr).\mathcal{F}^{\lambda}_{\mathfrak{a}}H^{0}(X,L^{r}):=H^{0}(X,L^{r}\otimes\mathfrak{a}_{r,-\lambda})\subseteq H^{0}(X,L^{r}). (47)
Proposition 4.8.

For any flag configuration 𝔞\mathfrak{a} of LL, 𝔞λ\mathcal{F}^{\lambda}_{\mathfrak{a}} defined in (47) is a filtration on R(X,L)R(X,L) in the sense of Definition 4.1.

Proof.

We need to verify the conditions in Definition 4.1. Condition (i) and (ii) are obvious. We verify Condition (iii), which says that for any λ,λ\lambda,\lambda^{\prime}\in\mathbb{R}, r,rr,r^{\prime}\in\mathbb{N}, we have

𝔞λH0(X,Lar)𝔞λH0(X,Lar)𝔞λ+λH0(X,Lr+r).\mathcal{F}^{\lambda}_{\mathfrak{a}}H^{0}(X,L^{ar})\mathcal{F}^{\lambda^{\prime}}_{\mathfrak{a}}H^{0}(X,L^{ar^{\prime}})\subseteq\mathcal{F}^{\lambda+\lambda^{\prime}}_{\mathfrak{a}}H^{0}(X,L^{r+r^{\prime}}).

By definition, this amounts to

H0(X,Lr𝔞r,λ)H0(X,Lr𝔞r,λ)H0(X,Lr+r𝔞r+r,(λ+λ)).H^{0}(X,L^{r}\otimes\mathfrak{a}_{r,-\lambda})\cdot H^{0}(X,L^{r^{\prime}}\otimes\mathfrak{a}_{r^{\prime},-\lambda^{\prime}})\subseteq H^{0}(X,L^{r+r^{\prime}}\otimes\mathfrak{a}_{r+r^{\prime},-(\lambda+\lambda^{\prime})}).

It suffices to argue

𝔞r,λ𝔞r,λ𝔞r+r,(λ+λ).\mathfrak{a}_{r,-\lambda}\cdot\mathfrak{a}_{r^{\prime},-\lambda^{\prime}}\subseteq\mathfrak{a}_{r+r^{\prime},-(\lambda+\lambda^{\prime})}. (48)

When λ>0\lambda>0 or λ>0\lambda^{\prime}>0 (48) is trivial. So we may assume that λ,λ0\lambda,\lambda^{\prime}\leq 0, and (48) follows from (46).

It remains to argue Condition (iv) in Definition 4.1. Namely, the filtration is linearly bounded. By definition, 𝔞λH0(X,Lr)=0\mathcal{F}^{\lambda}_{\mathfrak{a}}H^{0}(X,L^{r})=0 for λ>0\lambda>0. So it suffices to show that there is C>0C>0 so that

H0(X,Lr𝔞r,Cr)=H0(X,Lr).H^{0}(X,L^{r}\otimes\mathfrak{a}_{r,Cr})=H^{0}(X,L^{r}). (49)

We claim that it suffices to take C=NC=N. In this case, 𝔞r,Nr=𝔞Nr=(Vθ)r(rVθ)\mathfrak{a}_{r,Nr}=\mathfrak{a}^{r}_{N}=\mathcal{I}(V_{\theta})^{r}\supseteq\mathcal{I}(rV_{\theta}) by the multiplicativity of multiplier ideal sheaves [DEL00]. As VθV_{\theta} has minimal singularities, H0(X,Lr(rVθ))=H0(X,Lr)H^{0}(X,L^{r}\otimes\mathcal{I}(rV_{\theta}))=H^{0}(X,L^{r}), so (49) follows. ∎

Definition 4.9.

The Phong–Sturm ray {rtL,𝔞}tθ1\{r^{L,\mathfrak{a}}_{t}\}_{t}\in\mathcal{R}^{1}_{\theta} associated to a flag configuration 𝔞=𝔞0+𝔞1s+𝔞N1sN1+𝔞N(sN)\mathfrak{a}=\mathfrak{a}_{0}+\mathfrak{a}_{1}s+\dots\mathfrak{a}_{N-1}s^{N-1}+\mathfrak{a}_{N}(s^{N}) is the Phong–Sturm ray of the associated filtration 𝔞λ\mathcal{F}_{\mathfrak{a}}^{\lambda} (recall Definition 4.4). When there is no risk of confusion, we simply write {rt𝔞}t\{r^{\mathfrak{a}}_{t}\}_{t}.

Careful readers might notice that in the ample case there is a minor difference in our definition of filtration associated to a flag ideal and the one in [BHJ17], when LL is ample. Though this is the case, we now point out that the associated Phong–Sturm ray is going to be the same, regardless what filtration one works with, ours or the one in [BHJ17].

Assume that LL is ample and 𝔞\mathfrak{a} is a flag configuration with 𝔞N=𝒪X\mathfrak{a}_{N}=\mathcal{O}_{X}. We assume that L𝔞jL\otimes\mathfrak{a}_{j} is globally generated for all jj. As described in [Oda13, BHJ17] it is possible to associate to such 𝔞\mathfrak{a} a test configuration: let p:𝒳X×p:\mathcal{X}\rightarrow X\times\mathbb{C} be the normalized blow up of X×X\times\mathbb{C} with respect to flag ideal 𝔞:=𝔞0+𝔞1s++𝔞N(sN)\mathfrak{a}:=\mathfrak{a}_{0}+\mathfrak{a}_{1}s+\dots+\mathfrak{a}_{N}(s^{N}) and let EE be the exceptional divisor. Set =pp1L𝒪(E)\mathcal{L}=p^{*}p_{1}^{*}L\otimes\mathcal{O}(-E), where p1:X×Xp_{1}:X\times\mathbb{C}\rightarrow X is the natural projection. Then with respect to the obvious \mathbb{C}^{*}-action, (𝒳,)(\mathcal{X},\mathcal{L}) is a normal semi-ample test configuration of (X,L)(X,L) in the traditional sense. Let (𝒳,)\mathcal{F}_{(\mathcal{X},\mathcal{L})} denote the \mathbb{Z}-filtration of R(X,L)R(X,L) induced by (𝒳,)(\mathcal{X},\mathcal{L}), see [BHJ17, Section 2.5] for the precise definition. It is well-known that

(𝒳,)λH0(X,Lr)=H0(X,Lr𝔞r¯λ),\mathcal{F}_{(\mathcal{X},\mathcal{L})}^{\lambda}H^{0}(X,L^{r})=H^{0}(X,L^{r}\otimes\overline{\mathfrak{a}^{r}}_{-\lambda}), (50)

where 𝔞r¯\overline{\mathfrak{a}^{r}} is the integral closure of 𝔞r\mathfrak{a}^{r} and the subindex λ-\lambda denotes the coefficient of sλs^{-\lambda} in the expansion in ss ([BHJ17, Proposition 2.21]). We formally extend (𝒳,)\mathcal{F}_{(\mathcal{X},\mathcal{L})} to an \mathbb{R}-filtration by setting

(𝒳,)λH0(X,Lr)=(𝒳,)λH0(X,Lr).\mathcal{F}_{(\mathcal{X},\mathcal{L})}^{\lambda}H^{0}(X,L^{r})=\mathcal{F}_{(\mathcal{X},\mathcal{L})}^{\lceil\lambda\rceil}H^{0}(X,L^{r}).

From (50), we find immediately that

(𝒳,)λH0(X,Lr)𝔞λH0(X,Lr),\mathcal{F}_{(\mathcal{X},\mathcal{L})}^{\lambda}H^{0}(X,L^{r})\supseteq\mathcal{F}^{\lambda}_{\mathfrak{a}}H^{0}(X,L^{r}), (51)

and strict containment naturally occurs, however we now confirm that the associated Phong–Sturm rays do coincide.

Indeed, let {ψτ(𝒳,)}τ\{\psi^{(\mathcal{X},\mathcal{L})}_{\tau}\}_{\tau} be the test curve defined by the filtration (𝒳,)\mathcal{F}_{(\mathcal{X},\mathcal{L})}. Due to Theorem 3.12(ii) it is enough to show that ψ(𝒳,)=an(r^𝔞)an\psi^{(\mathcal{X},\mathcal{L})}{}^{\mathrm{an}}=(\hat{r}^{\mathfrak{a}})^{\mathrm{an}}. This is confirmed by the next lemma.

Lemma 4.10.

Under the assumptions above, we have

(r^𝔞)an(v)=ψ(𝒳,)(v)an=σ(v)(𝔞)=maxj(v(𝔞j)j)for vXdiv.(\hat{r}^{\mathfrak{a}})^{\mathrm{an}}(v)=\psi^{(\mathcal{X},\mathcal{L})}{}^{\mathrm{an}}(v)=-\sigma(v)(\mathfrak{a})=\max_{j}(-v(\mathfrak{a}_{j})-j)\quad\text{for }v\in X^{\mathrm{div}}. (52)
Proof.

Due to Fekete’s lemma we have that ϕτ𝔞ϕτ,2k𝔞ϕτ,1𝔞\phi^{\mathcal{F}^{\mathfrak{a}}_{\tau}}\nwarrow\phi^{\mathcal{F}^{\mathfrak{a}}_{\tau,2^{k}}}\geq\phi^{\mathcal{F}^{\mathfrak{a}}_{\tau,1}}. Due to global generatedness of L𝔞τL\otimes\mathfrak{a}_{-\tau} we get that v(ϕτ)v(τH0(X,L))=v(𝔞j),vXdiv-v(\phi^{\mathcal{F}}_{\tau})\geq-v\left(\mathcal{F}^{\tau}H^{0}(X,L)\right)=-v\left(\mathfrak{a}_{j}\right),\ v\in X^{\mathrm{div}}. As a result,

(r^𝔞)anmaxj(v(𝔞j)j).(\hat{r}^{\mathfrak{a}})^{\mathrm{an}}\geq\max_{j}(-v(\mathfrak{a}_{j})-j).

We argue the reverse direction. Due to (51) we have that ψˇt(𝒳,)rt𝔞.\check{\psi}^{(\mathcal{X},\mathcal{L})}_{t}\geq r^{\mathfrak{a}}_{t}. Hence, it suffices to show that

maxj(v(𝔞j)j)ψˇ(𝒳,),an(v)\max_{j}(-v(\mathfrak{a}_{j})-j)\geq\check{\psi}^{(\mathcal{X},\mathcal{L}),\mathrm{an}}(v) (53)

for any vXdivv\in X^{\mathrm{div}}. However, we actually have equality here. Indeed, by [RWN14, Theorem 9.2] and [BBJ21, Lemma 4.4] we have that ψˇ(𝒳,),an(v)=φ𝒳,(v)\check{\psi}^{(\mathcal{X},\mathcal{L}),\mathrm{an}}(v)=\varphi_{\mathcal{X},\mathcal{L}}(v), where φ𝒳,(v)\varphi_{\mathcal{X},\mathcal{L}}(v) is the non-Archimedean potential associated with (𝒳,)(\mathcal{X},\mathcal{L}). By definition, φ(𝒳,)(v)=σ(v)(E)\varphi_{(\mathcal{X},\mathcal{L})(v)}=-\sigma(v)(E). However, the pushforward of 𝒪(E)\mathcal{O}(-E) is just 𝔞¯,\overline{\mathfrak{a}}, the integral closure of 𝔞\mathfrak{a}. Hence, maxj(v(𝔞j)j)=σ(v)(𝔞)=σ(v)(𝔞¯)=ψˇ(𝒳,),an(v)\max_{j}(-v(\mathfrak{a}_{j})-j)=-\sigma(v)(\mathfrak{a})=-\sigma(v)(\overline{\mathfrak{a}})=\check{\psi}^{(\mathcal{X},\mathcal{L}),\mathrm{an}}(v). Putting everything together, we obtain equality in (53), finishing the proof. ∎

Now we come back to the general situation where LL is only assumed to be big. In this case, we note that when all L𝔞jL\otimes\mathfrak{a}_{j} are globally finitely generated, the first step in the proof of the above result implies that

(r^𝔞)an(v)maxj(v(𝔞j)j)=σ(v)(𝔞).(\hat{r}^{\mathfrak{a}})^{\mathrm{an}}(v)\geq\max_{j}(-v(\mathfrak{a}_{j})-j)=\sigma(v)(\mathfrak{a}). (54)

Next we show that the procedure of (33) associates a sequence of flag configurations to any geodesic ray:

Lemma 4.11.

Let ϕPSH(X,θ)\phi\in\mathrm{PSH}(X,\theta) be a model potential with positive mass. Let ΦPSH(X×Δ,p1θ)\Phi\in\textup{PSH}(X\times\Delta,p_{1}^{*}\theta) be the potential corresponding to the geodesic ray {ut}t\{u_{t}\}_{t} in PSH(X,θ)\mathrm{PSH}(X,\theta) emanating from ϕ\phi with supu10\sup u_{1}\leq 0. Given any m0m\geq 0, let

(mΦ)=𝔞0+𝔞1s++𝔞N1sN1+𝔞N(sN),\mathcal{I}(m\Phi)=\mathfrak{a}_{0}+\mathfrak{a}_{1}s+\dots+\mathfrak{a}_{N-1}s^{N-1}+\mathfrak{a}_{N}(s^{N}),

as in (33). Then we have 𝔞N=(mϕ)\mathfrak{a}_{N}=\mathcal{I}(m\phi).

Proof.

Recall (31) that (mΦ)|X×Δ=p1(mϕ)\mathcal{I}(m\Phi)|_{X\times\Delta^{*}}=p_{1}^{*}\mathcal{I}(m\phi). This allows to regard 𝔞0+𝔞1s++𝔞N1sN1+𝔞N(sN)\mathfrak{a}_{0}+\mathfrak{a}_{1}s+\dots+\mathfrak{a}_{N-1}s^{N-1}+\mathfrak{a}_{N}(s^{N}) as an algebraic coherent ideal sheaf on X×X\times\mathbb{C} by Lemma 3.8, so

(𝔞0++𝔞N1sN1+𝔞N(sN))|X×\displaystyle\left(\mathfrak{a}_{0}+\dots+\mathfrak{a}_{N-1}s^{N-1}+\mathfrak{a}_{N}(s^{N})\right)|_{X\times\mathbb{C}^{*}} =(𝔞0++𝔞N1sN1+𝔞N(sN))[s][s,s1]\displaystyle=\left(\mathfrak{a}_{0}+\dots+\mathfrak{a}_{N-1}s^{N-1}+\mathfrak{a}_{N}(s^{N})\right)\otimes_{\mathbb{C}[s]}\mathbb{C}[s,s^{-1}]
=p1𝔞N.\displaystyle=p_{1}^{*}\mathfrak{a}_{N}.

Comparing the fibers at s=1s=1 we conclude that 𝔞N=(mϕ)\mathfrak{a}_{N}=\mathcal{I}(m\phi). ∎

We argue that for the converse of (54), one does not even need global finite generation:

Lemma 4.12.

Let 𝔞=𝔞0+𝔞1s++𝔞N1sN1+𝔞N(sN)\mathfrak{a}=\mathfrak{a}_{0}+\mathfrak{a}_{1}s+\dots+\mathfrak{a}_{N-1}s^{N-1}+\mathfrak{a}_{N}(s^{N}) be a flag configuration of a big line bundle LL. Then

(r^𝔞)an(v)maxj(v(𝔞j)j)=σ(v)(𝔞).(\hat{r}^{\mathfrak{a}})^{\mathrm{an}}(v)\leq\max_{j}(-v(\mathfrak{a}_{j})-j)=\sigma(v)(\mathfrak{a}). (55)
Proof.

We first reduce the general case to the case when LL is very ample and L𝔞jL\otimes\mathfrak{a}_{j} is globally generated for all jj.

Suppose that AA is a very ample line bundle so that LAL\otimes A is ample and such that AL𝔞jA\otimes L\otimes\mathfrak{a}_{j} is globally generated for all jj. Such AA exists by [Har13, Theorem II.7.6] for example. Choose any C>0C\in\mathbb{N}_{>0}. Let 𝔟\mathfrak{b} be the following flag configuration of LAL\otimes A:

𝔟k={𝔞k,k=1,,N;𝔞N,k=N+1,,N+C1;𝒪X,k=N+C.\mathfrak{b}_{k}=\left\{\begin{aligned} \mathfrak{a}_{k},\quad&k=1,\dots,N;\\ \mathfrak{a}_{N},\quad&k=N+1,\dots,N+C-1;\\ \mathcal{O}_{X},\quad&k=N+C.\end{aligned}\right.

So 𝔟k𝔞k\mathfrak{b}_{k}\supseteq\mathfrak{a}_{k} for all kk\in\mathbb{N}. After choosing a smooth positive metric on AA, it is obvious that

(r^L,𝔞)an(r^LA,𝔟)an.(\hat{r}^{L,\mathfrak{a}})^{\mathrm{an}}\leq(\hat{r}^{L\otimes A,\mathfrak{b}})^{\mathrm{an}}.

If we managed to prove the result for 𝔟\mathfrak{b} and LAL\otimes A, then we would have

(r^L,𝔞)an(r^LA,𝔟)anmax{maxj=1,,N(v(𝔞j)j),NC}.(\hat{r}^{L,\mathfrak{a}})^{\mathrm{an}}\leq(\hat{r}^{L\otimes A,\mathfrak{b}})^{\mathrm{an}}\leq\max\left\{\max_{j=1,\dots,N}(-v(\mathfrak{a}_{j})-j),-N-C\right\}.

Letting CC\to\infty, we conclude (55). Therefore, the Lemma is reduced to proving (55) for ample bundles LL. But this was proved in Lemma 4.10. ∎

We will also need the following general approximation result for geodesic rays, reminiscent of [DX22, Theorem 3.19]:

Lemma 4.13.

Let {ϕt}tθ1\{\phi_{t}\}_{t}\in\mathcal{R}^{1}_{\theta} be a ray in a big class {θ}\{\theta\}. There exists an increasing sequence of subgeodesic rays {ψtj}tPSH(X,θ)\{\psi^{j}_{t}\}_{t}\subset\textup{PSH}(X,\theta), not necessarily emanating from VθV_{\theta}, such that θψtjεjω\theta_{\psi^{j}_{t}}\geq\varepsilon_{j}\omega for some εj0\varepsilon_{j}\searrow 0, and P[ψ^τj]P[ϕ^τ]=ϕ^τP[\hat{\psi}^{j}_{\tau}]\nearrow P[\hat{\phi}_{\tau}]=\hat{\phi}_{\tau} a.e. for any τ\tau\in\mathbb{R}.

Proof.

For any δ>0\delta>0, due to τ\tau-concavity of test curves, the potential ϕ^τϕ+δ\hat{\phi}_{\tau^{+}_{\phi}-\delta} has non-zero mass, so by [DX21, Proposition 3.6] there exists vδPSH(X,θ)v_{\delta}\in\textup{PSH}(X,\theta) such that vδϕ^τϕ+δv_{\delta}\leq\hat{\phi}_{\tau^{+}_{\phi}-\delta} and θvδεδω\theta_{v_{\delta}}\geq\varepsilon_{\delta}\omega for some εδ>0\varepsilon_{\delta}>0.

Let {ηtδ}t\{\eta^{\delta}_{t}\}_{t} be the subgeodesic associated with the test curve η^τδ:=(1δ)ϕ^τ+δvδ\hat{\eta}^{\delta}_{\tau}:=(1-\delta)\hat{\phi}_{\tau}+\delta v_{\delta} for ττϕ+δ\tau\leq\tau^{+}_{\phi}-\delta and η^τδ:=\hat{\eta}^{\delta}_{\tau}:=-\infty for τ>τϕ+δ\tau>\tau^{+}_{\phi}-\delta.

Since η^τδϕ^τ\hat{\eta}^{\delta}_{\tau}\leq\hat{\phi}_{\tau} for all τ\tau\in\mathbb{R}, we get that ηtδϕt\eta^{\delta}_{t}\leq\phi_{t} and θηtδδθvδδεδω\theta_{\eta^{\delta}_{t}}\geq\delta\theta_{v_{\delta}}\geq\delta\varepsilon_{\delta}\omega for all t0t\geq 0. Also, we have that

ϕ^τ=P[ϕ^τ]P[η^τδ](1δ)P[ϕ^τ]+δP[vδ](1δ)ϕ^τ+δP[vδ].\hat{\phi}_{\tau}=P[\hat{\phi}_{\tau}]\geq P[\hat{\eta}^{\delta}_{\tau}]\geq(1-\delta)P[\hat{\phi}_{\tau}]+\delta P[v_{\delta}]\geq(1-\delta)\hat{\phi}_{\tau}+\delta P[v_{\delta}].

Since supXP[vδ]=0\sup_{X}P[v_{\delta}]=0, letting δ0\delta\to 0 we obtain that P[η^τδ]ϕ^τP[\hat{\eta}^{\delta}_{\tau}]\to\hat{\phi}_{\tau} in the L1L^{1} topology.

However, δηtδ\delta\to\eta^{\delta}_{t} is not easily seen to be δ\delta-increasing. To address this, we introduce the sequence of subgeodesics ψtj:=maxk=1,,jηt1/k\psi^{j}_{t}:=\max_{k=1,\dots,j}\eta^{{1}/{k}}_{t}, that satisfies the requirements of the lemma and is additionally increasing. ∎

Finally, we arrive at the main result of this section:

Theorem 4.14.

Let {ϕt}t1(X,θ)\{\phi_{t}\}_{t}\in\mathcal{R}^{1}_{\mathcal{I}}(X,\theta) be an \mathcal{I}-maximal geodesic ray with potential ΦPSH(X×Δ,p1θ)\Phi\in\textup{PSH}(X\times\Delta,p_{1}^{*}\theta), normalized by supXϕ1=0\sup_{X}\phi_{1}=0. For m0m\geq 0, let

(2mΦ)=𝔞0m+𝔞1ms++𝔞Nm1msNm1+𝔞Nmm(sNm)𝒪X×Δ,\mathcal{I}(2^{m}\Phi)=\mathfrak{a}^{m}_{0}+\mathfrak{a}^{m}_{1}s+\dots+\mathfrak{a}_{N_{m}-1}^{m}s^{N_{m}-1}+\mathfrak{a}_{N_{m}}^{m}(s^{N_{m}})\subseteq\mathcal{O}_{X\times\Delta},

as in (33). Let {ϕtm}t:={2mrtL2m,(2mΦ)}tPSH(X,θ)\{\phi^{m}_{t}\}_{t}:=\big{\{}2^{-m}r^{{L^{2^{m}}},\mathcal{I}(2^{m}\Phi)}_{t}\big{\}}_{t}\subset\textup{PSH}(X,\theta) be a rescaled Phong–Sturm ray. We have that ϕtmϕt\phi^{m}_{t}\searrow\phi_{t} for all t0t\geq 0. In particular, d1c({ϕtm}t,{ϕt}t)0d_{1}^{c}(\{\phi^{m}_{t}\}_{t},\{\phi_{t}\}_{t})\to 0.

Proof.

Let {λ,m}λ\{\mathcal{F}^{\lambda,m}\}_{\lambda} be the filtrations of R(X,L2m)R(X,L^{2^{m}}) induced by the flag configurations {𝔞km}k\{\mathfrak{a}^{m}_{k}\}_{k} of L2mL^{2^{m}}, defined in (47).

By the subadditivity of multiplier ideals, we have that (2m+1Φ)(2mΦ)2\mathcal{I}(2^{m+1}\Phi)\subseteq\mathcal{I}(2^{m}\Phi)^{2}. In particular, 𝔞km+1j𝔞jm𝔞kjm\mathfrak{a}^{m+1}_{k}\subseteq\sum_{j}\mathfrak{a}^{m}_{j}\mathfrak{a}^{m}_{k-j}. As a result, 𝔞r,λm+1𝔞2r,λm\mathfrak{a}^{m+1}_{r,\lambda}\subset\mathfrak{a}^{m}_{2r,\lambda} (recall (45)).

We obtain that λ,m+1H0(X,L2m+1r)λ,mH0(X,L2m2r)\mathcal{F}^{\lambda,m+1}H^{0}(X,L^{2^{m+1}r})\subset\mathcal{F}^{\lambda,m}H^{0}(X,L^{2^{m}2r}). This implies that ϕ^τm+1ϕ^τm\hat{\phi}^{m+1}_{\tau}\leq\hat{\phi}^{m}_{\tau}, hence {ϕtm}t\{\phi^{m}_{t}\}_{t} is indeed mm-decreasing.

By Lemma 4.12,

ϕ^m,an(v)12mminj(v(𝔞jm)+i)=2mσ(v)((2mΦ))\hat{\phi}^{m,\mathrm{an}}(v)\leq-\frac{1}{2^{m}}\min_{j}(v(\mathfrak{a}_{j}^{m})+i)=-2^{-m}\sigma(v)(\mathcal{I}(2^{m}\Phi))